Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Chemical Engineering Science 154 (2016) 61–71

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Charge coup de fouet phenomenon in soluble lead redox flow battery


Mahendra Nandanwar, Sanjeev Kumar n
Department of Chemical Engineering, Indian Institute of Science, Bangalore 560012, India

H I G H L I G H T S

 Voltage dip after initial peaking of charging voltage, and slow return to normalcy.
 Effective conductivity is obtained using a percolation model with a critical fraction.
 Voltage drop across surface deposits included in model through reduced over-potential.
 Insulating regions on electrode expand and contract with discharge/charge cycle.
 Charging Coup de fouet increases with increase in depth of discharge in previous cycle.

art ic l e i nf o a b s t r a c t

Article history: The charge coup de fouet phenomenon, known in the context of lead-acid battery, refers to the presence
Received 15 February 2016 of a voltage dip shortly after charging of a fully discharged battery begins. While the attempts to relate
Received in revised form magnitude of coup de fouet phenomena with the state of health of battery have appeared in the literature,
30 June 2016
the phenomena continue to be poorly understood. The soluble lead redox flow battery (SLRFB), with
Accepted 3 July 2016
Available online 7 July 2016
potential for energy storage at large scale at low cost, also displays similar features. We report in this
work our modeling and experimental efforts aimed at understanding charge coup de fouet phenomenon
Keywords: in natural convection driven SLRFB. We present a model that incorporates the presence of non-con-
Flow battery ducting PbO in deposits through a percolation type model for conductivity. The associated potential drop
Effective conductivity
across the deposits is incorporated in the model through reduced overpotential available for driving
Percolation model
Butler–Volmer kinetics. The complete model with coupled natural convection induced by non-uniform
Overpotential
Natural convection concentration of Pb ions in electrolyte successfully captures charge coup de fouet phenomenon, and ex-
Soluble lead plains the measured variation of its magnitude with the depth of discharge in the previous cycle. The
model explains our earlier observation that during the discharge process, a receding zone of deposits is
seen only on cathode but not on anode. The approach used is applicable to electrochemical systems in
which solid conducting matrix evolves to non-conducting or poorly conducting and vice-versa, with a
change in composition.
& 2016 Elsevier Ltd. All rights reserved.

1. Introduction decrease slowly and steadily, as expected of discharging voltage


for a battery. These two phenomena, shown schematically in Fig. 1,
Coup de fouet phenomena are reported in the literature in the are named charge coup de fouet and discharge coup de fouet re-
context of conventional lead-acid batteries. When a fully dis- spectively. Depending on the operating conditions, magnitude of
charged battery is charged, the voltage first rises sharply in a ra- the voltage dip and voltage rise in charge and discharge coup de
ther short time, following which is decreases, goes through a fouet phenomena varies in range of 10–50 mV, and lasts from a few
comparatively shallow minimum, and then begins to rise slowly seconds to a couple of minutes (Delaille et al., 2006; Pascoe and
and steadily, as expected of charging voltage for a battery. Simi- Anbuky, 2002). These phenomena find use in battery diagnosis
larly, when a fully charged battery is discharged, the battery vol- through their correlation with state of health (SOH) of a battery
tage first drops sharply in a rather short time, following which it (Bose and Laman, 2000; Pascoe and Anbuky, 2002; Pascoe et al.,
rises, goes through a rather broad maximum, and then begins to 2002), an indicator of its ability to perform well in charge-dis-
charge cycles. The SOH tells a user when to replace a battery. The
n
Corresponding author. coup de fouet phenomena are also correlated to state of charge
E-mail address: sanjeev@chemeng.iisc.ernet.in (S. Kumar). (SOC) of a battery (Pascoe and Anbuky, 1999) which indicates how

http://dx.doi.org/10.1016/j.ces.2016.07.001
0009-2509/& 2016 Elsevier Ltd. All rights reserved.
62 M. Nandanwar, S. Kumar / Chemical Engineering Science 154 (2016) 61–71

Fig. 1. Schematics of (a) charge coup de fouet and (b) discharge coup de fouet observed in conventional lead-acid battery.

much more charge can be put in it. review of the current understanding of SLRFB, as this is needed for
Berndt and Voss (1965) has studied charge coup de fouet in model development. This is followed by sections covering ex-
conventional lead-acid battery, and suggested that a peak in po- perimental details, model development, and results.
tential could arise from insulating PbSO4 formed on discharge
products in lead-acid battery. Delaille et al. (2006) have contested
the above explanation. According to them, end-of-discharge does 2. Current state of understanding of SLRFB
not correspond to complete coverage of electrode by an insulating
layer of PbSO4. The authors instead proposed mass transport The continued investigations on SLRFB over the last ten years
limitations to be the reason for charge coup de fouet. No mathe- have brought up a number of their characteristics which need to be
matical model appears to have captured the peak associated with explained by models. Hazza et al. (2004) experimented with
charge coup de fouet yet. 10 cm  10 cm electrodes with inter-electrode separation of 1.2 cm
The recently proposed membraneless soluble lead-acid redox and found anode to be the limiting one. Collins et al. (2010) reported
flow battery (SLRFB) with only one electrolyte flow loop (Hazza that charging in the first cycle occurs at a nearly constant voltage,
et al., 2004, 2005; Li et al., 2009; Pletcher and Wills, 2004, 2005; followed by two step charging in the subsequent cycles. Beck and
Pletcher et al., 2008ab) is a promising candidate for large scale Bhn (1975) suggested that reduction of PbO2 during discharge leads
storage of energy, as it can potentially offer a number of ad- to its electro-dissolution and also formation of an insoluble deposit
vantages over the other widely used redox flow batteries (Verde which oxidizes more readily than Pb2 þ ions in the next charge cycle
et al., 2013; Nandanwar and Kumar, 2014). The active species Pb2 þ to form PbO2 deposits. The second cycle charging therefore starts at a
in SLRFB comes from lead methanesulfonate salt [(CH3SO3)2Pb: lower voltage. Pletcher and Wills (2005) carried out EDAX analysis of
PbMS] dissolved in aqueous solution of methanesulfonic acid insolubles on anode and found Pb and O in ratio close to 1:1, sug-
[CH3SO3H: MSA]. The redox reactions involved are: gesting PbO as insolubles. Oury et al. (2012) carried out mass change

charge ⎫
Negative electrode: Pb2 + + 2e− H ooooooooI Pb (s ) ( − 0.126 V /SHE )⎪
discharge ⎪
discharge ⎪
Positive electrode: PbO2 (s) + 4 H+ + 2e− H ooooooooI Pb2 + + 2H2 O ( + 1.455 V /SHE )⎬
charge ⎪
charge ⎪
Overall reaction: 2Pb2 + + 2H2 O H ooooooooI Pb (s ) + PbO2 (s) + 4H+ ( + 1.58 V /SHE ) ⎪

discharge ⎭ (1)

In charge cycle, lead ions are deposited from electrolyte on investigations at anode using quartz crystal micro-balance to confirm
cathode and anode in the form of Pb and PbO2 respectively. The the presence of solid state reaction on anode, and suggested for-
same are redissolved as Pb2 þ ions in discharge cycle. The electrode mation of non-stoichiometric PbOx. The authors suggested that such
reactions involved in soluble lead battery appear much simpler non-stoichiometric compounds may come from modification of in-
than those involved in the conventional lead-acid battery. The termediate compound Pb(OH)2 formed during PbO2 to Pb2 þ con-
measurements however show similar charge coup de fouet and version. These findings provided basis for Shah et al. (2010) to pro-
discharge coup de fouet phenomena (Verde et al., 2013). pose the following side reaction at anode to run concurrently to the
In this work, we investigate charge coup de fouet in SLRFB ex- main reactions given by Eq. (1):
perimentally, and develop a quantitative model to understand it. charge
The latter is necessary as the existing models do not capture it. The PbO + H2 O H ooooooooI PbO2 + 2e− + 2H+.
discharge (2)
experiments are carried out in a standard rectangular cell, with no
external flow, but in the presence of induced natural convection, Shah et al. (2010) developed the first model for SLRFB. The
explained in latter sections. We begin in the next section with a model considered Nernst-Planck equations to govern transport of
M. Nandanwar, S. Kumar / Chemical Engineering Science 154 (2016) 61–71 63

ions in the flow field prevailing in their flow-through cell, Butler– voltage towards the end of discharge cycle through partial baring
Volmer kinetics for electron-transfer reactions, and uniform de- of electrodes. It also predicts that alternating flow direction can
position/dissolution of solid Pb and PbO2 on cathode and anode. enhance performance of a SLRFB.
An empirical expression similar to the Butler–Volmer equations Nandanwar and Kumar (2016a) found that natural convection
but involving surface concentrations of solid deposits was pro- induced by difference in concentration of Pb2 þ ions near the
posed by them to account for the side reaction given by Eq. (2). electrode surface and the bulk of the electrolyte plays a dominant
The model captured the onset of two step charging from second role in the performance of a soluble lead battery with no external
charge cycle onwards very well. It required much smaller con- circulation. This has been modeled by Nandanwar and Kumar
centration of hydrogen ions than that present in the system to (2016b) by using a concentration difference driven body force
capture rapid decrease in discharge voltage towards the end (Na- term in the equations of motion in the model of Nandanwar and
danwar and Kumar, 2014). Bates et al. (2014) used similar equa- Kumar (2014). The augmented model of Nandanwar and Kumar
tions to study the effect of temperature and electrolyte con- (2016b), with all the parameter values same as those of Nandan-
centration, while Oury et al. (2014) investigated deposition of PbO2 war and Kumar (2014), very well explains the experimental data
alone in honeycomb shaped anode. obtained by them in a narrow rectangular cell with no external
Nandanwar and Kumar (2014) argued for varying mass transfer flow, shown schematically in Fig. 2.
resistance along the length of an electrode due to the growing The conductivity of electrodes and deposits in all the previous
concentration boundary layer, and accounted for non-uniform models is considered to be infinitely large compared to that of the
deposition and dissolution of PbO2 and Pb deposits in their model. electrolyte. These models therefore do not require potential dis-
Their model equations and the form used for Butler–Volmer ki- tribution in solid phase to be determined. On a microscopic level,
netics are the same as those of Shah et al. (2010). The electro- the assumption of high conductivity of deposits holds well for
dissolution of PbO2 in their model is permitted to continue in- cathode as current passes either through Pb deposits or graphite
dependent of the buildup of the non-conducting PbO locally. In the electrode, both of which are highly conducting. The situation at
charge cycle, the oxidation of solid PbO and Pb2 þ in the electrolyte anode is different though. The assumption of high conductivity of
commences simultaneously to form PbO2 everywhere on the anode holds for the first charge cycle as the deposit consists of
electrode. The model, with some refitting of parameters proposed only highly conducting PbO2. In the first discharge cycle, some of
by Shah et al. (2010) for reaction kinetics, successfully captures the PbO2 is electro-dissolved and some other is converted into
charge-discharge characteristics and rapid decrease in discharge non-conducting PbO solid (Eq. (2)). Oury et al. (2012) monitored

Fig. 2. Schematic of (a) closed rectangular cell with electrodes mounted on walls used to carry out experiments, and (b) the computational grid used to carry out simulations.
64 M. Nandanwar, S. Kumar / Chemical Engineering Science 154 (2016) 61–71

mass change at anode during discharge cycle using quartz crystal We adopt the model of Nandanwar and Kumar (2016b) which
micro-balance and impedance spectra. Their studies confirmed incorporates natural convection caused by electrochemical reac-
PbO2 to PbO conversion and formation of an insulating layer that tion on electrode surface and non-uniform current density on it as
eventually stops discharge process altogether. the base model, and incorporate in it electrical resistance of solid
The assumption of highly conducting deposits on anode irre- deposits on electrodes (Euler et al., 1980). The model set consists
spective of the proportion of non-conducting PbO in it appears to of transport equations for Pb2 þ , H þ , and ( CH3 SO−3 ), velocity field,
be an extreme one. In this work, we seek to incorporate in the
electrode kinetics, and relevant initial and boundary conditions.
model of Nandanwar and Kumar (2016b) changes occurring in
The commonly made assumptions, stated and discussed in detail
conductivity of deposits on anode, and study its impact on charge-
earlier (Shah et al., 2010; Nandanwar and Kumar, 2014) are: excess
discharge profiles, particularly in the context of charge coup de
water, dilute solution theory (Newman and Thomas-Alyea, 2004),
fouet which is not explained by any of the existing models.
laminar flow, and no gas evolution at electrodes. The transport
equation for ionic species are given by
3. Experimental ∂ci
= − ∇·Ni i = H+, Pb2 + , CH3 SO−3
∂t (4)
The experiments were carried out in a rectangular cell
where flux Ni is
(4 cm  1.5 cm  7 cm) made of transparent perspex, housing two
graphite electrodes (3 cm  3 cm) located symmetrically on the zi ci Di F
Ni = − Di ∇ci − ∇ϕ + ci u.
opposite walls, with an inter-electrode gap of 1.5 cm. We used RT (5)
smaller electrodes than those reported in the literature so as to
The electro-neutrality and charge conservation require
facilitate CFD simulations to be completed in reasonable time
(about 24 h). We expect electrode size to leave charge coup de ∑ zi ci = 0,
fouet phenomenon investigated in this work unaffected. The ex- i (6)
perimental procedure adopted is as follows. Briefly, the cell was
completely filled with electrolyte solution consisting of 0.5 M
methanesulfonic acid and 0.5 M lead methanesulfonate, and ∇·j = 0; j = F ∑ zi Ni .
aligned vertically. No additives were used. Electrochemical mea- i (7)
surements were carried out using Autolab Potentiostat/Galvano-
The time dependent velocity field u in Eq. (4) is obtained by sol-
stat-Model 30, in the absence of any external convection. The
battery cell was first galvanostatically charged (Banerjee et al., ving the Navier–Stokes equation
2013) for 900 s followed by relaxation (open circuit) for 300 s. The ∂u
ρ + ρ (u ·∇) u = − ∇p + μ∇2u + F
battery was then discharged galvanostatically. Unless stated ∂t (8)
otherwise, a cut-off voltage of 1.2 V and current density of
and the continuity equation for an incompressible fluid
200 A/m2 were used. The discharge phase was followed by a re-
laxation period of 300 s before starting the next charging phase. A ∇·u = 0. (9)
magnetic stirrer was used to homogenize electrolyte during the
relaxation phase (open circuit condition) to eliminate stratification The body-force F in Eq. (8) is given by Gebhart and Pera (1971)
resulting from charging/discharging. All the experiments were
F = gρ ∑ βi (ci0 − ci )
performed at room temperature. i (10)
þ
for i being species Pb2 þ and H . Here, c0i
is the initial concentra-
4. The model tion and βi is the expansion coefficient defined as

1 ⎛ ∂ρ ⎞
Euler et al. (1980) have proposed that when volume fraction of βi = ⎜ ⎟ .
ρ ⎝ ∂ci ⎠T , P, c j ≠ i (11)
conducting material in a matrix becomes less than a ‘critical limit’,
the conducting pathways cease to exist and conductivity of the The subscripts T, P, and cj indicate quantities held constant.
solid matrix drops to zero. Above the critical limit, the solid mix- The mass balance for deposition of solids (Csi) due to the local
ture behaves like a good conductor and below like an insulator.
electrochemical reaction, reflected by local current density (j), is
The conductivity of such a mixture (se), as given by effective
given by
medium theory (Euler et al., 1980), is
s s
dC PbO 2 dCPbO
1 ⎛ ⎞ at anode, 0<y<L = j PbO 2 /2F; = − jPbO /2F
σe = ⎜Q + Q2 + 2 (z − 2) σa σ b ⎟ dt dt (12)
z − 2⎝ ⎠ (3)

where
s
dCPb
⎛z ⎞ ⎛z ⎞ at cathode, 0<y<L = jPb /2F
Q = ⎜ A a − 1⎟ σa + ⎜ Ab − 1⎟ σ b, dt (13)
⎝2 ⎠ ⎝2 ⎠
The thickness of the deposits formed is several orders of magni-
Aa and Ab are fractional volumes of components a (PbO) and b
tude smaller than the inter-electrode gap. We therefore assume
(PbO2), sa and sb are specific conductivities of components, and z
is coordination number (number of contacts per particle). Re- the latter to remain unchanged. The local current density at anode
cently, Gandhi (2015) has incorporated formulation of Euler et al. jan is equal to jPbO2 + jPbO . The current density for the main reaction
(1980) to model charge/discharge of conventional lead-acid bat- on anode (Eq. (1)) is given by the classical Butler–Volmer reaction
teries. The author found that conductivity of the matrix plays a kinetics (Shah et al., 2010; Bates et al., 2014). We use here, without
major role in determining discharge capacity of battery. The au- refitting any parameters, the expression provided by Nandanwar
thor did not comment on coup de fouet phenomena. and Kumar (2014).
M. Nandanwar, S. Kumar / Chemical Engineering Science 154 (2016) 61–71 65

⎧ ⎛ ⎞⎛ ⎞⎡ ⎤ Vcell = V+ − V− (20)
⎪ ⎜ c 2+ ⎟ ⎜ cH + ⎟ ⎢ ⎛ 0.5Fη+ ⎞ ⎛ − 0.5Fη+ ⎞ ⎥
⎪ 2Fk PbO 2 ⎜ Pb
⎟⎜ ⎟ ⎢ exp ⎜ ⎟ − exp ⎜ ⎟ ⎥ for Case 1
jPbO2 = ⎨ ⎜ c ref2 + ⎟ ⎜ c ref+ ⎟ ⎢ ⎝ RT ⎠ ⎝ RT ⎠ ⎥ where V− is fixed at zero as the reference voltage.
⎪ ⎝ Pb ⎠⎝ H ⎠⎣ ⎦
⎪ An exact incorporation of spatially and temporally varying
⎩ 0, for Case 2 (14)
electrical resistance of deposits requires determination of voltage
The complexities associated with interconversion of solid PbO2 profile in the composite solid phase. The voltage on the surface of
to solid PbO and vice versa and the need for hydrogen ions in the deposits is coupled to the spatially distributed current density
electrolyte to participate in this reaction necessitate the use of an through rate expressions for electrochemical reactions. This in-
empirical approach. Shah et al. (2010) have proposed a form si- creases the computational complexity many fold. In view of the
milar to Butler–Volmer kinetics expression to model solid-solid high conductivity of the underlying graphite electrode and rela-
reaction. It involves surface concentrations of the deposited solids. tively small thickness of the deposit layer, we propose the fol-
The same approach is used by others subsequently (Bates et al., lowing. We treat the original electrode material (graphite) to be at
2014; Nandanwar and Kumar, 2014). In the present work, we have uniform potential everywhere, and consider the potential drop
used the expression provided by Nandanwar and Kumar (2014) required across the deposit, which supports the local current, to
without refitting any parameters: reduce the surface potential to which the reacting species are
exposed.
⎛ C s ⎞ν ⎛ ⎞
0.9Fη+ ⎟ In other words, the potential drop across the deposits reduces
jPbO = 2Fkf ⎜⎜ sPbO ⎟ exp ⎜
, ref ⎟ ⎜ RT ⎟ the overpotential available otherwise to drive Butler–Volmer ki-
⎝ CPbO ⎠ ⎝ ⎠
netics. We suitably modifying the value of overpotential on anode.
s ⎛ ⎞
CPbO 2 cH −0.9Fη+ ⎟ The voltage drop ΔVs required to support the flow of current from
exp ⎜⎜
+
− 2Fkb ⎟
s, ref
c ref+ the site of reaction on deposit surface to the underlying graphite
CPbO 2 H ⎝ RT ⎠ (15)
electrode, in the limit of one dimensional approximation for thin
The coefficient 0.9 used in the exponential terms is consistent with deposits, is
the constraint that in a multi-step reaction, a part of the energy lt
made available by over-potential is dissipated (Schmickler and ΔVs = jan × Rs = jan
σe (21)
Santos, 2010). The overpotential for the side reaction on anode is
considered to be the same as that for the main reaction, and any where jan = jPbO2 + jPbO and Rs = lt /σe is the resistance offered by
deviation is absorbed in constants kf and kb (Shah et al., 2010). The deposits of thickness lt with composition dependent electrical
value of exponent ν controls the onset of second step during conductivity se. The local overpotential available to drive the
charging from second cycle onwards, and is not expected to be an electrochemical reactions is suitably modified from η þ to η′+ as
integer for a non-elementary reaction.
The Butler–Volmer kinetics at cathode is given by (Shah et al.,
( )
η+′ = V+ − ΔVs − ϕ − E± = η+ − ΔVs (22)
2010; Bates et al., 2014; Nadanwar and Kumar, 2014): The initial and boundary conditions needed to simulate the
⎧ ⎛ ⎞⎡ ⎤ model, for the computational domain shown in Fig. 2, are as fol-
⎪ c 2+ ⎛ Fη ⎞ ⎛ −Fη− ⎞ ⎥
⎪ 2FkPb ⎜ Pb ⎟ ⎢ exp ⎜ − ⎟ − exp ⎜ ⎟ for Case 1 lows. The boundary conditions at anode and cathode are:
jPb =⎨ ⎜ c ref ⎟ ⎢ ⎝ RT ⎠ ⎝ RT ⎠ ⎥
⎪ ⎝ Pb2 + ⎠ ⎣ ⎦ 2F (NPb2 + ·n) = jPbO 2 F (NH + ·n) = 2jPbO 2 + jPbO 0<y<L (23)

⎩ 0, for Case 2 (16)

with all the parameter values held same as those provided by 2F (NPb2 + ·n) = jPb (NH + ·n) = 0 0<y<L (24)
Nandanwar and Kumar (2014). The two cases appearing in Eqs.
(14) and (16) are devised to prevent continued dissolution from a where n is the outward normal to surface. The applied average
local region even after all the active solids (Pb and PbO2) are current density at anode satisfies
electro-dissolved. These cases are: L
1
japp =
L
∫0 jan dy .
(25)
1. Case 1: (i) charging and Cis ≥ 0, and (ii) discharging and Cis > 0
2. Case 2: discharging and Cis = 0 The no-flux boundary condition ( j·n = 0) was applied at all other
boundaries. The initial conditions for concentrations of ionic spe-
where i stands for PbO2 on anode and Pb on cathode. cies in the electrolyte domain are specified as
For uniform potential in electrode and deposits, which corres-
c Pb2 +(x, y , t = 0) = c 0 2 ; c H +(x, y , t = 0) = c H0 +.
ponds to zero electrical resistance everywhere in solid phases, η þ Pb + (26)
and η  in Eqs. (14)–(16) are given by The initial concentration of solid species deposited on electrodes,
s (x = 0, y , t = 0) , C s
CPb , and CPbO (x = W , y, t = 0),
η± = V± − ϕ± − E±, (17) PbO2 (x = W , y , t = 0)
were set to zero. The initial velocity field u in the domain was also
where V± are electrode potentials, ϕ± are potentials in electrolyte specified to be zero everywhere. The no-slip condition was em-
next to the electrodes, and E± are Nernst potentials, given by (at ployed at all solid-electrolyte interfaces. The reference pressure
anode) pref was specified at the top left corner of the numerical domain.
RT ⎛ ⎞ The system of equations along with the initial and boundary
E+ = E+0 − ⎜ ln c Pb2 + − ln c H + ⎟. conditions can be solved in principle to obtain model predictions.
2F ⎝ ⎠ (18)
The presence of the unknown local flux jan in the exponential
and (at cathode) terms in Butler–Volmer equations makes the model highly cou-
pled and implicit. Even for the simplified 2-d configuration and
RT
E− = E−0 + ln c Pb2 +, relatively small size electrodes used in the work, the model re-
2F (19)
quires extensive computational effort in COMSOL to obtain pre-
The cell voltage is defined as dictions. We therefore simplified the model equations further
66 M. Nandanwar, S. Kumar / Chemical Engineering Science 154 (2016) 61–71

using the approximations discussed below. We rewrite Eqs. (14) Table 1


and (15), with η′+ replacing η þ , as Values of electrochemical parameters used for simulation of SLRFB.

⎡ ⎛ 0.5Fη ⎞ ⎛ −0.5FγjPbO Rs ⎞ Parameter Value Parameter Value


jPbO2 = i0 ⎢ exp ⎜ +
⎟ exp ⎜ 2

⎢⎣ ⎝ RT ⎠ ⎝ RT ⎠ D Pb2 + a 0.61e  9 m2/s kPbO2 6.25e  5 mol/m2/s
⎛ −0.5Fη ⎞ ⎛ 0.5FγjPbO Rs ⎞ ⎤ D H+b 5.66e  9 m2/s kPb 2.75e  3 mol/m2/s
− exp ⎜ +
⎟ exp ⎜ 2
⎟⎥ DCH3SO− c 2
2.0e  10 m /s kf 2.5e  4 mol/m2/s
⎝ RT ⎠ ⎝ RT ⎠ ⎥⎦ (27) 3
z Pb2 + 2 kb 2.5e  6 mol/m2/s

where z H+ 1 ν 1.4
zCH3SO− 1 s, ref
CPbO 1 mol/m2
⎛ ⎞⎛ ⎞ 3
c 2 + ⎟ ⎜ cH + ⎟ 3
1 mol/m2
i0 = 2Fk PbO 2 ⎜ Pb c0 500 mol/m s, ref
CPbO
Pb2 +
⎜ c ref ⎟ ⎜ c ref+ ⎟ 2
⎝ 2+ ⎠⎝ H ⎠
Pb c 0+
H
500 mol/m3 c ref2 + 500 mol/m3
Pb
T 300 K c ref+ 500 mol/m3
and H
pref 100 kPa E+0 1.56 V
⎛ C s ⎞ν ⎛ ⎞ ⎛ ⎞
0.9Fη+ ⎟ −0.9FλjPbO Rs ⎟
jPbO = 2Fkf ⎜⎜ sPbO ⎟ exp ⎜ exp ⎜⎜ μ 0.001 Pa s E−0 d  0.13 V
, ref ⎟ ⎜ ⎟ ⎟
⎝ CPbO ⎠ ⎝ RT ⎠ ⎝ RT ⎠ ρ0 1169 kg/m3 β Pb2 + 2.573e  4 m3/mol
F 96485 C/mol β H+ 3.328e  5 m3/mol
s
CPbO ⎛ ⎞ ⎛ ⎞
2 cH −0.9Fη+ ⎟ 0.9FλjPbO Rs ⎟
exp ⎜⎜ exp ⎜⎜
+ R 8.314 J/mol/k L 0.03 m
− 2Fkb s, ref ⎟ ⎟ W 0.015 m
CPbO c ref+
2 H ⎝ RT ⎠ ⎝ RT ⎠ (28)
a
Hazza et al. (2004).
Here, 1/γ and 1/λ represent fractional contribution of currents jPbO2 b
Soudijn (2012).
c
and jPbO to the total current jan, and satisfy the constraint: Telfah et al. (2010).
d
Zhang et al. (2011).
1 1
+ =1
γ λ (29) potential needed for the current to flow corresponds to an as-
along with γ, λ > 1. Further simplification was carried out by ex- sumed ratio of current carried by the two reactions (through fixed
panding exponential terms involving deposit resistance Rs and values of λ and γ). The values of these variables were changed
retaining only the first two terms. The new Butler–Volmer equa- later, while satisfying the constraint ( λ−1 + γ −1 = 1), to see their
tions for jPbO2 and jPbO take the following form: impact on model predictions.
The value of local deposit thickness lt, which also appears on
⎡ ⎛ 0.5Fη ⎞ ⎛ −0.5Fη+ ⎞ ⎤
i0 ⎢ exp ⎜ RT + ⎟ − exp ⎜ ⎟⎥ the right hand side through Rs, is uniquely related to the local and
⎣ ⎝ ⎠ ⎝ RT ⎠ ⎦ time varying values of C PbO2 and CPbO. In order to keep the equa-
jPbO2 =
0.5Fi γR ⎡ −0.5Fη+ ⎤
(
0.5Fη
1 + RT0 s ⎣⎢ exp RT + + exp ) RT (
⎦⎥ ) (30)
tions explicit, we have followed the same approach for lt as well,
and fixed its value to 250 μm. We will establish in the next section
and that the model predictions are not sensitively dependent on the

⎛ C s ⎞ν ⎛ ⎞ s
CPbO ⎛ ⎞
0.9Fη+ ⎟ c + −0.9Fη+ ⎟
2Fkf ⎜⎜ sPbO ⎟ exp ⎜ − 2Fkb s, ref2 Href exp ⎜⎜
, ref ⎟ ⎜ ⎟ ⎟
⎝ CPbO ⎠ ⎝ RT ⎠ CPbO 2 c H + ⎝ RT ⎠
jPbO =
⎛ C s ⎞ν ⎛ ⎞ s
CPbO ⎛ ⎞
0.9Fη+ ⎟ 2 c H 0.9FλRs ⎜ −0.9Fη+ ⎟
⎟ 0.9FλRs exp ⎜
+
1 + 2Fkf ⎜⎜ sPbO
, ref ⎟ ⎜ RT ⎟ + 2 Fk b s, ref
exp ⎜ RT ⎟
⎝ CPbO ⎠ RT ⎝ ⎠ CPbO c ref+ RT
2 H ⎝ ⎠ (31)

Eqs. (30) and (31) differ from the classical Butler–Volmer values of λ, γ, and lt over a wide range.
equation in the denominator which represents increased over- The above explicit equations for reaction kinetics were solved along
potential required to provide for ohmic drop across deposits. These with the other model equations using COMSOL4.3 under galvanostatic
equations allow us to go from one limiting case to other. When the conditions (Nandanwar and Kumar, 2016b). Boundary layer grid was
resistance offered by the deposits is zero (Rs ¼0), the above used near electrode surfaces where stiff gradients of velocity and
equations become identical to those used earlier in the literature concentrations are expected. A triangular mesh was used in the rest of
(Eqs. (14)–(16)). When the resistance offered by deposits ap- the numerical domain. A second order accurate discretization scheme
proached a large value, due to the presence of large proportion of was used for concentrations and potential in the electrolyte domain.
PbO, the current density approaches zero locally. P1þP1 discretization (Gresho et al., 1999) was used for velocity and
The above two equations are still implicit as γ = jan /jPbO2 and pressure fields. Streamline diffusion and crosswind diffusion methods
λ = jan /jPbO appear on the right hand side, and require highly were used to stabilize numerical solution. Mixing during the relaxa-
iterative and intensive computations even for the simplified 2-D tion period was simulated by setting buoyant force F to zero in Eq. (8)
geometry (because the cell voltage needs to be first guessed at and by increasing diffusion coefficients artificially for all the species to
every time instant). To obtain explicit equations, we assumed the let the concentration gradients in the domain die out. Mass con-
values of γ and λ to be the same and equal to 2. Thus, the re- servation in this process was monitored.
sistance Rs offered by the deposits is computed based on the In summary, we have developed a model which accounts for
proportion of PbO and PbO2 present on it locally, but the over- natural convection, spatially non-uniform deposition/dissolution
M. Nandanwar, S. Kumar / Chemical Engineering Science 154 (2016) 61–71 67

Table 2 literature (Bode, 1977). As the molar volumes of PbO2 (25.5 cm3/mol)
Values of natural convection related parameters used for simulation of SLRFB. and PbO (23.4 cm3/mol) are nearly the same, we replaced volume
proportions (Aa and Ab) in Eq. (3) with the respective molar pro-
Parameter Value Parameter Value
portions. The approximation introduces less than 10% error in the
μ 0.001 Pa s ρ0 1169 kg/m3 value of effective conductivity (se).
β H+ 3.328  10  5 m3/mol β Pb2 + 2.573  10  4 m3/mol The model, with composition dependent conductivity of de-
W 0.015 m L 0.03 m posits, was used to simulate voltage vs. time profile for current
density of japp = 200 A/m2. The experimental measurements under
these conditions are reported earlier (Nandanwar and Kumar,
Table 3
2016b). The deposits formed at the end of the first charge cycle are
Values of parameters of conductivity model used for simulation of SLRFB.
highly conducting PbO2 on anode and Pb on cathode. The model
Parameter Value Parameter Value therefore makes identical predictions as those obtained with the
model of Nandanwar and Kumar (2016b) based on infinite con-
z 8 lt 250 (μm) ductivity of deposits. We will henceforth refer to the latter pre-
sa 1.25  10  12 (S/m) sb 5  105 (S/m)
dictions as those from the base model with no conductivity effects.
Since the predictions of the earlier model are in good agreement
on electrode surface, and composition dependent conductivity of with the measurements, we have not shown any comparisons for
deposits on anode. It includes in it, over and above the model of the first charge cycle. Fig. 3 instead shows model predictions for
Nandanwar and Kumar (2016b), the effect of finite conductivity of the first discharge cycle (bottom) and the second charge cycle
deposits on anode which varies from highly conducting to non- (top) for z¼8 (the number of contacts per particle). The corre-
conducting with evolving composition of deposits. The new model sponding set of predictions from the base model are also shown.
has coordination number z alone as an unknown parameter. The The figure shows that the two sets of predictions for the first
values of all the other parameters in expressions for Butler–Volmer discharge cycle are nearly identical for the most part, except to-
kinetics and Nernst potential are taken from the literature. Al- wards the end, when the rapidly changing conductivity of deposits
though the deposit thickness lt, λ, and γ can be calculated using the on anode makes a difference. The overprediction seen in the early
model, these are introduced as parameters. The composition of stages for both the models is similar to the overprediction seen in
deposits which causes orders of magnitude change in resistance is the conventional lead-acid battery, which is explained by the over-
determined exactly from the model. potential required for nucleation of new phase. We conjecture a
similar explanation for SLRFB.
The predictions of the two models for the second charge cycle
5. Results and discussion (Fig. 3, top) are quite different at short times. In the base model,
the battery starts to charge at the voltage required to oxidize PbO
The model equations were solved for the rectangular cell shown (Nandanwar and Kumar, 2014). In the early stages of the second
in Fig. 2 with no external flow. The parameter values used to simulate charge cycle, most of the transferred charge on anode comes from
the model are shown in Tables 1–3. The parameter values are the the conversion of PbO to PbO2 which occurs as at lower potentials.
same as those reported by Nandanwar and Kumar (2014) for flow- With depletion of PbO, the transferred charge starts to come from
through SLRFB, also used by Nandanwar and Kumar (2016b) to deposition of Pb2 þ as PbO2, and the battery starts to charge at
successfully simulate natural convection in battery cell. The values of higher voltage. The charging voltage predicted by the base model
expansion coefficients βi are measured experimentally by Nandan- thus slowly increases with time, from the beginning.
war and Kumar (2016b). The conductivity values are taken from the The incorporation of composition dependent conductivity of
deposits in the present model leaves anode at the end of the
previous discharge cycle with very poorly conducting deposits.
The battery cell therefore needs higher voltage in the next charge
cycle to start the charging process. The formation of conducting
PbO2 even in small amount increases the conductivity of deposits.
The charging voltage rapidly comes down and joins the charging
voltage for the base model at around 100 s. The model based ex-
planation for charge coup de fouet phenomenon is the poor con-
ductivity of deposits towards the end of the previous discharge
cycle, and rapid increase in conductivity of deposits early in the
next charge cycle.

5.1. Parameter sensitivity

The model consists of parameters z, lt, γ, and λ. The latter two


satisfy the constraints λ−1 + γ −1 = 1, and λ , γ > 1. The sensitivity of
model predictions to these parameter values needs to be ascer-
tained. Fig. 4(a) shows insensitivity of predictions to coordination
number z for the most part and a slight increase in discharge time
(and charge efficiency) with an increase in coordination number:
91.1% for z ¼6, 94.1% for z¼ 8, and 95.1% for z ¼10. The coordination
number z is visualized as the number of current carrying contacts
Fig. 3. A comparison of experimentally measured cell voltage vs. time for first a particle has with its neighbors. The typical values for z are: 4 for
discharge cycle (bottom) and second charge cycle (top) with the predictions ob-
tained using the base model which did not consider conductivity effects and the
diamond packing, 6 for simple cubic packing, 8 for body centered
present model which incorporates the effect of finite conductivity of deposits on packing, and 12 for face centered and hexagonal closed packing
electrodes. (Euler et al., 1980). The value of z for complex mixtures needs to be
68 M. Nandanwar, S. Kumar / Chemical Engineering Science 154 (2016) 61–71

To understand the insensitivity of model prediction to these


parameters, we calculated voltage drop across a 250 μm thick
deposit film as a function of percentage of PbO2 in it, for two
current densities using Eq. 21. The results presented in Fig. 5 show
that at moderate fraction of PbO2, the required voltage drop is less
than 0.1 mV even at a current density of 1000 A/m2, which is
negligibly small compared with the typical values of overpotential
in the range of 100 mV (Fig. 1). The required voltage drop increases
steeply to infinity for both the current densities as the fraction of
PbO2 in deposits approaches the critical limit of 25% PbO2. Over a
short range of composition, the resistivity Rs increases by a factor
of 1015 and the deposits change from being highly conducting to
non-conducting. The value of ΔVs is thus either very low (ON) or
very high (OFF), which is largely controlled by the value of Rs
(through se). The values of parameters γ, λ, and lt therefore have
very little effect on model predictions.
In view of the extreme variation of Rs, Eqs. (30) and (31) can be
further approximated to:
⎡ ⎛ 0.5Fη+ ⎞ ⎛ −0.5Fη+ ⎞ ⎤
i0 ⎢ exp ⎜ ⎟ − exp ⎜ ⎟⎥
⎣ ⎝ RT ⎠ ⎝ RT ⎠ ⎦
jPbO2 =
1 + R¯ s (32)

and
⎛ C s ⎞ν ⎛ ⎞ s
C PbO ⎛ ⎞
0.9Fη+ ⎟ 2 c H+ −0.9Fη+ ⎟
2Fkf ⎜⎜ PbO ⎟⎟ exp ⎜ − 2Fkb exp ⎜
s , ref ⎜ ⎟ s , ref ref ⎜ ⎟
⎝ CPbO ⎠ ⎝ RT ⎠ C PbO c +
2 H ⎝ RT ⎠
jPbO = ,
1 + R¯ s (33)

( (
where R̄s is the scaled resistance =Rs / RT /kf F2 . The predictions))
obtained using the above equations are also shown in Fig. 4(b). The
Fig. 4. Effect of parameter values on discharge characteristics of battery, as pre- good agreement seen among all the predictions establishes that
dicted by the present model. charge coup de fouet is largely controlled by the model used to
represent rapid decrease in conductivity of deposits with forma-
tion of PbO.

5.2. Distribution of current and solids on electrodes

The inclusion of finite conductivity of deposits greatly influ-


ences model predicted composition of deposits on anode. Fig. 6
shows concentration profiles of PbO2 and PbO residue at the end of
the first discharge cycle for the base model and the present model.
The base model predicts formation of a patch which has no PbO2
in it. The present model in comparison predicts the presence of
PbO2 everywhere on anode. The PbO2 and PbO concentrations in
the upper half are non-uniform—larger near the edge due to the

Fig. 5. Voltage drop across 250 μm thick film of deposit as a function of % PbO2 in
the deposits at two current densities.

obtained experimentally. Gandhi (2015) has used z¼ 8 in his


model of lead-acid battery. The measured efficiency of 92.0% in the
present case supports a z value in the range of 6–8. We have
therefore used z ¼8 in all the subsequent model predictions.
Simulations were next carried out to test sensitivity of the
model to values of γ (and λ). The results presented in Fig. 4(b) for
three widely different values of γ (10, 2, and 100/99) show that the
discharge profiles are identical for the three cases. A ten fold in-
Fig. 6. The predicted concentration profiles of PbO2 and PbO deposits (in mol/m2)
crease and decrease in deposit layer thickness lt also did not on anode at the end of first discharge cycle. The predictions are obtained using the
change the predicted discharge profiles. base model (without the conductivity effects included in it) and the present model.
M. Nandanwar, S. Kumar / Chemical Engineering Science 154 (2016) 61–71 69

Fig. 8. Movement of location of peak current with time during first discharge cycle
(bottom) and charge second charge cycle (top), as predicted by the present model.
Fig. 7. Variation of ratio [PbO2]/[PbO] (top) and associated conductivity of the
mixture (bottom) on anode at various times during discharge cycle.
through SLRFB. The dynamics of charging in the next cycle, in
comparison, is quite different and interesting. When the con-
edge effect (Nandanwar and Kumar, 2014), but the fraction of PbO2 ductivity effects are not taken into account, the entire electrode
in deposits is uniform at 0.25 (the ratio of PbO2 to PbO is 0.33), the surface participates in chemical reactions and contributes to the
value at which deposit conductivity becomes zero for z¼ 8. The total current from the beginning itself (Nandanwar and Kumar,
early termination of discharge cycle at the same cut-off voltage 2014, 2016b). With only a part of the anode remaining conducting
leaves significantly larger PbO2 residue on anode. The amount of at the end of the previous discharge cycle, the present model
PbO formed on anode is the same for the two models, because the predicts a peak in current density in the next charge cycle at t¼0
formation rate is high in the beginning. at the interface of the conducting and non-conducting regions on
The charge efficiency for the first charge-discharge cycle as anode, as shown in Fig. 8 (top). The peak moves up rapidly to the
predicted by the present model (94.1%) is therefore smaller than top (right in the figure) and reaches the upper edge of the elec-
that for the base model (97.4%), and closer to the experimentally trode in less than 100 s, to mark conversion of the entire insulating
measured value of 92.0%. Interestingly, the charge efficiency ob- part of anode to conducting one. The movement is faster initially.
tained for the present model is far larger than 75% (corresponding Conversely, in the discharge cycle, the peak movement is slower in
to 25% PbO2 that remains trapped in residue) because a substantial the beginning. The latter is explained in detail by Nandanwar and
part of PbO2 dissolves as Pb2 þ as well. Kumar (2014). The explanation for the former is provided below.
Fig. 7 (top) shows the variation of ratio of surface concentration The conversion of non-conducting zone to conducting zone
of PbO2 to PbO as a function of distance from the bottom of anode requires addition of conducting PbO2 to it. This can be achieved
at different times during discharge. The corresponding con- through deposition of Pb2 þ as PbO2 or conversion of local PbO to
ductivity profile of deposits is shown at the bottom. The con- PbO2. Both the reactions require conduction of electrons and H þ
ductivity of the solid deposits lies in the range of 105 S/m for ratio ions next to the reaction site for Butler–Volmer kinetics. The
[PbO2]/[PbO] greater than 0.33. It drops quite abruptly by 15 or- highest concentration of H þ is in the electrolyte rising up, past the
ders of magnitude for the value of ratio approaching 0.33. A por- end point of the conducting zone, because H þ ions are produced
tion of the electrode thus turns into insulator. The voltage drop on the conducting part of anode. For a fixed applied current,
seen earlier in Fig. 3 beyond 600 s is due to the increase in current smaller the area of the conducting part, larger is the current
density on the remaining part of anode to sustain galvanostatic density on it and higher is the concentration of H þ in the rising
discharge. The non-conducting part of the electrode still has a liquid past the end point of the conducting zone. The conversion of
significant amount of PbO2 trapped in it. Thus, although the parts non-conducting part to conducting one occurs at this point which
of the electrode are conducting and non-conducting and are also moves up with time. As the surface area of the conducting part
separated clearly, PbO2 and PbO residue are present all over the increases with time, the current density on it decreases and so
electrode at the end of the discharge cycle. It may not be possible does the concentration of H þ in the rising liquid past the end
to visually distinguish the two regions. This possibly explains the point of conducting zone. The expansion of the conducting zone
visually uniform texture/morphology of anode reported by Nan- therefore slows down with time.
danwar and Kumar (2016b), which could not be explained by their
model based on zero resistance of deposits. Both the models 5.3. Effect of the depth of discharge (DOD)
predict formation of a bare patch on cathode on the upper side,
which is validated by the measurements (Nandanwar and Kumar, It is known that the magnitude of charge coup de fouet in
2016b). conventional lead acid batteries is influenced by the depth of
The step decrease in conductivity of deposits makes local cur- discharge (DOD) in the previous cycle. In order to see if a SLRFB
rent density zero. Thus, towards the end of the first discharge shows similar behavior, we experimentally studied charging be-
cycle, a receding current peak moves towards the bottom (left in havior after discharging the battery in the previous cycle to dif-
the figure), as shown in Fig. 8 (bottom). The movement is similar ferent cut-off voltages. The experiments were conducted in the
what is predicted by Nandanwar and Kumar (2014) for flow- rectangular cell, described earlier, in the absence of any forced
70 M. Nandanwar, S. Kumar / Chemical Engineering Science 154 (2016) 61–71

Fig. 10. The model predicted effect of depth of discharge in previous discharge
cycle on the magnitude of charge coup de fouet in the second charge cycle.

charge cycle of 900 s, followed by 5 min of relaxation, followed by


discharge cycle with decreasing cut-off voltages, and then char-
ging. The cut-off voltages for the first, second, and third discharge
cycles were 1.35 V, 1.3 V, and 1.2 V respectively. The results ob-
tained are presented in Fig. 9(b). The figure shows similar trend as
before, with improved control on the long time response during
charging. These measurements establish that charge coup de fouet
effect is related to anode, and its magnitude increases with de-
crease in cut-off voltage in the previous cycle.
The present model was simulated with different cut-off voltages
to compare model predictions with observations. Fig. 10 shows the
observed increase in the magnitude of charge coup de fouet as cap-
tured by the model. The experimentally measured magnitudes of
charge coup de fouet are about 0.03 V and 0.05 V for cutoff voltages of
1.3 V and 1.2 V respectively. The model predicted values of 0.03 V and
0.08 V respectively are obtained without fitting any parameters. Gi-
ven the complex nature of the underlying phenomenon, the agree-
ment obtained between the predictions and the observations is
considered good. The experimentally observed steep rise in voltage
in a very short time interval in the beginning of the experiment is a
well known feature (Budevski et al., 2000; Guo and Searson, 2010)
Fig. 9. Effect of depth of discharge on magnitude of charge coup de fouet observed
in individual experiments when cut-off voltage for previous discharge cycle is and is not attempted by the present set of models. It is believed to
varied. arise due to the overpotential required for formation of nuclei of the
depositing material on foreign substrate.

convection. The battery was first charged for 900 s, followed by


five minutes of relaxation with gentle mixing. In the discharge 6. Conclusions
cycle, different cut-off voltages were used to terminate the cycle. A
five minute relaxation period was provided with gentle mixing, The charge coup de fouet phenomenon, in which charging vol-
before the onset of the next charge cycle. The results obtained for tage of a fully discharged battery goes through a minimum, is
three cut-off voltages (1.2 V, 1.3 V and 1.35 V) in separate experi- systematically investigated in this work in the context of the so-
ments are presented in Fig. 9(a). The figure shows that the mag- luble lead redox flow battery. The experiments show that an in-
nitude of charge coup de fouet (difference between the peak and crease in the depth of discharge, effected by decreasing the cut-off
the plateau voltage) increases with a decrease in cut-off voltage in voltage for termination of discharge cycle, increases the magni-
the previous cycle, in agreement with the reported observations tude of charge coup de fouet. The models available in the literature
for the conventional lead-acid battery (Delaille et al., 2006). The in which the conductivity of deposits is assumed to be infinitely
magnitudes of charge coup de fouet are 0.047 V, 0.026 V, and 0.01 V large fail to explain the observed phenomenon. A new model is
for cut-off voltages of 1.2 V, 1.3 V, and 1.35 V respectively in the developed which incorporates in it variation of conductivity of
previous cycle. deposits with composition, from highly conducting to non-con-
The measurements do not show any regular pattern in charging ducting when conducting pathways cease to percolate beyond a
voltage at later times, however. This could be caused by varying critical volume fraction of non-conducting phase. The orders of
external contact resistance when new connections are made for magnitude change in conductivity of a thin layer of deposits on
separate experiments. In order to establish that the observed effect conducting graphite electrodes is used to develop an approxima-
is due to anode only, a pre-deposited cathode (for 3600 s at tion that permits model predictions to be obtained without de-
200 A/m2) was used. It also eliminated any complications arising termining voltage profile in the composite solid phase. The model
out of partial baring of cathode during the experiments. A single explains the observed charge coup de phenomenon and its var-
experiment without removing the contacts was performed next: a iation with depth-of-discharge well.
M. Nandanwar, S. Kumar / Chemical Engineering Science 154 (2016) 61–71 71

The model is general and applicable to situations involving a Guo, Lian, Searson, Peter C., 2010. On the influence of the nucleation overpotential
conducting electrode turning itself into a non-conducting one due on island growth in electrodeposition. Electrochim. Acta 55 (13), 4086–4091.
Hazza, A., Pletcher, D., Wills, R., 2004. A novel flow battery—A lead acid battery
to the formation of an insulating phase, such as in the conven- based on an electrolyte with soluble lead(II): I. Preliminary studies. Phys. Chem.
tional lead-acid battery. Chem. Phys. 6, 1773–1778.
Hazza, A., Pletcher, D., Wills, R., 2005. A novel flow battery—A lead acid battery
based on an electrolyte with soluble lead(II): IV. The influence of additives. J.
Power Sources 149, 103–111.
Acknowledgments Li, X., Pletcher, D., Walsh, F.C., 2009. A novel flow battery—A lead acid battery based
on an electrolyte with soluble lead(II): VII. Further studies of the lead dioxide
positive electrode lead acid batterybased on an electrolyte with soluble lead(II)
The authors are grateful to Prof. K.S. Gandhi for many useful Part VII. further studies of the lead dioxidepositive electrode. Electrochim. Acta
discussions. The authors also acknowledge financial support re- 54, 4688–4695.
ceived from the Department of Science and Technology, New Nandanwar, M.N., Kumar, S., 2014. Modelling of effect of non-uniform current
density on the performance of soluble lead redox flow batteries. J. Electrochem.
Delhi, India through its IRHPA scheme. Soc. 161 (10), A1602–A1610.
Nandanwar, M.N., Kumar, S., 2016a. A soluble lead redox flow battery with no
externalflow during charging/discharging. J. Power Sources, Submitted for
publication.
Appendix A. Supplementary data
Nandanwar, M.N., Kumar, S., 2016b. Soluble lead redox (natural convection) flow-
battery. J. Electrochem. Soc., Submitted for publication
Supplementary data associated with this article can be found in Newman, J., Thomas-Alyea, K.E., 2004. Electrochemical Systems, 3rd edition. John
Wiley & Sons, Hoboken, N.J., p. 672.
the online version at http://dx.doi.org/10.1016/j.ces.2016.07.001.
Oury, A., Kirchev, A., Bultel, Y., 2014. A numerical model for a soluble lead-acid flow
battery comprising a three-dimensional honeycomb-shaped positive electrode.
J. Power Sources 246, 703–718.
References Oury, A., Kirchev, A., Bultel, Y., Chainet, E., 2012. PbO2/Pb2 þ cycling in methane-
sulfonic acid and mechanisms associated for soluble lead-acid flow battery
applications. Electrochim. Acta 71, 140–149.
Banerjee, A., Saha, D., Guru Rao, T.N., Shukla, A.K., 2013. A soluble-lead redox flow Pascoe, Phillip E., Anbuky, Adnan H., 2002. The behaviour of the coup de fouet of
battery with corrugated graphite sheet and reticulated vitreous carbon as po- valve-regulated lead-acid batteries. J. Power Sources 111 (2), 304–319.
sitive and negative current collectors. Bull. Mater. Sci. 36 (1), 163–170. Pascoe, P.E., Anbuky, A.H., 1999. Estimation of vrls battery capacity using the ana-
Bates, A., Mukerjee, S., Lee, S.C., Lee, D., Park, S., 2014. An analytical study of a lead- lysis of the coup de fouet region. In: Proceedings of the Telecommunication
acid flow battery as an energy storage system. J. Power Sources 249, 207–218. Energy Conference, 1999. INTELEC'99. The 21st International, IEEE, pp. 114–122.
Beck, F., Bhn, H., 1975. Bleidioxid als lsungselektrode. Berichte der Bunsenge- Pascoe, P.E., Sirisena, H., Anbuky, A.H., 2002. Coup de fouet based VRLA battery
sellschaftfr physikalische Chemie, vol. 79(3), pp. 233–244. capacity estimation. In: Proceedings of the Electronic Design, Test and Appli-
Berndt, D., Voss, E., 1965. The voltage characteristics of a lead-acid cell during cations. The First IEEE International Workshop on, IEEE, pp. 149–153.
charge and discharge. In: Collins, D.H. (Ed.), Batteries 2: Research and devel- Pletcher, D., Wills, R., 2004. A novel flow battery—A lead acid battery based on
opment in non-mechanical electrical power sources. In: Proceedings of the 4th anelectrolyte with solublelead(II): II. Flow cell studies. Phys. Chem. Chem. Phys.
International Symposium held at Brighton, September 1964, Pergamon Press, 6, 1779–1785.
pp. 17–27. Pletcher, D., Wills, R., 2005. A novel flow battery—A lead acid battery based on an
Bode, H., 1977. Lead-Acid Batteries', first edition. J. Wiley& Sons, New York electrolyte with soluble lead(II): III. The influence of conditions on battery
(translated by RJ Brodd and KV Kordesch). performance. J. Power Sources 149, 96–102.
Bose, C.S., Laman, F.C., 2000. Battery state of health estimation through coup de Pletcher, D., Zhou, H., Kear, G., Low, C.T.J., Walsh, F.C., Wills, R., 2008a. A novel flow
fouet. In: Preceedings of the Telecommunications Energy Conference, 2000. battery—A lead-acid battery based on an electrolyte with soluble lead(II): VI.
INTELEC. Twenty-second International, IEEE, pp. 597–601. Studies of the lead dioxide positive electrode. J. Power Sources 180, 621–629.
Budevski, E., Staikov, G., Lorenz, W.J., 2000. Electrocrystallization: Nucleation and Pletcher, D., Zhou, H., Kear, G., Low, C.T.J., Walsh, F.C., Wills, R., 2008. A novel flow
growth phenomena. Electrochim. Acta 45 (15), 2559–2574. battery—A lead-acid battery based on an electrolyte with soluble lead(II):
Collins, J., Kear, G., Li, X., Low, C.T.J., Pletcher, D., Tangirala, R., Stratton-Campbell, D., Studies of the lead negative electrode. J. Power Sources 180, 630–634.
Walsh, F.C., Zhang, C., 2010. A novel flow battery—A lead acid battery based on Schmickler, W., Santos, E., 2010. Interfacial Electrochemistry, 2nd edition. Springer-
an electrolyte with soluble lead(II): VIII. The cycling of a 10 cm  10 cm flow Verlag, Heidelberg, pp. 145–150.
cell. J. Power Sources 195, 1731–1738. Shah, A.A., Li, X., Wills, R.G.A., Walsh, F.C., 2010. A mathematical model for the
Delaille, A., Perrin, M., Huet, F., Hernout, L., 2006. Study of the coup de fouet of lead- soluble lead-acid flow battery. J. Electrochem. Soc. 157 (5), A589–A599.
acid cells as a function of their state-of-charge and state-of-health. J. Power Soudijn, M., 2012. Proton Transport in Aqueous Ionic Solutions. University of Am-
Sources 158 (2), 1019–1028. sterdam, Master’s thesis, 53-54.
Euler, K.J., Kirchhof, R., Metzendorf, H., 1980. Electronic conductivity of battery Telfah, A., Majer, G., Kreuer, K.D., Schuster, M., Maier, J., 2010. Formation andmo-
active masses–a limiting factor? J. Power Sources 5, 255–262. bility of protonic charge carriers in methyl sulfonic acidwater mixtures: a
Gandhi, K.S., 2015. Role of electrical resistance of electrodes in modeling of dis- modelfor sulfonic acid based ionomers at low degree of hydration. Solid State
charging and charging of flooded lead-acid batteries. J. Power Sources 277, Ion. 181, 461–465.
124–130. Verde, M.G., Carroll, K.J., Wang, Z., Sathrum, A., Meng, Y.S., 2013. Achieving high
Gebhart, B., Pera, L., 1971. The nature of vertical natural convection flows resulting efficiency and cyclability in inexpensive soluble lead flow batteries. Energy
from the combined buoyancy effects of thermal and mass diffusion. Int. J. Heat. Environ. Sci. 6, 1573–1581.
Mass. Transf. 14, 2025–2050. Zhang, C.P., Sharkh, S.M., Li, X., Walsh, F.C., Zhang, C.N., Jiang, J.C., 2011. The per-
Gresho, P.M., Sani, R.L., Engelman, M.S., 1999. Incompressible Flow and the Finite formance of a soluble lead-acid flow battery and its comparison to a static lead-
Element Method. John Wiley and Sons, Chichester, pp. 457–470. acid battery. Energy Convers. Manag. 52, 3391–3398.

You might also like