You are on page 1of 14

Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017) A3117

Tortuosity of Composite Porous Electrodes with Various


Conductive Additives in an Alkaline System
Mehdi M. Forouzan, Michael Wray, Logan Robertson, and Dean R. Wheeler∗,z
Department of Chemical Engineering, Brigham Young University, Provo, Utah 84602, USA

The role of carbon additives in improving the electronic conductivity of composite porous electrodes is well understood. However,
there has been little work studying the effect of various carbon additives on effective ionic transport in porous electrodes. This work
determines effective ionic conductivities and associated tortuosities of composite cathodes with various types of carbon additive
and porosities in an alkaline system. A two-compartment direct-current method was developed to make these measurements and
was validated with multiple electrolyte solutions. This experimental method was modeled using COMSOL Multiphysics in order
to understand the effect of design parameters on the polarization curve. Empirical correlations were developed to predict the effect
of porosity and various carbon additives on tortuosity. As expected, the results show that tortuosity decreases with porosity and
increases with carbon amount. Cathodes containing BNB90 and KS6 carbon additives have the highest and the lowest tortuosity,
respectively. Furthermore, for cathodes containing BNB90, tortuosity in the direction orthogonal to the direction of compression
(in-plane tortuosity) was found to be less than tortuosity in the direction parallel to compression (out-of-plane tortuosity).
© The Author(s) 2017. Published by ECS. This is an open access article distributed under the terms of the Creative Commons
Attribution Non-Commercial No Derivatives 4.0 License (CC BY-NC-ND, http://creativecommons.org/licenses/by-nc-nd/4.0/),
which permits non-commercial reuse, distribution, and reproduction in any medium, provided the original work is not changed in any
way and is properly cited. For permission for commercial reuse, please email: oa@electrochem.org. [DOI: 10.1149/2.0911713jes]
All rights reserved.

Manuscript submitted August 11, 2017; revised manuscript received September 18, 2017. Published October 14, 2017. This was
Paper 21 presented at the Chicago, Illinois, Meeting of the Society, May 24–28, 2015.

The performance of batteries is strongly affected by the composi- developed an empirical relationship to predict tortuosity as a func-
tion and microstructure of the electrodes because of these variables’ tion of porosity. Continuing that work, Zacharias et al.14 used the
connection to transport and reaction conditions. To identify this con- polarization-interrupt method to quantify tortuosity and ionic con-
nection, several works have focused on predicting and understanding ductivity for different amounts of carbon additive and porosity in
microstructure of porous media.1–11 To understand and optimize mi- lithium-ion batteries. They showed that tortuosity increases with the
crostructure of porous electrodes, effective ionic conductivity is an amount of carbon and binder. Finally, they provided a Bruggeman-
important factor to accompany electronic conductivity. Although there type correlation relating tortuosity to porosity with both exponent
has been much work regarding transport characteristics in lithium-ion and scaling factor being functions of composition. Stephenson and
batteries, few attempts have been made to describe transport behav- coworkers15 performed both experiments and modeling to connect
ior of primary alkaline batteries. This is despite the fact that alkaline electrode microstructure to ionic conductivity. They confirmed that
batteries represent about a fourth of the worldwide battery market although carbon additive is necessary to improve electronic conduc-
(namely both primary and secondary cells). It is estimated that ap- tivity, it restricts ionic conductivity and can therefore hurt battery
proximately 6 × 109 alkaline and dry batteries are consumed yearly.12 performance. Hence, there is an optimal point for the amount of con-
This work focuses on understanding tortuosity and effective ionic ductive additive in the cathode.
conductivity of alkaline electrolyte in porous electrodes. Recently our group investigated the effect of several types of car-
Alkaline batteries use Zn as the anode, KOH as electrolyte, and bon additive on electronic conductivity within cathodes of primary
a mixture of electrolytic manganese dioxide (EMD) and graphite as alkaline batteries.4 The additives used were identical to those used in
the cathode. In the cathode, EMD particles are not sufficiently elec- this work, namely Timcal BNB90, MX15, KS15, KS6, and graphene.
trically conductive, so the carbon is needed to improve electronic Amongst these conductive additives, BNB90 was shown to produce
conductivity.4,7 The carbon additionally acts as a lubricant and binder the highest electronic conductivity, correlated with its large particle
during manufacturing. However, this additive reduces the capacity of size (90 μm), low Scott density, and high BET surface area. It was
the battery because it occupies a portion of the volume that would oth- also shown that the electronic conductivity of a wet cathode with con-
erwise contain active material. In addition, as is shown in this work, centrated KOH is reduced by around 30% relative to the dry value.
carbon additives reduce the ionic transport in the cathode, because As expected, electronic conductivity decreased with porosity and rose
when compressed they form dense regions that tend to plug the ion- with the amount of carbon additive. Because knowledge of both ionic
containing pores of the electrode. There is a trade-off between ionic and electronic conductivity is needed to optimize the electrode, there-
and electronic conductivity in batteries generally, meaning there is an fore we undertook the present work, which is a parallel study of ionic
optimal point for the amount of carbon additive. transport properties.
Another important factor is porosity of porous electrodes. Previous The only other recent work we are aware of dealing with changing
works have shown that electronic conductivity decreases and ionic additives in alkaline battery cathodes is due to Dong et al.12 They in-
conductivity increases with porosity.13–19 Thus, finding the optimum vestigated the effect on discharge performance of fluorinated graphite
point with respect to electronic and ionic conductivity and capacity (FG) as a cathode additive. They found that FG improves battery per-
is quite challenging and requires detailed knowledge of the physical formance significantly and that the optimal weight percentage of FG
and chemical behavior of porous electrodes. To gain this knowledge, is 5 wt%.
much work has been done to explicate kinetic, thermodynamic and The main purpose of this study is to understand how tortuosity (τ)
transport phenomena in porous electrodes.14,20–26 of alkaline battery cathodes changes with porosity, composition, and
Our group has a history of studying ion transport for lithium-ion the direction of compression. To that end, a common definition of tor-
cathodes. Thorat et al.13 performed both experiments and modeling tuosity is needed. Tortuosity is a geometric parameter, which accounts
to determine tortuosity in porous electrodes of lithium-ion batter- for irregularity and nonuniformity of pores, and can be considered a
ies. They used AC impedance and polarization-interrupt methods and dimensionless resistance to ionic transport. In battery-related fields
the following definition of tortuosity is commonly used:13–15, 21, 27–36
∗ Electrochemical Society Member. kε
keff = , [1]
z
E-mail: dean_wheeler@byu.edu τ

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A3118 Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017)

or equivalently,

Deff = , [2]
τ
where k and keff are the intrinsic and effective ionic conductivities, and
D and Deff are intrinsic and effective electrolyte diffusivities, respec-
tively. Thus, one value of tortuosity is that it can predict both effective
conductivity and diffusivity independent of the exact electrolyte used.
Likewise, an experiment that measures either effective conductivity or
effective diffusivity can determine the tortuosity.13,14,18,21,30,37–39 The
porosity of the cathode is represented above by ε.
The structure of the porous network, including the distribution of
particles and pores, contribute to τ in complex ways. Therefore τ is a
function of such geometrical variables, particularly the porosity. An
empirical correlation between ε and τ is provided by a generalized
Bruggeman-type relationship:13,14,18,21,30,37–39
τ = γε1−α , [3]
where α is the Bruggeman exponent, and γ is a scaling factor. In
battery simulations it has been common to assume that α ≈ 1.5 and
that γ = 1.18,40,41 These parameters have been shown to describe the
tortuosity of a system composed of spherical insulating particles of Figure 1. Experimental apparatus (not to scale) with porous electrode (pellet)
uniform size and random distribution in a homogeneous conductive secured in cylindrical aperture between two electrolyte-filled chambers.
medium. However, for realistic systems with non-ideal particle sizes
and shapes, values of α and γ can differ from this convention. In
particular, values of α up to 4 have been used.39 compartment. The orthogonal-direction experiments differed in the
Some researchers have used Archie’s Equation instead to explain means of mounting the pellet and are discussed below.
conductivity in porous electrodes,42–46 although it has historically been Both compartments were filled to the same height with electrolyte
used to describe the conductivity of sedimentary porous rocks. Nev- solution of the same composition. One reference electrode (RE) was
ertheless, when applied to battery applications, Archie’s Equation placed in each compartment adjacent to the pellet. In addition, each
is mathematically equivalent to the above generalized Bruggeman compartment contained a working electrode (WE) secured to the in-
Equation. side wall of the container. A constant current was then passed through
One obstacle in directly obtaining effective ionic transport prop- the system using a potentiostat (Arbin MSTAT). The voltage drop
erties of porous electrodes is that such materials facilitate both ionic across the REs was measured over time (additional details below).
and electronic transport. In many cases, electronic conductivity is
greater than the ionic conductivity4 and in any case is convoluted with Composite electrode.—Multiple cathode compositions and car-
ionic conductivity when performing electrochemical impedance spec- bon additives with different porosity were investigated, as given by
troscopy (EIS) or similar methods.13,14,47 For this reason, we prefer Table I and previously discussed.4 The primary carbon additives were
other methods for determining electrode ionic conductivity. the TIMCAL graphites BNB90, MX15, KS15, KS6. In addition, a
The remainder of this paper discusses the measurement proce- mixed graphene-type material (Graphene Supermarket) was used.
dure and results for tortuosity and effective ionic conductivity of The composition of pellets were varied from 0 to 100 wt% carbon
compressed composite porous electrodes (i.e. pellets) representing al- additive, with the remaining solid material being EMD (TRONOX).
kaline battery cathodes of different compositions and porosities. In Because BNB90 is the best-performing additive electronically,4 it is
addition, experiments for tortuosity in the direction orthogonal to the more attractive for commercial use. Therefore, compositions contain-
compression direction are conducted for one carbon additive type. In ing BNB90 and EMD received more attention than others in this
each case empirical fitting correlations are developed that smooth the research.
data and could be used in cell-level performance models. Furthermore, In general, EMD particles by themselves cannot hold together well
a COMSOL ion-transport model is developed to understand observed as a pellet; carbon additives act as a binder in cathodes in addition
concentration polarization and validate the experimental tests. to increasing electronic conductivity. It was found qualitatively that
BNB90 functions better in this role than other carbons, with KS6
being the worst binder. For small carbon fractions, pellets could be
Experimental
quite fragile when removed from the compression die (see below), and
Electrochemical cell.—A two-compartment apparatus for mea- this limited the ability to perform experiments on some compositions
suring effective ionic conductivity of cathode pellets was developed at higher porosities. This difference in carbon mechanical behavior
for this study, as shown in Fig. 1. Also, galvanostatic control mode was also apparent, if less noticeable, when making pellets of 100%
was used in order to characterize transport in the porous electrodes. carbon. The beneficial role of graphites can be explained by their
To that end, ionic current is flowed from one reservoir, through the malleability and morphology. It appears that expanded flakes like
pellet, and into another reservoir. A 150-mL polypropylene container BNB90 and graphene can hold pellets together well, while spheroidal
(Nalgene) was divided into two compartments with 6.35-mm-thick additives like KS15 and KS6 make less strong pellets.4
sheet of polyethylene, through which a circular hole was milled to Each sample was compressed in a cylindrical die (20-mm diame-
connect the two compartments. Multiple hole diameters were used, ter) placed in a benchtop press (Carver Model 3851-0). The mass of
but most commonly were 20 mm. The partition was glued in place pellets varied from 1 to 4 g, depending on the composition and type
with silicone to form a tight seal against the container walls. The pellet of graphite used in the pellet. Different porosities were achieved gen-
was secured inside the aperture. To provide a tight fit that would in- erally by using a compression force between 16 and 107 kN for about
hibit any shunt currents and assist in maintaining structural integrity, 1 minute. However, for lower amounts of carbon (less than 3 wt%) a
pellets were generally placed inside an elastomeric sleeve (i.e. a thick compression force of 107 kN was applied for longer periods of time to
rubber band) with a slightly longer width than the pellet’s axial length, densify the porous electrodes and make them stronger. Because EMD
then inserted into the hole, leaving a flat pellet face exposed to each electrodes are not able to hold together well without carbon additive,

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017) A3119

Table I. Evaluated properties of the pellets with varying compositions and porosities.

Composition MacMullin number Individual fit


Type (Graphite wt%) Porosity (%) Tortuosity ( τε ) (γ, α) τ = γε1−α R 2 (%)
TIMCAL BNB90 100 10–24 11.3–19.5 46–186 4.47, 1.64 99
Parallel 75 15–25 10.5–13.7 42–90 4.97, 1.54 99
direction 50 24–29 9–10.4 31—44 3.57, 1.74 99
25 24–30 7.9–9.4 26–39 2.51,1.96 93
10 23–36 4.5–7.8 13–33 1.56, 2.09 96
4.5 27–32 3.2–6.0 10–22 0.1, 3.99 81
3 24–31 3.3–5.0 11–21 0.35, 2.92 98
2.5 25–30 2.9–4.7 10–19 0.17, 3.35 94
2 24–29 2.9–4.2 10–18 0.33, 2.79 98
1 24–26 2.5–3.4 10–14 0.004, 5.82 92
0 21–24 1.7–2.4 7–11 0.047, 3.54 99
TIMCAL BNB90 Orthogonal direction 100 8–18 11.4–16.4 62–213 5.67, 1.42 97
10 22–33 4.6–5.9 14–27 2.13, 1.69 94
4.5 25–29 3.3–4.6 12–18 0.22, 3.16 95
3 24–29 3.1–4. 0 11–16 0.64, 2.26 91
2 21–25 3.3–3.6 13–16 1.21, 1.71 69
1 23–24 3.2–3.4 13–15 0.56, 2.22 94
0 24–26 1.4–1.7 5–7 0.026, 3.92 1
TIMCAL MX15 100 9–21 7.9–14.3 38–159 2.46, 1.74 99
10 21–26 5.4–7.7 21–36 0.64, 2.56 74
4.5 15–20 5.5–7.0 28–47 1.77, 1.71 84
TIMCAL KS15 100 10–19 7.5–11.7 41–115 2.35, 1.71 98
10 22–25 4.6–6.7 19–31 0.04, 4.34 91
4.5 21–25 3.9–5.0 16–24 0.70, 2.24 98
TIMCAL KS6 100 11–24 6.7–9.4 28–85 3.44, 1.46 97
10 14–17 6.1–8.1 35–60 1.08, 2.0 96
4.5 21–25 3.6–4.3 14–21 1.0, 1.94 92
Graphene 4.5 26–29 3.9–5.5 14–21 0.03, 4.93 99

high-porosity samples with lower amounts of carbon could not be was oxidized for approximately 3 minutes. The resulting AgCl-coated
made. wire was then placed in a fresh 1 M KCl solution. Silicone or hot-melt
The porosities of the pellets were determined using the following adhesive was used to seal one end of the wire inside a glass capillary
equation:4 tube (1.5 mm inside diameter, 100 mm length) filled with 1 M KCl.
m The seal ensured that the electrolyte would remain inside the capillary
ε=1− , [4] tube and around the wire.
L Aρ
Because AgCl deteriorates in light, the electrodes were stored in
where the pellet mass, thickness, and cross-sectional area are given the dark in a 1 M KCl solution when not being used for measurement.
by m, L, and A, respectively. The density (ρ) is the composite density Nevertheless, it became necessary to remake the electrodes periodi-
of the solids, which is a mass-average density of the EMD and carbon cally, using the following procedure. The silver wire from old REs
crystalline densities: ρ−1 = wEMD ρ−1 −1
EMD + wC ρC , where each w is a was left in ammonium hydroxide for approximately 5 minutes to strip
component dry mass fraction. The porosity calculated this way con- any residual AgCl from the surface of the electrode. Electrodes were
siders all pores including intra-particle and inter-particle pores (for then rinsed with distilled water. Abrasive paper was used to remove
instance, EMD contains approximately 25% intra-particle nanoporos- any remaining oxidant on the surface of the wire. The wire was again
ity). In previous work by Nevers et al., inter-particle porosity was rinsed in distilled water. The stripped wire was then recoated with
used for instance in correlating electronic conductivity.4 Crystalline AgCl and placed in a capillary as described in the previous paragraph.
densities for EMD and the different types of graphite were assumed
to be 4.9 g cm−3 and 2.15–2.26 g cm−3 (depending on the graphite
type), respectively.48,49 Polarization procedure.—The pellet was mounted in the divider
between the two cell compartments. REs and WEs were placed in each
Working and reference electrodes.—Platinum mesh WEs were compartment and connected externally to the potentiostat. Immedi-
used to provide current flow through the pellet, though preliminary ately after adding electrolyte to both compartments, constant current
work with stainless steel mesh as WEs gave comparable results. (2 mA) was passed in one direction until the system approximately
Luggin-type capillary Ag/AgCl reference electrodes were used to reached quasi-steady state. The required amount of time depended
measure the voltage drop across the pellet, in which the open end of upon the thickness, porosity, and composition of the pellet, with a
the capillary was placed against each of the two exposed surfaces of typical time being 20 minutes. The temperature of the electrolyte was
the pellet. Consistent positioning of the REs as close as possible to then measured. The current was interrupted and the cell relaxed for
the pellet in all tests serves to minimize erroneous voltage losses not a modestly longer length of time (e.g. 25 minutes) until the potential
taking place inside the porous electrode (what we call pellet) itself, across the pellet had decayed essentially to zero. The polarization and
although voltage losses due to pure electrolyte is negligible compared relaxation steps were then repeated with the direction of current flow
to voltage drop in the porous electrode, meaning that the experiment (i.e. polarity) reversed during this second cycle.
has little sensitivity to the exact positioning of the REs. The effective ionic conductivity of the pellet is then calculated
The REs were formed from 0.5-mm diameter silver wires that were from:
plated with silver chloride. To make new electrodes, silver wire and
a small silver sheet were placed in 1 M KCl. The potential between LP I
keff = , [5]
the two was set to 2 volts with an external power supply, and the wire VR E A

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A3120 Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017)

Figure 3. Double-layer charging transient for a pellet with 4.5% BNB90 at


short times.

As described by Gomadam et al.,52 the double-layer charging


process is essentially exponential with a characteristic time given
by
 
1 1
τ DL = L 2P aC + , [6]
σ keff
where σ is solid-phase electronic conductivity, a is specific interfa-
cial area, and C is double-layer areal capacitance. For purposes of
estimation and example, we use the following reasonable values ob-
tained from prior literature: C = 2 F · m2 , σ = 150 S · m−1 . 4,53
Furthermore, based on the experimental results given below, we take
a = 70000 m−1 , keff = 0.5 S · m−1 , and L p = 2.5 mm, again as
Figure 2. Effect of varying the mass of a pellet (4.5 wt% BNB90) (a) on reasonable estimates. This results in τ DL = 0.9 s. faradaic reaction
tortuosity and (b) on effective ionic conductivity for multiple porosities. An rate is also assumed to be negligible, consistent with the use of KCl
electrolyte of 7 M KOH was assumed. electrolyte.
The result from the short-time model described in Ref. 52 is plotted
in Figure 3. The double-layer is essentially fully charged by 0.7 sec-
where I is the current, VR E is the initial potential drop (at t ≈ ond, current no longer follows the electronic pathway, and steady-state
τ DL , see below) between the REs, or across the cathode, L P is the VR E (in the absence of concentration variations) is approximately
thickness of pellet, and A is cross-sectional area of the pellet. In 32 mV. This illustrative example is consistent with the case for 4.5%
order to compute tortuosity from Eq. 1, the intrinsic conductivity BNB90 shown in Figure 6. Furthermore, in this specific case the VR E
of the electrolyte solution (k) is also needed. Intrinsic conductivities used to determine keff was measured at t = 1.06 s, that is, t ≈ τ DL
for the electrolytes used here, as a function of temperature, have been and the measurement occurs after the double-layer transient shown in
determined previously.50,51 Note that Eq. 5 is a modified form of Ohm’s Fig. 3.
law without any concentration polarization or electronic current flow. Because exact τ DL values could be somewhat uncertain, a further
To satisfy these assumptions, the voltage is collected around 1 s to 1.5 s validation was performed to ensure that the measured initial VR E
after the onset of current flow. was collected as intended at t ≈ τ DL . This was done by checking that
As explained below, most experiments were conducted with 1 M the system is not exhibiting exponential-decay-like behavior for the
KCl electrolyte, which is adequate for determining both keff and tor- few VR E values collected immediately following the initial value.
tuosity, and hence Bruggeman-type relations (Eq. 20). However, in
Fig. 2, keff is presented using the synthetic k value for 7 M KOH, to Anisotropic tortuosity.—Most of the work here was done with
make the keff values more representative of what would be observed ionic conductivity determined in the same axial direction as the cylin-
in a commercial alkaline cell. drical pellets were compressed (out-of-plane tortuosity). In cylindrical
commercial cells, the annular cathode segments are likewise com-
Double-layer charging analysis.—In general there are two path- pressed primarily in the axial direction, however this is orthogonal
ways for current to flow through our sample: electronic and ionic. to the radial direction of current flow. Therefore, some samples were
In order for our experiment to succeed as designed, most or all of tested in this orthogonal current direction to assess the magnitude of
the current must flow through the ionic pathway during the period of this effect.
voltage measurement and application of Eq. 5. However, at extremely In developing the procedure for measuring the conductivity in the
short times after current begins to flow, some will flow through the orthogonal direction, it was decided to keep the sample fabrication and
electronic pathway. This is well described by the analysis performed testing processes as close to the original tests as possible. At the same
by Gomadam et al.52 In the language of that work, our experiment con- time, simply rotating the cylindrical pellet inside the cell would lead
stitutes Configuration III, for which there is a well-developed math- to multiple problems. Therefore, longer pellets were made using the
ematical model that will function at short times, i.e. in the absence same die and compression process. These longer cylindrical pellets
of concentration polarization. That analysis shows that once the ca- were, as before, compressed in the axial direction. They were then en-
pacitive double-layer throughout the pellet has fully charged, all the cased in Lucite acrylic resin (LECO 32 mounting press, LECO Corp).
current will flow through the ionic pathway and Eq. 5 will be valid. This provided mechanical support to allow a relatively fragile pellet
On the other hand, if one waits for much longer periods of time, then to be sectioned using a diamond saw (4-inch diameter, water cooled,
concentration polarization effects become significant. Thus, there is LECO VC-50). Two parallel cuts were made in the axial direction
a small window of time in which to analyze the system according to (the direction of compression) on either side of the pellet central axis,
Eq. 5. Our experimental times for the collection of VR E , namely leaving behind a 1- to 4-mm-thick quasi-rectangular section taken
1–1.5 s, are within that window as demonstrated below. from the center of the pellet. This section still contained an acrylic

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017) A3121

shell around its circumference, allowing it to be mounted in a modi-


fied cell partition using hot-melt glue to secure it inside a rectangular
aperture. Because these orthogonal sections were of similar thickness
to the previous cylindrical samples (1–4 mm), the time it took to reach
equilibrium was comparable to the previous experiments. Represen-
tative surface area A was determined using a harmonic mean of both
exposed faces of the section. The thicknesses at 5 different locations
across the pellet were averaged to obtain L used in Eq. 5.
There were concerns that the encasing and slicing procedure might
moderately change the morphology of the pellet. A validation exper-
iment was run to alleviate these concerns. This experiment followed
the above procedure for mounting and slicing the pellet but did so
in a manner that the finished sample was simply a thinner circular
cross-section of the original pellet, now encased in the acrylic resin.
It was found that there was an acceptably small difference (less than
2%) between the effective axial conductivities of the samples made
by the two methods. Therefore, it was concluded that the effect of the
mounting and cutting procedures for orthogonal samples was accept-
ably small.

Choice of electrolyte.—Fig. 4 shows polarization curves of pellets


with zero, low, and high additive fraction in 1 M KCl electrolyte.
According to our observations, for pellets with low tortuosity or high
porosity (pure EMD pellets) the polarization curve included an initial
jump followed by almost no additional voltage increase (Fig. 4a).
However, by increasing additive fraction, a large transient polarization
appeared (Fig. 4c). In order to understand this difference between
Figs. 4a, 4b, and 4c, a model was undertaken as described in the
Model development section.
Although initial polarization (Ohmic polarization) was used for
tortuosity determination, all of the experiments were performed for
15–75 minutes, depending on composition of the pellet, to generate
a more complete polarization curve. One purpose of running the ex-
periments for additional time was we wanted to understand what was
causing the large transient polarization at higher additive fractions.
Another purpose was to confirm that tortuosities were accurate, based
on the hypothesis that the transient part was due to concentration po-
larization and would be subject to the same tortuosity. In other words,
the tortuosity used for effective ionic conductivity (initial voltage
increase) should be consistent with the tortuosity used for effective
diffusivity which represents the transient part.
Multiple electrolytes were used to measure the transport proper-
ties of cathodes in initial experiments. Naturally, the first electrolyte
chosen for the experiments was KOH, the electrolyte used in real al-
kaline batteries. However, it was discovered that it took the system
approximately 3 hours to reach quasi-steady state using KOH versus Figure 4. Representative experimental results using an EMD/graphite pellet
5–30 minutes with other electrolytes such as KCl and NaCl. NaOH in 1 M KCl.
was also investigated as an electrolyte. It behaved similarly to KOH
and offered no advantages. The observed time variations cannot be
explained solely due to transport property differences and may well tube where the pellet had been, while blocking all other current paths.
be due to adsorption or wettability of the different ions with the pellet Passing current through the capillary tube gave the expected tortuosity
materials. value of 1.0 and the system reached quasi-steady state in less than
For our subsequent experiments, KCl was the electrolyte of choice. 0.05 seconds. Then a metal wire was placed inside the capillary tube
While identical pellets in KCl and KOH gave consistent tortuosity to imitate the electronic conductivity that exists in the pellet. The
values, pellet tortuosities evaluated in NaCl were approximately 15% result was the same: The system reached equilibrium almost instantly
less than those obtained while using KCl and KOH. This may be due and the tortuosity of the tube was calculated to be 1.0. These tests
to the presence of nano-pores in EMD in which ion-specific surface were done with multiple electrolytes. Next, a pellet was formed from
interactions have an effect on ion transport.54–56 copper powder compressed in the same manner as the other pellets.
In addition, the pellets appeared to show less mechanical degra- The result was slightly different. The voltage drop across the pellet
dation over time when using KCl as electrolyte. With the exception stepped up initially followed by a much slower transient increase.
of graphite additive, the cathodes used did not contain any binder or The pressed copper thus seemed to behave like a hybrid between
other additives that would help them remain intact in a freestanding the alkaline cathode pellets used in other experiments and the glass
geometry. Though all cathodes showed some crumbling or loss of capillary tube.
material from the surface to a certain extent over extended periods Both graphite and EMD are known for having a fairly high surface
of time, the amount of pellet deterioration in KCl solution over the area.4 BET surface area tests were run on pure KS6 graphite, EMD,
experimental time scale was considered negligible. and EMD/graphite mixtures. The EMD and EMD/graphite mixtures
had a surface area of 31 m2 g−1 , and the pure graphite had a surface area
Other tests.—Several experiments were run to further validate our of 18 m2 .g−1 . Comparable values For BET surface area are reported
experimental design. One test was run by placing a small capillary in work by Cericola and Spahr.57 For such large surface areas the

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A3122 Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017)

Figure 5. Schematic of the proposed ion transfer and reaction mechanism (top), reaction equations (middle), and 1D model geometry (bottom). OH− is generated
on the negative working electrode (WE) and Cl− is mainly consumed on the positive WE. V R E is measured by 2 reference electrodes right outside the pellet.
Standard potentials are given with respect to the Standard Hydrogen Electrode (SHE). The 1D geometry shows relevant domain lengths, transport parameters, and
dependent variables. The WEs are considered to be on the extreme left and right sides of the model geometry. Experimentally measured pH on both sides are also
shown in the picture.

wettability of the carbon and EMD surfaces to particular electrolytes Assumptions.—As mentioned before, the main electrolyte used in
are likely to differ, and could contribute to observed electrolyte effects. this work was KCl. While this was useful experimentally, KCl added
The time constant for capacitive effects, as another possible cause some complexity to the model, in order to account for the reaction and
of the transient polarization, is less than a second for our electrodes, transport mechanisms. There are several possible mechanisms of ion
based on reasonable values for surface area, area-specific capacitance, reaction in the experimental system. According to the observations and
and conductivities.58 On the other hand, the time constant for diffu- modeling results, it was decided to use concentrated ternary electrolyte
sion is from a few minutes to a few hours. Comparing these two theory.59–62 Fig. 5 shows the proposed mechanism. Based on the low
time constants, it appears that double-layer-charging time is negligi- potential needed for K+ reduction, the most likely reaction on the neg-
ble and the time-dependent part of the polarization curve is coming ative WE is water reduction that generates hydroxide ion (OH− ) and
from concentration polarization, i.e. diffusion potential. In the next hydrogen gas. On the positive WE, OH− oxidation would normally
section this conclusion is confirmed by a COMSOL Multiphysics occur at a lower (thermodynamically preferred) voltage compared
model. to Cl− oxidation reaction. However, the low concentration of OH−
(∼ 10−7 M) compared to the concentration of Cl− (∼1 M) results in
reaching the limiting flux for OH− transport quickly, while Cl− serves
Model Development
as the principal ion for current flow. Thus, the large polarization poten-
As described in the previous section, although multiple experi- tial observed for higher-graphite pellet formulations is concentration
ments helped to validate the method of determining tortuosity from polarization in which the system is approaching the limiting current.
the initial (ohmic) polarization, the subsequent transient polarization For model simplicity, only OH− generation on the positive WE and
with increasing amounts of carbon was not understood merely by the Cl− consumption on the negative WE were considered. Notably, this
experiments. In order to better understand what happens and con- mechanism is consistent with experimental observations including a
firm our interpretation of the experiments, a cell model was needed. faint smell of chlorine from the positive WE and a relatively high
Therefore, a transport model using COSMOL Multiphysics 5.3 was measured pH on the negative WE (see Fig. 5).
developed and is reported in this section. Another assumption in this model is axial 1D transport, which
Since this work is not focused on modeling, the primary pur- is more accurate inside the cylindrical pellet than in the reservoirs
pose of the model is to understand the physical basis of the observed on either side of the pellet. In order to imitate the real 3D transport
polarization curves including ohmic and concentration polarization outside the pellet, ε R , τ R , and M R are defined in the model. εR and τR
and validate the method used for tortuosity measurement (using only respectively account for extra volume and extra ion pathways from WE
ohmic polarization for effective ionic conductivity and tortuosity). A to the pellet in the 3D reservoirs. M R is a mixing factor that increases
perfectly quantitative match between experiment and model is not effective diffusion coefficient of each reservoir, turning diffusivity
required to fulfill this purpose. into a dispersion coefficient accounting for natural convection. Hence,

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017) A3123

using Eq. 2 effective transport parameters in the reservoir are defined where i is ionic current density, ϕ is potential in ionic phase, keff is
as: effective ionic conductivity in the pellet and reservoir, T is temper-
kε R ature, Rg is gas constant, F is Faraday constant, tCl− is transference
R =
k eff , [7] number of Cl− ion, tOH− is transference number of OH− ion, c0 is
τR
water concentration (55 M), f KCl and f KOH are molarity-based activ-
and ity coefficients defined in concentrated electrolyte theory. The fact
Dε R that concentration polarization terms for both salts (KCl, KOH) oc-
R =
D eff MR . [8] cur in Eq. 11 means that depletion of either species can lead to large
τR
overpotentials.
In our experience, the voltage across the pellet is relatively insen- Boundary conditions for potential and current are on the negative
sitive to any manual (i.e. additional) stirring of the electrolyte in the WE
reservoirs. We chose a single large coarse-grain value for M R to reflect
this and found that the model matched the experiment for a range of ϕ = 0, [13]
conditions. Other similar values for M R effected little change in the and on the positive WE
model, and so no further effort was expended to fit the M R value.
The following additional assumptions were used in the model i = −I, [14]
1) The liquid junction potential for the reference electrodes is small where I is applied current density in the experiments (based on pellet
since K+ and Cl− have similar transport properties. Furthermore, cross-sectional area).
because of similar electrolyte concentration on both sides, any Boundary conditions for concentration PDEs at WEs are given as:
junction potentials of the two reference electrodes will tend to ∂cKCl (RCl − tCl− )
cancel when taking the potential difference between the two. = I, [15]
∂x eff
F DKCl
2) Electrochemical reaction in the pellet is not considered. The driv-
ing force for reaction is small because of the experimental design. ∂cKOH (1 − RCl − tOH− )
3) Double layer charging and electronic conductivity are not affect- = I, [16]
ing polarization voltage in the time scale of experimental tests. ∂x eff
F DKOH
This assumption is validated as described in Double-layer charg- − +
where RCl has values of RCl for negative WE and RCl for positive WE
ing analysis section. boundary conditions (see Table II). Note that no chemical-kinetic-
4) Concentration of KCl does not vary substantially from initial −
type expression is used in the boundary conditions. Instead, RCl and
value (1 M) and concentration of KOH is close to zero, so that +
RCl are defined as the fraction of fixed current coming from the
concentration dependence of all transport properties can be ne- chloride reaction on the negative WE and positive WE, respectively
glected with the exception of anion transference numbers. (see Table II).
Based on the above assumptions, partial differential equations In the interior boundaries (at the pellet boundaries), continuity
(PDEs) and boundary conditions (BCs) resulted from material and of superficial fluxes and of concentration and potential is considered
charge balances are described below. The equations and associated automatically by COMSOL.
BCs are solved simultaneously in the pellet and reservoir. Initial conditions are given as:
Model equations.—In this model OH− and Cl− ions are generated ϕ = 0, [17]
or consumed on WE boundaries and K+ is the counter-ion moving
around to satisfy electroneutrality of the electrolyte. Therefore, two
equations for concentration of OH− and Cl− and one equation for cKCL = 1 M, [18]
electrolyte potential are needed.
The material balance for the electrolyte-filled pores and electrolyte
in the reservoir are given by: cKOH = 10−6 . . . 10−5 M. [19]
  Note that a range of values for cKOH was used, depending on the
∂cKCl ∂ eff ∂cKCl
ε = DKCl , [9] simulation; initial cKOH values were kept as small as numerical stability
∂t ∂x ∂x
allows (10−7 M made the model numerically unstable) and to produce
  reasonable results for a system close to the limiting current.
∂cKOH ∂ ∂cKOH
ε = eff
DKOH , [10]
∂t ∂x ∂x Model parameters.—Most of the parameters used in this model
eff eff come from literature51,55,61,62,64 or are measured experimentally.
where DKCl and DKOH are effective diffusion coefficients inside the
Table II summarizes the parameters used in this model.
pellet or and cKOH are concentration of KCl and KOH,
reservoir, cKCl
Note that the activity correction factor of KOH is taken to be 1.
respectively, and ε is porosity of the pellet (ε P E < 1) or reservoir
From the Debye-Huckel electrostatic theory61,65 the activity coeffi-
volume correction factor (ε R > 1). Effective properties are determined
cient of KOH is mostly determined by ionic strength, in this case
from Eqs. 1 and 2 in the pellet and Eqs. 6 and 7 in the reservoir. Note
determined by the much higher concentration of KCl, close to 1 M.
that there is no source term in these PDEs, due to a combination of
In the absence of a ternary activity coefficient model for this system,
given assumptions.
we examined the behavior of binary KOH electrolyte from 0 to 1 M
Charge balance in the electrolyte phase is based on modified Ohm’s
concentration. The activity correction factor in that case is always
law and is given by concentrated ternary electrolyte theory (with the
close to unity, therefore we assume it is also unity for the ternary
water reduction reaction as the reference reaction):59,62,63
   system.61
∂ϕ 2Rg T keff cKCl ∂ln f KCl ∂
i = −keff − −tCl− + 1+ lncKCl
∂x F c0 ∂lncKCl ∂ x
    Results and Discussion
cKOH ∂ln f KOH ∂
+ 1 − tOH− + 1+ lncKOH , [11] Model and experiment polarization curves.—Experimental and
c0 ∂lncKOH ∂ x
modeling results suggested larger concentration polarization and
and longer quasi-steady state time for thicker and higher tortuosity elec-
∂i trodes (higher amounts of carbon additives or less porous electrodes).
= 0, [12] Further, the ratio of concentration polarization to ohmic polarization
∂x
Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A3124 Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017)

Table II. Model parameters.

Symbol Parameter Value Reference


LP Thickness of pellet 2.5–3 mm ∗

LR Reservoir length 60−L P


mm ∗
2
T Temperature 298 K ∗

F Faraday constant 96485C mol−1


A Pellet cross-sectional area 3.1416 cm2 ∗

I Applied Current 2 mA ∗

1 + ∂ln f KCl
∂lncKCl Activity coefficient correction factor of KCl 0.8575 51,61,62
∂ln f KOH
1 + ∂lncKOH Activity coefficient correction factor of KOH 1 51,61,62
DKCl Diffusion coefficient of KCl 1.89 × 10−5 cm2 s−1 51
DKOH Diffusion coefficient of KOH 2.85 × 10−5 cm2 s−1 51,61,62
k Ionic conductivity of KCl 11.1 S m−1 ∗

tK+ Transference number of K+ 0.4886 51,64


tOH− Transference number of OH− 1.766 × 10−6 (cKOH /M) ∗

tCl − Transference number of Cl− 1 − tK+ − tOH− ∗

τP Tortuosity of pellet 3.46-7.33 ∗

τR Tortuosity of reservoir 1 ∗∗

εP Porosity of pellet 0.23-0.27 ∗

εR Reservoir volume correction factor 4.5 ∗∗

MR Reservoir mixing factor 500 ∗∗


− ∗∗
RCl Fraction of current coming from chloride on the negative WE 0
+ ∗∗
RCl Fraction of current coming from chloride on the positive WE 1

∗ Experimentally measured or calculated.


∗∗ Supposed in this model for simplicity or physical justification.

increases with increased tortuosity and thickness as seen in Fig. 6. Ac- calculation (Eq. 5). One purpose for developing the model is to ensure
cordingly, the quasi-steady state voltage can be considered as a further that this VR E value does not erroneously include a significant de-
indicator of ionic resistance (tortuosity). However, quasi-steady state gree of concentration polarization, so that the corresponding tortuosity
voltage was not considered an accurate means to solely determine values are accurate.
ionic resistance due to the need for a more complicated transport As shown in Fig. 6, the model polarization curves are shown to
model. be in satisfactory agreement with the experiment, particularly the
Fig. 6 shows characteristic polarization curves. The initial voltage magnitude of the transient region. The concentration polarization is
after application of current corresponds to the ohmic polarization, strongly affected by tortuosity, just as is ohmic polarization. Therefore,
but because of concentration variation in the pellet and reservoirs the the substantial agreement between model and experiment implies that
voltage increases with time, observed in both experimental and model the tortuosity values used in the model are realistic. However, there
results. is a brief delay in modeled concentration polarization at small times
The design input parameters in the model include porosity, tor- (see Fig. 6 inset). The causes of this are the approximations in the
tuosity, and pellet thickness. The experimental procedure is that the model boundary conditions and 1D treatment of the reservoir, but
measured voltage difference at t ≈ τ DL is used for the tortuosity these do not affect the tortuosities determined by comparing model to
experiment for ohmic polarization values.
From the OH− concentrations predicted in the model, pH of the
negative WE reservoir is approximately 9.5 and pH of the positive
WE reservoir is approximately 7.5, which are in satisfactory agree-
ment with experimental values (see Fig. 5 for experimental pHs). In
addition, the model assumption that concentration of KCl electrolyte
remains close to 1 M throughout the test was confirmed by the model
results.
Another possible approach to compute tortuosity would be ana-
lyzing the relaxation part of the polarization curve after the current
is set to zero (i.e. interrupted), as studied by our group in two prior
publications.13,14 However, in the present ternary electrolyte system
with two ions reacting (compared to Li-ion batteries where only Li+
reacts), and with the complication of the reservoirs being part of
the relaxation process, obtaining an accurate tortuosity on this basis
alone was considered less reliable. Nevertheless, as discussed above,
our use of a model to compare to the concentration polarization part
of the experimental curve is an effort to use effective diffusivity of
the electrolyte to validate the tortuosity determined from keff at short
time. Again, getting tortuosity from keff is considered more reliable
than from Deff for this system. Furthermore, because the information
contained in the concentration polarization process and a possible
Figure 6. Experimental (solid lines) and model (dotted lines) polarization subsequent relaxation process are essentially the same, there is no
curves of pellets with 10%, 4.5%, and 2% BNB90 additive. This inset shows need to additionally compare the model to such a relaxation event at
additional detail at short times. the conclusion of the experiment.

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017) A3125

Figure 7. Voltage vs. time curves for wet pellets showing the correspondence
of the 20-hour test to the test run after a 3-hour delay.

Ionic conductivity of partially deteriorated pellet.—As previously


discussed, there was some concern about pellets slowly disintegrating Figure 8. a) Tortuosity of composite electrodes made of differing dry mass
over time when contacted by electrolyte. We wished to know if this fraction of BNB90, as a function of porosity. Symbols represent experimental
behavior changed with passing of current. In order to understand data while lines of matching color represent fits to a modified Bruggeman-type
the behavior of a pellet submerged in electrolyte for longer times, model.
a pellet with 10 wt% BNB90, 90 wt% EMD, and porosity of 22%
was tested in the system for 20 hours. During this time, voltage drop
across the pellet was recorded and it was observed that, after reaching structural changes upon contact with electrolyte, meaning that the dry
apparent quasi-steady state (in Fig. 7 this appears as a voltage peak on electrode must be overdesigned with respect to electronic conductivity
this timescale), voltage began slowly decreasing, indicating decreased compared to ionic conductivity.
ionic resistances. To see the possible causes of this phenomenon, a
pellet having the same composition and porosity was tested with a Experimental variations .—Varying the pellet’s mass (i.e. thick-
3-hour delay, meaning that the pellet had been put in electrolyte for ness) does not seem to systematically affect the calculated tortuosity or
3 hours before the current was flowed. As shown in Fig. 7, the two effective conductivity (Fig. 2). This suggests that contact or interfacial
pellets showed the same electrochemical response to current flow, resistances, not included in our calculations, are negligible.
after accounting for the delay on the second pellet. This suggests that As seen in Fig. 8, increasing the mass fraction of BNB90 in the pel-
pellet internal structure is affected only by electrolyte contact and not let increased tortuosity and caused its effective ionic conductivity to
by the passing of current. decrease. Pellets with higher proportions of BNB90 held together bet-
Additionally, the electrolyte of both compartments was manually ter under experimental conditions, likely because compression forced
mixed at times during the 20-hour experiment and no change was BNB90 into the pores between EMD particles, clogging pores while
seen in the voltage signal. This shows that the long-time change in also effectively functioning as a binder.
the voltage is caused by slow changes to the pellet itself, rather than Varying the amount and type of carbon additive also signifi-
diffusion limitations in the reservoirs. As a reminder, in this work all cantly affected structural properties. Pellets fabricated with BNB90
of the effective conductivities were measured within a few minutes and graphene additives held together quite well under experimen-
of the pellet being placed in contact with electrolyte so that tortuosity tal conditions while those fabricated with KS6 did not hold to-
values reflect the tortuosity of the intact pellet (before any structural gether well unless compressed above 3100 bar. Pellets made with
changes occurred due to partial deterioration of pellet). pure EMD were left under 3400 bar of pressure for a day or two
The increase in effective ionic conductivity of a soaked cathode (occasional adjustment of the press was necessary to maintain pres-
is most likely due to structural rearrangement. In order to understand sure) in order to ensure structural integrity. Fig. 8 demonstrates that
this effect, the volume of a wetted sample was measured before and graphitic additives obstruct ion flow much more than does EMD active
after wetting and then after being left in open air for a day to com- material.
pletely dry. For a typical pellet sample (10% BNB90 and weighing 3 g) We propose that the volume fraction of the carbon additive in
soaked with electrolyte for 4 hours, thickness increased approximately the pellet plays a significant role in pellet stability (the binder ef-
2% (initial thickness was 3.03 mm) and calculated porosity increased fect) as well as tortuosity (the pore-plugging effect). As seen in
from 21.4% to 22.8%. After letting the same sample dry completely, Table III, BNB90 and graphene exhibit comparably low densities.
the thickness and porosity did not further change by consequential Pellets formed with slightly denser KS6 additive were found to be
amounts. This illustrates that pellets swell moderately when left in less tortuous than pellets fabricated with graphene and BNB90 and
the electrolyte—permanently rearranging their microstructure. The did not hold together as well when placed in electrolyte solution.
electrolyte probably creates new pathways for ions, increases the size Therefore, it is reasonable to generalize that additives with lower den-
of pores, and/or changes the contacts among adjacent carbon-carbon, sities and therefore larger volume fractions result in pellets with higher
carbon-EMD, and EMD-EMD particles. It should be noted that these tortuosities. Similarly, an increase in carbon mass fraction increases
structural changes depend on how long the pellet is in the electrolyte; the tortuosity and stability of pellets.
a longer time results in greater changes in the pellet. However, there Volume fraction of additive, though important, is not the only
may be less structural change in commercial electrodes due to confine- factor determining ion transport. In order to understand the shape,
ment in the metal can. Nevers et al. considered electronic conductivity structure, and particle size of the additives, scanning electron micro-
of wet electrodes and came to similar conclusions about the structure scope (SEM) images were taken of graphite powders tested in this
of soaked pellets.4 In commercial primary alkaline batteries, the com- work (Fig. 9). Most notably, in spite of the fact that TIMCAL reports
position appears to be chosen so that electronic conductivity, rather spheroid particle shapes for KS15 and KS6 (see Table II), the observed
than ionic conductivity, is preferred. This is most likely due to the fact shapes resemble small flakes rather than spheroids. BNB90 looks to
that electronic conductivity decreases and is more sensitive to small have large flat surfaces (around 90 μm) possibly resulting in good

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A3126 Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017)

Table III. Morphological and structural properties of various graphitic additives and EMD.4,48,66,67

Carbon Additive d90 ∗ (μm) Shape As-Poured (Scott) Density (g.cm−3 )


BNB90 85 Expanded Anisometric 2D flakes 0.03
MX15 17 Anisometric flakes 0.065
KS6 6.5 Isometric, irregular spheroids 0.07
KS15 17 Isometric, irregular spheroids 0.1
Graphene 4.5 12 nm thick, 4.5 μm wide 0.04

∗ 10% of the particles have diameters more than reported value.

electronic connection between particles, improved electronic conduc- square (RMS) residual values, with residuals defined in a normalized
tivity, and very large barriers for ion transport. On the other hand, the fashion:
small flakes of KS6 (around 7 μm) seem to make weaker electronic
connection, fewer blockages, and less tortuous paths. Because of the |τexp − τmodel |
similar structures of MX15, KS15, and KS6, it is expected that they residual = . [21]
τexp
show similar behavior in transport and structural properties.
As seen in Table IV, RMS residual is less than 11%, which can
Correlations between tortuosity, porosity, and composition.— be considered the average relative error of the model. To show how
Accurate battery models require accurate transport data especially well Eq. 20 fits experimental data, corresponding curves are included
at high charge/discharge rates, as tortuosity plays a more significant in Figs. 8, 10–12.
role in this case. To aid in future battery models and optimize bat- Tortuosity of pellets containing BNB90, MX15, KS15, and KS6
tery performance, it is necessary to know how ionic and electronic from both experiment and correlation are included in Fig. 10, showing
conductivity change with porosity and carbon fraction. To that end, the differences between the tortuosities of various additives for three
we provide a series of empirical fits to Bruggeman-type equations. amounts of graphitic mass fraction. KS6 exhibits the lowest tortuosity,
Firstly, one may use Table I and Eq. 3 to calculate tortuosity from followed by KS15, MX15, graphene, and BNB90. Furthermore, all
individual least-squares fits of tortuosity at various compositions. In types of graphitic material show increasing tortuosity with decreasing
this case the R2 values, given in Table I, are quite high. porosity and increasing additive mass fraction. With our model it was
We also endeavored to perform fits with a more generalized em- difficult to fit the baseline case of BNB90 additive for the full (0–
pirical model that accounts for composition variations. What follows 100%) composition range, so the fit was divided into two ranges as
is a modified Bruggeman-type equation in which γ and α are each shown in Table IV. The curves shown in Figs. 10b and 10c were for
polynomial functions of composition: the smaller (0–10%) range, whereas the curve for BNB90 in Fig. 10a
  uses the upper range.
τ = γ0 + γ1 wC + γ2 wC 2 + γ3 wC3 ε(α0 +α1 wC ) , [20]
Fig. 11 illustrates how tortuosity orthogonal to the compression
where ε is porosity and wC is the mass fraction of the graphite. This changes with porosity and the amount of additive (BNB90). This
relation has six adjustable parameters, which were determined by a shows the same trends as the baseline case of tortuosity parallel to the
combined least squares fit. Table IV gives the parameters, which were compression.
fit for four different types of graphite along with an orthogonal config- Fig. 12 compares MacMullin numbers of the pellets made with
uration (BNB90). The lowest R2 is 0.96, which shows an acceptable fit BNB90 and EMD in orthogonal and tangential compressions. Mac-
between the data and the model. Another way to assess the accuracy Mullin number is defined as the ratio of tortuosity to porosity.18,68–73
of the model is through the maximum relative residual and root mean The results illustrate that the orthogonal MacMullin numbers are lower
than tangential MacMullin numbers. This difference is more signifi-
cant at lower porosities and higher amounts of graphite. For example,
a pellet (3% BNB90) with porosity of 30% has an orthogonal ionic
conductivity 7.6% greater than that found in the parallel direction.
However, a pellet with the same composition but a porosity of 20%
has an orthogonal ionic conductivity 31.3% greater than that found
in the parallel direction. This effect is even more pronounced for
pellets with greater carbon mass fraction. Pellets with 10% BNB90
show an ionic conductivity increase of 30.8% for 30% porosity and
47.8% for 20%. This tortuosity heterogeneity effect has been also ob-
served previously for calendared thin-film electrodes used in Li-ion
batteries.31
The difference between orthogonal and tangential tortuosity seems
to be due to the change in the pathways caused by compression: the
pathways in the compression direction are made more tortuous while
those in the orthogonal direction are less affected. In the case of Li-
ion electrodes and in particular Li-ion anodes made with graphite, a
similar anisotropy has been attributed to particle alignment, i.e. non-
spherical particles adopting a preferred orientation 14,57 . In our work
the investigation of tortuosity in the two directions is not specifically
intended to quantify the influence of particle alignment, though there
is a significant degree of anisotropy evidenced in Figs. 12 and 13. Even
though they are fabricated by different processes than those made here,
Li-ion electrodes still follow the same trend in which calendering to
create higher density or lower porosity increases tangential tortuosity
Figure 9. SEM images of powder of graphitic additives used in this work. (out-of-plane tortuosity in flat Li-ion electrodes).

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017) A3127

Table IV. Parameters to be used in Eq. 20 for tangential (to compression) tortuosity of porous electrodes with different types of graphitic additives,
and orthogonal (to compression) tortuosity for those with BNB90.

10 wt% and below


Parameters BNB90 Above 10 wt% BNB90 Orthogonal BNB90 MX15 KS15 KS6
α0 −1.97 −0.67 −1.71 −1.01 −2.17 −1.06
α1 5.95 0.11 1.39 0.28 1.48 0.59
γ0 0.11 2.76 0.17 0.46 0.08 0.43
γ1 9.49 5.99 7.81 15.77 2.98 13.00
γ2 0 −8.60 −37.95 −59.67 −11.44 −70.99
γ3 0 5.18 37.26 45.95 10.81 60.94
R 2 (τexp /τmodel ) 0.96 0.99 0.99 0.99 0.99 0.99
Max. residual 14% 5% 25% 12% 10% 11%
RMS residual 7% 2% 11% 5% 5% 2%
fitted range for wt% 0–10 11–100 0–100 0–100 0–100 0–100

Unlike Li-ion batteries, in which ions transport normal to current much as in Li-ion batteries. Thus, for decreased out-of-plane tortu-
collector (i.e. out-of-plane tortuosity is most critical), for alkaline bat- osity in Li-ion cells (which is ideal for high-rate performance) some
teries ion movement mainly takes place in the orthogonal direction modification is needed in the manufacturing process in order to pro-
to the compression. For that reason, compression during electrode duce electrodes that do not require high levels of calendering for ideal
fabrication does not reduce the ionic transport in alkaline batteries as porosity.
Although only BNB90 was tested for orthogonal tortuosity in this
work, it is instructive to examine cross-sections of other compressed
graphitic additives to qualitatively compare orthogonal and tangential
pathways. An FEI Helios Nanolab 600 instrument was used to mill
pellet samples by focused ion beam (FIB) and to take SEM images
of cross-sections. As seen in Fig. 13, for multiple graphite types pore
paths are more aligned to the orthogonal direction. It was not possible
to test all pellet compositions using this method, due to cost and pellet
fragility.

Pellet morphology .—The results in this work together with work


by Nevers et al.4 show that various carbon additives affect ionic and
electronic conductivity differently. This behavior corresponds to the
particle shape, size, aspect ratio, surface area, Scott density, and other
morphological features. As previously mentioned, carbon additives
with lower Scott density generate higher tortuosity. It can also be seen
that expanded graphite with big flakes, like BNB90 and Graphene,
increase the restriction of ion transfer, resulting in lower ionic con-
ductivity (higher tortuosity), while graphite with small flakes, like
KS6 and KS15, provide higher ionic conductivity (lower tortuosity).
In addition, it appears that additives with larger particle sizes tend

Figure 10. a) Tortuosity versus porosity of composite electrodes with (a) Figure 11. Tortuosity orthogonal to the compression for BNB90/EMD com-
100%, (b) 10%, and (c) 4.5% graphite (BNB90, MX15,KS15, KS6, and posite electrodes as a function of porosity. Symbols represent experimental
graphene). Symbols represent experimental data while lines of matching color data while lines of matching color represent fits to a modified Bruggeman-type
represent fits to a modified Bruggeman-type model. model.

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A3128 Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017)

Figure 12. MacMullin numbers of BNB90/EMD composite electrodes, com- Figure 14. SEM images of pellet surfaces with pure (a) BNB90, (b) MX15,
paring values for parallel and orthogonal directions of compression. Symbols (c) KS15, and (d) KS6.
represent experimental data while lines of matching color represent fits to a
modified Bruggeman-type model.

electronically conductive, has a large ionic conductivity, and results in


to agglomerate to a higher degree, reducing ionic conductivity, but such a fragile pellet. Large particle size and particular particle shapes
graphene is an exception in this regard. (images a and c) create these pores, providing convenient paths for ions
To clarify the effects of particle shape and size on ionic conductiv- to traverse and resulting in low electronic contact between particles.
ity and tortuosity, SEM images were taken of the surfaces of prepared This characteristic of EMD is also evident in images b and c which are
pellets with pure BNB90, MX15, KS15, KS6, EMD, 10% BNB90, pellets with 10% BNB90 and KS6, respectively. In these images, large
and 10% KS6. (Porosities were 10% for pure graphitic samples and pores introduced by the large EMD particles are creating ion paths and
25% for the mixtures.) These images supplement the images of the therefore increasing ionic conductivity (decreased tortuosity). Further,
four carbon additives taken as powders shown in Fig. 9. in image b, EMD particles are better connected and the graphite
Fig. 14 compares four types of graphite, showing a very dense media is denser and more compact, correlating with the lower ionic
and compact media for BNB90 (image a), resulting in lower ionic conductivity of pellets made with BNB90.
conductivity (higher tortuosity) and higher electronic conductivity due Among all carbon additives tested, BNB90 exhibits the highest
to larger flakes and smaller pores. For the other three types of graphite, electronic conductivity and the lowest ionic conductivity. This is prob-
one sees smaller flakes and more obvious pores, which result in higher ably because of its large flakes (85 μm), low Scott density, high surface
ionic conductivity and lower electronic conductivity. The similarity area, high aspect ratio, particle shape (expanded flakes), and well-
between the images of KS15 and KS6 is consistent with their similar connected, interspersed carbon pathways.4 Since expanded graphites
transport properties. like BNB90 and graphene are more expensive, it is worth studying
Fig. 15 shows the surfaces of additional pellets, with image a being the characteristics of cathodes with various kinds of additives in order
a pellet made with pure EMD. The macroscopic pores seen in this to apply the best composition, leading to better performance at lower
image can satisfactorily explain why pure EMD is not significantly cost.

Figure 13. Cross-sectional view of pellets with a) 10% BNB90, 24% porosity, b) Pure MX15, 13% porosity, c) Pure KS15, 16% porosity, d) Pure KS6, 16%
porosity. Note that the direction of compression is from the top to the bottom of the images.

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017) A3129

the negative working electrode and Cl− is consumed on the positive


working electrode. However, for model simplicity reaction terms on
the working electrodes were replaced by a term (RCl ) defined as the
fraction of current coming from the chloride reaction. Since there are
three relevant ions in the experimental system, concentrated ternary
electrolyte theory was used in the model.
According to our findings, it is best to use large-flake additives like
BNB90 to improve electronic conductivity, but the amount should not
be so high as to block the pores and impair pathways for ion transfer.
Other avenues of research could examine a combination of several
carbon additives, the amount of carbon additives, and the amount of
compression.
The polarization method used in this work could be adapted to
examine thin-film electrodes for other types of batteries, such as Li-
ion. In one manifestation, cells could be fabricated inside a glove box
with a symmetric series of layers: lithium foils as WEs, separator
films between WEs and REs, and separators between the REs and a
delaminated porous electrode film. The reference electrodes would
need to be either wires or porous grids containing lithium.

Acknowledgment

Figure 15. SEM surface images of compressed pellets with (a) Pure EMD,
A portion of this work was supported by the BMR program of the
(b) 10% BNB90, (c) 10% KS6 surface, as well as (d) EMD powder. U. S. Department of Energy.

ORCID
Conclusions
Accurate battery models require transport properties such as ef- Mehdi M. Forouzan https://orcid.org/0000-0002-3542-9371
fective ionic conductivity (or tortuosity) and electronic conductivity
through porous electrodes. Ionic and electronic conductivity change References
in opposing ways with changes in carbon additive fraction, carbon
type, and porosity. Thus, understanding both ionic and electronic con- 1. M. Smith, R. E. Garcı́a, and Q. C. Horn, J. Electrochem. Soc., 156(11), A896 (2009).
ductivity is necessary for battery optimization. Although electronic 2. P. R. Shearing, L. E. Howard, P. S. Jørgensen, N. P. Brandon, and S. J. Harris, Elec-
trochem. Commun., 12(3), 374 (2010).
conductivity has been explored relatively thoroughly in alkaline cath- 3. P. R. Shearing, R. S. Bradley, J. Gelb, S. N. Lee, A. Atkinson, P. J. Withers, and
odes, few attempts have been made to determine the corresponding N. P. Brandon, Meeting Abstracts, MA2011-01(10), 443 (2011).
ionic conductivity. 4. D. R. Nevers, S. W. Peterson, L. Robertson, C. Chubbuck, J. Flygare, K. Cole, and
In this work, a two-compartment DC method was developed for D. R. Wheeler, J. Electrochem. Soc., 161(10), A1691 (2014).
5. Z. Liu and P. P. Mukherjee, J. Electrochem. Soc., 161(8), E3248 (2014).
tortuosity measurement. Effective ionic conductivity and tortuosity 6. J. Joos, T. Carraro, A. Weber, and E. Ivers-Tiffée, J. Power Sources, 196(17), 7302
of several types of graphite, namely TIMCAL BNB90 (orthogonal (2011).
and parallel to the compression direction), MX15, KS15, KS6, and 7. M. M. Forouzan, C.-W. Chao, D. Bustamante, B. A. Mazzeo, and D. R. Wheeler, J.
graphene (only parallel to the compression direction) were measured, Power Sources, 312 172 (2016).
8. Q. Cai, C. S. Adjiman, and N. P. Brandon, Electrochim. Acta, 56(16), 5804 (2011).
each in several porosities and as varying fractions of graphitic ad- 9. M. I. J. Beale, J. D. Benjamin, M. J. Uren, N. G. Chew, and A. G. Cullis, J. Cryst.
ditives in order to understand the effects of porosity, graphite type, Growth, 73(3), 622 (1985).
and composition on structural and transport properties. Calculated 10. M. M. Forouzan, C.-W. Chien, D. Bustamante, W. Lange, B. A. Mazzeo, and
tortuosities were fit to a modified Bruggeman-type relation, giving an D. Wheeler, Meeting Abstracts, MA2015-01(31), 1798 (2015).
11. M. M. Forouzan, A. Gillespie, N. Lewis, B. A. Mazzeo, and D. R. Wheeler, Meeting
independent correlation for each type of graphite tested. Abstracts, MA2016-02(3), 271 (2016).
By understanding effects of conductive additives on microstructure 12. Q. F. Dong, Y. W. Zhu, M. S. Zheng, Y. D. Zhan, and Z. G. Lin, J. Electrochem. Soc.,
and transport properties, taking into account cost and application, bat- 153(3), A459 (2006).
tery performance may be optimized. Our results reveal that effective 13. I. V. Thorat, D. E. Stephenson, N. A. Zacharias, K. Zaghib, J. N. Harb, and
D. R. Wheeler, J. Power Sources, 188(2), 592 (2009).
ionic conductivity increases with porosity and decreases with increas- 14. N. A. Zacharias, D. R. Nevers, C. Skelton, K. Knackstedt, D. E. Stephenson, and
ing graphite fraction. KS6, KS15, MX15, Graphene, and BNB90 were D. R. Wheeler, J. Electrochem. Soc., 160(2), A306 (2013).
found to have decreasing ionic conductivities in that order. Tortuosity 15. D. E. Stephenson, B. C. Walker, C. B. Skelton, E. P. Gorzkowski, D. J. Rowenhorst,
experiments in orthogonal direction to the direction of compression and D. R. Wheeler, J. Electrochem. Soc., 158(7), A781 (2011).
16. I. V. Thorat, T. Joshi, K. Zaghib, J. N. Harb, and D. R. Wheeler, J. Electrochem. Soc.,
(corresponding to in-plane tortuosity in Li-ion cells) on porous alka- 158(11), A1185 (2011).
line electrodes were performed, showing better ionic transport in the 17. J. H. Cao, B. K. Zhu, and Y. Y. Xu, J. Membr. Sci, 281(1–2), 446 (2006).
direction orthogonal to the direction of compression. This is advanta- 18. K. K. Patel, J. M. Paulsen, and J. Desilvestro, J. Power Sources, 122(2), 144 (2003).
geous in manufactured alkaline cells because ions primarily transport 19. M. M. Forouzan, L. Robertson, M. Wray, and D. Wheeler, Meeting Abstracts,
MA2015-01(1), 21 (2015).
in orthogonal direction to the compression. 20. P. Ramadass, B. Haran, R. White, and B. N. Popov, J. Power Sources, 123(2), 230
It was observed in the experiments that contact with electrolyte (2003).
over time tended to cause swelling or disintegration of the pressed 21. M. Doyle, J. Newman, A. S. Gozdz, C. N. Schmutz, and J. M. Tarascon, J. Elec-
electrodes. This could result in improved ionic transport in real cells trochem. Soc., 143(6), 1890 (1996).
22. T. Kotaka, Y. Tabuchi, and P. P. Mukherjee, J. Power Sources, 280(0), 231 (2015).
over time, while at the same time harming electronic transport. On the 23. A. Gully, H. Liu, S. Srinivasan, A. K. Sethurajan, S. Schougaard, and B. Protas, J.
other hand, loss of solvent over time (i.e. drying out of the cell) would Electrochem. Soc., 161(8), E3066 (2014).
tend to inhibit ion transport. 24. J. Guan and M. Liu, Solid State Ionics, 110(1–2), 21 (1998).
A COMSOL Multiphysics model was developed and results were 25. W. Du, N. Xue, W. Shyy, and J. R. R. A. Martins, J. Electrochem. Soc., 161(8), E3086
(2014).
in semi-quantitative agreement with experimental polarization data. 26. M. Andersson, J. Yuan, and B. Sundén, Applied Energy, 87(5), 1461 (2010).
According to the observations and modeling results, an ion trans- 27. D. Kehrwald, P. R. Shearing, N. P. Brandon, P. K. Sinha, and S. J. Harris, J. Elec-
fer and reaction mechanism proposed so that OH− is generated on trochem. Soc., 158(12), A1393 (2011).

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A3130 Journal of The Electrochemical Society, 164 (13) A3117-A3130 (2017)

28. B. Vijayaraghavan, D. R. Ely, Y.-M. Chiang, R. Garcı́a-Garcı́a, and R. E. Garcı́a, J. 48. Improved manganese dioxide for alkaline cells, Pat. #: EP1490918,
Electrochem. Soc., 159(5), A548 (2012). http://www.google.com/patents/EP1490918B1?cl=en (2009).
29. C. Ding-Wen, E. Martin, R. E. David, W. Vanessa, and R. E. Garcı́a, Modell. Simul. 49. M. Wissler, J. Power Sources, 156(2), 142 (2006).
Mater. Sci. Eng., 21(7), 074009 (2013). 50. R. J. Gilliam, J. W. Graydon, D. W. Kirk, and S. J. Thorpe, Int. J. Hydrogen Energy,
30. T. Hutzenlaub, A. Asthana, J. Becker, D. R. Wheeler, R. Zengerle, and S. Thiele, 32(3), 359 (2007).
Electrochem. Commun., 27 77 (2013). 51. T. Chapman, in “Chemical Engineering”, Vol. Ph.D. University of California, Berke-
31. S. J. Cooper, D. S. Eastwood, J. Gelb, G. Damblanc, D. J. L. Brett, R. S. Bradley, ley, 1967.
P. J. Withers, P. D. Lee, A. J. Marquis, N. P. Brandon, and P. R. Shearing, J. Power 52. P. M. Gomadam, J. W. Weidner, T. A. Zawodzinski, and A. P. Saab, J. Electrochem.
Sources, 247(0), 1033 (2014). Soc., 150(8), E371 (2003).
32. T. DuBeshter, P. K. Sinha, A. Sakars, G. W. Fly, and J. Jorne, J. Electrochem. Soc., 53. A. Yuan and Q. Zhang, Electrochem. Commun., 8(7), 1173 (2006).
161(4), A599 (2014). 54. Y. Chabre and J. Pannetier, Prog. Solid State Chem., 23, 1 (1995).
33. M. Ebner, D. W. Chung, R. E. Garcı́a, and V. Wood, Adv. Energy Mater., 4(5), 55. T. W. Farrell, C. P. Please, D. L. S. McElwain, and D. A. J. Swinkels, J. Electrochem.
1301278 (2014). Soc., 147(11), 4034 (2000).
34. M. Ebner and V. Wood, J. Electrochem. Soc., 162(2), A3064 (2015). 56. T. W. Farrell and C. P. Please, J. Electrochem. Soc., 152(10), A1930 (2005).
35. H. Iwai, N. Shikazono, T. Matsui, H. Teshima, M. Kishimoto, R. Kishida, D. Hayashi, 57. D. Cericola and M. E. Spahr, Electrochim. Acta, 191 558 (2016).
K. Matsuzaki, D. Kanno, M. Saito, H. Muroyama, K. Eguchi, N. Kasagi, and 58. I. J. Ong and J. Newman, J. Electrochem. Soc., 146(12), 4360 (1999).
H. Yoshida, J. Power Sources, 195(4), 955 (2010). 59. W. G. Sunu and D. N. Bennion, J. Electrochem. Soc., 127(9), 2007 (1980).
36. Z. Fishman and A. Bazylak, J. Electrochem. Soc., 158(2), B247 (2011). 60. Z. Mao and R. E. White, J. Electrochem. Soc., 139(4), 1105 (1992).
37. M. J. Martinez, S. Shimpalee, and J. W. Van Zee, J. Electrochem. Soc., 156(1), B80 61. E. J. Podlaha and H. Y. Cheh, J. Electrochem. Soc., 141(1), 15 (1994).
(2009). 62. E. J. Podlaha and H. Y. Cheh, J. Electrochem. Soc., 141(1), 28 (1994).
38. D. M. Bernardi, J. Electrochem. Soc., 137(11), 3344 (1990). 63. R. B. Smith and M. Z. Bazant, J. Electrochem. Soc., 164(11), E3291 (2017).
39. F. M. Delnick and R. A. Guidotti, J. Electrochem. Soc., 137(1), 11 (1990). 64. A. Cañas and J. Benavente, J. Colloid Interface Sci., 246(2), 328 (2002).
40. A. Suzuki, T. Hattori, R. Miura, H. Tsuboi, N. Hatakeyama, H. Takaba, 65. J. Newman and K. E. Thomas-Alyea, Electrochemical systems, John Wiley & Sons
M. C. Williams, and A. Miyamoto, Int. J. Electrochem. Sci., 5(12), 1948 (2012).
(2010). 66. Carbon Powder-Based Solutions for Alkaline Batteries (TIMCAL). www.timcal.ch/
41. A. Z. Weber, R. M. Darling, and J. Newman, J. Electrochem. Soc., 151(10), A1715 scopi/group/timcal/timcal.nsf/pagesref/MCOA-7S6JFG/$File/TIMCAL_brochure_
(2004). alkaline.pdf (2017).
42. H. Munakata, D. Yamamoto, and K. Kanamura, J. Power Sources, 178(2), 596 67. Graphene supermarket. “graphene nanopowder: 12 nm flakes-5 g”,https://
(2008). graphene-supermarket.com/Graphene-Nanopowder-12-nm-Flakes-5-g.html (2017).
43. P. O. Lopez-Montesinos, A. V. Desai, and P. J. A. Kenis, J. Power Sources, 269(0), 68. M. M. Hantel, M. J. Armstrong, F. DaRosa, and R. l’Abee, J. Electrochem. Soc.,
542 (2014). 164(2), A334 (2017).
44. A. Unemoto, T. Matsuo, H. Ogawa, Y. Gambe, and I. Honma, J. Power Sources, 69. K. M. Abraham, Electrochim. Acta, 38(9), 1233 (1993).
244(0), 354 (2013). 70. R. B. MacMullin and G. A. Muccini, AIChE J., 2(3), 393 (1956).
45. D. Marrero-López, L. dos Santos-Gómez, L. León-Reina, J. Canales-Vázquez, and 71. J. Landesfeind, J. Hattendorff, A. Ehrl, W. A. Wall, and H. A. Gasteiger, J. Elec-
E. R. Losilla, J. Power Sources, 245(0), 107 (2014). trochem. Soc., 163(7), A1373 (2016).
46. H. S. Chen, S. J. Lue, Y. L. Tung, K. W. Cheng, F. Y. Huang, and K. C. Ho, J. Power 72. D. E. Stephenson, E. M. Hartman, J. N. Harb, and D. R. Wheeler, J. Electrochem.
Sources, 196(8), 4162 (2011). Soc., 154(12), A1146 (2007).
47. S. P. Sheu, C. Y. Yao, J. M. Chen, and Y. C. Chiou, J. Power Sources, 68(2), 533 73. D. Djian, F. Alloin, S. Martinet, H. Lignier, and J. Y. Sanchez, J. Power Sources,
(1997). 172(1), 416 (2007).

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

You might also like