Sei sulla pagina 1di 10

E3122 Journal of The Electrochemical Society, 164 (11) E3122-E3131 (2017)

JES FOCUS ISSUE ON MATHEMATICAL MODELING OF ELECTROCHEMICAL SYSTEMS AT MULTIPLE SCALES IN HONOR OF JOHN NEWMAN
Performance Modeling and Design of Ultra-High Power
Microbatteries
James H. Pikul,a,∗,z Paul V. Braun,b,c and William P. Kingb,c
a Department of Mechanical Engineering and Applied Mechanics, University of Pennsylvania, Philadelphia,
Pennsylvania 19104, USA
b Department of Mechanical Science and Engineering, University of Illinois at Urbana - Champaign, Illinois, 61801,
USA
c Department of Materials Science and Engineering, University of Illinois at Urbana-Champaign, Illinois, 61801, USA

High power density microbatteries could enable new capabilities for miniature sensors, radios, and industrial electronics. There is,
however, a lack of understanding on how battery architecture and materials limit power performance when battery discharge rates
exceed 100 C. This paper describes the development and application of an electrochemical model to predict the performance of
microbatteries having interdigitated bicontinuous microporous electrodes, discharged at up to 600 C rates. We compare predicted
battery behavior with measurements, and use the model to explore the underlying physics. The model shows that diffusion through
the solid electrodes governs microbattery power performance. We develop design rules that could guide the development of improved
batteries.
© The Author(s) 2017. Published by ECS. This is an open access article distributed under the terms of the Creative Commons
Attribution 4.0 License (CC BY, http://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse of the work in any
medium, provided the original work is properly cited. [DOI: 10.1149/2.0151711jes] All rights reserved.

Manuscript submitted February 1, 2017; revised manuscript received April 24, 2017. Published May 4, 2017. This paper is part of
the JES Focus Issue on Mathematical Modeling of Electrochemical Systems at Multiple Scales in Honor of John Newman.

High power density microbatteries would enable new capabilities from high power microbatteries we previously reported.5 The model
for miniature sensors, radios, and industrial electronics.1–4 Recent im- enables the study of local lithium concentration and overpotentials,
provements in electrode architectures, materials, and fabrication tech- which indicate that diffusion through the solid electrodes most limits
nologies have enabled microbatteries with power densities as high as the microbattery performance. Using the validated model, we propose
7.4 mW cm−2 μm−1 , which is about 100 times greater than power den- design rules for high power batteries that optimize battery perfor-
sities provided by larger conventional format batteries.5–9 The ultra- mance.
high power densities were achieved by the simultaneous reduction of
ion and electron transport resistances across the anode, cathode, and
electrolyte. Fabricating electrodes with increasingly fine nanostruc- System Description and Modeling Approach
tures that provide shorter ion and electron transport paths has been the Figures 1a and 1b show a schematic and SEMs of the high power
main strategy for reducing transport resistances.2,5–7,10–18 However, as battery modeled in this work.5 The battery has interdigitated elec-
the electrode dimensions decrease, electrode fabrication and incorpo- trodes composed of a highly porous bicontinuous nickel current col-
rating large volume fractions of high capacity materials into the nano lector (blue) conformally coated with electrochemically active mate-
architectures (important for obtaining high energy densities) become rials (active materials). The active material is nickel-tin (red) in the
more difficult. Additionally, the larger surface area leads to increased anode and lithiated manganese oxide (yellow) in the cathode. The
SEI formation during battery fabrication. To realize both high power electrolyte fills the volume between the electrodes and inside the elec-
and high energy density, it is important to understand how battery ar- trode pores. Figure 1c shows the lithium and electron transport paths
chitecture and materials affect the physical processes that limit power during discharge. Lithium is stored at a high energy state in the anode.
density and improve energy density, and to develop experimentally An oxidation reaction at the anode-electrolyte surface releases lithium
validated design rules that address the many engineering constraints ions and electrons that flow to the lower energy state cathode where
in full battery assemblies. they undergo reduction. Electrons do not travel through the electrolyte
Simulations of battery operation, considering ion transport across and instead travel from the anode active material surface through the
both the anode and cathode regions, can be used to assist in battery anode active material, anode current collector, external circuit where
design and optimization. A key parameter in understanding battery they power a load, cathode current collector and cathode active mate-
discharge is the C rate, where the time it takes to discharge a battery rial until they react at the cathode surface (black). Lithium ions flow
in one hour is 1 C rate. An X C rate discharge corresponds to a current from the anode to the cathode through the electrolyte (light blue). As
density X times the 1 C rate current density. The validity of battery lithium ions are released or inserted into the active material, lithium
discharge models has rarely been explored above discharge rates of stored in the active material bulk diffuses toward or away from the
25 C,19–32 with only a few exceptions.33–37 To our knowledge, there is active material surface according to the concentration gradient (dark
no published work that validates models for batteries discharged at the blue). The microbattery architecture achieves high power density by
100–1000 C rates relevant to high power microbatteries. There is thus simultaneously reducing the ion and electron transport lengths.
a need for an experimentally validated battery model that describes The battery power performance depends upon the battery voltage,
the physical processes that limit battery performance at high C rates. which is the difference in anode and cathode electrochemical poten-
Such a model would provide valuable insights for future high power tials minus any internal voltage drops. The electrochemical potential
microbattery designs and fabrication, as well as guide the design of difference depends on the lithium concentrations at the active material
macroscopic battery architectures. surfaces. Internal voltage drops are due to the ohmic conduction of
In this paper we present a 1-D electrochemical model that ac- electrons through the electrodes, the ohmic conduction of ions across
curately predicts the power performance of batteries discharged at the electrolyte, and the electrochemical kinetics at the anode and cath-
up to 600 C rates. The predictions were compared to discharge data ode surface. Energy density is the product of the battery voltage and
amount of charge transferred between electrodes per battery volume.
∗ Electrochemical Society Student Member. Power density is the product of the battery voltage and charge transfer
z
E-mail: pikul@seas.upenn.edu rate between electrodes per battery volume. As the charge transfer

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 164 (11) E3122-E3131 (2017) E3123

Figure 1. a) A diagram of a microbattery with interdigitated electrodes that consist of an electrochemically active layer (red and yellow) coated on an electrically
conductive porous bicontinuous nickel scaffold (blue). b) Electron microscope image of the interdigitated battery electrodes. c) Diagram of a unit cell of the lithium
ion microbattery electrodes showing the key transport paths. d) The simulation domain of this microbattery.

rate increases, the electrochemical potential difference decreases and model inputs that determine the energy density of the batteries and
the internal voltage drops increase, so that the battery voltage falls impact the transport of ions and electrons. ε N i , εact and εe depend on the
below its equilibrium value and reaches a cutoff voltage before all bicontinuous electrode structure and are calculated from a geometric
of the energy can be fully extracted. The reduced battery voltage at model of self-assembled polystyrene (PS) spheres organized in a FCC
high charge transfer rates decreases power performance, or amount of unit cell. The current collector volume is the cubic unit cell volume
energy extracted at a given power density. minus the void volume left by the sintered PS spheres after PS etching,
Here, we model the power performance of high power interdigi- Vvoid . The active material volume, Vact , is the volume of a thin layer
tated bicontinuous microbatteries using a 1-D model of ion and elec- coated on the current collector, calculated from thickness t. Figures
tron transport.19,20 Figure 1d shows the model space, which includes 2a and 2b show an inverse opal unit cell and the simplified geometry
the porous anode, porous cathode, and separator. The model space sim- of two neighboring spheres used to calculate the volume fractions.
plifies the interdigitated electrode design to two electrodes bounded PS spheres in contact after opal self-assembly are represented by
by the electrode centerlines with a symmetric boundary condition. circles with radius R and have 0.74 volume fractions. Sintering the PS
The electrodes are approximated as rectangular, with widths (Wneg increases the radius from R to Rn such that neighboring radii overlap
and Wpos ), height (H), and cross sectional areas equal to the mea- by a length b. Vvoid is the volume of 4 sintered PS spheres of radii Rn
sured cross sectional areas of the experimental electrodes. Fig. 1d in a FCC unit cell minus the volume of 48 overlapping spherical caps
shows the 1-D transport of ions and electrons. Ion transport through of height h or
the electrolyte is governed by migration and diffusion using concen-
trated solution theory,19,20 and electron transport is ohmic. The total
4 1
current through each differential length, dx, along the battery is the Vvoid = 4 πRn 3 − 48 πh 2 (3Rn − h) , [1]
sum of the current due to local conduction of electrons and ions and is 3 3
constant across the entire pitch. At each differential length in the elec-
trodes, 1-D diffusion models predict lithium concentration through 
where Rn = (b/2)2 − R 2 and h is Rn – R. Vact is the volume of
the active materials. The active material coating is approximated as a
4 spheres with radius Rv subtracted from Vvoid , where Rv is Rn – t.
thin film with zero lithium flux to the nickel scaffold and flux at the
Additionally, the volume where the active material does not deposit,
electrode-electrolyte interface determined by Fick’s law. The amount
marked with hatching, is integrated and subtracted out of Vact , but not
of lithium inserted or removed from the active material is determined
including the volume of the spherical cap marked by h2.
by Butler-Volmer kinetics which link the electrolyte simulation to the
electrochemical potential determined by the active material surface
  2
concentration. The model space and boundary conditions are cho- b 2
Rv R
4 Rn
sen to match conventional battery formats so that the results can be Vact = Vvoid − 4 π(Rv )3 − 48 π − t − x2
generalized to other electrode architectures. 3 0 2
The nickel current collector volume fraction, ε N i , active material 1
volume fraction, εact , and electrolyte volume fraction, εe , are important d x + 48 πh22 (3Rv − h2) , [2]
3
Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
E3124 Journal of The Electrochemical Society, 164 (11) E3122-E3131 (2017)

Figure 3. Equilibrium voltage profiles used for the anode and cathode mate-
rials in the simulation. During discharge, the cathode voltage decreases and
anode voltage increases. The black line shows the difference between the
cathode and anode equilibrium voltages.

normalized to 1, such that


εN i εact
σe f f = σN i + σact . [7]
ε N i + εact ε N i + εact
σepos
ff
is 3.3 × 106 S m−1 and σneg ef f
is 1.1 × 107 S m−1 . k e f f is then
corrected for the porosity of the electrode with a Bruggeman exponent
of 1.5.
The model uses electrolyte material properties from available elec-
trolyte data. The experimental electrolyte is 1:1 ethylene carbonate:
dimethyl carbonate (EC:DMC) with initial 1000 mol m−3 LiClO4
concentration, ce0 . The electrolyte interdiffusion coefficient is approx-
imated as 2.6 × 10−10 m2 s−1 based on 1 M LiClO4 in propylene
Figure 2. Model of the electrode showing key parameters. a) Unit cell of carbonate (PC).38 The electrolyte conductivity depends on concen-
the microbattery electrode with some geometric parameters. b) Model used to tration and is taken from data on LiClO4 in EC:PC,39 where the 8.5
calculate the polystyrene radius (R), active material thickness (t), and volume × 10−3 S cm−1 maximum conductivity, σe , is close to that of 1 M
fractions of the current collector and electrochemically active materials. LiClO4 in EC:DMC.40,41 The lithium transference number, t+ , is 0.363
based on LiPF6 in EC:DMC because the transference number depends
on the solvation radius of the anion, which is similar for PF6 − and
where h2 is Rv (1 – R / Rn ). ε N i , εact and εe are calculated
√ from the Vact ClO4 − .19,40,42 The electrolyte conductivity and diffusivity through the
and Vvoid normalized by the unit cell volume, (2R 2)3 . porous electrodes are corrected for the increased path length using a
Vvoid Bruggeman exponent of 1.5.19,20 Changes to the pore shape or volume
εN i = 1 −  √ 3 , [3] fraction in the microbattery electrodes would change the tortuosity
2R 2 and affect the Bruggeman exponent.8
The following are material properties for the electrochemically
active materials. Figure 3 shows the open circuit voltage (OCV),
Vact
εact =  √ 3 , [4] versus lithium, as a function of the state of charge for the anode
2R 2 and cathode active materials. The OCV is measured from batter-
ies discharged at low rates. The maximum capacity of the cathode,
−3
and cmax
pos , is 23,000 mol m based on a 145 mAh g−1 manganese oxide
cathode. The initial cathode concentration, c0pos , is approximated as
43
εe = 1 − εact − ε N i . [5]
50 mol m−3 for all batteries. The nickel-tin anode is composed of
Equations 3–5 calculate ε N i , εact and εe from fitting parameter t and 80% tin. At a full state of charge, the anode OCV corresponds to
experimental measurements of R and b. ε N i = 0.12 for the experimen- Li3 Sn with a 55,000 mol m−3 volumetric capacity.44 cneg 0
is set so that
tal batteries where b is 200 nm and R is 500 nm. the low rate voltage plateau matches between predicted and measured
An effective electrode conductivity across the electrode width, batteries.
k e f f , is approximated by equating the conductance across the width Table I shows the active material thickness, t, and diffusivity, D,
to the conductance down the length, using used to simulate each battery in addition to experimentally measured
parameters. The active material thickness and diffusivity are fitting
ke f f L H σe f f W H
= , [6] parameters bound by half the interconnect pore size (100 nm) and
W L PITT diffusivity measurements in the anode and cathode materials.
as electrical conduction occurs primarily down the length (normal to Measured anode diffusivities varied between 9.34 × 10−18 and 8.51
the model space in Fig. 1d), L, in the interdigitated architecture. The × 10−17 m2 s−1 , with an average of 3.35 × 10−17 m2 s−1 . Cathode
bicontinuous electrode conductivity, σe f f , is approximated by con- diffusivities varied between 1.57 × 10−19 and 1.14 × 10−16 m2 s−1 ,
duction through the two media in parallel with their volume fractions with an average of 1.83 × 10−17 m2 s−1 . The anode and cathode

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 164 (11) E3122-E3131 (2017) E3125

Table I. Parameters used for the simulation of batteries 1 through 5.

Battery

Parameter 1 2 3 4 5
Dneg [m2 s−1 ] 3.0 × 10−17 1.0 × 10−17 3.0 × 10−17 3.5 × 10−17 4 × 10−17
Dpos [m2 s−1 ] 1.3 × 10−17 5.0 × 10−18 6.0 × 10−18 3.9 × 10−17 6.0 × 10−18
tneg [nm] 30.0 17.0 30.0 21.0 27.0
tpos [nm] 55.0 37.0 20.0 29.5 13.3
cneg 0 [mol m−3] 54,500 54,500 21,000 54,500 15,700
H [μm] 14.9 9.9 14.9 12.6 14.9
Wneg , Wpos [μm] 12.6 8.4 12.6 10.7 12.6
P [μm] 45 27 45 45 45

diffusivities in the simulation vary by a maximum of 3.35 and 3.66 in Battery 3 indicate that the anode was not fully charged before
X the average measured diffusivities for each battery. The model was cycling, which caused to cathode to be overcharged when Battery 3
simulated in COMSOL using a 1-D battery module. was charged to a 4.0 volt cutoff. The overcharging caused a higher
voltage for the first 0.15 Ah m−2 of experimental discharge. The
overall good agreement between predicted and measured discharge
Modeling Results curves indicates that the model captures the major physics for high
Figure 4 shows predicted and experimental discharge curves of rate discharging and that the 1-D transport model provides an accurate
Batteries 1–4, which have the highest combined energy and power foundation for design studies of batteries discharged at up to 600 C
densities of our previously tested microbatteries.5 At low C rates, rates.
predictions and measurements compare very well. At greater than 30 We use the calculated lithium concentration throughout the battery
C rate discharges, there is a slight departure between prediction and to understand how different parameters affect battery performance.
measurement, up to 3 percent of total capacity. The measured and Figure 5a shows the electrolyte concentration in Battery 1 at the end of
predicted discharge curves for Batteries 2 and 3 closely overlap at low discharge at multiple C rates. The change in electrolyte concentration
and high discharge rates, but the higher predicted capacities at 8.6 increases as the C rate increases, except at 457 C, where the battery
and 20 C rates are likely due to error in the thin film approximation quickly shuts off before a significant amount of ions can be removed
as the amount of capacity extracted from a thin film is larger than from the electrolyte. This quick shut off is due to limits in the solid
the concave geometries in the curved electrode pores at moderate active material diffusion and will be discussed later. The maximum
discharge rates. The lower average voltages of the discharge curves concentration change is 250 mol m−3 at 146 C, which decreases the

Figure 4. Comparison of predicted and measured discharge curves in Batteries 1 through 4. The capacity is normalized by the separator area.

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
E3126 Journal of The Electrochemical Society, 164 (11) E3122-E3131 (2017)

Figure 6. Diagrams of the microbattery electrodes showing important design


parameters. a) Cross-section of the microbattery showing electrode width,
W, and electrode pitch, P. b) Microbattery electrode architecture showing the
electrode pore size and active material thickness, as well as electron transport
through the bicontinuous nickel scaffold. c) Microbattery top view showing
the interdigitated layout and the electrode length, L.

increases. The surface and center concentrations in the cathode are


22,200 mol m−3 and 2,600 mol m−3 at 30.5 C, which corresponds to a
1.26 volt difference, or 63% of the battery voltage window. At 457 C,
the cathode surface concentration reaches a maximum in 0.17 seconds
and limits the capacity to 0.058 Ah m−2 . The large concentration
differences across the anode and cathode active materials and the
associated overpotential at moderate and high discharge rates show
that diffusion in the active material limits the power performance of
Battery 1. Diffusive limitations in the active materials limit the power
performance of all simulated batteries.

Design and Optimization Considerations


The design process that follows informs how microbattery geom-
etry influences the trade-offs between energy and power density, and
develops design rules that guide these trade-offs. The power density
of a battery is proportional to the energy density times the C rate,
Power ∼ Crate × Energy, so an improved power performance re-
Figure 5. Predicted lithium concentrations for Battery 1 at various discharge sults from an improved C rate and energy density. Figure 6 shows the
rates. a) Lithium concentration in the electrolyte. b) Lithium concentration in microbattery geometries that govern ion and electron transport. The
the solid. c) Lithium concentration across the cathode active material. electrode width, W, governs diffusion through the electrolyte because
the largest electrolyte concentration gradients occur from depletion
and generation of lithium ions across the electrode width. The elec-
electrolyte conductivity from 8.6 × 10−3 S cm−1 to 8.2 × 10−3 S cm−1 trode pitch, P, governs ionic conduction in the electrolyte. The active
in the cathode region and 8.5 × 10−3 S cm−1 in the anode region. material thickness, t, governs active material diffusion. The particle
The minimal changes in conductivity indicate that diffusion in the radius could function as the thickness in other electrode designs. The
electrolyte did not limit the power performance of Battery 1. Figures bicontinuous electrode architecture and the electrode length, L, have
5b and 5c show the final lithium concentrations in the solid electrode the largest impact on electrode electron conduction in this work. Sim-
active materials in Battery 1 at each discharge rate. The differences ulating the energy density of Battery 1 at various C rates for a range
in surface and center concentrations increase as the discharge rate of values of a single geometric component or governing transport

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 164 (11) E3122-E3131 (2017) E3127

Figure 7. a) Ragone plot of Battery 1 for various values of electrode width (5, 12.6, and 50 μm) and diffusivity (scaling of 0.04, 0.2, 1, 10, and 25X). b) Predicted
energy density versus electrode width for Battery 1 at 1 to 146 C rate discharges. The dashed line is the design parameter, Wc , for the electrode width, which
corresponds to the complete depletion of lithium ions at the cathode centerline for each C rate. c) Predicted energy density versus electrolyte peak conductivity for
Battery 1 at 1 to 457 C rate discharges. The dashed lines are the design parameter, Ve , for the electrode pitch calculated with 0.1 and 0.3 voltage drops across the
electrolyte for each C rate. d) Ragone plot of Battery 1 showing the effect of a change in electrolyte conductivity by 0.01, 0.1, 1, and 10 for each electrode pitch.
The red dots are experimental data from Battery 1. Black diamonds correspond to inflection points calculated from design parameters.

property isolates the influence of each geometric component on the electrolyte with constant and uniform depletion is
power performance. Data from Battery 1 provides the primary com-
parison between experiment and simulation in many of the design ∂ 2c
D + q̇ = 0. [8]
plots because the experimental microbatteries cover a small subset ∂x2
of geometries compared to the large variation required to cover the
The depletion rate in the electrolyte, q̇, is assumed constant when
simulation range.
discharged at a constant C rate and uniformly distributed throughout
the electrode region. q̇ is the amount of lithium ions that enter the
cathode from the electrolyte, per electrode volume, or
Electrode width (W).—Figure 7a shows the predicted energy and
power density of Battery 1 for various values of the electrode width  Crate
pos − c pos ε pos
q̇ = − cmax 0
. [9]
(5, 12.6 and 50 μm) and electrolyte diffusivity (scaling factor 0.04, 3600
0.2, 1, 5, and 25X from the value in Table I). For batteries with
Equation 9 is valid when the capacities in the battery electrodes
wide electrodes and low electrolyte diffusivities, the Ragone curves
are balanced, or when the cathode limits capacity. The discharge time
show large drops in power density for a given energy density. This
is approximated as the time it takes to fully discharge a battery at a
results from lithium ions severely depleting in the cathode region
constant current density (3600 / Crate). Solving Equation 8 yields the
such that the concentration is near zero in the electrolyte close to the
concentration profile
electrode center. The near zero concentration significantly reduces the
local electrolyte conductivity and causes a large voltage to develop q̇ q̇ W pos
across the electrolyte, which brings the battery to the shut off voltage c=− x2 + x + ce0 , [10]
2De εe 1.5 De εe 1.5
before more energy can be extracted. At higher C rates, the Ragone
curves return to normal because diffusion in the active material causes where x is the distance between the separator, x = 0, and the electrode
the battery to shut off before severe electrolyte depletion occurs. We centerline, x = Wpos . The diffusivity is corrected using the Brugge-
conclude that diffusion in the electrolyte has a minimal effect on the man approximation. The initial concentration of the electrolyte, ce0 , is
battery performance unless the concentration in the cathode region the concentration at x = 0, which is valid for most batteries because
depletes to near zero, at which point the effect is dramatic. the concentration change across the separator is small. A symmetric
The point when the concentration in the cathode region depletes to boundary condition exists at x = Wpos . Equation 10 solves for the
zero is predicted by solving for steady state diffusion in the cathode critical electrode width, Wc , where the steady state electrolyte con-
electrolyte region. The governing equation for 1-D diffusion in the centration depletes to zero by setting x = Wc = W pos when c = 0. Wc

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
E3128 Journal of The Electrochemical Society, 164 (11) E3122-E3131 (2017)

is then

−2ce0 De εe 1.5
Wc = . [11]

Figure 7b shows the predicted energy densities of Battery 1 dis-
charged at multiple C rates with varying electrode width (3–100 μm)
and constant Wsep = 19.8 μm. The dashed line shows the energy
density of batteries with electrode widths equal to Wc . A maximum
energy density occurs for each C rate because of competition between
the increases in electrode volume fractions (W/P) that improve energy
density and the increase in ion diffusion length through the electrode
width. The energy density at 12, 20 and 30.5 C dramatically decreases
after the maximum energy density is reached because the concentra-
tion of lithium is depleted to near zero in the cathode electrolyte. An
electrolyte with increased or decreased diffusion coefficient would
increase or decrease the electrode width required to deplete to zero
concentration. We conclude that Wc is a good design parameter for the
cathode width as it predicts the largest allowable cathode electrode
width before severe ion depletion occurs.

Electrode pitch (P).—Figure 7c shows the predicted energy den-


sity of Battery 1 discharged at 1–500 C rates with electrolyte conduc-
tivities varied between 4 × 10−5 S cm−1 and 1 × 10−1 S cm−1 . The
electrolyte conductivity was scaled by 1 × 105 to study the effect of
the pitch on power performance without affecting other physics in the
system, and also represents the affect of changing the battery elec-
trolyte. The dashed lines represent a constant potential drop across
the electrolyte calculated using Ohm’s law, where the voltage drop,
Ve , and electrolyte conductivity, σe , are related to the pitch, P, by
I I P
Ve = I R = ρl = m . [12]
A pos,neg,sep A σe ε1.5
e

The current density is



I F cimax − ci0 εi Wi Crate
= . [13]
A 3600 Figure 8. a) Predicted energy density versus electrode conductivity scaling
factor for Battery 1 at 1 to 457 C rate discharges. The dashed lines are the
i is the electrode with the lowest total capacity. m = 0.5 in the positive design parameter, Vs , for the electrode length and architecture calculated with
and negative electrode region because on average half the current 0.1 and 0.3 voltage drops across the electrode for each C rate. b) Ragone plot
is transported by ionic conduction and the other half by electron of Battery 1 as the electrode conductivity is scaled from 10−7 to 100 times the
conduction in the electrode. m = 1 in the separator region. εe is the reference conductivity.
volume fraction of the electrolyte in each region. Equation 12 is valid
when the electrolyte concentration has little variance during discharge,
which is true if W pos < Wc . The energy density at each C rate in Fig.
7c remains constant as the conductivity decreases until the voltage I L2 L2
= 0.5 ef f
+ ef f
, [14]
drop in the electrolyte, calculated from Equation 12, is greater than A 2σ pos W pos ε1.5
pos 2σneg Wneg ε1.5
neg
0.1 volts.
Figure 7d further demonstrates how the electrolyte conductance
affects power performance by plotting power density versus energy where the current density is calculated from Equation 13. Figure 8b
density for various electrode pitches (30, 45, and 120 μm). The elec- shows predicted power density versus energy density for Battery 1
trode separation is held constant. At each pitch, the peak electrolyte with varying electrode conductivities. The point at which the power
conductivity is scaled by a factor of 0.01, 0.1, 1 or 10X. At low power performance begins to rapidly change (marked by a black diamond
density, battery performance has minimal dependence upon electrode for each battery) is predicted by Equation 14 when Vs = 0.1 volts.
pitch or conductivity. The point at which the power performance be- We conclude that high power performance microbatteries should be
gins to rapidly change (marked by a black diamond for each battery) designed so that Vs < 0.1 volts.
is predicted by Equation 12 when Ve = 0.1 volts. We conclude that
high power performance microbatteries should be designed so that Ve Active material thickness (t).—The active material thickness in-
< 0.1 volts. fluences the fraction of lithium extracted or inserted into the active
material at a discharge rate. A smaller active material thickness allows
Electrode electron path.—Figure 8a shows how the electrode con- more energy to be extracted in less time; there are however practical
ductivity affects the energy density of Battery 1 at 1–500 C discharge limits on the thickness. Figure 9a shows how the cathode active ma-
rates. The conductivity is varied by up to 1 × 10−8 times the conduc- terial thickness in Battery 1 affects the energy density at 1–500 C rate
tivity in Table I. The dashed lines in Fig. 8a correspond to the energy discharges. A maximum energy density occurs because of competing
density of batteries where the voltage drop across the electrodes, Vs , effects between an increase in the total volume fraction of the ac-
is 0.3 or 0.1 V. The voltage drop is estimated using Ohm’s law, tive material and an increase in the distance that lithium travels. The
maximum can be predicted by comparing the diffusion time, τ, to the
I I W pos Wneg diffusion distance, l. The active material thickness corresponding to
Vs = ρl = 0.5 + ef f
A A k epos
f f 1.5
ε pos kneg ε1.5
neg the maximum power performance, to , is predicted from the diffusivity,

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 164 (11) E3122-E3131 (2017) E3129

Figure 9. a) Predicted energy density versus active material thickness for Battery 1 at discharge rates of 1 to 457 C. The dashed line is the design parameter to
for the active material thickness. b) Ragone plot of Battery 1 as active material thickness is varied. The 15.5 C power curve is the energy and power density of
batteries with varied thicknesses discharged at 15.5 C rate. Colored dots are experimental data. Black diamonds correspond to inflection points calculated from
design parameters. c) Predicted energy density versus pore diameter for Battery 1 at discharge rates of 1 to 457 C. The active material thickness is varied with the
pore size to maintain a constant volume fraction. Dc is the pore diameter corresponding to to . d) Ragone plot of Battery 1 as active material thickness is varied and
volume fraction maintained.

Ds , and C rate using Dτ/l2 , or with the pore size to maintain a constant volume fraction so that the
 pore diameters in Figs. 9c and 9d are approximately proportional to
Ds 3600 the active material thickness. The dashed line in Fig. 9c shows bat-
to = n , [15]
Crate tery performances when the active material thickness is to . Increasing
√ the active material thickness past to dramatically reduces the power
where n is 1 for a slab and 5 for a sphere. The dashed line in Fig. 9a is
the energy density corresponding to to , when n = 1, which predicts the performance at all C rates, including 1 C. Decreasing the pore size
maximum with reasonable accuracy. The amount of lithium extracted (PS diameter) improves the power performance because the diffusion
or inserted into the active material is constant for a constant Dτ/l2 distance through the solid active material is reduced (higher power)
when the discharge is stopped at a set surface concentration, which is while the active material volume fraction remains the same (constant
valid for batteries discharged at a constant C rate.45 It follows that the energy). A battery fabricated with 109 nm PS diameter particles can
extracted capacity of a microbattery with to active material thickness be discharged at 300 C and maintain 55% of its total energy density.
is a constant fraction of the total capacity if the shutoff concentration In the simulated batteries, the electrode pores are FCC closed
is set to the maximum concentration in the cathode or minimum in the packed. If the pore distribution deviated from FCC close packed to-
anode. The fraction of extracted capacity is 2/3 the maximum when ward random close packed, the nickel volume fraction would increase
n = 1 and the active material thickness is to . Figure 9b is a simulated and the active material volume fraction would decrease. If the active
Ragone plot for Battery 1 with cathode active material thicknesses material thickness was maintained, and thus the C rate performance
varied between 13 and 100 nm. The power curve for 15.5 C shows the held constant, the power density would still decrease because the
energy and power density of each battery when discharged at a 15.5 C reduced active material volume fraction would decrease the energy
rate. The maximum power performance occurs at the apex of the 15.5 density.
C power curve and corresponds to to = 55 nm, which is the optimal
active material thickness predicted from Equation 15 at 15.5 C rate.
Battery Design Summary
The energy and power density corresponding to to for each battery is
marked with a diamond. We conclude that power performance is near The design parameters (to , Wc , Ve , and Vs ) provide a simple tool
optimal when the active material thickness is to . to optimize the power performance of a battery using only battery
Figures 9c and 9d show how the active material thickness af- geometry and material properties. To illustrate the design utility, we
fects the power performance independent of the volume fraction by optimize Battery 1 for a 50 C rate discharge using the design param-
changing the pore diameters. The active material thickness is varied eters and compare the optimized performance to experimental data.

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
E3130 Journal of The Electrochemical Society, 164 (11) E3122-E3131 (2017)

Figure 10. a) Schematic of a battery architecture optimized for a 50 C discharge rate using the design rules presented here. b) Predicted discharge curve of the
battery. c) Ragone plot of the battery performance compared to previously fabricated batteries. The 50 C rate discharge point is marked. d) Sensitivity analysis of
the optimized battery. The electrochemically active material thickness and electrode width are varied by plus and minus 10 and 25 percent.

Figure 10a shows a diagram of the optimized microbattery. The micro- Conclusions
battery geometry was chosen by setting the active material thickness
to to and cathode width to Wc , while decreasing the pore size to 250 In conclusion, the performance of microbatteries with up to 7.4
nm to increase the active material volume fraction. Additionally, the mW cm−2 μm−1 power densities were predicted using 1-D transport
pitch was set so Ve ≤ 0.1, and the electrode length so that Vs ≤ 0.1. models combined with a geometric model of the bicontinuous elec-
Figures 10b and 10c show the power performance of the 50 C rate trodes. The 1-D model showed good agreement between predicted
optimized microbattery. The microbattery capacity at 50 C is 56% of and measured discharge curves at up to 600 C rate discharges and
the 1 C rate capacity, which agrees well with the 2/3 performance indicated that diffusion through the electrochemically active material
predicted by Equation 15. The 50 C rate optimized design has ∼10 X limited battery power performance. The accuracy of predictions could
increase in energy density compared to Battery 1 at 1,000 mW cm−2 be improved by accounting for material property changes, including
μm−1 power density and ∼10 X increase in power density at 10 μWh capacitive transport, and accounting for pore wall or surface film mass
cm−2 μm−1 energy density. The most significant power performance transport limitations. The microbattery geometry was related to the
improvements came from increasing the cathode width and optimiz- power performance using four design parameters, to , Wc , Ve , and Vs .
ing the active material thickness while reducing the pore size to 250 The microbattery geometry can be optimized for a desired output
nm. Figure 10d shows sensitivity analysis results for the 50 C rate op- power by setting the active material thickness to to , the cathode width
timized design when all combinations of the active material thickness to Wc , the pitch so Ve ≤ 0.1, and the electrode length so that Vs ≤ 0.1.
and electrode width are varied by ±10 and 25%. Microbatteries with A 50 C rate optimized microbattery based on the interdigitated bicon-
better power performance than the 50 C optimized design are to the tinuous architecture showed a ∼10 X improvement in power per-
right of the red line. A majority of the microbatteries have lower power formance over the experimental batteries, indicating that significant
performance than the 50 C rate optimized design. We conclude that improvements in power performance are possible with advances in mi-
the design parameters can be used to optimize microbattery geometry crobattery design and fabrication. The 1-D electrochemical model and
for maximum power performance. design parameters provide a new tool for developing novel material
The design optimization presented here is for a single porous anode architectures, nano composites, and full battery cells with improved
and cathode separated by electrolyte. Stacking many of these anode, power performance, which is critical for the maximum exploitation
cathode, and electrolyte sections together would increase the total of energy storage technologies in many applications. In addition, the
battery energy and power, but maintain the same energy and power performance limitations and design parameters developed in this work
density, and thus performance, as long as the electrodes are stacked are based on the same physical models used for macroscopic batteries;
so that the geometric properties that control transport (represented by therefore, these design principles might be applicable for improving
to , Wc , Ve , and Vs ) remain the same. the performance of macroscopic batteries.

Downloaded on 2017-11-09 to IP 190.111.30.254 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 164 (11) E3122-E3131 (2017) E3131

Acknowledgments References
This research was supported by the National Science Foundation 1. D. Golodnitsky, M. Nathan, V. Yufit, E. Strauss, K. Freedman, L. Burstein,
Engineering Research Center for Power Optimization of Electro Ther- A. Gladkich, and E. Peled, Solid State Ionics, 177, 2811 (2006).
2. T. S. Arthur, D. J. Bates, N. Cirigliano, D. C. Johnson, P. Malati, J. M. Mosby,
mal Systems (POETS) with cooperative agreement EEC-1449548 E. Perre, M. T. Rawls, A. L. Prieto, and B. Dunn, MRS Bulletin, 36, 523
and in part by the Department of Energy Office of Science Gradu- (2011).
ate Fellowship Program (DOE SCGF), made possible in part by the 3. J. W. Long, B. Dunn, D. R. Rolison, and H. S. White, Chemical Reviews, 104, 4463
American Recovery and Reinvestment Act of 2009, administered by (2004).
4. V. Narain, Global thin-film batteries (TFB) market, in, p. 1, Frost and Sullivan
ORISE-ORAU under contract no. DE-AC05-06OR23100. (2013).
5. J. H. Pikul, H. Gang Zhang, J. Cho, P. V. Braun, and W. P. King, Nat. Commun., 4,
List of Symbols 1732 (2013).
6. P. V. Braun, J. Cho, J. H. Pikul, W. P. King, and H. Zhang, Curr. Opin. Solid State
1-D One dimensional Mater. Sci., 16, 186 (2012).
A Area 7. C. Liu, E. I. Gillette, X. Chen, A. J. Pearse, A. C. Kozen, M. A. Schroeder,
b Interconnect diameter K. E. Gregorczyk, S. B. Lee, and G. W. Rubloff, Nat Nano, 9, 1031
(2014).
c Concentration 8. H. Ning, J. H. Pikul, R. Zhang, X. Li, S. Xu, J. Wang, J. A. Rogers, W. P. King,
Crate C rate of discharge and P. V. Braun, Proceedings of the National Academy of Sciences, 112, 6573
D Diffusion coefficient (2015).
Dc Polystyrene diameter corresponding to to 9. J. H. Pikul, J. Liu, P. V. Braun, and W. P. King, J. Power Sources, 315, 308
(2016).
DMC Dimethyl carbonate 10. Y. Tang, Y. Zhang, W. Li, B. Ma, and X. Chen, Chemical Society Reviews
EC Ethylene carbonate (2015).
F Faraday constant [96,485.3365 A s mol−1 ] 11. S. Lee, Y. Cho, H. -K. Song, K. T. Lee, and J. Cho, Angewandte Chemie International
FCC Face centered cubic Edition, 51, 8748 (2012).
12. S. Chen, Y. Xin, Y. Zhou, Y. Ma, H. Zhou, and L. Qi, Energy & Environmental
h, h2 Spherical cap height Science, 7, 1924 (2014).
H Electrode height 13. M. A. Martin, C. -F. Chen, P. P. Mukherjee, S. Pannala, J. -F. Dietiker, J. A. Turner,
i,I Current and D. Ranjan, J. Electrochem. Soc., 162, A991 (2015).
k Electrode conductivity 14. A. Shah, M. N. Ates, S. Kotz, J. Seo, K. M. Abraham, S. Somu, and A. Busnaina, J.
Electrochem. Soc., 161, A989 (2014).
l Length 15. Y. Tang, Y. Zhang, J. Deng, D. Qi, W. R. Leow, J. Wei, S. Yin, Z. Dong, R. Yazami,
L Electrode length Z. Chen, and X. Chen, Angewandte Chemie International Edition, 53, 13488
n Shape factor (2014).
OCV Open circuit voltage 16. A. S. Westover, D. Freudiger, Z. S. Gani, K. Share, L. Oakes, R. E. Carter, and
C. L. Pint, Nanoscale, 7, 98 (2015).
P Electrode pitch 17. X. Wang, D. Liu, Q. Weng, J. Liu, Q. Liang, and C. Zhang, NPG Asia Mater, 7, e171
PC Propylene carbonate (2015).
PITT Potentiostatic intermittent titration technique 18. H. Zhang, X. Yu, and P. V. Braun, Nat. Nanotechnol., 6, 277 (2011).
PS Polystyrene 19. M. Doyle, J. Newman, A. S. Gozdz, C. N. Schmutz, and J. -M. Tarascon, J. Elec-
trochem. Soc., 143, 1890 (1996).
q̇ Depletion rate (in electrolyte) 20. T. F. Fuller, M. Doyle, and J. Newman, J. Electrochem. Soc., 141, 1
R Polystyrene radius, or resistance (1994).
ρ Resistivity 21. M. Doyle and J. Newman, J. Power Sources, 54, 46 (1995).
SEI Solid electrolyte interphase 22. Á. G. Miranda and C. W. Hong, Applied Energy, 111, 681 (2013).
23. C. Min and G. A. Rincon-Mora, Energy Conversion, IEEE Transactions on, 21, 504
SEM Scanning electron microscopy (2006).
t Active material thickness 24. I. Papic, Energy Conversion, IEEE Transactions on, 21, 608 (2006).
τ Time 25. P. Arora, M. Doyle, A. S. Gozdz, R. E. White, and J. Newman, J. Power Sources, 88,
t+ Transference number 219 (2000).
26. O. Tremblay, L. A. Dessaint, and A. I. Dekkiche, in Vehicle Power and Propulsion
V Voltage or volume Conference, 2007. VPPC 2007. IEEE, p. 284 (2007).
W Electrode width 27. K. Smith and C. -Y. Wang, J. Power Sources, 161, 628 (2006).
Greek 28. M. Xiao and S. -Y. Choe, J. Power Sources, 218, 357 (2012).
29. J. N. Reimers, M. Shoesmith, Y. S. Lin, and L. O. Valøen, J. Electrochem. Soc., 160,
σ Conductivity A1870 (2013).
ε Volume fraction 30. D. N. Wong, D. A. Wetz, J. M. Heinzel, and A. N. Mansour, J. Power Sources, 328,
81 (2016).
Subscripts 31. Y. Tang, M. Jia, J. Li, Y. Lai, Y. Cheng, and Y. Liu, J. Electrochem. Soc., 161, E3021
(2014).
act Active material 32. S. De, P. W. Northrop, V. Ramadesigan, and V. R. Subramanian, J. Power Sources,
c Critical 227, 161 (2013).
33. D. Cericola, P. W. Ruch, R. Kötz, P. Novák, and A. Wokaun, J. Power Sources, 195,
e Electrolyte 2731 (2010).
i Electrode with the lowest total capacity 34. S. Buller, M. Thele, R. W. A. A. De Doncker, and E. Karden, Industry Applications,
n Sintered polystyrene IEEE Transactions on, 41, 742 (2005).
neg Negative electrode 35. B. Schweighofer, K. M. Raab, and G. Brasseur, Instrumentation and Measurement,
IEEE Transactions on, 52, 1087 (2003).
Ni Nickel 36. P. M. Gomadam, J. W. Weidner, R. A. Dougal, and R. E. White, J. Power Sources,
o Optimal 110, 267 (2002).
pos Positive electrode 37. K. Smith and C. -Y. Wang, J. Power Sources, 160, 662 (2006).
s Solid component 38. J. M. Sullivan, D. C. Hanson, and R. Keller, J. Electrochem. Soc., 117, 779
(1970).
sep Separator or region between electrodes 39. S.-i. Tobishima and T. Okada, Electrochim. Acta, 30, 1715 (1985).
v Central void 40. K. Xu, Chemical Reviews, 104, 4303 (2004).
void Non-solid pore space in between polystyrene particles 41. B. Klassen, R. Aroca, M. Nazri, and G. A. Nazri, The Journal of Physical Chemistry
B, 102, 4795 (1998).
Superscripts 42. M. Ue, J. Electrochem. Soc., 141, 3336 (1994).
43. N. J. Dudney, J. B. Bates, R. A. Zuhr, S. Young, J. D. Robertson, H. P. Jun, and
0 Initial condition S. A. Hackney, J. Electrochem. Soc., 146, 2455 (1999).
eff Effective 44. M. Winter and J. O. Besenhard, Electrochim. Acta, 45, 31 (1999).
max Maximum 45. J. Crank, The mathematics of diffusion, Clarendon press Oxford (1975).

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

Potrebbero piacerti anche