The following article is Open access

A Discussion of the Cell Voltage during Discharge of an Intercalation Electrode for Various C-Rates Based on Non-Equilibrium Thermodynamics and Numerical Simulations

Published 19 November 2019 © The Author(s) 2019. Published by ECS.
, , Citation Manuel Landstorfer 2020 J. Electrochem. Soc. 167 013518 DOI 10.1149/2.0182001JES

1945-7111/167/1/013518

Abstract

In this work we discuss the modeling procedure and validation of a non-porous intercalation half-cell during galvanostatic discharge. The modeling is based on continuum thermodynamics with non-equilibrium processes in the active intercalation particle, the electrolyte, and the common interface where the intercalation reaction occurs. The model is in detail investigated and discussed in terms of scalings of the non-equilibrium parameters, i.e. the diffusion coefficients and of the active phase and the electrolyte, conductivity and of both phases, and the exchange current density , with numerical solutions of the underlying PDE system. The current density i as well as all non-equilibrium parameters are scaled with respect to the 1-C current density of the intercalation electrode. We compute then numerically the cell voltage E as function of the capacity Q and the C-rate Ch. Within a hierarchy of approximations we provide computations of E(Q) for various scalings of the diffusion coefficients, the conductivities and the exchange current density. For the later we provide finally a discussion for possible concentration dependencies.

Export citation and abstract BibTeX RIS

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.

Lithium ion batteries (LIBs) are vital today for many branches of modern society and especially for electro-mobility. The german national platform electro-mobility aims one million electric vehicles by 2020, as well as the U.S., while China targets about five million zero emission cars. To achieve these goals, substantial knowledge on the effectively non-linear behavior of LiBs is required in order to reduce cost, increase their efficiency, safety, durability and further. The interpretation of experimental data requires a versatile and predictive mathematical model of a LIB, which accounts for the many physicochemical processes occurring simultaneously during charge and discharge, e.g. Li+ diffusion in the electrolyte, surface reactions at the electrode/electrolyte interface, solid state diffusion in the active particles, and electrical conductivity.

First academic steps to model the functional principle of LIBs with the purpose of simulating their charge/discharge behavior were carried out by Newman et al. around 1993.1 This electrochemical model became a central tool to interpret measured data of intercalation batteries. One of the central ingredients of the Newman model is the Butler–Volmer-type reaction rate for the intercalation reaction occurring at the interface between an intercalation electrode (particle) and the electrolyte . The actual functional dependency of on the different variables of the equation system, e.g. the electrolyte concentration , the electrostatic potential in the electrolyte, the concentration of intercalated ions, and the electrostatic potential of the active phase, is, however, rather stated then derived. Especially the so called exchange current density and its functional relationship to the cation concentration is doubtable.

From a non-equilibrium thermodynamics (NET) point of view, the functional dependency can be consistently derived and NET restricts this functional dependency in a very specific manner. We discuss in this work the modeling procedure of a single transfer reaction at the interface between an active intercalation phase and some electrolyte based on the framework of NET for volumes and surfaces and draw some conclusions regarding thermodynamic consistent models of the reaction rate. We account also for diffusion processes in the adjacent active particle and the electrolyte, as well es electrical conductivity, and state the corresponding balance equations. Then we consider galvanostatic discharge in half cell of some cathode intercalation material, electrolyte, and a lithium reference electrode, which is considered as ideally polarizable counter electrode.

We introduce the C1-current density, i.e. the current at which the electrode is completely discharged during one hour, and scale all non-equilibrium parameters based on the C-rate Ch, i.e. multiples of the C1 current density. It is then possible to derive a general relation between the measured cell voltage E, the capacity Q, and the C-rate Ch based on the reaction rate . Since, however, actually the concentrations at the interface of intercalated cations and electrolytic cations enter the surface reaction rate , we need to solve necessarily the diffusion equations in the adjacent phases. We discuss various approximation regimes and parameter scalings of the non-equilibrium parameters which allows us to compare numerical simulations of cell voltage E = E(Q, Ch) to some representative experimental examples, especially of (NMC). Fig. 1 shows the measured cell voltage E as function of the capacity (or status of charge) for various discharge rates of thin of NMC half cell.2

Figure 1.

Figure 1. Discharge curves (lower part) for various C-rates (Data of Fig 1b from Ref. 2, reprinted with permission of The Electrochemical Society).

We show that a rather simple (but thermodynamically consistent) model of the surface reaction rate , or more precise of the exchange current density, is sufficient to understand and predict the complex non-linear behavior of the cell voltage as function of the capacity Q and the C-Rate Ch. We provide also computations of E = E(Q, Ch) for the exchange current density introduced by Newman et al., draw some conclusions regarding thermodynamic consistency, and compare computations based on this expression to the cell voltage based on our simple expression of the current density.

Modeling

We consider an active intercalation particle in contact with some electrolyte . The interface captures the actual surface of the active particle as well as the electrochemical double layer forming at the interface, i.e. . The domains and are thus electro-neutral, and we refer to Refs. 35 for details on the derivation. The electrolyte is on the right side in contact to some metallic counter electrode , where the interface captures also the double layer forming at the interface between the electrolyte and the counter electrode .

We consider a 1D approximation, where the electrode-electrolyte interface is positioned at , the left boundary of is denoted by x = 0 and the right boundary of is , with and . The counter electrode is positioned at and spans to .

For some quantity u(x, t), we denote with

Equation ([1])

the evaluation at the respective side of the interface and . If u is present only on one phase, we drop the superscript ±.

The active particle is a mixture of electrons e, intercalated cations C and lattice ions M+, and the electrolyte a mixture of solvated cations C+, solvated anions A and solvent molecules S. The respective species densities are denoted with nα(x, t), x ∈ Ωi. We denote with

Equation ([2])

the chemical potential of the constituents, which are derived from a free energy density6,7 with of the active particle and of the electrolyte phase.

For the surface Σ we have surface chemical potentials4,6,8,9

Equation ([3])

which are derived from some general surface free energy density .

Material functions

For the electrolyte we consider exclusively the material model911 of an incompressible liquid electrolyte accounting for solvation effects, i.e.

Equation ([4])

with mole fraction

Equation ([5])

molar concentration nα, and total molar concentration of the mixture (with respect to the number of mixing particles9)

Equation ([6])

In (4) T denotes temperature, kB  the Boltzmann constant, gRα denotes the reference molar Gibbs free energy (or chemical potential of the pure substance), pR the reference pressure and vRα the partial molar volume of constituent α in the mixture. Throughout this manuscript we assume an isothermal temperature of .

Note that denotes the number of free solvent molecules, whereas and are the densities of the solvated ions. This is crucial for various aspects of the thermodynamic model, and we refer to Refs. 9,10,12,13 for details. Overall, the material model for the electrolyte corresponds to an incompressible mixture with solvation effects. We assume further

Equation ([7])

whereby the incompressibility constraint911 implies also a conservation of mass, i.e.

Equation ([8])

The molar volume of the solvent is related to the mole density of the pure solvent as

Equation ([9])

Note further that the partial molar volumes vRα and the molar masses mα of the cation and anion are related to the solvation number and , respectively.

We assume that the partial molar volume of the ionic species is mainly determined by the solvation shell, which seems reasonable for large solvents like DMC in comparison to the small ions like Li+. We proceed thus with the assumption

Equation ([10])

For the active particle, we consider an extension of a classical lattice mixture model1421 which accounts for occupation numbers as well as a Redlich–Kister type enthalpy term22,23 for the intercalation material Liy(Ni1/3Mn1/3Co1/3)O2 (NMC). We refer to Ref. 24 for a detailed discussion and derivation based on a free energy . The chemical potential of intercalated lithium is derived as

Equation ([11])

with

Equation ([12])

and mole fraction

Equation ([13])

of intercalated cations in the active phase. The number density of lattice sites is constant, which corresponds to an incompressible lattice, and the enthalpy parameter . Note that entails a phase separation20 and requires an additional term in the chemical potential. However, we assume throughout this work that no phase separation occurs, whereby in diffusional equilibrium of the intercalation phase the concentration is homogeneous. An extension of this discussion toward phase separating materials will given in a subsequent work.

For the electrons we consider9,25

Equation ([14])

and for the lattice ions

Equation ([15])

where is the molar volume of the lattice ions, the partial pressure and the constant molar Gibbs energy. The material functions of the active intercalation electrode is essentially model an incompressible solid with a sub-lattice for the intercalated cations .

The explicit surface chemical potentials

Equation ([16])

are not required throughout this work since we will assume that the double layer is in equilibrium and that the double layer capacity (and thus also adsorption), is negligible for the sake of this work. However, we refer to Ref. 9 for the explicit functions of and the surface free energy of a surface lattice mixture with solvation effects.

Electroneutrality condition

The electroneutrality condition of , and can be obtained by an asymptotic expansion of the balance equations in the electrochemical double layer at the respective surface Σ. We only briefly recapture the central conclusions and refer to Refs. 35,9,26 for details on the modeling, validation and the asymptotics. Most importantly, we have that

  • the double layer is in thermodynamic equilibrium, i.e. ∇μα + e0zα∇φ = 0 in and , where e0 is the elementary charge, zα the charge number and φ the electrostatic potential
  • there exists a potential drop between the active particle surface Σ and the hyper-surface outside of the respective space charge layers which is denoted by
    Equation ([17])
    where is the electrostatic potential right outside the space charge layer in the electrolyte or the active particle, respectively, and the (continuous) potential at the surface Σa. The whole potential drop across the double layer at is denoted by
    Equation ([18])
  • the chemical potential at the surface can be pulled back through the double layer, i.e.
  • the condition entails that the potential drop is constant (with respect to some applied voltage) and determined by
    Equation ([19])
  • the charge density in the electrolyte vanishes and that for monovalent electrolytes the cation mole fraction (or number density) is equal to the anion mole fraction, i.e.
    Equation ([20])
  • in the active phase the electroneutrality entails
    Equation ([21])
    whereby we abbreviate
    Equation ([22])
    which is basically the Fermi energy of the solid material.

Transport equations

In the electrolyte we have two balance equations determining the concentration (or the mole fraction ) and the electrostatic potential in the electrolyte,2732 i.e.

Equation ([23])

Equation ([24])

with (dimensionless) thermodynamic factor

Equation ([25])

where

Equation ([26])

is the thermodynamic driving force for diffusion,11 and the chemical diffusion coefficient, the cation transfer number, and the molar conductivity.

Note that we assumed and which yields the representation 26. Note further that the total number density in the electrolyte writes as

Equation ([27])

which is determined from the incompressibility constraint 8

Equation ([28])

and the electrolyte concentration in terms of as

Equation ([29])

If we consider a simple Nernst–Planck-flux relation for the cation and anion fluxes,11,33 respectively, i.e.

Equation ([30])

with constant diffusion coefficients for the anion and for the cation, we obtain (in the electroneutral electrolyte)

Equation ([31])

Equation ([32])

Note, however, for general Maxwell-Stefan type diffusion2932,34 or cross-diffusion coefficients7,24,35 in the cation and anion fluxes lead to more complex representations of the transport parameters . In general, three of the transport parameters are independent, and , and are related to each other via

Equation ([33])

Further, depend in general non-linearly on the electrolyte concentration . However, it is sufficient for the sake of this work to assume constant values for the transport parameters , together with relation 33.

In the active particle we have two balance equations determining the concentration (or mole fraction ) and the electrostatic potential in the active particle, i.e.

Equation ([34])

Equation ([35])

and (dimensionless) thermodynamic factor

Equation ([36])

Note that in principle can be dependent on the amount of intercalated ions, i.e. .

Reaction rate based on surface thermodynamics

We want to investigate the non-equilibrium thermodynamic modeling of the intercalation reaction

Equation ([37])

Surface thermodynamics dictates that the reaction rate of this process can in general be written as4,5,13,36,37

Equation ([38])

with α ∈ [0, 1]. Note that a non-negative function in 38 ensures a non-negative entropy production due to reactions on the surface, i.e. .

The quantity can be considered as surface affinity of the Reaction 37. The surface reaction rate vanishes when the affinity vanishes, which is the actually the thermodynamic equilibrium condition of 37, i.e. .

Since the electrochemical double layer is in equilibrium, we can pull back the surface chemical potentials through the double layer to the respective points (in an asymptotic sense) outside of the double layer, whereby we obtain for the surface affinity

Equation ([39])

With the material models 4 and 11 we can rewrite the surface affinity as

Equation ([40])

with

Equation ([41])

and

Equation ([42])

Equation ([43])

with according to 12. Note again that denotes the evaluation of at the interface and that the surface affinity 40 is dependent on the chemical potential (or the mole fraction) evaluated at the interface.

Cell Voltage

We consider the cell voltage in a half cell with metallic lithium as counter electrode, denoted by and position at (see Fig. 2. The cell voltage in such a cell is

Equation ([44])

where is the potential drop in the bulk active particle due to the electron transport, is the potential drop across the double layer at the interface between the active particle and the electrolyte, and the bulk potential drop due to cation electric current.

Figure 2.

Figure 2. Sketch of an active intercalation phase in contact with some electrolyte . The electrode-electrolyte interface covers the space charge layer of the electrolyte and of the electrode as well as the actual electrode surface Σ. Several processes occur simultaneously, i.e. the intercalation reaction, electrolyte diffusion and solid state diffusion as well es electrical conductivity.

We assume that the counter electrode is ideally polarizable,28 whereby the reaction

Equation ([45])

at the the interface positioned at is in thermodynamic equilibrium and . The equilibrium condition of 45 entails

Equation ([46])

Equation ([47])

where is the chemical potential of the metallic lithium.

For the surface affinity 40 we obtain the compact typeface

Equation ([48])

with

Equation ([49])

and

Equation ([50])

Current–Voltage relation

For the single intercalation reaction we have the following expression4

Equation ([51])

for the current density i flowing out of the electrode , where is the double layer capacity. Note that the reaction rate is

Equation ([52])

Since g(x) is a strictly monotone function, we can introduce the inverse of g, i.e. g− 1. For we have and . For values α ≠ 0.5 the inverse function g− 1 is only implicitly given, however, can easily be calculated numerically. Fig. 3 display the functions g and g− 1 for various values of α. We call g(x) the reaction rate function and g− 1 the inverse reaction rate function.

Figure 3.

Figure 3. Reaction rate function and its inverse g− 1 for various values of α.

Note that in the Tafel approximation Eq. 51 yieldsb

Equation ([53])

The term can be considered as the exchange current density.28

Onsager coefficient of the intercalation reaction

The Onsager coefficient (or the exchange current density ) of the surface reaction 37 could in principle be a function of the surface chemical potentials (or surface concentrations), i.e. or or the surface affinity, i.e. , as long as the condition is ensured.4,8,26 Note, however, that surface thermodynamics dictates the dependency of on the surface chemical potentials and not the bulk chemical potentials μα.

For a general relation we can pull back the surface chemical potentials through the double layer to obtain

Equation ([54])

Note that this necessarily restricts the functional dependency of on the mole fractions at the interface .

Consider, for example a model , where the exchange current density is dependent on the electrolyte concentration at the interface. This would be, however, thermodynamically inconsistent since the general functional dependency of 54 requires for the electrolyte concentration at the interface

Equation ([55])

Another commonly used model is a functional dependency of on the concentration of intercalated ions at the interface, i.e. . Since the space charge layer in the active particle is essentially constant (because is constant), we can indeed write

Equation ([56])

We discuss this aspect as well as various models for in Discussion of the exchange current density section. Meanwhile we assume and proceed the following derivation and the discussion based on this assumption since it turns out to be very reasonable.

Discussion of the model parameters

At this stage, it is illustrative to discuss the explicit value of the parameters.

  • For the electrode geometry we consider for a planar surface of area A and a thickness which yields and . The electrolyte is considered with a thickness of . This corresponds to the cell dimensions of the cell MX-6 in Ref. 2.
  • Throughout this work we consider DMC as solvent with and assume for the solvation number . The reference electrolyte concentration is and average amount of electrolyte is and a parameter of the model.
  • Average concentrations (or mole fractions) are abbreviated as
    Equation ([57])
    for the electrolyte species and
    Equation ([58])
    for the amount of intercalated ions in the active phase.
  • For the active particle phase we consider Li(Ni1/3Mn1/3Co1/3)O2 (NMC) whereby
    Equation ([59])
    which is simply computed from the density and stoichiometry of the bulk material.38 As parameters for the chemical potential we consider an occupation number of and a Redlich–Kister interaction energy of .24
  • The differential capacity has a prescribed value (actually is a function of , but we proceed here with a constant approximation for the sake of simplicity9 of about
    Equation ([60])
  • The electrode capacity Q is
    Equation ([61])
    This yields the non-dimensional capacity
    Equation ([62])
    which is sometimes also called status of charge (SOC) or depth of discharge (DOD).Note that during discharge of a complete battery the cathode is actually filled up with lithium. In a half cell with metallic lithium as counter electrode, discharge thus actually means filling up the intercalation electrode, here the NMC cathode material. Hence corresponds to a fully charged cathode (i.e. no lithium in the intercalation compound, ) while corresponds to a fully discharged cathode (i.e. the intercalation compound is completely filled with lithium, ).
  • From the charge balance 35 of the active particle we can deduce
    Equation ([63])
    where I is the current flowing into the intercalation electrode during discharge and Q(t = 0) the initial charge state. For a galvanostatic discharge I > 0 we obtain thus
    Equation ([64])
  • The C-Rate defines (implicitly) the current at which after h-hours the intercalation cathode is completely filled during galvanostatic discharge. C1 is thus the rate at which the battery is charged within one hour and commonly abbreviated just as C-rate C, i.e.
    Equation ([65])
    We can hence express the current I in multiples of the C-rate, i.e.
    Equation ([66])
    which yields
    Equation ([67])
    The only parameter for the current density i = I/A is thus Ch.
  • For the time t we consider the interval of one discharge cycle, i.e. t ∈ [0, tend] with
    Equation ([68])
    We can thus introduce the non-dimensional time
    Equation ([69])
    whereby the capacity rewrites as
    Equation ([70])
  • For the current density i at the planar electrode we have thus
    Equation ([71])
Discussion of the scaling

Consider the non-dimensional voltage

Equation ([72])

and abbreviate

Equation ([73])

which yields

Equation ([74])

with

Equation ([75])

The parameters and yield

Equation ([76])

and

Equation ([77])

whereby

Equation ([78])

The double layer contribution in Eq. 51 is thus almost negligible whereby 51 reduces to

Equation ([79])

We consider for the exchange current density the rescaling

Equation ([80])

This is the crucial decomposition throughout this work and the parameter of the surface reaction rate .

For the current density and the inverse function g− 1 we obtain thus with Eq. 48 for the surface affinity the general expression

Equation ([81])

for the cell voltage E.

Discussion

Throughout the manuscript, we assume that the initial state is completely uncharged, i.e. Q0 = 0 and . If not stated otherwise, we abbreviate

Equation ([82])

as well as the respective densities , fluxes , and chemical potential in the following.

We seek to discuss the general relation 81 of the cell voltage E as function of the capacity

Equation ([83])

during discharge of an intercalation electrode. Note that necessarily Ch > 0 (discharge) and (Onsager constraint of 38), whereby , which entails that any current decreases the cell voltage E during discharge.

We will discuss consecutively the following hierarchy of approximations:

  • BV 0: infinite slow discharge - the open circuit potential
  • BV 1: infinite fast diffusion and conductivity in the active particle and the electrolyte
  • BV 2: finite conductivity in the active particle, infinite diffusion in the active particle, infinite fast diffusion and conductivity the electrolyte
  • BV 3: finite conductivity and diffusion in the active particle, infinite fast diffusion and conductivity the electrolyte
  • BV 4: finite conductivity and diffusion in the active particle, finite conductivity in the electrolyte, infinite fast diffusion the electrolyte
  • BV 5: finite conductivity in the active particle and the electrolyte, finite solid state diffusion in the intercalation electrode as well as finite diffusion in the electrolyte

BV 0: Open circuit potential

The open circuit potential (OCP) is obtained from 81 as

Equation ([84])

for Ch = 0 (infinite slow discharge), which entails also as well as . Hence we have

Equation ([85])

For Liy(Ni1/3Mn1/3Co1/3)O239 as intercalation electrode, the two parameters of the chemical potential function are the occupation number and the interaction energy of the Redlich–Kister type enthalpy contribution. This yields an absolute ℓ2-error of and a relative error of 1.860% vs. experimental data of P. Bruce et al.,39 and Fig. 4 shows a comparison to two experimental data sets of measured OCP data.

Figure 4.

Figure 4. OCP of Liy(Ni1/3Mn1/3Co1/3)O2. Comparison between the material model 11 and experimental data of P. Bruce (Data of Fig. 3 in Ref. 39) and N. Nitta et al. (Data of Fig. 4e in Ref. 52).

BV 1: Infinite fast diffusion and conductivity in the active particle and the electrolyte

Infinite conductivity within the active particle phase as well as within the electrolyte yields

Equation ([86])

and infinite fast diffusion in the active particle and the electrolyte entails

Equation ([87])

Hence is directly related to the capacity via

Equation ([88])

whereby the cell voltage of BV 1 is

Equation ([89])

It is a simple algebraic relation between the measured cell voltage E, the C-rate Ch, the capacity Q and the (non-dimensional) exchange current density .

In order to compare the cell voltage E computed in the approximation BV 1 with other approximations, we abbreviate the voltage computed from 89 as E(1). Note that cell voltage 89 is actually independent of the electrolyte. For we obtain the voltage/capacity relation given in Fig. 5 for a variation of Ch from 0 (open circuit potential) to Ch = 100 (extremely fast discharge).

Figure 5.

Figure 5. Computed voltage E as function of the capacity according to Eq. 89 for various values of and Ch.

Reaction overpotential

We define the reaction overpotential as

Equation ([90])

which is actually independent of the status of charge or capacity. Measured voltage data would thus allow to determine and the parameter α ∈ (0, 1).

Fig. 6 shows computations of the reaction overpotential ηR for various values of α and as function of Ch.

Figure 6.

Figure 6. Reaction overpotential ηR as function of the C-Rate Ch with parameter variations of α and .

BV 2: Contribution of finite active phase conductivity

Finite conductivity within active particle phase entails from Eq. 35

Equation ([91])

Employing the scaling 71 of the current density i, i.e. , as well as the decomposition

Equation ([92])

yields

Equation ([93])

The quantity is the specific conductivity of the active particle phase at C-rate of one. For the parameters given in Discussion of the model parameters section computes as

Equation ([94])

The measured cell voltage is then

Equation ([95])

which is (yet again) a simple algebraic relation between E, the C-rate Ch, and the capacity . E(2) is additionally parametrically dependent on the conductivity .

We define the active phase conductivity overpotential as

Equation ([96])

BV 3: Contribution of the solid-state diffusion in the active particle phase

Constant diffusion coefficient

Reconsider that we have assumed yet with respect to space in the intercalation particle. In general, however, we have to solve a (here 1D) diffusion equation

Equation ([97])

with

Equation ([98])

This yields at the interface some solution

Equation ([99])

which will also impact the cell voltage

Equation ([100])

Equation ([101])

In order to discuss this impact systematically, we apply the following scaling

Equation ([102])

as well as

Equation ([103])

which leads to

Equation ([104])

The dimensionless flux

Equation ([105])

yields the dimensionless diffusion coefficient

Equation ([106])

and thus

Equation ([107])

At the interface we have thus

Equation ([108])

Overall we may write

Equation ([109])

with

Equation ([110])

Note that we can analytically compute from Eq. 11 (see also appendix B1) as

Equation ([111])

Since the problem 109 is non-linear, a classical separation Ansatz is not meaningful. We proceed thus with solving the problem 109 with 110 numerically with MATLAB and the pdepe() function. The syntax for pdepe() of the problem 109 with 110 is given in appendix B2.

Based on the numerical solution we compute then numerically for various values of Ch and . The (global) capacity is yet .

We assume the same parameters as before, now additionally with two values of the diffusion coefficient , i.e. slow diffusion and fast diffusion , and compute as function of the capacity (or time τ). Fig. 7 shows computations of for various discharge rates and diffusion coefficients in the active particle phase as function of the cell capacity. The angle bisection in black corresponds to the open circuit potential situation, where . For increasing discharge rates, the concentration at the interface is larger than the average concentration in since the evacuation of intercalated ions is delayed by the finite diffusion. This effect becomes even stronger for smaller values of , i.e. slow diffusion in the active particle.

Figure 7.

Figure 7. Concentration of intercalated ions at the interface as function the status of discharge for various values of Ch and .

The cell voltage E is then computed a posteriori from 100 based on the numerical solution of . Fig. 8 displays the cell voltage for various discharge rates as well as slow () and fast () diffusion in the intercalation phase. Finite diffusion in the active particle has an enormous impact on the cell voltage and changes qualitatively the shape due to the non-linear feedback. This effect is also found experimentally, see Fig. 1, and extremely important since it determines the maximum amount of charge that can be withdrawn from an intercalation electrode.

Figure 8.

Figure 8. Cell voltage E for BV 3 as function of the status of discharge for various values of Ch and with numerical computation of from the PDE 109 with boundary conditions 110.

Two important measures serve to discuss the impact of the diffusion coefficient ,

  • the cell voltage at 50% discharge, i.e. ,
  • and the capacity at the cut off voltage Eoff, here with .

Fig. 9 shows numerical computations of and for various values of the C-rate Ch and diffusion coefficients in the range of 10− 3 − 102. For slow discharge rates, i.e. Ch < 1 a diffusion coefficient of is sufficient to achieve a voltage of at 50% discharge and capacity of 90% at the the cutoff voltage. However, for higher C-rates, e.g. Ch = 50, the impact of the solid state diffusion becomes enormous, requiring a diffusion coefficient of to discharge the electrode to 50%.

Figure 9.

Figure 9. Cell voltage and Capacity for various discharge rates and diffusion coefficients. (a) Cell voltage at 50% state of discharge. (b) Capacity at the cutoff voltage Eoff = 2.6/V.

Overpotential

The overpotential due to finite diffusion in the active particle phase can be defined as

Equation ([112])

which computes as

Equation ([113])

Fig. 10 shows computations of for slow and fast diffusion.

Figure 10.

Figure 10. Overpotential as function of the status of discharge for slow () and fast () diffusion in the intercalation phase.

Concentration dependent diffusion coefficient

The diffusion coefficient in 34 was yet assumed to be a constantc with respect to the mole fraction of intercalated ions in the solid phase. This assumption might be inappropriate or over-simplified, and we seek to discuss this hypothesis again on the cell voltage E as function the capacity Q and the C-rate C. For this sake, we consider in the following a concentration dependent diffusion coefficient

Equation ([114])

where the constant is scaled equally than before, yielding again dimensionless diffusion coefficient . The term models that the diffusion on a lattice reduces when the lattice becomes more occupied (i.e. ).

We obtain hence the non-dimensionalized transport equation

Equation ([115])

with boundary conditions

Equation ([116])

This equation system is again solved numerically with MATLAB and the pdepe() function yielding numerical solutions whereby we compute then for various values of Ch and . The (global) capacity is yet . The cell voltage E is then computed a posteriori from 100 based on the numerical solution of .

We assume the same parameters as before and two values of the diffusion coefficient , i.e. slow diffusion and fast diffusion . Based on the numerical solution of 115 and 116 we compute as function of the capacity (or time τ), and subsequent the cell voltage E from 100 based on the numerical solution of .

Fig. 11 displays the cell voltage for various discharge rates computed numerically with a constant (dashed line) and the concentration dependent (solid line) diffusion coefficient according to 114.

Figure 11.

Figure 11. Cell voltage E for BV 3 as function of the status of discharge for various values of Ch and with numerical computation of for a constant diffusion coefficient (PDE 109 with boundary conditions 110) and a concentration dependent diffusion coefficient (PDE 115 with boundary condition 116).

Expectably, the two models behave similar when the intercalation electrode is empty, i.e. close zero. However, when the electrode gets filled with lithium (i.e. ), the -term of the concentration dependent diffusion coefficient reduces significantly the diffusivity, which affects the cell voltage in a surprisingly non-linear behavior.

Most important, Fig. 11 shows that the concentration dependent diffusion coefficient 114 could explain the decrease of the capacity at the end of discharge, e.g. here at , for increasing discharge current densities (c.f. the experimental data in Fig. 1). Wu et al. state in Ref. 2 that "the electrode capacity measured at the end of discharge decreases with increasing rate, as a result of transport limitations in the electrode". This can be confirmed on the basis of concentration dependent diffusion coefficient with

BV 4: Finite conductivity in the electrolyte

First note that an infinite fast diffusion in the electrolyte yet entails , whereby the (coupled) transport equation system 2324 of the electrolyte reduces to

Equation ([117])

which yields

Equation ([118])

Employing the scaling 71 of the current density i, i.e. yields

Equation ([119])

which motivates the decomposition

Equation ([120])

Here is a constant reference electrolyte concentration, e.g. , and is the corresponding reference conductivity. Hence

Equation ([121])

whereby the cell voltage is

Equation ([122])

Equation ([123])

Hence, finite conductivity in the electrolyte linearly decreases the cell voltage and scales also with the ratio of the electrode width to the electrolyte width, i.e. . The quantity accounts for concentration dependence of the electrolyte conductivity.

Correspondingly we define the electrolyte conductivity overpotential as

Equation ([124])

BV 5: Finite diffusion in the electrolyte phase

The final contribution to the surface reaction is the space dependent electrolyte concentration. We have yet assumed with respect to space, however, in general the (coupled) equation system 2324 has to be solved.

Note that simplifies the (coupled) equation system 2324 to

Equation ([125])

Equation ([126])

Further, entails at the interface the condition

Equation ([127])

We assume that the average electrolyte concentration is constant in time, i.e.

Equation ([128])

which yields at the right boundary the condition

Equation ([129])

The concept of an ideally polarizable counter-electrode , positioned at , delivers (or consumes) hence exactly the amount of ions in the electrolyte which flow in (or out) of at , with keeping the reaction 45 in thermodynamic equilibrium. The initial value is

Equation ([130])

where is the prescribed average electrolyte concentration.

We introduce the scalings

Equation ([131])

Equation ([132])

with and . This yields for 125

Equation ([133])

The corresponding non-dimensional boundary conditions at ξ = 0 reads

Equation ([134])

which introduces (implicitly) the scaling

Equation ([135])

leaving

Equation ([136])

The balance Equation 125 then reads

Equation ([137])

with

Equation ([138])

Note that the 2 in accounts for the charge of cations and anions. The charge capacity of a electrolyte is

Equation ([139])

whereby

Equation ([140])

The dimensionless transference number and . The PDE is solved with MATLAB's pdepe function, and details are given in the appendix A3. We denote the numerical solution of with and emphasize that the capacity is yet . The numerical solutions at the respective boundaries and are

Equation ([141])

We discuss now briefly the concentration distribution in the electrolyte as function of the C-rate Ch and the diffusion coefficient based on numerical solutions of 125 with boundary conditions 127 and 129.

Fig. 12 displays computations of the electrolytic cation concentration at the interface of the intercalation electrode, i.e. , and at the interface of the counter electrode, i.e. for slow electrolytic diffusion (, left) and fast diffusion (, right) for various values of the C-rate.

Figure 12.

Figure 12. Numerical computation of the cation interface concentrations and in the electrolyte for slow () and fast () diffusion and various C-rates Ch.

After a short time the concentration yields a stationary state and Fig. 13 displays the stationary concentration in the electrolyte, again for slow and fast diffusion as well as for various C-rates.

Figure 13.

Figure 13. Numerical computation of the stationary cation distribution in the electrolyte for slow () and fast () diffusion and various C-rates Ch.

Note, however, that the concentration variation of has additionally an impact on the voltage drop . First reconsider that 126 rewrites as

Equation ([142])

since

Equation ([143])

We abbreviate

Equation ([144])

and insert the scaling which yields

Equation ([145])

The overall cell voltage of BV 5 is then

Equation ([146])

Equation ([147])

In order to show the impact of the electrolyte concentration variation on the cell voltage E, assume as well infinite fast diffusion in the active particle phase . This yields

Equation ([148])

and numerical computations of the cell voltage for slow and fast diffusion are shown in Fig. 14.

Figure 14.

Figure 14. Computed cell voltage according to 148 with numerical solutions shown in Fig. 12 of the interface concentrations and for various discharge rates.

Overpotential

Due to the (stationary) concentration gradients in the electrolyte (c.f. Fig. 13) we have a diffusional overpotential , which can be defined as

Equation ([149])

Equation ([150])

Fig. 15 displays numerical computations of the overpotential for slow and fast diffusion in the electrolyte.

Figure 15.

Figure 15. Diffusional overpotential of the electrolyte for slow and fast diffusion computed from numerical solutions of the interface concentrations and and Eq. 149.

The recursive definition of the various overpotentials allows us to write

Equation ([151])

with one overpotential for each non-equilibrium process, measuring the deviation from the equilibrium of the respective process. This decomposition is hence a useful tool to systematically investigate the contribution of each process in broadly conceived experimental or numerical studies of a cell batch with varying parameters.

Internal resistance

Note that we can also compute the internal resistance Rint of the electrochemical cell via the implicit definition

Equation ([152])

With we obtain for the specific resistance

Equation ([153])

with

Equation ([154])

and

  • intercalation reaction resistance
    Equation ([155])
  • active phase diffusional resistance
    Equation ([156])
  • active phase conduction resistance
    Equation ([157])
  • electrolyte diffusional resistance
    Equation ([158])
  • electrolyte conduction resistance
    Equation ([159])

Conclusions

Validation

Equation 146 for the general relation for the cell voltage E in a simple, non-porous intercalation electrode. Note, however, that the scalings, discussion and parameter study of Discussion section can be straight forward adapted to porous electrodes.

We provide finally a validation study for NMC with the following set of the non-dimensional parameters

Equation ([160])

and a concentration dependent diffusion coefficient . In absolute values, these translate to:

  • exchange current density ,
  • NMC electric conductivity of ,
  • lithium diffusion coefficient in NMC ,
  • electrolyte conductivity ,
  • electrolyte diffusion coefficient

Figure 16 displays the numerical computation of the cell voltage. In comparison to experimental data for a cell of the same dimension (however, neglecting porosity), we obtain a good qualitative and quantitative agreement to Fig. 1.

Figure 16.

Figure 16. Computed cell voltage E as function of the capacity with parameters of the non-equilibrium processes according to 160.

This is especially remarkable since we assumed essentially for all non-equilibrium parameters constant values, i.e. no concentration dependence of the diffusion coefficient , the cation transference number , and the conductivities. In particular we assumed that the exchange current density is also constant, yielding the reasonable results of the last section. Note, however, that it is frequently assumed that the exchange current density is dependent on cation concentration at the interface . We discuss this aspect in the next section and emphasize again that a consistent thermodynamic modeling as well as coupling through the surface reaction rate yields the reasonable results of the last sections. The scaling of all non-equilibrium parameters to the C-rate is quite illustrative for the sake of galvanostatic discharge and especially for the systematic search of the parameters of a specific battery.

Discussion of the exchange current density

The preceding discussion of the cell voltage E was based on the model 38 of the surface reaction rate , i.e.

Equation ([161])

with In Discussion of the model parameters section we showed that double layer charging effects are negligible under galvanostatic conditions, whereby the measurable current density i is directly related to , i.e. . The surface affinity is related to the cell voltage E via 48, i.e.

Equation ([162])

with

Equation ([163])

Equation ([164])

Note that we have introduce the open circuit potential E(0) in BV 0: Open circuit potential section as

Equation ([165])

We can also evaluate the open circuit potential function E(0) with the interface concentration , i.e.

Equation ([166])

Note that this is a crucially different to 165 when finite diffusion in the active particle phase is considered, see BV 1: Infinite fast diffusion and conductivity in the active particle and the electrolyte section. However, this allows us rewrite the surface affinity as

Equation ([167])

Equation ([168])

with

Equation ([169])

and

Equation ([170])

This yields

Equation ([171])

This is the general, thermodynamic consistent version of the Butler–Volmer equation.4,26 The specific form 171 of the current density i in terms of the surface overpotential1 , the open circuit potential , and the exchange current density is widely employed in the literature28,40 and thus feasible to discuss various material models of .

In Refs. 1,41,42 as well as subsequent work we find

Equation ([172])

with η = Φ1 − Φ2, where Φ2 "is measured with a lithium reference electrode",1 p.1527 and Φ1 the electrostatic potential in the active phase. For the exchange current density iBV0 we find various models:

  • In Ref. 41 we find
    Equation ([173])
  • In Ref. 41 we find
    Equation ([174])
  • In Refs. 35,43 we find
    Equation ([175])

This model for the Butler–Volmer-reaction rate became a standard in the literature of modeling intercalation batteries4448 and is implemented in various software packaged to simulate battery cycles (i.e. COMSOL, Battery Design Studio,49 BEST35,43) as well as a basis for the interpretation of experimental data.50

We compare the Butler–Volmer Equation 172 to the surface reaction rate 171 and discuss the thermodynamic consistency of the three models 173175 for the exchange current density. Latz et al.,51 Bazant,21 and Dreyer et al.26 also point out the importance of thermodynamic consistency of the Butler–Volmer equation to achieve some overall predictive model since it couples the different thermodynamic bulk models. Dreyer et al.26 use the same structure 38 of the reaction rate, in combination with a constant exchange current density, however do not study intercalation electrodes.

First of all we mention again that the Butler–Volmer Equation 171 is derived from surface thermodynamics (see Reaction rate based on surface thermodynamics section) and that the exchange current density is the Onsager coefficient of the surface reaction Onsager coefficient of the intercalation reaction. This yields some necessary constraints on in terms of the functional dependency on the concentrations (or mole fractions) evaluated at the interface , which are discussed in Onsager coefficient of the intercalation reaction.

By comparison of Eqs. 171 and 172 we obtain

Equation ([176])

For a metallic lithium counter electrode, where the reaction

Equation ([177])

is in thermodynamic equilibrium we have 45 entails

Equation ([178])

which somehow justifies the interpretation of Φ2 as the potential "is measured with a lithium reference electrode",1 p.1527. However, from a thermodynamic point of view this re-definition of the potential is not necessary and could lead to inconsistencies when not applied in all balance equations (e.g. of the electrolyte transport) and boundary conditions of the intercalation battery model.

For the exchange current density we showed in Onsager coefficient of the intercalation reaction section that if is dependent on the concentrations at the interface , the dependency for the electrolyte species is necessarily

Equation ([179])

and for the intercalated ions in the active phase

Equation ([180])

or overall

Equation ([181])

with . Comparing these constraints with the models of the exchange current densities 173175 clearly shows that the dependency of iBV0 on the mole fraction (or concentration ) of the electrolyte concentration is not compatible with a reaction rate based on non-equilibrium surface thermodynamics. The concentration dependence is already embedded in the term of the surface affinity 167. A dependency of the exchange current density on is in principle compatible with surface thermodynamics. All three models propose

Equation ([182])

which in terms of the surface Onsager coefficient would be

Equation ([183])

In order to discuss the validity, predictability and finally the necessity (or non-necessity) of a concentration dependent surface Onsager coefficient (or exchange current density), we pursue the same strategy and scalings as in Discussion section, however, now with the model 183. We compute the cell voltage E as function of the capacity and the C-rate Ch in the hierarchy of approximations BV 1BV 5 and compare it to the computations based on the constant Onsager coefficient.

Eq. 51 reduces with negligible double layer contributions to

Equation ([184])

We consider again the scaling

Equation ([185])

which yields the cell voltage

Equation ([186])

Consider the approximation of infinite conductivity in both phases as well as infinite fast diffusion in the electrolyte, i.e. the approximation BV 3. Fig. 17 shows computations of cell voltage with constant exchange current density as well as concentration dependent exchange current density, for slow () and fast () diffusion in the active particle phase.

Figure 17.

Figure 17. Comparison of the compute cell voltage for the exchange current density according to Eq. 183 (–) and a constant surface Onsager coefficient for various C-rates and slow diffusion (left) as well as fast diffusion (right).

The impact of the model 183 for the Onsager coefficient (or the exchange current density) on the cell voltage is surprisingly small. Quite similar to the assumed concentration independence of the diffusion coefficients and we can conclude that is a rather good approximation for the overall modeling procedure. However, well defined and reproduce experimental data sets to compute absolute and relative model errors are rare throughout the literature and the deviations in 17 within the experimental variability. We conclude hence that the model 183 is in principle thermodynamically consistent, when embedded rigorously as stated in Modeling section, however, a constant exchange current density produces also very reasonable results and is thus the first choice.

Summary

In this work we discuss the cell voltage E of a non-porous intercalation half-cell during galvanostatic discharge with a continuum model for the active intercalation phase, the adjacent electrolyte, and boundary conditions coupling the phases. Based on non-equilibrium surface thermodynamics a reaction rate for the intercalation reaction is stated and the measured cell voltage E subsequently derived. We emphasize some necessary restrictions on the exchange current density of the surface reaction rate in terms of concentration dependence to ensure surface thermodynamic consistency.

For the detailed investigation of the non-equilibrium processes, scalings of all non-equilibrium parameters, i.e. the diffusion coefficients and of the active phase and the electrolyte, conductivity and of both phases, and the exchange current density of the intercalation reaction, with respect to the 1-C current density are introduced. The current density i, entering the model via the boundary conditions, is then expressed as multiple of , i.e. i = Ch · i, where Ch is the C-rate. Further we derive an expression for the capacity Q of the intercalation cell, which allows us to compute numerically the cell voltage E as function of the capacity Q for various C-rates Ch. Within a hierarchy of approximations, e.g. open circuit potential, infinite conductivity, infinite fast diffusion, and so forth, we provide simulations of E = E(Ch, Q) for various values of the (non-dimensional) parameters (), scaled with respect to the material constant . This provides an overall view of the processes and scalings within a lithium ion half cell which is validated at experimental data of LixNi1/3Mn1/3Co1/3O2(NMC).

List of Symbols

A Area of the electrode
Double layer capacity
C-rate
Active phase diffusion coefficient
Dimensionless diffusion coefficient in the active phase
Electrolyte diffusion coefficient
E Measured cell voltage according to 44
Elementary charge
Intercalation reaction energy
Electrolyte reaction potential (see 42)
Reaction rate function
g− 1 Inverse reaction rate function
Molar enthalpy function according to 12
i Current density flowing out of the electrode
1C current density.
I = A · i Measurable current
Current with which the battery is charged within one hour
Boltzmann constant
Dimensionless exchange current density
Onsager coefficient of the intercalation reaction
mα Molar mass of constituent α in solution
nα Molar concentration
Molar concentration of mixing particles in the electrolyte
Mole density of the pure solvent
Mass charge density of NMC
Volumetric charge density of NMC
Electrode capacity
Surface reaction rate of the intercalation reaction according to 38
Charge diffusion coefficient
Cation transfer number
Isothermal temperature
Potential drop across the double layer of
Space charge layer drop between the surface and the adjacent points outside the space charge layer
vRα Partial molar volume of constituent α in solution
Volume of the electrode
x = 0 Left boundary of in 1D approximation
Position of in 1D approximation
Right boundary of in 1D approximation
Mole fraction of intercalated cations in the active phase
Average mole fraction
Mole fraction (with respect to the number of mixing particles)
zα Charge number of constituent α

Greek

Solvation number of cations and anions in the electrolyte
Spatial domain of the active intercalation phase
Spatial domain of the electrolyte
Interface between active phase and electrolyte (including electrochemical double layers)
μα Chemical potential (function) for constituent
μα Surface chemical potential (function) for constituent
φ(x, z) Electrostatic potential
Thermodynamic factor electrolyte (see Eq. 25)
Molar conductivity
Thermodynamic factor active phase (see Eq. 36)
Active phase conductivity
Surface affinity (see Eq. 40)

Acknowledgments

The author acknowledges the financial support by the Federal Ministry of Education and Research of Germany (BMBF) in the framework mathematics for innovation (project number 05M18BCA). The author thanks the reviewers for their beneficial remarks.

ORCID

Manuel Landstorfer 0000-0002-0565-2601

: Appendix A: Electrolyte

Appendix. Mole fractions

We consider complete dissociation of the electrolyte and can thus express the mole fractions yα in terms of , i.e.

Equation ([A1])

Equation ([A2])

and according to the electroneutrality condition. Note we assume

Equation ([A3])

whereby

Equation ([A4])

Equation ([A5])

We can also express yα as function of of , i.e.

Equation ([A6])

Appendix. Thermodynamic factor

Equation ([A7])

Further

Equation ([A8])

whereby

Equation ([A9])

and thus

Equation ([A10])

Finally we have also

Equation ([A11])

Appendix. PDEPE syntax for the electrolyte phase

We want to solve numerically the problem

Equation ([A12])

with boundary conditions

Equation ([A13])

and

Equation ([A14])

Equation ([A15])

Equation ([A16])

Note that it is ever convenient for the numerical computation of to introduce the variable

Equation ([A17])

which yields

Equation ([A18])

The parameter a can be adjusted for numerical computations.

Correspondingly, we obtain

Equation ([A19])

Equation ([A20])

Equation ([A21])

and

Equation ([A22])

with

Equation ([A23])

This yields

Equation ([A24])

with boundary conditions

Equation ([A25])

Equation ([A26])

and

Equation ([A27])

The initial value is

Equation ([A28])

and transfers as

Equation ([A29])

PDEPE takes the form

Equation ([A30])

with boundary conditions

Equation ([A31])

We have hence

Equation ([A32])

Equation ([A33])

and

Equation ([A34])

Equation ([A35])

: Appendix B: Active Particle

Appendix. Thermodynamic factor

We consider for the chemical potential in the active particle phase

Equation ([B1])

with

Equation ([B2])

Hence

Equation ([B3])

with

Equation ([B4])

The thermodynamic factor is then

Equation ([B5])

Appendix. PDEPE notation

We seek to solve 109, i.e.

Equation ([B6])

with boundary conditions 110

Equation ([B7])

and

Equation ([B8])

Note that it is ever convenient for the numerical computation of to introduce the variable

Equation ([B9])

which yields

Equation ([B10])

We have hence

Equation ([B11])

and

Equation ([B12])

with

Equation ([B13])

PDEPE takes the form

Equation ([B14])

with boundary conditions

Equation ([B15])

We have hence

Equation ([B16])

Equation ([B17])

and

Equation ([B18])

Equation ([B19])

Note that we introduce the stop-event for the time-integration of pdepe.

Footnotes

  • Note that the continuity of φ across Σ is an assumption.

  • Note again that and that the space charge layer drop is constant due to the material model whereby .

  • Note, however, that the effective diffusion coefficient as pre-factor of in the flux relation 34 is inherently concentration dependent.

Please wait… references are loading.
10.1149/2.0182001JES