Department of Chemical and Biomolecular Engineering, University of California, Berkeley, CA, 94720

Similar documents
Development of First Principles Capacity Fade Model for Li-Ion Cells

Electrochemical Cell - Basics

Capacity fade studies of Lithium Ion cells

Effect of Intercalation Diffusivity When Simulating Mixed Electrode Materials in Li-Ion Batteries

Comparison of approximate solution methods for the solid phase diffusion equation in a porous electrode model

Lithium Batteries. Rechargeable batteries

Modeling the next battery generation: Lithium-sulfur and lithium-air cells

A Boundary Condition for Porous Electrodes

Modeling as a tool for understanding the MEA. Henrik Ekström Utö Summer School, June 22 nd 2010

ELECTROCHEMICAL SYSTEMS

Capacity fade analysis of a lithium ion cell

A Mathematical Model for a Lithium-Ion Battery/ Electrochemical Capacitor Hybrid System

Supporting Information. Fabrication, Testing and Simulation of All Solid State Three Dimensional Li-ion Batteries

Battery Design Studio Update

Distributed Thermal-Electrochemical Modeling of a Lithium-Ion Battery to Study the Effect of High Charging Rates

Multidimensional, Non-Isothermal, Dynamic Modelling Of Planar Solid Oxide Fuel Cells

Stress analysis of lithium-based rechargeable batteries using micro and macro scale analysis

Modeling lithium/hybrid-cathode batteries

FUEL CELLS in energy technology (4)

Parameter Estimation and Life Modeling of Lithium-Ion Cells

Experimental identification and validation of an electrochemical model of a Lithium-Ion Battery

Basic overall reaction for hydrogen powering

Electrochemical Impedance Spectroscopy of a LiFePO 4 /Li Half-Cell

III. Reaction Kinetics Lecture 15: Ion Adsorption and Intercalation

lect 26:Electrolytic Cells

Understanding Impedance of Li-Ion Batteries with Battery Design Studio

Advanced Analytical Chemistry Lecture 12. Chem 4631

Modeling of Liquid Water Distribution at Cathode Gas Flow Channels in Proton Exchange Membrane Fuel Cell - PEMFC

THE EFFECTS OF PULSED CHARGING ON LITHIUM ION BATTERIES

Overview of electrochemistry

Enhanced Lithium-Ion Mobility Induced by the Local Piezoelectric

Micro-Macroscopic Coupled Modeling of Batteries and Fuel Cells Part 1. Model Development

A Single Particle Thermal Model for Lithium Ion Batteries

Introduction to Dualfoil 5.0

Multi-physics Simulation of a Circular-Planar Anode-Supported Solid Oxide Fuel Cell

Effect of Varying Electrolyte Conductivity on the Electrochemical Behavior of Porous Electrodes

Computational model of a PEM fuel cell with serpentine gas flow channels

Basic Concepts of Electrochemistry

IV. Transport Phenomena. Lecture 23: Ion Concentration Polarization

Mikaël Cugnet, Issam Baghdadi, and Marion Perrin OCTOBER 10, Excerpt from the Proceedings of the 2012 COMSOL Conference in Milan

Modeling and Experimental Studies of Miniature Lithium Air Batteries and Alkaline Direct Ethanol Fuel Cells

Electrolytes for Innovative Li-Batteries

STRESS ANALYSIS OF THE SEPARATOR IN A LITHIUM-ION BATTERY

Studying On Capacity Fade Mechanisms of Li-Ion Batteries Through Modeling

Simulation and Analysis of Inhomogeneous Degradation in Large Format LiMn 2 O 4 / Carbon Cells

Supplementary Information

Ugur Pasaogullari, Chao-Yang Wang Electrochemical Engine Center The Pennsylvania State University University Park, PA, 16802

SCIENCES & TECHNOLOGY

FINITE ELEMENT METHOD MODELLING OF A HIGH TEMPERATURE PEM FUEL CELL

INTRODUCTION CHAPTER 1

Basic overall reaction for hydrogen powering

SUPPLEMENTARY INFORMATION

FUEL CELLS: INTRODUCTION

Supplementary Information Supplementary Figures

Tertiary Current Distributions on the Wafer in a Plating Cell

i i ne. (1) i The potential difference, which is always defined to be the potential of the electrode minus the potential of the electrolyte, is ln( a

Lecture 14. Electrolysis.

A chemo-mechanical model coupled with thermal effect on the hollow core shell electrodes in lithium-ion batteries

LITHIUM ION BATTERIES

Rechargeable Lithium-Air Batteries Using Mathematical Modelling

Experimental Validation of a Lithium-Ion Battery State of Charge Estimation with an Extended Kalman Filter

Battery System Safety and Health Management for Electric Vehicles

Electrical Conduction. Electrical conduction is the flow of electric charge produced by the movement of electrons in a conductor. I = Q/t.

Contents. 2. Fluids. 1. Introduction

Nanoscale Interface Control of High-Quality Electrode Materials for Li-Ion Battery and Fuel Cell

surface c, c. Concentrations in bulk s b s b red red ox red

Analysis of Charge Redistribution During Self-discharge of Double-Layer Supercapacitors

Fernando O. Raineri. Office Hours: MWF 9:30-10:30 AM Room 519 Tue. 3:00-5:00 CLC (lobby).

Solid State electrochemistry

CHM 213 (INORGANIC CHEMISTRY): Applications of Standard Reduction Potentials. Compiled by. Dr. A.O. Oladebeye

Figure 2: Simulation results of concentration by line scan along the x-axis at different partitioning coefficients after 0.48 ms.

Effect of electrode physical and chemical properties on lithium-ion battery performance

Amperometric biosensors

Modeling, Validation and Analysis of Degradation Processes of Lithium Ion Polymer Batteries. Rujian Fu

Nernst voltage loss in oxyhydrogen fuel cells

Computational Fluid Dynamics Modeling of a Lithium/Thionyl Chloride Battery with Electrolyte Flow

Supplemental Information. An In Vivo Formed Solid. Electrolyte Surface Layer Enables. Stable Plating of Li Metal

Tutorials : Corrosion Part 1: Theory and basics

Modeling the Effects of Ion Association on Alternating Current Impedance of Solid Polymer Electrolytes

Parameter Estimation and Model Discrimination for a Lithium-Ion Cell

Transport (kinetic) phenomena: diffusion, electric conductivity, viscosity, heat conduction...


Figure 1. Schematic of Scriber Associates Model 850C fuel cell system.

Diffusion. Spring Quarter 2004 Instructor: Richard Roberts. Reading Assignment: Ch 6: Tinoco; Ch 16: Levine; Ch 15: Eisenberg&Crothers

Principles and Applications of Electrochemistry

Journal of Power Sources

Electrochemistry objectives

Simulation of MEA in PEMFC and Interface of Nanometer-Sized Electrodes

Numerical Modeling of the Bistability of Electrolyte Transport in Conical Nanopores

Designing Thermal Management Systems For Lithium-Ion Battery Modules Using COMSOL. Emma Bergman

sensors ISSN by MDPI

Supporting Information for Dynamics of Lithium Dendrite Growth and Inhibition - Pulse Charging. Experiments and Monte Carlo Calculations

ONE DIMENSIONAL COMPUTER MODELING OF A LITHIUM-ION BATTERY

Electrochemical Cells

Performance Analysis of a Two phase Non-isothermal PEM Fuel Cell

Bruno Bastos Sales, Joost Helsen and Arne Verliefde

Oxygen Transfer Model in Cathode GDL of PEM Fuel Cell for Estimation of Cathode Overpotential

Review of temperature distribution in cathode of PEMFC

Hydrodynamic Electrodes and Microelectrodes

EEC 503 Spring 2009 REVIEW 1

Transcription:

Supporting Information for The Most Promising Routes to a High Li + Transference Number Electrolyte for Lithium Ion Batteries Kyle M. Diederichsen, Eric J. McShane, Bryan D. McCloskey* Department of Chemical and Biomolecular Engineering, University of California, Berkeley, CA, 94720 Energy Storage and Distributed Resources Division, Lawrence Berkeley National Laboratory, Berkeley, CA, 94720 *bmcclosk@berkeley.edu Dual Lithium Ion Insertion Cell Model Development Geometry. We implemented a 1-D model to simulate the galvanostatic charge of the Li x C 6 electrolyte Li y CoO 2 cell shown in Figure 1A using COMSOL Multiphysics 5.2a. We defined the LiCoO 2 cathode length to be 100 µm and determined the graphite anode length according to the expression below, ϵ s,neg L neg x =1.2 (1) ϵ s,pos L pos y c, c, where c, is the maximum theoretical solid phase lithium concentration in electrode material j, ϵ s,j is the active material volume fraction of electrode j, and L j is the length of electrode j. We assumed the anode, Li x C 6, operates with a lithium stoichiometric coefficient between 0 and 0.99 ( x = 0.99), and the cathode, Li y CoO 2, operates with a lithium stoichiometric coefficient between 0.50 and 1.00 ( y = 0.50). By implementing 20% excess anode material and setting a cutoff potential of 4.2 V, we avoided the onset of lithium deposition due to overcharging. 1

Separator Region. As is typically done when modeling transport through porous media, we adjusted the electrolyte diffusivity (D l ) and conductivity (σ l ) to an effective electrolyte diffusivity (D l,eff ) and conductivity (σ l,eff ) according to the Bruggeman correction, D l,eff =ϵ 1.5 D l (2) σ l,eff =ϵ 1.5 σ l (3) where ϵ is liquid phase volume fraction. Ion transport in the separator was modeled using transport equations derived from Newman s work 1 contained in the Batteries and Fuel Cells Module of COMSOL. Beginning with concentrated solution theory, we can express the velocity of each component (v i ) in terms of the electrochemical potential gradient ( µ i ) and concentration (c i ) of that component, c i µ i = K ij (v j -v i ) i j (4) where K ij is the frictional coefficient describing interactions between components i and j. Assuming a binary electrolyte and a solvent species with no net velocity, we can invert the above equation to arrive at an expression for the molar flux of the salt species in the liquid phase (N l ), N l =-D l,eff c l + i lt + F (5) where D l,eff is the effective diffusion coefficient of the salt, c l is the concentration of the salt in the liquid phase, i l is the liquid phase current density, F is Faraday s constant, and t + is the cation transference number as defined in the body of this manuscript. We note that c l is defined such that the cation and anion concentrations are equal due to electroneutrality. If we let υ + and υ - be the number of cations and anions, respectively, produced by the dissociation of one mole of the electrolyte salt, we can define c l as such: 2

c l = c + υ + = c - υ - (6) With an expression for the molar flux of the salt, the mass balance becomes ϵ l c l t + N l=r l (7) where R l is the cation source term, which only contributes in the porous electrode region. Finally, we employ an expression for the current density, 1 which may be used to solve for the electrical potential (ϕ l ), i l =-σ l ϕ l + 2σ lrt 1+ lnf 1-t F lnc + lnc l (8) l where σ l is the electrolyte conductivity, R is the ideal gas constant, T is the temperature in Kelvin, and f is the mean salt activity coefficient. Porous Electrodes. The salt flux and conservation equations for the porous electrode regions are analogous to those shown in the previous section. However, we must now additionally account for reactions which produce or consume lithium ions and the solid phase diffusion of lithium. We apply Butler-Volmer kinetics, manifested in the expression below for the local current density at the electrode-electrolyte interface (i loc ), i loc =i 0 exp α afη RT -exp -α cfη (9) RT where i 0 is the exchange current density, α i is the anodic or cathodic transfer coefficient, and η is the local overpotential. The exchange current density is defined generally for electrode j below, i 0 =Fk j c max s,j -c s,j αa c s,j αc c l α a (10) 3

where k j is the reaction rate constant and c s,j is the solid phase lithium concentration. Further, the local overpotential is defined as η=ϕ s -ϕ l -E eq (11) where ϕ s and ϕ l are the solid and liquid phase electrical potentials, respectively, and E eq is the equilibrium potential for the electrode material at a given state of charge. Empirical data is used to define E eq as a function of state of charge for both electrodes. 2, 3 We assume the electrode particles are spherical, and thus we are able to define the reaction source term (R l ), R l = 3ν Li+i loc r p F (12) where r p is the average electrode particle radius. The factor of 3/r p is the specific area of the spherical particles, used to correct the units on the source term. Upon reacting at the particle surface, the lithium must subsequently diffuse through the solid particle. The material balance for this process is c s t = D s c s (13) where c s is the lithium concentration in the solid particle and D s is the diffusion coefficient of lithium in the particle. This solid-state diffusion process is subject to the boundary conditions and c s r =0 (14) r=0 c s -D s r = ν Li+i loc r=r p F (15) where r p is the radius of the particle. 4

Initial Conditions and Boundary Conditions. We set the initial electrolyte concentration to 1 M. The initial electrode potentials were determined from the respective equilibrium potential functions according to the state of charge of each electrode, and the initial electrolyte potential was assumed to be equal to the initial anode potential. The system was assumed to be isothermal at 298 K. No flux boundary conditions were imposed on the salt at the two ends of the cell, and a current density (i set ) was specified at the cathode/current collector interface to set the C-rate. Assumptions. A few simplifying assumptions were necessary to enable a broad comparison of electrolytes of varying conductivity, transference number, and salt diffusion coefficient. 1. The binary salt liquid electrolyte was modeled with the following properties: =10mS/cm, t + =0.40, D l =3x10-6 cm 2 /s. These properties are consistent with those reported by Valo en and Reimers for a 1 M LiPF 6 electrolyte in propylene carbonate/ethylene carbonate/dimethyl carbonate. 4 2. A generalized activity coefficient concentration dependence was applied for all tested transport properties. The concentration dependence of the salt activity coefficient appears in the expression for current density. To incorporate this factor in our model, we used empirical activity coefficient data obtained for LiPF 6 in propylene carbonate/ethylene carbonate/dimethyl carbonate, 4 and we assumed that this activity coefficient data applied to all tested sets of σ, t +, and D l. In reality, this dependence will change if the solvent or salt are changed, but we assumed it was valid over all simulated transport properties for simplicity. 5

3. Conductivity, diffusion coefficient, and transference number were independent of salt concentration. The parameters that were varied between simulations (t +, D l, σ) were assumed to be independent of salt concentration. That is, the parameters did not vary across the length of the cell during a simulation. 4. The salt diffusion coefficient (D l ) was varied between simulations as t + was varied. To determine the salt diffusion coefficient for a given t +, we used the expressions below, 5 D l = c T c 0 1+ lnf lnc D 0+D 0- (z + -z - ) z + D 0+ -z - D 0- (16) z + D 0+ t + = (17) z + D 0+ +z - D 0- where c T is the total solution concentration (including solvent, cation, and anion), c 0 is the concentration of solvent, f is the mean molar salt activity coefficient, D 0+ and D 0- are the selfdiffusion coefficients of the cation (+) and anion (-) in the solvent (0), and z + and z - are the charge numbers of the cation and anion, respectively. We set D 0+ and D 0- to 2.5x10-6 cm 2 /s and 3.75x10-6 cm 2 /s, respectively, for the binary salt liquid electrolyte case in accordance with data presented by Valo en and Reimers. 4 We then held D 0+ constant at 2.5x10-6 cm 2 /s for all simulated t + and σ in the main manuscript, such that a new D l and D 0- were calculated for each tested t +. We additionally assumed that the factor c T /c 0 was equal to 1 for all simulations. 6

Figure S1: Schematic of dual lithium ion insertion cell with 1-D coordinate system shown. Li + Concentration Profiles. Lithium ion concentration gradients that arise during charge limit the utilization of active graphite material in the anode. At early times during charge, steep concentration gradients develop for electrolytes with low t +, and the concentration gradients become gradually less steep with increasing t +, becoming completely flat as t + approaches one. In Figure S2, lithium ion concentration gradients are shown for a cell charged at 2C with σ=10ms/cm and varying t +. At 50 seconds, the lithium ion concentration has already plummeted to 0.4 mol/dm 3 in the anode for the t + =0.40 case, making the back end of the anode (near x=0) relatively less utilized than the front end (near x=l neg ). In turn, the lithium ions continue to insert primarily at the front end of the anode (because lithium ions are more abundant there compared to the back end) despite the thermodynamic penalty incurred as the front end graphite becomes more lithiated. In contrast, higher t + electrolytes afford more even utilization of the anode material, as the lithium ion concentration gradients are less steep, resulting in a higher concentration of lithium ions towards the back end of the anode at short times. By 1350 seconds 7

(the end of the charge cycle for the t + =0.40 case), the lithium ion concentration profiles for the nonunity transference number cases converge, but the more even utilization of the anode afforded by higher t + electrolytes reduces the overpotential required during charging, allowing the high t + electrolytes to attain a higher SOC before reaching the cutoff voltage. Figure S2: A-C) Lithium ion concentration profiles for 2C charge rate with σ=10ms/cm and varying t + at A) 50 seconds B) 100 seconds and C) 1350 seconds. D) Local graphite state of charge with varying t + for 2C charge rate at 1350 seconds. 8

Graphite Porosity. We explored the effect of varying the graphite anode porosity (and adjusting the anode length accordingly to retain 20% excess anode capacity) on charge performance across a range of t + and σ. Although the attainable state of charge decreases with decreasing porosity for all tested t + and σ, the high transference number electrolytes decrease less relative to standard liquid electrolyte. This insight may be useful for optimizing cell weight, as less porous electrodes in combination with high t + electrolytes may prove economically superior despite the loss in attainable capacity. Figure S3: Attainable state of charge versus C-rate with σ=5ms/cm and varying t + using graphite anode porosity of A) 0.15, B) 0.20, C) 0.25, D) 0.30, and E) 0.35. Cutoff voltage was set to 4.2 V for all simulations, and the graphite electrode length was adjusted to retain 20% excess anode capacity. The liquid electrolyte was assumed to have σ=10ms/cm and t + =0.4. 9

Cutoff Voltage. We applied a cutoff voltage of 4.2 V for all simulations in the main manuscript, as we found that we could avoid lithium plating by doing so. If we raise the cutoff voltage to 4.3 V (see Figure S4), we observe a considerable improvement in charge performance, but we run the risk of plating lithium on the anode. Similar to Suthar et al., we define the overpotential for the plating reaction as η plating =Φ s x,t -Φ l x,t -E plating (18) where η plating is the overpotential for the lithium plating reaction, Φ s is the solid phase potential in the anode, Φ l is the electrolyte phase potential in the anode, and E plating is the equilibrium potential for the plating reaction, which is taken to be 0 V vs Li/Li +. 6 When η plating is negative, the plating reaction is thermodynamically favorable. We found that, in order to avoid a negative η plating with a 4.3 V cutoff voltage at 2C, a t + of 0.73 or above coupled with a σ of 10 ms/cm would be required. Figure S4: Attainable state of charge versus C-rate with σ=5ms/cm and varying t + using cutoff voltage of A) 4.2 V and B) 4.3 V. The liquid electrolyte was assumed to have σ=10ms/cm and t + =0.4. 10

Lithium Plating. As mentioned in the previous section, lithium plating can limit the operating voltage of batteries, and one must be especially diligent to avoid lithium plating at high charge rates. Below, we show a breakeven analysis which accounts for lithium plating. The breakeven curves in the main manuscript are superimposed with two curves representing the t + and σ needed to ensure η plating 10 mv and η plating 20 mv at every point in the anode throughout the entire charge cycle at 2C. Interestingly, low transference number electrolytes run a higher risk of lithium deposition, such that the low t +, high σ cases in the original breakeven analysis become unviable if we require a 20 mv safety window to avoid lithium plating. However, the high transference number cases in the original breakeven analysis remain viable even after implementing the 20 mv safety window. High t + electrolytes present an attractive option for suppressing lithium plating at high charge rates. Figure S5: Red triangles showing minimum σ and t + values required to achieve 75% state of charge at 2C with 4.2 V cutoff voltage. Black squares showing minimum σ and t + values required to remain 10 mv and 20 mv above plating overpotential of 0 V. Negative plating overpotential indicates thermodynamically favorable lithium plating. 11

LiCoO 2 Electrode Length. If we increase the LiCoO 2 electrode length from 100 to 150 µm (and adjust the graphite electrode length accordingly), we see that the charging performance suffers heavily, especially at high rates. However, the relative improvement offered by a high transference number electrolyte over a standard liquid electrolyte increases slightly as the electrode length increases. Figure S6: Attainable state of charge versus C-rate with σ=5ms/cm and varying t + using LiCoO 2 length of A) 100 µm and B) 150 µm. Graphite anode length was also varied to retain 20% excess anode capacity. The liquid electrolyte was assumed to have σ=10ms/cm and t + =0.4. Graphitic reaction rate constant. The charge performance is sensitive to the graphitic reaction rate constant used, yet this constant is difficult to measure accurately. If the rate constant is increased from the base case (1x10-9 m/s), only slight improvements in charge performance are observed. However, lower rate constants cause a precipitous reduction in attainable SOC, severely limiting the attainable capacity at rates above 2C. 12

Figure S7: Attainable state of charge versus C-rate with σ=5ms/cm and varying t + using graphitic reaction rate constant of A) 10-8 m/s, B) 10-9 m/s, C) 10-10 m/s, and D) 10-11 m/s. The liquid electrolyte was assumed to have σ=10ms/cm and t + =0.4. Lithium Self-Diffusion Coefficient, D 0+. The self-diffusion coefficient of the lithium ion (D 0+ ) was held constant at 2.5x10-6 cm 2 /s, corresponding to D 0+ for the binary salt liquid electrolyte case, for all simulations in the main manuscript. In Figure S8, we allow D 0+ to be reduced below 2.5x10-6 cm 2 /s (which may be expected in a concentrated polyelectrolyte solution or in an electrolyte with a more viscous solvent), and track the t + and σ needed to reach 75% SOC at 2C 13

with a 4.2 V cutoff voltage. As expected, slightly higher σ and t + are required to break even with the binary salt liquid electrolyte case (75% SOC at 2C) as D 0+ is decreased. Figure S8: Minimum σ and t + values required to achieve 75% state of charge at 2C with 4.2 V cutoff voltage using D 0+ of A) 2.5x10-6 cm 2 /s, B) 2.0x10-6 cm 2 /s, and C) 1.5x10-6 cm 2 /s. Table 1. Base case model parameters Symbol Value Units Reference L pos 100 µm assumed L sep 25 µm assumed L neg 91.8 µm calculated ϵ s,pos 0.62 assumed ϵ l,pos 0.2 assumed ϵ f,pos 0.18 assumed ϵ s,neg 0.65 assumed ϵ l,neg 0.25 assumed ϵ f,neg 0.10 assumed ϵ l,sep 0.39 assumed 14

c, 50051 mol/m 3 assumed c, 50051 mol/m 3 assumed c, 0 mol/m 3 assumed c, 31507 mol/m 3 assumed c 1 mol/l assumed r p,pos 8 µm assumed r p,neg 11 µm assumed D s,pos 5x10-8 cm 2 /s 7 D s,neg 3x10-8 cm 2 /s 8 σ s,pos 0.1 S/m 9 σ s,neg 100 S/m 9 k pos 6.67x10-11 mol -1/2 m -3/2 s -1 10 k neg 1x10-9 mol -1/2 m -3/2 s -1 10, 11 intermediate value α a 0.5 assumed α c 0.5 assumed T 298 K assumed V cut 4.2 V assumed D 0+ 2.5x10-6 cm 2 /s 4, calculated Nomenclature L ϵ c r D σ k α Length volume fraction, porosity concentration Radius diffusion coefficient conductivity reaction rate constant transfer coefficient 15

T V temperature voltage Subscripts pos neg sep s l f p a c cut cathode anode separator solid phase liquid phase filler (binder, carbon black) particle anodic cathodic cutoff 0+, 0- cation, anion with respect to solvent Superscripts 0 initial max maximum theoretical References (1) J. Newman and K. E. Thomas-Alyea, Electrochemical Systems, John Wiley & Sons, New York 2012. 16

(2) Tang, M.; Albertus, P.; Newman, J. Two-dimensional modeling of lithium deposition during cell charging. Journal of the Electrochemical Society. 2009, 156 (5), A390-A399. (3) Srinivasan, V.; Newman, J. Design and optimization of a natural graphite/iron phosphate lithium-ion cell. Journal of the Electrochemical Society. 2004, 151 (10), A1530-A1538. (4) Valo en, L. O.; Reimers, J. N. Transport Properties of LiPF 6 -Based Li-Ion Battery Electrolytes. J. Electrochem. Soc. 2005, 152 (5), A882. (5) Doyle, M.; Fuller, T.F.; Newman, J. The importance of the lithium ion transference number in lithium/polymer cells. Electrochimica Acta. 1994, 39 (13), 2073-2081. (6) Suthar, B.; Northrop, P. W. C.; Rife, D.; Subramanian, V. R. Effect of Porosity, Thickness and Tortuosity on Capacity Fade of Anode. Journal of The Electrochemical Society. 2015, 162 (9), A1708-A1717. (7) Thomas, M.; Bruce, P; Goodenough, J. B. Lithium mobility in the layered oxide Li1 xcoo2. Solid State Ionics. 1985, 17 (1), 13-19. (8) Persson, K.; et al. Lithium diffusion in graphitic carbon. The journal of physical chemistry letters. 2010, 1 (8), 1176-1180. (9) Park, M.; et al. A review of conduction phenomena in Li-ion batteries. Journal of Power Sources. 2010, 195 (24), 7904-7929. (10) Kumaresan, K.; Sikha, G.; White, R. E. Thermal model for a Li-ion cell. Journal of the Electrochemical Society. 2008, 155 (2), A164-A171. (11) Levi, M. D.; Aurbach, D. The mechanism of lithium intercalation in graphite film electrodes in aprotic media. Part 1. High resolution slow scan rate cyclic voltammetric studies and modeling. Journal of Electroanalytical Chemistry. 1997, 421 (1-2), 79-88. 17