Journal of Power Sources: Praphulla Rao, Sreenivas Jayanti

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Power Sources 482 (2021) 228988

Contents lists available at ScienceDirect

Journal of Power Sources


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

Influence of electrode design parameters on the performance of vanadium


redox flow battery cells at low temperatures
Praphulla Rao , Sreenivas Jayanti *
Department of Chemical Engineering and DST-Solar Energy Harnessing Centre IIT Madras, Chennai, 600036, India

H I G H L I G H T S

• Experimental studies have been conducted in a flow battery cell of 426 cm2 area.
• Influence of electrode design parameters has been studied at 25, 10 and − 10 ◦ C.
• Results show 50% or more loss of discharge capacity at − 10 ◦ C.
• Loss of performance is attributed to increased ohmic and charge transfer resistance.
• Thinner electrode and increased compression improve low temperature performance.

A R T I C L E I N F O A B S T R A C T

Keywords: Performance of electrochemical batteries suffers at cold temperatures. In this paper, we report on an experi­
Vanadium redox flow batteries mental investigation of sensitivity of the vanadium redox flow battery (VRFB) to the cell operating temperature.
Low temperature Experiments have been carried out on a VRFB cell with an active area of 426 cm2 at cell temperatures of 25, +10
Polarization
and − 10 ◦ C. Measured parameters include cycle efficiency and discharge capacity at three current densities and
Energy efficiency
Impedance spectroscopy
two flow rates, peak power and cell impedance obtained through electrochemical impedance spectroscopy (EIS)
Electrode design studies. The influence of electrode compression and electrode thickness has also been studied through
comparative experimental protocols. Results indicate significant loss of discharge capacity and cell efficiency at
low temperatures; this can be attributed to substantial increase in ohmic resistance and even higher increase in
charge transfer resistance of the cell. Two methods known to reduce cell resistance, namely, increasing
compression ratio of the electrode and using thinner electrode, have led to restoring much of the lost discharge
capacity and energy efficiency. Optimizing the electrode parameters such as its compression, permeability and
thickness for industrial scale cells, together with possible use of electrocatalysts, seems to be a promising way of
improving low temperature performance of VRFB cells.

1. Introduction everywhere and are seen as the best of the class, they too suffer from
many problems when operated at low temperatures. These include slow
Depletion of fossil fuels and concerns of global warming call for the charging, faster ageing, poor capacity and energy [1–3]. In their review
immediate and widespread of harnessing solar energy and wind energy article, Zhu et al. [1] report a sharp drop in power and energy capa­
worldwide. Rechargeable batteries play a vital role in effective utiliza­ bilities when LIBs are operated at sub-zero temperatures, and most of the
tion of solar and wind power. Since a large fraction of the population energy capacity is lost when operated at − 10 ◦ C or lower due to high
lives in cold climates for part of the year, sensitivity of the battery charge transfer resistance and lowered mass transport leading to
performance to ambient temperature is important. Extreme cold or high dendrite formation and its attendant risks. In very cold climates, LIBs
heat reduces the performance of batteries. Most of the batteries used in can only be operated by heating up the cell, which limits their suitability
present days suffer at cold conditions and require active temperature for large scale applications. Significant amount of research is going on
management system. Even though lithium ion batteries (LIB) are used redox flow batteries (RFBs) for use in cold temperature conditions.

* Corresponding author.
E-mail address: sjayanti@iitm.ac.in (S. Jayanti).

https://doi.org/10.1016/j.jpowsour.2020.228988
Received 27 April 2020; Received in revised form 31 August 2020; Accepted 22 September 2020
Available online 3 October 2020
0378-7753/© 2020 Elsevier B.V. All rights reserved.
P. Rao and S. Jayanti Journal of Power Sources 482 (2021) 228988

However, RFBs are typically used above 10 ◦ C. As compared to aqueous evaluated the suitability of the different cation and anion exchange
RFBs, non-aqueous RFBs can use electrolytes that can operate at lower membranes for the practical application of the VRFB over wide tem­
temperature and with better cell voltage, thereby making them appro­ perature range from − 20 to 50 ◦ C. They evaluated the performance of
priate for use in cold regions. New chemistries such as non-aqueous single cell VRFB for different types of membranes. They reported that
organic RFBs, aqueous zinc manganese-dioxide battery are being increase of temperature accelerates the vanadium crossover, but also
developed to find the suitability of operating at sub-zero temperature. improves the transport of ions through the membrane, which leads to
But these studies are still in nascent stages of research. Besides, increase of membrane conductivity and electrolyte conductivity. Nafion
non-aqueous organic RFBs still face challenges such as the low solubility 115, 212, SPEEK and APS membranes can be operated stably at − 20 to
of organic active materials, restricted solution stability, and capacity 50 ◦ C. Anion exchange membranes were not suitable for VRFB as they
decay [4,5]. would accelerate the precipitation of the VO+ 2 electrolyte in the long run.
Vanadium redox flow batteries (VRFB) have gained huge attention During a long run test over a broad temperature range from − 15 ◦ C to
over the last decade due to safe operation, long cycling potential, 50 ◦ C, Nafion 115 membrane showed the lowest capacity fading rate. Yu
chemical stability, recyclability of electrolyte etc. [6]. Plenty of research et al. [14] developed ultrathin Nafion XL membrane by reinforcing
work has been carried out over the past several years to improve the microporous PTFE middle layer between the two dense Nafion outer
energy density and power density of the VRFB by working on electrolyte layers. This membrane showed excellent long-term cycling stability over
and electrode materials, and membranes respectively. There is 2200 cycles with stable and good electrochemical performance due to
comparatively little amount of work on VRFB performance on at low enhanced the fast proton transport through the membrane and reduced
temperatures, and much of it has been reported in the last five years. Vanadium cross-over. Kim and Park [15] found from their 25 cm2 VRFB
Xiao et al. [7] demonstrated the effect of temperature from − 35 to cell study that the exchange current density increased from 38.8 to 49.1
+50 ◦ C on the stability, physicochemical and electrochemical properties A m− 2 as the cell temperature increased from +5 ◦ C to 45 ◦ C. The open
of vanadium electrolyte. They found that V2+, V3+, V3.5+ electrolytes circuit voltage was inversely proportional to the flow rate at tempera­
precipitate out at − 35 ◦ C at different time intervals. VO+2 electrolyte will ture of +5 ◦ C and directly proportional to the flow rate from 25 ◦ C to
precipitate out above 35 ◦ C. Conductivity of the vanadium electrolyte 45 ◦ C.
increases with increase in temperature but viscosity decreases. Diffu­ Most of the literature related to VRFB applications at low tempera­
sivity of the positive and negative electrolytes was estimated through ture attributes the degradation in performance to an increase in the
the cyclic voltammetry (CV) and the value decreased four folds from resistance for the reaction kinetics. One strategy to reduce the resistance
30 ◦ C to − 10 ◦ C. Electrochemical impedance spectroscopy (EIS) was for reaction kinetics is using electrocatalysts to increase active sites in
used to estimate ohmic and charge transfer resistance of positive and electrode. Surface modification of the electrode is commonly done using
negative electrolytes. It was found that charge transfer resistance was thermal or acid treatment. Significant improvement in power density is
seven times higher at − 10 ◦ C compared to that at 30 ◦ C for both elec­ shown by using electrocatalyst at low temperature in literature. Carbon
trolytes. But the increase in ohmic resistance was only two to three times nano materials are mostly used in non-metallic type of electrocatalysts
higher over the same temperature range. Xi et al. [8] evaluated the which are easy to fabricate but get detached from the surface during
performance of the VRFB (5 cm × 5 cm) single cell over the temperature prolonged usage. Wu et al. [16] found that nitrogen-doped carbon
range of − 20 to 50 ◦ C in the current density of 40–200 mA.cm− 2. They nanosphere graphite felt shows good stability, durability and electro­
found that the cell could be operated at a current density of 100 mA. chemical performance over the temperature range from − 15 to 50 ◦ C.
cm− 2 over the range of − 20 to 50 ◦ C with energy efficiencies (EE) of 65 Yu et al. [17] demonstrated the outstanding performance of
and 78% at the two extreme temperatures, respectively. Evaluation of phosphorous-doped graphite felt with respect to high rate capacity and
cell resistance was done through EIS at different temperatures in the broad temperature adaptability in the range from − 20 to 60 ◦ C. Some of
initial and after pre-charge state. Charge transfer resistance was found to the metal based catalysts like Au, Ir, Pt show good performance but add
be 50% higher at initial state than after the pre-charge state at − 20 ◦ C. high cost to the system. In recent years, researchers reported that some
Pan et al. [9] studied the performance of the VRFB single cell (approx­ of the cheap and non-toxic metals are showing promising results to in­
imately 10 cm × 10 cm) and a stack with five cells with an electrode area crease the power density of VRFB. Liu et al. [18] used low cost highly
of 250 cm2 under controlled environment to maintain the temperature active bismuth electrocatalyst to modify the graphite felt. This electrode
in the range of − 10 to 30 ◦ C. They found that the voltage efficiency was used on the negative side to inhibit hydrogen evolution and was
decreased from 87% to 83.5% as the environment temperature dropped found to exhibit the excellent performance at high current density of
from 20 ◦ C to 5 ◦ C at 60 mA cm− 2. Capacities at 5 ◦ C and 10 ◦ C were only 400 mA cm− 2. It also showed broad temperature adaptability in the
slightly less than those at 20 ◦ C. During operation, the static resistance of range from − 10 ◦ C and 50 ◦ C. Yu et al. [19] used high active
the stack was about 38.1 mΩ on average at − 2 ◦ C. According to CV ceria-zirconia coated graphite felt electrode and high proton selective
results, anodic and cathodic peak current decreased with temperature. SPEEK membrane to obtain efficient and low cost VRFB. This electro­
Also, the system became more irreversible at low temperatures. Wang catalyst enhanced the voltage efficiency of the VRFB significantly by
et al. [10] studied the static stability of the vanadium electrolytes with increasing the wettability and activity of the electrode. CexZr1-­
different total vanadium concentrations (0.4–2.2 M) and various sulfuric xO2/GF-SPEEK combination showed excellent efficiency and durability.
acid concentrations (1.5–3.0 M) over a wide temperature range of − 35 Mehboob et al. [20] used SnO2 nanoparticle decorated graphite felt as
to 60 ◦ C. They found that electrolyte was stable for − 25 to 60 ◦ C. VRFB the electrode for VRFB, which showed 77% EE at current density 150
single cell study was performed with 1.5 M vanadium in 2.0 M total mAcm− 2.
sulfate in the temperature range from − 20 to 50 ◦ C. According to the Another way of improving reaction kinetics is through changes of
results, VRFB with the selected electrolyte could operate steadily for electrode properties. Specifically, recent studies [21–23] of electrode
more than 150 cycles over the temperature interval of − 20 ◦ C–60 ◦ C. thickness and compression have shown good improvement in the per­
Research work has also been reported on materials to improve the formance of VRFB at room temperature. Several studies reported that
performance of VRFB at low temperatures. Li et al. [11] used vanadium compression of graphite felt electrode was helpful by decreasing the
electrolyte and mixed sulfate chloride acid system in VRFB to ensure the ohmic resistance of the cell. But increase in compression leads to
stability of the electrolyte over the temperature range of − 5 to 50 ◦ C. decrease in the porosity of the electrode which may hinder electrolyte
Their results showed that energy efficiency increased with increase in flow in the cell. Gundlapalli and Jayanti [23] have studied the perfor­
temperature from 80% at 0 ◦ C to 90% at 40 ◦ C. A team from Pacific mance of 426 cm2 and 936 cm2 single cell VRFB with compressions of
Northwest National Laboratory developed mixed acid electrolyte which 20, 35, 50 and 70%. They found that at higher current densities, 35–50%
could work with 70% increase in capacity up to − 5 ◦ C [12]. Xi et al. [13] compression would be optimal for both the sizes at room temperature.

2
P. Rao and S. Jayanti Journal of Power Sources 482 (2021) 228988

Compression beyond 50% would lead to high mass transfer resistance choosing a gasket of 4 mm thickness, 35% compression of the 6 mm-
due to lessened permeability. Kumar and Jayanti [24] have performed a thick electrode was achieved. Similar cell construction was used in our
study on effect of thickness of electrode on 100 and 400 cm2 single cell earlier experimental studies [23–25].
VRFB and found improved performance at 6 mm thickness compared to The electrolytes were prepared in the following way. Vanadium
that at 3 mm or 9 mm. oxysulfate crystals with 99.5% wt. purity were procured from Noah
Much of the research work on low temperature VRFB cell has been Technologies to prepare the electrolyte solution. Thermogravimetric
done with small cells (active area <100 cm2). Typically, these cells are analysis of the as-received sample showed it to contain 2.3 mol of water
operated at very high electrolyte circulation rates compared to indus­ per mole of VOSO4. Required amount of 5 M H2SO4 and demineralized
trial size cells (typically in the range of 800–1600 cm2 active area) where water were added to make a 1.7 M VOSO4 solution containing vanadium
the cells are operated at area-specific flow rates [23] < 1 ml cm− 2. atom in the oxidation state of +4 in the form of VO2+ ions. The solution
min− 1. Under these conditions, mass transfer (concentration) polariza­ was then double-charged to a voltage of 1.8 V to produce electrolytes
tion is expected to influence the discharge capacity of the cell. Due to with mixtures of VO2+ and VO+ 2 ions on one side and V
2+
and V3+ ions
reduced diffusivities of active species at low temperatures, this aspect on the other side. From the current-time graph, the initial state of charge
becomes more important at low temperature operations. The objective (SoC) was estimated to be 75%. In the experiments, 340 ml of electrolyte
of the present work is to establish the sensitivity of the low temperature was taken on each side. In order to regulate the cell temperature, both
performance of VRFB cells to electrode parameters. To this end, sys­ the VRFB cell and the electrolyte solutions were kept inside an envi­
tematic electrochemical studies have been conducted on a single cell ronmental test chamber (Scientific Innovations, − 50 to 150 ◦ C) as
VRFB with active area of 426 cm2 at different compression ratios (CR), shown in Fig. 1. Peristaltic pumps used for the circulation of the elec­
current densities and flow rates at cell temperatures of 25, 10 and trolytes were kept outside the chamber and the tubing connecting the
− 10 ◦ C. The studies include the measurement of capacity, columbic, cell and the tanks was thermally insulated. Experiments were conducted
voltaic and energy efficiencies and estimation of ohmic and charge by keeping the set up at the target temperature for at least for 30 min
transfer resistances for a reference case. Additional, comparative ex­ prior to the start. The electrolyte temperature going into the cell was
periments have been conducted for a different electrode thickness and monitored using pre-calibrated thermocouples and was found to vary
for different compression ratios. Details of the studies and the results within 1 ◦ C of the target temperature.
obtained are discussed below. Silicone gaskets with different thicknesses were used to provide the
required compression for the electrode of the cell. In the present study,
2. Materials and methods compression ratio (CR) is defined as (1-tc/tu) x 100%, where tc and tu are
the thickness of the compressed and the uncompressed electrode,
Single cell VRFB was fabricated with Sigracell graphite felt electrode respectively. Since the silicone gasket was quite hard and not easily
(with active area of 426 cm2), SGL graphite plate with serpentine flow deformable, the thickness of the gasket determines the effective thick­
field, Nafion 117 membrane and copper current collectors. The elec­ ness of the compressed felt. By choosing gasket thicknesses appropri­
trode was placed on the graphite plate covering the active area; the ately, cells with compression ratios of 17, 35 and 50% with 6 mm
membrane was placed on top of it to make one half of the cell. The electrode and only 35% compression ratio with an electrode thickness of
second electrode and the second graphite plate corresponding to the 4.6 mm were assembled. For each of these, charge-discharge studies
other half cell were placed in a symmetrical way over the membrane. were carried out at cell temperatures of 25 ◦ C, 10 ◦ C and − 10 ◦ C and at
Each graphite plate had a single serpentine flow field of 4 mm width, 3 current densities 40, 60 and 80 mA.cm− 2 for five cycles at each condi­
mm depth and 3 mm rib width grooved using a computer-controlled tion over a pre-set voltage limit of 1.8 V during charging and 0.8 V
milling machine. Silicone gasket of appropriate thickness to obtain the during discharging (Although there is suggestion in the literature [26,
target compression ratio was cut into shape and was placed around the 27] that accelerated degradation of the cell can occur at cell voltages of
electrode on each side. The membrane was the largest in size and pro­ 1.8 V due to CO2 generation on the positive side, especially at very high
truded on all the four sides of the rectangular cell by about 2 cm so as to SoCs [28], leading to significant capacity fade, no such drop in energy
prevent mixing of the anolyte and the catholyte. Copper plates, placed in capacity has been observed in our earlier studies [23,29] of over 100
contact with the graphite plate on each side, were used as current col­ successive charge-discharge cycles in the same voltage limits. The ca­
lectors. Aluminium end plates with multiple bolt and nut pairs were used pacity loss in these studies was only of the order of 0.25% per cycle in
to tighten the entire set-up so as to ensure a leak-proof arrangement. By cells of size range of 100–900 cm2 and the coulombic efficiency was

Fig. 1. Photograph of experimental set up used for low temperature VRFB studies: (a) the environmental test chamber with potentiostat and pump arrangement
outside the chamber, and (b) the positioning of the cell (below) and the electrolyte tanks (above) inside the chamber.

3
P. Rao and S. Jayanti Journal of Power Sources 482 (2021) 228988

maintained at 97% or higher over this voltage range.). For each case 3.1. Effect of cell temperature on performance for the baseline case
corresponding to a combination of compression ratio/current densi­
ty/operating temperature, data were collected for two volumetric flow The sensitivity of the cell performance to cell operating temperature
rates of the electrolytes. The flow rate can be expressed in terms of the can be captured through the polarization curve during charge-discharge
non-dimensional stoichiometric flow factor (SF) defined as follows: cycling studies. Results obtained at constant current cycling of 60 mA
cm− 2 are shown in Fig. 2a and b for the low flow rate (SF50 of 6.6) and
Q I
SF = where Qth = ( ) (1) the high flow rate (SF50 of 13.2) condition, respectively. The data are the
Qth n F C SoCref
averaged values of cycles 2, 3 and 4 of five consecutive cycles for which
measurements were taken. Important features of the polarization curve
where Iis the current, n is the number of electrons transferred in the
under constant current mode are the total discharge (or charge) capac­
stoichiometric reaction, Fis the Faraday constant, C is concentration of
ity, the voltage gap between the charging and the discharging parts and
electrolyte, SoCref is a reference state of charge, which in the present case
the onset of mass transfer (concentration polarization) in discharging (or
is taken as 50%. Thus, the flow factor at 50% reference SoC can be
charging) which is evidenced by a highly non-linear variation of the cell
expressed as
voltage with change in capacity. With reference to Fig. 2a, the part of the
n F C( SoC50 ) curve above a voltage of about 1.4 V is the charging part and the part
SF50 = (2)
I below that is the discharging part. For the discharging part of the curve,
the abscissa should be interpreted as the amount of discharged capacity
It may be noted that in a typical charge-discharge test at constant
and not as the remaining capacity. As a general observation, one sees
current density and constant circulation rate, the stoichiometric flow
that the charging and the discharging capacities are very close in all the
factor changes as charging or discharging progresses. If the actual SoC is
cases. As the cell temperature decreases, one sees a widening voltage gap
less than 50%, then the effective SF will be less than the nominal value
between charging and discharging parts of the polarization curves. Thus,
given by eq. (2). Using 50% SoC as the reference, the effective stoi­
the upper limit of 1.8 V is attained well before the onset of mass transfer
chiometric flow factor, SFeff is given by
polarization in the case of charging. However, the discharge curve is
SFeff = 2 SF50 (SoC) (3) clearly marked by a well-defined mass transfer polarization part while
this is absent in the charging part. Thus, the capacity here does not
Thus, towards the end of discharge of a constant circulation rate, SFeff
reflect the true capacity of the solution which can be achieved only for
may drop down to values much lower than the nominal value and
extremely low charging current densities. Under cycling conditions, it is
thereby hasten the onset of mass transfer polarization if the circulation
the discharging capacity that determines the overall capacity of the cell.
rate is low. In view of this, experiments have been conducted at circu­
Similarly, the charged or discharged energy is the integral of the product
lation rates in the range of 82 and 328 mL min− 1 which correspond,
of cell voltage and the current density over the time period of charging
respectively, to SF50 factor of 6.6 at a current density of 40 mA cm− 2 and
or discharging, as the case may be. In the polarization curves given in
to 13.2 at a current density of 80 mA cm− 2.
Fig. 2a, this amounts to the area under curve of the voltage vs capacity
In order to find the effect of thickness of electrode on the electro­
curve. Since the latter is limited by the discharge capacity, the energy
chemical behaviour of the cell, another cell was assembled with 4.6 mm
that can be stored in and drawn from a cell under cyclic conditions is
thick SGL carbon felt with 35% CR. Charge-discharge studies were
also determined by the polarization under discharge conditions.
carried out as described earlier at 25, 10 and − 10 ◦ C.
Comparing the polarization curves from different cell temperatures
Electrochemical impedance spectroscopy (EIS) was performed at 10
in terms of the above-mentioned features, one can see from Fig. 2 that
mV amplitude with frequency range from 10 mHz to 10 MHz at the
cell temperature has a significant effect on the discharge capacity,
temperatures at 25, 10 and − 10 ◦ C before beginning the charge-
especially when operating at − 10 ◦ C. For the high flow rate case
discharge cycles. EIS was SF50 performed for the cells with CR 35%
(Fig. 2a), there is about 15% drop in the discharge capacity when the cell
and at the electrolyte circulation rate corresponding to SF50 of 13.2 for a
temperature is brought down to +10 ◦ C from 25 ◦ C, and close to 50%
current density of 60 mA.cm− 2.
drop when the temperature is brought down to − 10 ◦ C. For the low flow
The power density curve at a given cell temperature was obtained as
rate case (Fig. 2b), the discharge capacity decreases in all the three cases.
follows. Once the target cell temperature was achieved through circu­
The loss of discharge capacity is more than 60% for the low flow rate
lation of the cold electrolytes inside the environmental test chamber, the
case at − 10 ◦ C which shows that slightly beneficial effect of higher
cell was charged up to 1.8 V at 60 mA cm− 2 at SF50 of 13.2. Then the cell
circulation rate on discharge capacity. It is to be noted that the loss of
was discharged in a stepwise manner by increasing current density in
capacity should be attributed to the high overpotential in the cell
steps of magnitude of 20 mA cm− 2. In each step, current density was
coupled with the onset of mass transfer polarization when operating at
maintained constant for 30 s, and the average power over this period
low temperatures. This is also evidenced by the high coulombic effi­
was used to indicate the discharge power under those conditions.
ciency (defined as the ratio of the discharge capacity to the charge ca­
pacity), which is in excess of 95% in all cases. This high value is
3. Result & discussion
attributed to the fact that this too is calculated based only on the energy
charged and discharged in the 2nd, 3rd and 4th cycles.
A large number of experiments, amounting to 54 different conditions
Further evidence of the influence of overpotential and mass transfer
covering three cell temperatures, one high and one low flow rate con­
polarization on the discharge capacity is brought out in Fig. 3 where the
dition, three current densities and three compression ratios have been
polarization curves are shown for a constant current charge/discharge of
for a cell with 6 mm thick electrode felt. In addition, 18 data sets
40 mA cm− 2 (Figs. 3a) and 80 mA cm− 2 (Fig. 3b) for the three cell
covering the same cell temperature, flow rate and current density con­
temperatures for the same circulation rate as in Fig. 2a. The effective
ditions, were obtained for a felt thickness of 4.6 mm at a compression
SF50 is 19.8 for the current density of 40 mA cm− 2 and 9.9 for the current
ratio of 35%. For both felt thickness cases, power density and EIS
density of 80 mA cm− 2. One can see that, when compared with data
measurements were carried out at three cell temperatures for a fixed
shown in Fig. 2a, the discharging capacity and hence the charging ca­
compression ratio of 35%. The results from these are presented below in
pacity has increased in all the cases in Fig. 3a and has decreased in all the
a structured manner.
cases in Fig. 3b. While the decrease is rather small (about 15–20%)
between cell temperatures of 25 and 10 ◦ C, very little charging and
discharging has been possible for the high current case at − 10 ◦ C. From
Fig. 3b, it is evident that mass transfer polarization sets in immediately

4
P. Rao and S. Jayanti Journal of Power Sources 482 (2021) 228988

Fig. 2. Effect of cell temperature on polarization curves during charge-discharge test for the baseline case of electrode of 6 mm thickness, compression ratio of 35%
and a constant current density of 60 mA cm− 2:(a) high circulation rate condition of SF50 of 13.2 and (b) low circulation rate condition of SF50 of 6.6.

in this case, presumably because low mass diffusivity of the reactant contributing to the cell overpotential, EIS studies have been conducted
species. at the three cell temperatures and the results are summarized in Fig. 5 in
Fig. 4 shows the useful parameter of power density of the cell at three the form of Nyquist plots. Usually, an electrochemical process encoun­
operating temperatures. The power density curve has been obtained by ters three types of resistances, namely, ohmic, charge transfer and mass
monitoring the cell power while increasing the current density in short transfer resistances. Ohmic resistance represents the resistance offered
steps while maintaining a constant circulation rate corresponding to that by the electrode, electrolyte, graphite plate and current collectors for
in Fig. 2a. The peak value of the power density obtained through this transport of electrons, and the electrode, the electrolyte and the mem­
curve can be interpreted as the peak power density achievable during a brane for the transport of ionic species at the rate required by the rate of
rapid power transient. One can see that, at cell temperature of 25 ◦ C, the the electrochemical reaction. Charge transfer resistance is associated
cell could be operated at a current density of 200 mA cm− 2 with a with activation kinetics of the charge transfer at the electrodes to the
creditable peak power in excess of 200 mW cm− 2 (albeit for a short required rate. Mass transfer resistance is associated with the transport of
duration of 30 s) at a flow rate designed for operation at 60 mA cm− 2. At inert reactant or product species to and from the reaction sites on the
cell temperature of +10 ◦ C, the peak current density is reduced to 180 electrodes, and is usually dependent on the state of charge (SoC)
mA cm− 2, which gets further reduced to 140 mA cm− 2 when the cell reflecting the concentration of the active species and electrolyte circu­
temperature reduces to − 10 ◦ C. It must be noted that these curves need lation rate. Here, sufficient flow rate of the electrolyte maintained to
to be interpreted carefully as the SoC changes during these tests, i.e., as overcome the mass transfer resistance. The x-axis intercept of Nyquist
one approaches higher current densities. Thus, the peak power density is plot corresponding to high frequency represents the ohmic resistance
also influenced by the SoC. Despite this, one can infer, by comparing the and diameter of the semi-circle represents the charge transfer resistance
data of +25 ◦ C and − 10 ◦ C, that the peak power density for the latter of the VRFB. We can see that both the resistances increase with decrease
case is reduced by about 50%. The cell could not be operated at higher in temperature. Table 1 shows the numerical values obtained by model-
current densities even for a short duration even though the effective flow fitting. One can see that the ohmic resistance is 4.19 mΩ at 25 ◦ C, 4.92
factor would have been about SF50 of 5.7 at the peak current density of mΩ at 10 ◦ C and 7.17 mΩ at − 10 ◦ C. The charge transfer resistance is
140 mA cm− 2. The low mass diffusivities of the reactant and product lower at 1.81 mΩ at 25 ◦ C, 3.48 mΩ at 10 ◦ C but becomes more pre­
species would have led to rapid depletion of the species in the reaction dominant at a value of 9.83 mΩ at − 10 ◦ C. Thus, a typical VRFB cell
zone leading to unstable operation. It was observed in these cases that suffers from high ohmic and charge transfer resistances at low operating
the power discharged would vary rapidly within the short step period. temperatures. These results explain the widening of the voltage between
The data of cell parameters of engineering interest shown in Figs. 2–4 the charging and discharging polarization curves as the cell temperature
above show that decreasing the cell temperature has a deleterious effect is decreased. When this is coupled with reduced diffusivity of reactant
on the overpotential at the same operating conditions such as current species at low temperatures, then mass transfer polarization sets in early
density, flow rate and SoC. In order to shed further light on the factors leading to significantly compromised charge and energy storage

5
P. Rao and S. Jayanti Journal of Power Sources 482 (2021) 228988

Fig. 3. Effect of cell temperature on polarization curves during charge-discharge test for the baseline case of electrode of 6 mm thickness, compression ratio of 35%
and at the same circulation rate as in Fig. 2a:(a) low current density of 40 mA cm− 2 (effective SF50 of 19.8) and (b) high current density of 80 mA cm− 2 (effective SF50
of 9.9).

Fig. 4. Effect of cell temperature on the power density for the baseline case:
electrode of 6 mm thickness, compression ratio of 35% and constant circulation Fig. 5. Effect of cell temperature on the cell impedance for the baseline case:
rate corresponding to that of Fig. 2a. electrode of 6 mm thickness, compression ratio of 35% and constant circulation
rate corresponding to that of Fig. 2a.

capacity of the cell.


Table 1
Resistances obtained from EIS at different thickness of electrode of VRFB.
3.2. Influence of electrode design parameters
Temperature Ohmic resistance Charge transfer resistance
Recent studies, in the form of in situ measurements of cell pressure ( oC) (mΩ) (mΩ)
drop and electrochemical performance characteristics coupled with 6 mm 4.6 mm 6 mm 4.6 mm
calibrated computational fluid dynamics simulations, by the present +25 4.19 3.16 1.81 2.05
authors [23–25], among others, have highlighted the important role of +10 4.92 3.85 3.48 3.65
under-the-rib convection of electrolyte in serpentine flow fields and its -10 7.17 6.20 9.83 4.18
dependence on electrode parameters such as its compression and
thickness. Specifically, they show that the onset of mass transfer polar­
ization can be influenced by optimization of these parameters. Further,
increasing electrode compression is shown [22,23] to reduce ohmic

6
P. Rao and S. Jayanti Journal of Power Sources 482 (2021) 228988

resistance and porosity of cells. In view of this, experiments have been


conducted in the present study for electrode compression ratios of 17%,
35% and 50% while keeping other parameters constants. The results
from these are shown in Fig. 6 wherein polarization curves for three
compression ratios are compared separately for the three cell tempera­
tures. In all these cases, the current density was kept constant at 60 mA
cm− 2 and the electrolyte circulation rate maintained constant at SF50 of
13.2. One can see that in all cases, compression ratio of 50% gives
improved results as far as the discharge efficiency is concerned. The
difference is more pronounced at cell temperature of − 10 ◦ C (Fig. 6c)
where one can expect the highest mass transfer resistance due to reduced
mass diffusivity of species at low temperatures. This is partly due to
decrease in the viscosity of the electrolyte; Xiao et al. [7] have reported
that electrolyte viscosity increases nearly 10 times at − 20 ◦ C compared
to 50 ◦ C. Increase in electrode compression also has a strong effect on the
electrode porosity, especially at low SoCs. Park et al. [22] reported a
three-fold decrease in area specific resistance when the electrode
compression was increased from 10% to 30% compression although the
porosity decreased only by 3% over the same range of compression.
Previous room temperature studies in large cells [23] found that
compression ratio of the range 35–50% was beneficial by contributing
significantly to increased performance of VRFB in cells of active areas of
426 and 936 cm2; at 70% compression, the effect of increase in con­
ductivity appeared to be nullified by decrease of the permeability
leading to earlier onset of mass transfer polarization. These results show
that compression ratio of the electrode is an important parameter in
improving the cell performance at low temperatures.

Fig. 7. Data obtained three cell temperatures with 4.6 mm thick electrode with
35% compression ratio: (a) low circulation rate (SF50 of 6.6) polarization curves
at 60 mA cm− 2, (b) high current density (80 mA cm− 2) polarization curves at
SF50 of 9.9, (c) power density curves, and (d) cell impedance.

Fig. 7 shows the results obtained with an alternative approach,


namely, a thinner electrode (4.6 mm uncompressed thickness) but one
which is compressed only by 35%. Although a fuller set of evaluation has
been done, see Table 2 to be presented later, key results are summarized
in Fig. 7. The cell polarization curve at the three temperatures is shown
in Fig. 7a for the more critical case of low circulation rate (corre­
sponding to SF50 of 6.6) at a current density of 60 mA cm− 2. These re­
sults can therefore be compared with those shown in Fig. 2b for a 6 mm
thick electrode compressed by 35%. One can see significantly improved
discharge energy with the thinner electrode. The extent of discharge
energy gain compares with that obtained with 50% compression of a 6
mm electrode. The polarization curves for the higher current density of
Fig. 6. Effect of compression ratio (CR) on the polarization curve during 80 mA cm− 2 at SF50 of 13.2 are shown in Fig. 7b for the thinner cell.
charge-discharge testing for an electrode thickness of 6 mm, current density of Comparing these with the corresponding results in Fig. 3b for the thicker
60 mA cm− 2 and a circulation of SF50 of 13.2 at cell temperature of (a) +25 ◦ C, electrode, one can once again see vast improvement in the discharge
(b) +10 ◦ C and (c) − 10 ◦ C.

7
P. Rao and S. Jayanti Journal of Power Sources 482 (2021) 228988

Table 2
2
Consolidated data of discharge capacity and cell efficiencies for various conditions; current densities corresponding to A, B and C are 40, 60 and 80 mAcm−
respectively.
Compression ratio & Electrode Cell tempera- Stoichio-metric Discharge capacity (mAh) Coulombic efficiency Voltage efficiency Energy efficiency
thickness ture factor

(oC) (%) (%) (%)

A B C A B C A B C A B C

17% & 6.0 mm 25 6.6 9111 9241 8474 94.9 95.0 94.9 81.8 75.5 73.4 77.6 71.8 69.7
13.2 9563 9699 8876 94.8 95.0 95.0 83.2 75.7 74.0 78.9 71.9 70.3
10 6.6 7946 7366 7076 90.3 95.0 94.9 81.8 75.5 73.4 68.4 65.6 64.1
13.2 8460 7822 6830 94.8 95.0 95.0 83.2 75.7 74.0 72.9 66.6 62.7
− 10 6.6 7945 3953 – 97.1 94.7 NC 69.8 57.2 – 67.8 54.2 –
13.2 8846 4893 – 97.1 94.4 NC 72.8 58.6 – 70.7 55.4 –
25 6.6 9903 9807 9528 95.7 95.8 95.6 82.3 77.9 73.8 78.8 74.6 70.5
13.2 10,597 10,325 9880 96.0 95.9 95.6 83.0 78.7 74.7 79.7 75.5 71.3
35% & 6.0 mm 10 6.6 8728 8344 7603 96.3 96.3 95.4 78.9 73.5 68.6 75.9 70.7 65.4
13.2 9276 8864 7999 96.3 96.0 95.4 80.1 74.8 69.6 77.1 71.8 66.4
− 10 6.6 7627 3940 414 97.8 94.5 76.6 66.3 57.3 50.5 64.8 54.2 38.7
13.2 8443 5482 570 97.1 95.7 70.0 67.9 60.1 48.3 65.9 57.5 33.8
50% & 6.0 mm 25 6.6 10,404 10,851 10,614 93.6 95.7 97.0 83.2 79.7 78.0 77.9 76.3 75.5
13.2 11,573 11,190 11,112 95.3 95.4 97.0 83.7 80.7 78.0 79.7 77.0 76.1
10 6.6 7461 8539 7889 93.9 95.6 97.2 75.8 73.4 67.3 71.2 70.2 65.5
13.2 9415 9590 8756 96.5 96.3 95.6 78.7 73.2 69.3 76.0 70.5 66.2
− 10 6.6 8298 7185 4498 96.5 96.0 93.5 72.6 66.4 58.4 70.1 63.8 54.6
13.2 9595 8325 4522 96.7 96.3 93.2 71.9 67.8 58.2 69.6 65.3 54.3
25 6.6 10,683 10,713 10,458 96.8 96.6 96.3 80.7 79.4 72.4 78.0 76.7 69.7
13.2 11,289 11,438 10,728 96.5 95.9 96.3 82.1 81.3 73.3 79.2 77.9 70.6
35% & 4.6 mm 10 6.6 9918 10,230 9558 96.9 96.6 95.7 82.9 78.9 74.6 80.4 76.2 71.4
13.2 10,719 10,711 9615 96.4 96.7 96.0 84.1 77.0 75.4 81.1 74.5 72.3
− 10 6.6 8519 8557 4921 97.8 97.2 94.5 73.5 69.0 60.1 71.9 67.1 56.8
13.2 9516 8899 6572 97.6 97.1 95.7 75.5 69.0 63.1 73.7 67.0 60.4

capacity. Whereas Fig. 3b shows little discharge capacity (<10% of the cm− 2 at +25 ◦ C and reduces to lower values as the cell temperature
discharge capacity at +25 ◦ C), Fig. 7b shows that the thinner electrode decreases. Increasing the compression ratio to 50% or reducing the
retains nearly 60% of the discharge capacity even at − 10 ◦ C. Fig. 7c electrode thickness to 4.6 mm while keeping the compression ratio at
shows that the peak power density too has improved by about 15%. A 35% have the effect of restoring much of the lost discharge capacity and
key to understanding the improved performance lies in the EIS data energy efficiency when the cell operates at low temperatures. Studies
shown in Fig. 7d. One can see that, as in the 6 mm electrode case, the [23] in a large area cell typical of industry-scale have shown that for
ohmic and the charge transfer resistance do increase with decreasing cell large cells 50% compression is too high from a pressure drop point of
temperature, but not by as much. The numerical data are compared in view. Thus, reducing the electrode thickness appears to be the appro­
Table 1 with those obtained with the thicker electrode. One can observe priate way of making the cell function closer to its true potential when
a decrease of ohmic resistance by 15–20% at each cell temperature with its operating at sub-zero cell temperatures.
the thinner electrode. While the charge transfer resistance is about the Cell- and stack-level optimization should also include the effect of
same for +25 ◦ C and +10 ◦ C, it is reduced by more than half at − 10 ◦ C. increased viscosity of the electrolyte at low temperatures and the
This seems to have improved the overall cell performance. consequent increase in pressure drop. Extensive studies [23–25] of flow
distribution and pressure drop in large area cells show that the flow
distribution in a serpentine flow field remains essentially uniform over a
3.3. Comparison of consolidated data of discharge capacity and cell wide range of pressure drops. Thus, with decreases in temperature, one
efficiencies can expect the pressure drop to increase but the electrolyte distribution
over the cell to remain uniform if a serpentine flow field is used. Overall
The consolidated data of the discharge capacity of the cell and the optimization should thus include possible role of electrocatalysts, elec­
three efficiencies, namely, coulombic, voltaic and energy efficiency, are trode compression and thickness as well as the flow field parameters to
listed in Table 2. The evaluation of the coulombic efficiency and the arrive at the most optimal solution.
energy efficiency (also known as the round-trip energy efficiency) has
been discussed earlier. Voltaic efficiency is obtained from these as it is 4. Conclusions
the ratio of the energy efficiency to the coulombic efficiency. In constant
current operations, the voltage efficiency is a measure of the voltage The polarization characteristics of a single cell VRFB with 17%, 35%
window during charging and discharging relative to the cell voltage, and and 50% compression ratio of 6 mm thick electrodes have been studied
is thus a measure of the average overpotentials during charging and experimentally at 25, 10 and − 10 ◦ C in a cell of semi-industrial size (426
discharging. cm2). Comparative charge-discharge cycle studies, EIS analysis and
Included in Table 2 are the discharge capacities and the three effi­ power density estimation have been carried out for cells with 6 mm and
ciencies averaged over the 2nd, the 3rd and the 4th cycle of a five-cycle, 4.6 mm thick electrodes. The following conclusions can be drawn from
constant current charge-discharge cycle test for various combination of these studies:
parameters, namely, electrode thickness, electrode compression, current
density, electrolyte circulation rate and cell temperature. One can see • Significant loss of performance is observed when the cell is operated
that, on the average, doubling the electrolyte circulation increases the at − 10 ◦ C. This is in the form of reduced discharge capacity (by up to
discharge capacity by about 5% only at +25 and + 10 ◦ C cell temper­ 50% compared to the baseline case), reduced energy efficiency (by
atures. The effect is more significant (~10–20%) at − 10 ◦ C. The overall about 15%) and inability to operate in a sustained manner at high
energy efficiency is reasonably high (~75% or higher) in the 40–80 mA

8
P. Rao and S. Jayanti Journal of Power Sources 482 (2021) 228988

current densities (of the order of 100 mA cm− 2) that are normally [6] Chanyong Choi, Hyungjun Noh, Soohyun Kim, Riyul Kim, Juhyuk Lee, Jiyun Heo,
Hee-Tak Kim, Understanding the redox reaction mechanism of vanadium
possible at the baseline condition. The loss of performance is far less
electrolytes in all-vanadium redox flow batteries, J. Energy Storage 21 (2019)
at cell temperature of +10 ◦ C. 321–327.
• EIS studies show that both ohmic resistance and charge transfer [7] Shuibo Xiao, Lihong Yu, Lantao Wu, Le Liu, Xinping Qiu, Jingyu Xi, Broad
resistance increase significantly in comparison with the baseline temperature adaptability of vanadium redox flow battery—Part 1: electrolyte
research, Electrochim. Acta 187 (2016) 525–534.
case, by about 75% and 400%, respectively, when the cell operating [8] Jingyu Xi, Shuibo Xiao, Lihong Yu, Lantao Wu, Le Liu, Xinping Qiu, Broad
temperature is − 10 ◦ C. When this is coupled with reduced diffusivity temperature adaptability of vanadium redox flow battery—Part 2: cell research,
of reactant species at low temperatures, mass transfer polarization Electrochim. Acta 191 (2016) 695–704.
[9] Jianxin Pan, Mianyan Huang, Li Xue, Shubo Wang, Weihua Li, Tao Ma,
sets in early leading to significantly compromised charge and energy Xiaofeng Xie, Vijay Ramani, The performance of all vanadium redox flow batteries
storage capacity of the cell. at below-ambient temperatures, Energy 107 (2016) 784–790.
• Two methods known to reduce cell resistance, namely, increasing [10] Ke Wang, Yunong Zhang, Le Liu, Jingyu Xi, Zenghua Wu, Xinping Qiu, Broad
temperature adaptability of vanadium redox flow battery-Part 3: the effects of total
compression ratio of the electrode and using thinner electrode, offer vanadium concentration and sulfuric acid concentration, Electrochim. Acta 259
promising results for improving low temperature performance. (2018) 11–19.
Increasing the compression ratio to 50% or reducing the electrode [11] Liyu Li, Soowhan Kim, Wei Wang, M. Vijayakumar, Zimin Nie, Baowei Chen,
Jianlu Zhang, Guanguang Xia, Jianzhi Hu, Gordon Graff, Jun Liu, Zhenguo Yang,
thickness to 4.6 mm while keeping the compression ratio at 35% A stable vanadium redox-flow battery with high energy density for large-scale
have the effect of restoring much of the lost discharge capacity and energy storage, Adv. Energy Mater. 1 (2011) 394–400.
energy efficiency when the cell operates at low temperatures. [12] David Reed, Wei Wang Thomsen, Zimin Nie, Bin Li, Kizewski James,
Vilayanur Viswanathan, Alasdair Crawford, Sprenkle Vincent, Advances in PNNL’s
mixed acid redox flow battery stack, Pacific Northwest National Laboratory, OE
In summary, we can conclude that optimizing the electrode param­ Energy Storage Systems Program Review Washington DC September 26–28th
eters is a promising way for improving low temperature performance of (2016).
VRFB cells. Overall stack optimization should thus include possible role [13] Jingyu Xi, Bo Jiang, LihongYu, Le Liu, Membrane evaluation for vanadium flow
batteries in a temperature, range of 20–50 ◦ C, J. Membr. Sci. 522 (2017) 45–55.
of electrocatalysts, electrode compression and thickness as well as the [14] Lihong Yu, Feng Lin, Ling Xu, Jingyu Xi, Structure–property relationship study of
flow field parameters to arrive at the most optimal solution for a given Nafion XL membrane for high-rate, long-lifespan, and all climate vanadium flow
range of cell operation. batteries, RSC Adv. 7 (2017) 31164–31172.
[15] Jungmyung Kim, Heesung Park, Experimental analysis of discharge characteristics
in vanadium redox flow battery, Appl. Energy 206 (2017) 451–457.
CRediT authorship contribution statement [16] Lantao Wu, Yi Shen, Lihong Yu, Jingyu Xi, Xinping Qiu, Boosting vanadium flow
battery performance by nitrogen-doped carbon nanospheres electrocatalyst,
Nanomater. Energy 28 (2016) 19–28.
Praphulla Rao: Conceptualization, Methodology, Software, Data [17] Lihong Yu, Feng Lin, Lin Xu, Jingyu Xi, P-doped electrode for vanadium flow
curation, Visualization, Writing - original draft. Sreenivas Jayanti: battery with high-rate capability and all-climate adaptability, J. Energy Chem. 35
Conceptualization, Data curation, Validation, Writing - review & edit­ (2009) 55–59.
[18] Yuchen Liu, Feng Liang, Yang Zhao, Lihong Yu, Le Liu, Jingyu Xi, Broad
ing, Supervision, Funding acquisition. temperature adaptability of vanadium redox flow battery–part 4: unraveling wide
temperature promotion mechanism of bismuth for V2+/V3+ couple, J. Energy
Declaration of competing interest Chem. 27 (2018) 1333–1340.
[19] Lihong Yu, Feng Lin, Wangdong Xiao, Ling Xu, Jingyu Xi, Achieving efficient and
inexpensive vanadium flow battery by combining CexZr1− xO2 electrocatalyst and
The authors declare that they have no known competing financial hydrocarbon membrane, Chem. Eng. J. 356 (2019) 622–631.
interests or personal relationships that could have appeared to influence [20] Sheeraz Mehboob, Ghulam Ali, Hyun-Jin Shin, Jinyeon Hwang, Saleem Abbas,
the work reported in this paper. Kyung Yoon Chung, Heung Yong Ha, Enhancing the performance of all-vanadium
redox flow batteries by decorating carbon felt electrodes with SnO2 nanoparticles,
Appl. Energy 229 (2018) 910–921.
Acknowledgements [21] Q. Xu, T.S. Zhao, C. Zhang, Performance of a vanadium redox flow battery with
and without flow fields, Electrochim. Acta 142 (2014) 61–67.
[22] S.K. Park, J. Shim, J.H. Yang, C.S. Jin, B.S. Lee, Y.S. Lee, K.H. Shin, J.D. Jeon, The
The work described in this paper was funded by grants from DST- influence of compressed carbon felt electrodes on the performance of a vanadium
Solar Energy Harnessing Centre (Grant reference no. DST/TMD/SERI/ redox flow battery, Electrochim. Acta 116 (2014) 447–452.
HUB/1(C)), Department of Science and Technology, Government of [23] R. Gundlapalli, S. Jayanti, Effect of electrode compression and operating
parameters on the performance of large vanadium redox flow battery cells,
India. The first author (PR) would like to acknowledge B M S College of J. Power Sources 427 (2019) 231–242.
Engineering, Bengaluru, India for sponsoring her Ph.D. studies under the [24] S. Kumar, S. Jayanti, Effect of electrode intrusion on pressure drop and
Quality Improvement of Programme of the Government of India. electrochemical performance of an all-vanadium redox flow battery, J. Power
Sources 360 (2017) 548–558.
[25] R. Gundlapalli, S. Jayanti, Effect of channel dimensions of serpentine flow fields on
References the performance of a vanadium redox flow battery, J. Energy Storage 23 (2019)
148–158.
[1] Gaolong Zhu, Kechun Wen, Weiqiang Lv, Xingzhi Zhou, Yachun Liang, Fei Yang, [26] Igor Derr, Michael Bruns, Joachim Langner, Abdulmonem Fetyan, Julia Melke,
Zhilin Chen, Minda Zou, Jinchao Li, Yuqian Zhang, Weidong He, Materials insights Christina Roth, Degradation of all-vanadium redox flow batteries (VRFB)
into low-temperature performances of lithium-ion batteries, J. Power Sources 300 investigated by electrochemical impedance and X-ray photoelectron spectroscopy:
(2015) 29–40. Part 2 electrochemical degradation, J. Power Sources 325 (2016) 351–359.
[2] Bharat Kumar Suthar, Dayaram Sonawane, Richard D. Braatz, Venkat [27] Igor Derr, Abdulmonem Fetyan, Konstantin Schutjajew, Christina Roth,
R. Subramanian, Optimal low temperature charging of lithium-ion batteries, IFAC- Electrochemical analysis of the performance loss in all vanadium redox flow
Papers Online 48–8 (2015) 1216–1221. batteries using different cut-off voltages, Electrochim. Acta 224 (2017) 9–16.
[3] Asma Mohamad Aris, Bahman Shabani, An experimental study of a lithium ion cell [28] Ao Tang, Jie Bao, Maria Skyllas-Kazacos, Dynamic modelling of the effects of ion
operation at low temperature conditions, Energy Procedia 110 (2017) 128–135. diffusion and side reactions on the capacity loss for vanadium redox flow battery,
[4] Ting Ma, Pan Zeng, Licheng Miao, Chengcheng Chen, Mo Han, Zhenfeng Shang, J. Power Sources 196 (2011) 10737–10747.
Jun Chen, Porphyrin-based symmetric redox-flow batteries towards cold-climate [29] S. Kumar, S. Jayanti, High energy efficiency with low pressure drop configuration
energy storage, Angew. Chem. Int. Ed. 57 (2018) 3158–3162. for an all-vanadium redox flow battery, J. Electrochem. Energy Conv. Storage,
[5] Funian Mo, Guojin Liang, Qiangqiang Meng, Zhuoxin Liu, Hongfei Li, Jun Fan, Trans. ASME 13 (2017), 041005.
Chunyi Zhi, A flexible rechargeable aqueous zinc manganese-dioxide battery
working at -20◦ C, Energy Environ. Sci. 12 (2019) 706–715.

You might also like