Next Article in Journal
An Overview on Co-Pyrolysis of Biodegradable and Non-Biodegradable Wastes
Previous Article in Journal
Stability Analysis of a Typical Salt Cavern Gas Storage in the Jintan Area of China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Simulation of the Thermal Runaway Onset in Li-Ion Cells—Influence of Cathode Materials and Operating Conditions

Department of Chemical, Materials and Environmental Engineering, “Sapienza” University of Rome, Via Eudossiana 18, 00184 Rome, Italy
*
Author to whom correspondence should be addressed.
Energies 2022, 15(11), 4169; https://doi.org/10.3390/en15114169
Submission received: 11 May 2022 / Revised: 28 May 2022 / Accepted: 1 June 2022 / Published: 6 June 2022
(This article belongs to the Section D2: Electrochem: Batteries, Fuel Cells, Capacitors)

Abstract

:
Li-ion batteries are already being used in several applications, from portable devices to the automotive industry, and they represent a promising option also for other critical uses, such as in the storage of energy from renewable sources. However, two of the main concerns that still hinder their massive introduction in these further areas, are their safety and reliability. Depending on cell characteristics and operating conditions, the heat generated within the cell can exceed that dissipated from its surface, and the cell will fail, possibly with catastrophic consequences. To identify the hazardous working conditions of a cell, a simulation model including the main exothermic reactions was set up to investigate the onset of thermal runaway in several Li-ion cell configurations under various operating conditions. The behavior of four different cathodes under thermal abuse and the influence of external factors such as the environmental temperature and the cooling system efficiency were assessed. It was found that among those investigated, the lithium iron phosphate cathode is characterized by a higher thermal stability and that an efficient superficial heat exchange can prevent thermal runaway in most of the cases.

1. Introduction

Nowadays, Li-ion batteries are commonly applied in a wide range of applications, from mobile devices (cell phones, laptops, etc.), to automotive (electric and hybrid electric vehicles) and even larger stationary applications in the power industry [1,2,3,4], where continuously increasing energy densities need to be stored. However, in order to allow a more stable and widespread use of these batteries, especially in critical applications requiring high levels of safety and reliability, some critical issues need still be addressed, most of them being associated with the occurrence of thermal runaway [5,6,7,8,9,10]. This phenomenon is associated with the internal generation of heat, due to a variety of exothermic reactions occurring within the cell, especially under abuse conditions (overcharge/overdischarge, thermal or mechanical abuse, and others). When this heat is not adequately dissipated toward the exterior of the cell, it causes an internal temperature rise and thus the acceleration of the exothermic reactions themselves or the ignition of additional reactions, with further uncontrolled overheating. This will finally lead to the failure of the cell, often catastrophic, with the generation of fires and explosions and the possible involvement of additional cells in a pack with even more severe consequences [7,11,12,13,14].
The identification of these hazardous conditions [7,8,9,10,15,16,17] is of the utmost importance to prevent the occurrence of thermal runaway and to design adequate protection systems and devices (PTC, BMS, etc.), thus allowing to prevent accidents or power shutdowns [13,18,19,20,21,22,23,24,25,26,27]. However, an efficient analysis requires detailed knowledge of the dynamic phenomena involved in the thermal runaway onset and its development.
The main components of a Li-ion battery are the cathode, the anode, and the electrolyte. At present, in most practical applications, the anode consists almost exclusively of graphite; the cathode usually contains a metal oxide, such as lithium cobalt oxide LiCoO2, lithium manganese oxide LiMn2O4, and lithium iron phosphate LiFePO4, and the electrolyte is usually based on a solvent containing a mixture of organic carbonates (mainly ethylene carbonate EC, diethyl carbonate DEC, and dimethyl carbonate DMC). However, there is a continuous search for better-performing and safer materials, so a wider combination of components is possible, making a generalized analysis more complex [28,29,30,31].
In order to have better knowledge of the failure modes of a Li-ion cell, both experimental and theoretical analyses have been carried out in the literature. Due to the difficulties of performing an extended experimental investigation, especially on a large scale, only a few field data are available in the literature [32,33,34,35,36,37,38,39,40], often referred to a limited number of system configurations (cell size, components, operating, and abuse conditions, and so on), or to the assessment of specific parameters [41,42,43,44,45,46,47].
A larger number of papers report theoretical analyses; however, due to the complexities of a rigorous mathematical representation of the system, several simplifying hypotheses are often introduced to make the calculation effort more reasonable [11,48,49,50,51,52,53].
It is generally recognized that most of the heat produced during thermal runaway is caused by the decomposition of the solid electrolyte interface (SEI), by reactions between the electrodes (anode and cathode) and the electrolyte, by the electrolyte decomposition, and by some other phenomena [11,54,55,56,57,58,59].
In the present paper, several configurations of cylindrical Li-ion cells subjected to abuse conditions of high current were analyzed, and their behavior was compared in terms of safe operational range, maximum temperature reached at a steady state, and operation time before the start of thermal runaway, as a function of operating conditions. In particular, four different cathodes were simulated, namely LiMn2O4 (LMO), LiFePO4 (LFP), LiNi0.4Co0.2Mn0.4O2 (NMC), and LiCoO2 (LCO), while graphite and LiPF6 lithium salt in 1:1 EC:EMC solvent were always assumed as anode and electrolyte, respectively. Each cell was simulated under continuous cycles of charge/discharge phases at 80 A/m2 and exposed to varying external cooling conditions. Unlike most previous studies, the physical properties of the materials involved were considered as a function of the changing temperature over time but were assumed constant in space. In addition, the direct comparison of several combinations of materials and external conditions, rather than focusing on single configurations, allowed to derive more clear conclusions in terms of thermal stability and safety of the cells.
The results allow to identify the optimal operating conditions to avoid a thermal runaway for each cell type, and their general trend can reasonably be extended to other configurations/materials, although the absolute values of the parameters involved (air temperature, flow rate, etc.) still need to be experimentally assessed. In addition to the analysis of single cells with different chemical characteristics, the case of a cell exposed to an external heat flux due to other neighboring cells experiencing a thermal runaway in a battery pack was also investigated.

2. Model Overview

Given the variety of interacting phenomena involved in the operation of a Li-ion cell, its modeling requires both a multiphysics and multiscale approach, and although this is already a simplified approach, it is still computationally demanding. In this paper, the model was implemented in COMSOL Multiphysics v5.5 [60,61], where both the electrochemical and the thermal processes were simultaneously represented making use of the Battery and Fuel Cell module. Additional information on modeling and a critical comparison of models can be found in the useful literature reviews [62].
The electrochemical model is based on the theory of the porous electrode: the current flows both in the solid electrode and across its pores containing the electrolyte. The equations required to describe this system are the materials balance in the solid and in the electrolyte, the charge conservation within the solid and the pores, and the kinetic reactions equations. The separator will be described by the materials balance and charge conservation for the liquid phase.
The transport of the Li particles within the electrode during charge and discharge is described by the Fick’s law in spherical form [63,64,65],
c s , i t = D s r 2 r r 2 c s r
with cs [mol/m3] and Ds [m2/s] as the lithium concentration and diffusivity, respectively.
At the particle center and on its external surface (particle radius Rp, [m]), the conditions of no flow and of local current density will apply:
c s , i r | r = 0 = 0
D s , i c s , i r | r = R p = i loc F
The transport of Li-ions within the electrolyte is governed by Fick’s diffusion law and electromigration. In addition, according to the electroneutrality assumption [66], it can be written as:
c l ε l t = D l , eff c l + a s 1 t + F i loc
with cl [mol/m3] as the concentration of Li-ions within the electrolyte; εl is the volume fraction of the electrolyte; as is the specific interface area (cm2/cm3); iloc is the local current density at the electrode (A/cm2), and t+ is the transference number of Li-ions with respect to the velocity of the solvent (−). The diffusion coefficient Dl,eff is derived from the Bruggeman relation:
D l , eff = ϵ l , i b D l
with the Bruggeman factor b usually equal to 1.5.
The boundary conditions represent a null flux, corresponding to no ion penetration at the interface of the current collectors, on the metal, on the external surfaces, and on the symmetry axis at r = 0:
c l r | r = 0 = 0
c l r | r = Rp = 0
c l z | z = Las = c l z | z = Lsc = 0
c l z | z = Lca = c l z | z = Lcc = 0
c l r | r = Rm = 0
Charge conservation within the solid electrode is governed by the equation
i s = a s i loc
where as is the specific interfacial area characterizing the electrode
a s = 3 ε s R p
with is as the current density in the solid phase, given by an Ohm expression:
i s = σ s , eff ϕ s
In the above equation, ϕs [V] is the potential at the electrodes, and σs,eff is the effective electric conductivity in the solid phase, which depends on the electric conductivity of the material and its porosity, according to the Bruggeman relation:
σ s , eff = ϵ s , i b σ s
A nocurrent condition is applied at the SEI
ϕ s z | z = Las = ϕ s z | z = Lsc = 0
while a symmetry condition for the potential of the solid phase for r = 0 and an isolation condition at the external surface of the electrode solid phase apply:
ϕ s r | r = 0 = 0
ϕ s r | r = Rp = 0
Within the electrolyte, the current flow is due to the Li-ion migration and diffusion
i l = σ l , eff ϕ l + 2 σ l , eff RT F 1 + lnf lnc l 1 t + lnc l
where ϕl [V] is the electrolyte potential; f represents the average molar coefficient of activity, which results in a concentration gradient due to the effect of the polarization potential, and (1 + ∂lnf/(∂lncl) is a thermodynamic factor.
A symmetry condition also applies for r = 0, and a null flux condition at the electrode–collector interface and on the external surface of the electrode separator.
ϕ s r | r = 0 = 0
ϕ l z | z = Lca = ϕ l z | z = Lcc = 0
The kinetics for the electrochemical reactions are represented by a Butler–Volmer equation, providing the following equation for the current density:
i loc , m = i 0 exp α 0 F η RT exp α c F η RT
where i0 is the exchanged current density; α0 the anodic transfer coefficient; αc is the cathodic transfer coefficient; F is the Faraday constant; T is the absolute temperature; R is the universal gas constant, and η is the overcharge:
η m = ϕ s ϕ l E eq , m
Eeq,m depends both on the temperature and on the state of charge of the cell:
E eq , m = E eq , i + U eq , i T T T ref
The current density is thus calculated as
i 0 = Fk 0 cl α 0 c s , max c s α a c s α c
With k0 as the kinetic constant and cs,max as the maximum Li concentration in the solid phase.
As far as the energy balance of the cell is concerned, it must be observed that heat generation inside the cell originates from three different sources: (1) heat generated by entropy variations during charge/discharge cycles, usually referred to as entropic reversible heat qrev = asilocT(∂Ueq)/∂T; (2) heat generated by irreversible electrochemical reactions, qirr = as iloc η; and (3) heat generated by electron transfer in the solid phase and by Li-ion migration and diffusion within the electrolyte. The latter term, often denoted as Ohmic heat qohm, is usually split into three different components: electric Ohmic heat qohm,s, ionic Ohmic heat qohm,i, and Ohmic heat due to an internal short circuit in the metal or the collector qohm,m.
The resulting overall balance equation is given by:
ρ c p T t = k 2 T + q rev + q irr + q ohm

3. Simulation Implementation

In order to optimize the calculation burden [40,49,63,67,68,69,70,71,72,73,74], the whole model was set up using two different geometries connected to each other to exchange the required information. One-dimensional geometry was used to characterize the basic chemical reactions of the specific type of cell (LCO, LMO, NMC, or LFP), and it ran in all simulations. Two-dimensional geometry was needed to simulate the thermal behavior of the cell; at each incremental step, it provided the temperature variations to the 1D component and then retrieved the physical properties calculated by the 1D geometry at the updated T.

3.1. One-Dimensional Component

The one-dimensional component handles the chemical characteristics of the cell, and to this end, use was made of the physical interface Lithium Ion Battery present in COMSOL Multiphysics. Three options associated with this interface were adopted to represent the different issues of the model:
  • Porous Electrode to define the charge balance for the electrodes and the electrolyte inside the pores;
  • Porous Electrode Reaction to define the charge transfer reactions at the interface between the electrodes and the electrolytes;
  • Separator to represent the separator properties (conductivity, diffusivity, etc.).
The geometrical dimensions of the one-dimensional component were assumed in accordance with previous analyses from the literature [75,76,77,78], and they are reported in Table 1.
As far as the chemical characterization of the cells is concerned, the main properties of the four cathodes adopted for the simulations were already present in the COMSOL library. Each cell was simulated for continuous cycles of charge/discharge phases, with a duration of 500 s for each full cycle, at a constant current of i = ± 80 A/m2, without any relaxation time (Figure 1).

3.2. Two-Dimensional Component

The two-dimensional component was used to represent the internal heat generation. The main items setting up a whole cylindrical cell are: (a) the mandrel, a low conductivity material supporting the active layers; (b) the external metal protective can; and (c) the internal active materials. The latter item consists of a series of alternating layers of cathode, anode, separator, etc. spirally wound around the mandrel. In order to simplify the calculations, the various active layers of a real cell were modeled as a single pseudo-homogeneous active material with constant average properties. This also allows the adoption of an axial symmetry on an x–y plane with respect to x = 0, as represented in Figure 2, where the main components of the cell are shown; their corresponding dimensions are reported in Table 2, while their physical properties are reported in Table 3 [75,76,77,78].
Given the structure of the active material, the thermal conductivity is anisotropic, and its radial and axial values can be derived by the following equations, respectively [11,77]:
k T , r = L i L i k T , i
and
k T , a = L i k T , i L i
where Li and kT,i are the thickness and the thermal conductivity of the single layers composing the active element.
Similarly, the average heat capacity and density were calculated as
ρ a = L i ρ i L i
and
c p , a = L i c p , i L i
The mandrel and the external were assumed to be composed of nylon and steel, respectively.
The physical interface Heat Transfer in Solids present in COMSOL was adopted to introduce the heat sources in the model. The heat generated by the Joule effect in the active material (Qh,3D [W/m3]) was based on the results obtained by the 1D model and integrated in the whole cell volume
Q h , 3 D = Q h , 1 D L an + L sep + L cat L cell r cell d can 2 r mandrel 2 h cell 2 d can r cell 2 r mandrel 2 h cell
The temperature distribution calculated by the 2D model was then recycled to the 1D model to update it and recalculate the generated heat and all other temperature-dependent parameters in a continuous circular exchange of information. Unlike most of the previous papers, all physical parameters here were calculated as a function of the variable temperature.
The other source of heat generation was associated with the exothermic reactions, which gradually occurred with increasing temperature, and were conventionally grouped in the following four categories [11,75,76,77,78]:
  • SEI decomposition in the range between 90 and 120 °C;
  • Reactions between anode and electrolyte at T > 120 °C;
  • Reactions between cathode and electrolyte at T > 170 °C;
  • Electrolyte decomposition at T > 200 °C.
The common approach of introducing an average value of the heat generated by the many, and often interacting, exothermic reactions [11,75,76,77,78] was also adopted here, and the specific values of the involved parameters are presented below.
The total exothermic heat was calculated as the sum of all the above contributions:
Q = Q SEI + Q anode + Q cathode + Q electrolyte
where for each of the above terms, an Arrhenius-like kinetics was assumed [11,32,39,77]
Q i x , t = q i R i x , t
R i x , t = A i c i m i x , t exp E a , i RT x , t
Adopting the Constant Fuel Model, the dependence on time and space can be neglected [79], and mi can be assumed as 1, so that the concentration will only depend on the temperature
c i x , t = c i , 0 ,   | t > 0 ,   x   ϵ   Ω i
Ωi represents the whole volume where the exothermic reactions occur.
The resulting equation for each exothermic reaction heat is thus:
Q i x , t = q i A i c i , 0 exp E a , i RT x , t
and the values of the kinetic parameters used in the simulations are reported in Table 4.

4. Model Results and Discussion

4.1. Single Cells

The results obtained by applying the present model to different single cells under several operating conditions are presented below. The model was preliminarily validated against the literature data comparing the trend of the thermal runaway ignition time (tTR) for different external heat exchange coefficients and different ambient/initial temperatures, adopting the LCO cathode configuration (Figure 3). It is apparent from Figure 3 that the agreement is quite good.
The results are also reported in Table 5 for the sake of completeness. As might be expected, the thermal runaway ignition time markedly increases at higher heat transfer coefficients (higher thermal stability of the cell), while it decreases at higher initial temperatures, also because of a reduced cooling effect by the warmer external environment.
In order to clearly highlight the influence of the exothermic reactions on the cell temperature, reference simulations were preliminarily carried out neglecting those reactions. Figure 4, Figure 5, Figure 6 and Figure 7 show the trend of the internal temperature as a function of time, for the LCO, LMO, NMC, and LFP configurations, respectively, with values of the external heat transfer coefficients in the range of 0–20 W/m2K, which correspond to typical stationary or low-velocity flow of gases or air (natural convection) [77,80,81] in the cooling system.
It can be seen that, as expected, a more or less pronounced continuous increase in the temperature with time was calculated, depending on the heat dissipation capacity of the cell; however, in all cases the cell operates for a relatively long-time interval, without experiencing thermal runaway. In particular, with high enough heat exchange coefficients, the cell temperature initially rises, but after some time (in the order of several hours), it reaches a stable value for the rest of the cell operation.
If the exothermic reactions are included in the model, a different behavior of the cell temperature is obtained: depending on the cell type and operating conditions, at a given time, a sudden dramatic increase in the temperature is observed, corresponding to the thermal runaway ignition.
In the following Figure 8, Figure 9, Figure 10, Figure 11, Figure 12 and Figure 13, the time profile of the cell temperature at different values of the heat transfer coefficient (h = 0–10 W/m2K) is reported; on each graph, the results for all the cathodes are shown to allow an easier comparison among them. In all cases, an initial ambient temperature of 293 K was assumed. As a general observation, the cell temperature almost linearly increases for some time until it reaches the ignition temperature of the thermal runaway, when the heat production cannot be controlled any longer by the external heat exchange, and the failure of the cell will occur [78]. For all cell types, this will happen earlier for lower heat exchange coefficients, with minimum values under adiabatic conditions (h = 0 W/m2K), while for higher values of h, longer operation times can be obtained; in particular, for h = 10 W/m2K, the produced heat is always adequately dissipated from the cell surface, and thermal runaway is prevented (Figure 13). This result confirms and extends to a larger number of chemistries the results reported by Melcher et al. [75] and Lopez et al. [77] for LCO cells. In the case of the LFP cell, this favorable condition was established already for h = 8 W/m2K, even though reaching a final temperature of 357 K vs. 347 K for h = 10 W/m2K. The latter considerations denote a reduced susceptibility of the LFP type of chemistry to runaway conditions with respect to the other ones. Conversely, from this point of view, the LCO cathode cell is always more prone to runaway conditions, with the ignition time for thermal runaway always lower and often much lower than those found for the other cathodes (often nearly half those for LFP). Similar to the present results and based on oven tests, Fouchard et al. [82] observed that LiMn2O4 cathodes offer better thermal stability than LiCoO2 or LiNiO2 cathodes. After the accelerating rate calorimetry (ARC) measurements on 18,650 Li-ion cells at different state of charges (SOCs), Mendoza-Hernandez et al. [83] found that LMO cells presented a higher thermal stability than LCO cells and also put forward the hypothesis that this might be due to the higher amount of oxygen possibly released by the latter cathode and thus available for combustion reactions. Using ARC data, Jiang and Dahn [84] also detected a much higher thermal stability of the LiFePO4 cathode, compared to the LiCoO2 and Li[Ni0.1Co0.8Mn0.1]O2 cathodes, especially in a LiBoB EC/DEC electrolyte.
Limiting the analysis to the conditions where a linear temperature increase is calculated (i.e., approximately up to h = 6 W/m2K), Figure 14 shows the overall influence of h on cell heating. As already mentioned, LCO cells always show a higher absolute value of the temperature rise (dT/dt) at any h, but the influence of a variation of h is very similar for all chemistries, as represented by the almost equal slopes of the fitting lines for the different cathodes.
As observed in the validation step (Figure 3), in addition to the heat exchange coefficient, ambient temperature also has an influence on the thermal behavior of the cell. Based on this, the thermal runaway ignition time was investigated for two additional ambient temperatures (namely, 273 and 313 K) for all chemistries. Temperature trends qualitatively similar to those reported in Figure 8, Figure 9, Figure 10, Figure 11, Figure 12 and Figure 13 were obtained so that the corresponding curves are not repeated here, but the ignition times are synthetically reported altogether in Table 6.
As might be expected, it is apparent that a lower temperature of the external environment will improve the thermal stability of the cells [75], increasing the thermal runaway ignition time and preventing it already for h < 8 W/m2K for all cells at 273 K (for the LFP cell, even less than 6 W/m2K is enough for a stable operation). This agrees well with the results obtained from oven tests by Lopez et al. [77] for an LCO cell.
In terms of cathode chemistry, it was also found that the LFP type is characterized by a higher thermal stability; the NMC cathode also behaves rather well, while the LCO is still characterized by a higher sensitivity to temperature increases.
At any temperature, tTR has an exponential-like trend (Figure 15) [75] so that beyond a given value of h its effect increases more markedly. It is hardly the case highlighting that a combination of a low temperature and a high heat transfer coefficient will provide the most beneficial conditions for a safe operation of the cells; for example, at 273 K and h = 10 W/m2K, the maximum constant value of the temperature ranges between 340 K (LCO) and 320 K (LFP).
It might be expected that a temperature gradient would establish within the cell during operation; based on this consideration and adopting a conservative approach, the maximum temperature calculated by the model was reported in the above figures. However, up to the onset of the thermal runaway, an almost homogeneous temperature distribution was actually calculated within the cell: Figure 16a,c report the internal temperature distribution within an LCO cell after one full charge–discharge cycle for h = 0 and h = 10 W/m2K, respectively, for the run at 313 K initial temperature.
It can be seen that a very small temperature variation is present within the active volume of the cell, no matter where the temperature is estimated. Onda et al. [85] calculated a maximum difference between the surface temperature and that at the center of the cell of less than 2 °C at the end of 3C discharge cycles for an LCO cell. Guo et al. [48] calculated small gradients in two directions of a prismatic LFP cell, whereas they estimated a difference of up to 30 °C between the center of the cell and its surface in the thickness direction; this was due to low thermal conductivity materials in that direction. Small variations of the internal temperature were also reported by Kim et al. [11] from simulations of a cobalt oxide cathode in an oven. Chen and Evans [12] calculated small temperature gradients also in large prismatic cells; however, Kim et al. [11] suggested that larger cells are “intrinsically more vulnerable to thermal runaway” because of their lower surface/volume ratio, “which in turn reduces cooling area per volumetric heat generation”.
It is worth highlighting here that all the above observations are applicable to “new” cells only, with optimal and constant physical properties, while due to several causes (such as aging, manufacture defects, etc.), unpredicted “hot spots” can develop within a cell, where the ignition of the thermal runaway can occur much earlier than expected and even under theoretically safe conditions [86,87,88,89,90].
Closer to the onset of the thermal runaway, larger temperature gradients are observed, with a variability in the order of 15–20 K in the active area (see Figure 16b,d). This finding would introduce some uncertainty in the definition of the ignition time, which would thus depend on the location of the adopted reference temperature; however, as seen above, the temperature increase in the proximity of the ignition time is so rapid that only small variations of tTR would derive.
From the data reported so far, it must be highlighted that in addition to the above considerations, an efficient cooling system brings also another very important beneficial result: besides delaying or even preventing thermal runaway, in any case, an efficient cooling process will lead to a lower average temperature of the cell so that in case of an accident (e.g., due to a mechanical impact or other temperature-unrelated causes), the consequences of the failure are expected to be milder than in case of a higher energy content of the cell, associated with its higher final temperature.

4.2. Propagation

Since in most of the applications batteries are grouped into large packs, the problem of a generalized thermal runaway, originated from a single cell failure and potentially involving adjacent cells, is present [67,91]. Given the influence of the external temperature highlighted above, in the present section, the simulation of a cell exposed to the heat generated by one or more adjacent cells was investigated. In such a condition, the heat entering the cell through the external surface can be calculated as
Q = h · T ext T
where Text was assumed at 420 K; T is the variable cell surface temperature, and h [W/m2K] is the external heat transfer coefficient with air. Several values were assumed in the literature for the latter parameter, commonly in the range of 4–10 W/m2K [75,81], in addition to the adiabatic condition h = 0 W/m2K; however, larger ranges were also adopted [75,77]. Here, h was calculated from fundamental parameters assuming natural convection with the external air around a cylindrical cell and calculating the nondimensional parameters of interest by the conventional relations for the Rayleigh (Ra), Prandtl (Pr), and Nusselt (Nu) numbers, respectively:
Ra = β Δ TgL 3 α ν
Pr = ν α
Nu = 0.825 + 0.387 Ra 1 / 6 1 + 0.492 Pr 9 / 16 8 / 27 2
β is the volumetric expansion coefficient (calculated as β = 1/T [K–1] for an ideal gas); ΔT is the temperature difference between cell surface and air temperature [K]; g is the acceleration of gravity [m/s2], and L is the cell height [m]. The kinetic viscosity ν [m2/s] and the thermal diffusivity α [m2/s] of air were calculated from
ν = μ ρ
α = k ρ c p
Assuming the values reported in Table 7 for the physical parameters [92], an average value of h = 7.5 W/m2K was obtained, which was adopted as a reference value for the following simulations.
As might be expected, because of the additional source of heat from the exterior, the ignition times for the cell thermal runaway are lower than those calculated above for an isolated cell. In Figure 17, the temperature–time profiles for such a system with all cathode types are shown.
It must be observed that in real conditions, the actual ignition times can be rather different from those calculated under the ideal and simplified configuration adopted in the present study. In fact, the initial temperature of the cell can be higher than the ambient temperature, depending on the operation modality of the cell, on the service time of the pack, on the number of cycles, etc. Furthermore, the number of cells already in runaway conditions can vary so that a nonhomogeneous distribution of the temperature around the analyzed cell can be generated. As a consequence, the values presented here can be assumed only as indicative and are to be compared with the previous results for a stand-alone cell.
For a direct comparison with the previous results, the onset times for the thermal runaway of a cell under an external heat flux are synthetically reported in Table 8 for all cathode materials, along with the values obtained for an isolated cell under adiabatic conditions (h = 0 W/m2K).
It can be seen that even when compared with the worst conditions previously investigated, i.e., no refrigeration with the external air in the presence of adjacent cells already under a thermal runaway, a cell will fail much more rapidly than predictable based on the internal heat generation only. In practice, once a thermal runaway has occurred in a given cell, it is likely that it will propagate to the whole pack in a short time. This conclusion is drawn on heat exchange considerations only; however, due to the consequences of the failure of a cell after a thermal runaway (i.e., a fire and/or an explosion), the propagation to the whole battery pack can occur even earlier than calculated in Table 8, thus further stressing the critical importance of thermal runaway prevention in most of the applications.
The effect of the incoming heat is apparent from Figure 18, where the temperature distribution inside a cell exposed to an external runaway, as outlined above, is reported.
Initially, the external surface of the cell is warmer than its inner part because of the exposure to a hotter environment; however, once the exothermic reactions have started inside the cell, this heat becomes more significant, and hot spots are generated in the inner volume, quickly leading to its failure.

5. Conclusions

In the present paper, a simulation model was set up to investigate the onset of thermal runaway in Li-ion cells under thermal abuse and various operating conditions. The model consists of both the electrochemical and the thermal submodels of the cell, thus including the exothermic reactions at the origin of the overheating of the cell. The adoption of interconnected interfaces allowed to properly take into account the variability of the involved parameters with the temperature and to discharge some simplifying assumptions, which were often assumed in the previous literature. The model provides the significant advantage of allowing to quantitatively assess the behavior of several types of cells under abuse conditions without the need for an experimental investigation.
To summarize the present conclusions, as far as the influence of the cell structure is concerned, it was found that the presence of a lithium–iron–phosphate cathode provides the cell with higher thermal stability than that assessed with the other cathodes. Conversely, a LiCoO2 cathode is characterized by much lower stability with the thermal runaway phenomenon occurring much earlier and for a wider range of conditions.
An efficient cooling system plays a decisive role in thermal runaway prevention, significantly increasing the operation time for all cells and making them stable beyond given values of the external heat transfer coefficient. As might be expected, the required cooling efficiency increases with decreasing intrinsic stability of the cell, and a synergic effect is also associated with the initial temperature of the cooling medium.
Based on the same considerations, it was also shown that a cell in a pack will experience thermal failure much more quickly than a single stand-alone cell if exposed to the heat flux due to the thermal runaway of one or more neighboring cells.

Author Contributions

M.C.: software, performing simulations, and validation; G.V.: resources; B.M.: methodology and supervision; R.B.: conceptualization, methodology, writing—original draft preparation, review, and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Berckmans, G.; Messagie, M.; Smekens, J.; Omar, N.; Vanhaverbeke, L.; Van Mierlo, J. Cost Projection of State of the Art Lithium-Ion Batteries for Electric Vehicles Up to 2030. Energies 2017, 10, 1314. [Google Scholar] [CrossRef] [Green Version]
  2. Dunn, J.B.; Gaines, L.; Kelly, J.C.; James, C.; Gallagher, K.G. The significance of Li-ion batteries in electric vehicle life-cycle energy and emissions and recycling’s role in its reduction. Energy Environ. Sci. 2015, 8, 158–168. [Google Scholar] [CrossRef]
  3. Park, J.K. Principles and Applications of Lithium Secondary Batteries, 1st ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2012. [Google Scholar]
  4. Santhanagopalan, S.; Smith, K.; Neubauer, J.; Kim, G.H.; Pesaran, A.; Keyser, M. Design and Analysis of Large Lithium-Ion Battery Systems; Artech House: Miami, FL, USA, 2015. [Google Scholar]
  5. Feng, X.; Ouyang, M.; Liu, X.; Lu, L.; Xia, Y.; He, X. Thermal runaway mechanism of lithium ion battery for electric vehicles: A review. Energy Storage Mater. 2018, 10, 246–267. [Google Scholar] [CrossRef]
  6. Wen, J.; Yu, Y.; Chen, C. A Review on Lithium-Ion Batteries Safety Issues: Existing Problems and Possible Solutions. Mater. Express 2012, 2, 197–212. [Google Scholar] [CrossRef]
  7. Wang, Q.; Ping, P.; Zhao, X.; Chu, G.; Sun, J.; Chen, C. Thermal runaway caused fire and explosion of lithium ion battery. J. Power Sources 2012, 208, 210–224. [Google Scholar] [CrossRef]
  8. Hendricks, C.; Williard, N.; Mathew, S.; Pecht, M. A failure modes, mechanisms, and effects analysis (FMMEA) of lithium-ion batteries. J. Power Sources 2015, 297, 113–120. [Google Scholar] [CrossRef] [Green Version]
  9. Soares, F.J.; Carvalho, L.; Costa, I.C.; Iria, J.P.; Bodet, J.M.; Jacinto, G.; Lecocq, A.; Roessner, J.; Caillard, B.; Salvi, O. The STABALID project: Risk analysis of stationary Li-ion batteries for power system applications. Reliab. Eng. Syst. Saf. 2015, 140, 142–175. [Google Scholar] [CrossRef] [Green Version]
  10. Bubbico, R.; Greco, V.; Menale, C. Hazardous scenarios identification for Li-ion secondary batteries. Saf. Sci. 2018, 108, 72–88. [Google Scholar] [CrossRef]
  11. Kim, G.H.; Pesaran, A.; Spotnitz, R. A three-dimensional thermal abuse model for lithium-ion cells. J. Power Sources 2007, 170, 476–489. [Google Scholar] [CrossRef]
  12. Chen, Y.J.W.; Evans, J.W. Thermal Analysis of Lithium-Ion Batteries. J. Electrochem. Soc. 1996, 143, 2708–2712. [Google Scholar] [CrossRef]
  13. Balakrishnan, P.G.; Ramesh, R.; Prem Kumar, T. Safety mechanisms in lithium-ion batteries. J. Power Sources 2006, 155, 401–414. [Google Scholar] [CrossRef]
  14. Chen, S.C.; Wan, C.C.; Wang, Y.Y. Thermal analysis of lithium-ion batteries. J. Power Sources 2005, 140, 111–124. [Google Scholar] [CrossRef]
  15. Zhang, L.; Zhao, P.; Xu, M.; Wang, X. Computational identification of the safety regime of Li-ion battery thermal runaway. Appl. Energy 2020, 261, 114440. [Google Scholar] [CrossRef]
  16. Mikolajczak, C.; Kahn, M.; White, K.; Long, F.T. Lithium-Ion Batteries Hazard and Use Assessment; Springer: New York, NY, USA, 2011. [Google Scholar]
  17. Zhang, Z.J.; Ramadass, P.; Fang, W. Safety of Lithium-Ion Batteries. In Lithium-Ion Batteries: Advances and Applications; Pistoia, G., Ed.; Elsevier: Amsterdam, The Netherlands, 2014; pp. 409–436. [Google Scholar]
  18. Kaliaperumal, M.; Dharanendrakumar, M.S.; Prasanna, S.; Abhishek, K.V.; Chidambaram, R.K.; Adams, S.; Zaghib, K.; Reddy, M.V. Cause and Mitigation of Lithium-Ion Battery Failure—A Review. Materials 2021, 14, 5676. [Google Scholar] [CrossRef] [PubMed]
  19. Xu, B.; Lee, J.; Kwon, D.; Kong, L.; Pecht, M. Mitigation strategies for Li-ion battery thermal runaway: A review. Renew. Sustain. Energy Rev. 2021, 150, 111437. [Google Scholar] [CrossRef]
  20. Jeevarajan, J. Safety of commercial lithium-ion cells and batteries. In Lithium-Ion Batteries; Pistoia, G., Ed.; Elsevier: Amsterdam, The Netherlands, 2014; pp. 387–407. [Google Scholar]
  21. Spotnitz, R.M.; Weaver, J.; Yeduvaka, G.; Doughty, D.H.; Roth, E.P. Simulation of abuse tolerance of lithium-ion battery packs. J. Power Sources 2007, 163, 1080–1086. [Google Scholar] [CrossRef]
  22. Chen, Y.S.; Hu, C.C.; Li, Y.Y. The importance of heat evolution during the overcharge process and the protection mechanism of electrolyte additives for prismatic lithium ion batteries. J. Power Sources 2008, 181, 69–73. [Google Scholar] [CrossRef]
  23. Kise, M.; Yoshioka, S.; Kuriki, H. Relation between composition of the positive electrode and cell performance and safety of lithium-ion PTC batteries. J. Power Sources 2007, 174, 861–866. [Google Scholar] [CrossRef]
  24. Jossen, A.; Spath, V.; Doring, H.; Garche, J. Reliable battery operation—A challenge for the battery management system. J. Power Sources 1999, 84, 283–286. [Google Scholar] [CrossRef]
  25. Al-Hallaj, S.; Selman, J.R. Thermal modeling of secondary lithium batteries for electric vehicle/hybrid electric vehicle applications. J. Power Sources 2002, 110, 341–348. [Google Scholar] [CrossRef]
  26. Sabbah, R.; Kizilel, R.; Selman, J.R.; Al-Hallaj, S. Active (air-cooled) vs. passive (phase change material) thermal management of high power lithium-ion packs: Limitation of temperature rise and uniformity of temperature distribution. J. Power Sources 2008, 182, 630–638. [Google Scholar] [CrossRef]
  27. Bubbico, R.; D’Annibale, F.; Mazzarotta, B.; Menale, C. Analysis of passive temperature control systems using PCM for application to secondary batteries cooling. J. Thermal Sci. Eng. Appl. 2018, 10, 061009. [Google Scholar] [CrossRef]
  28. Claus, D.; Besenhard, J.O. Handbook of Battery Materials; Wiley-VCH: Weinheim, Germany, 2011. [Google Scholar]
  29. Yuan, X.; Liu, H.; Zhang, J. Lithium-Ion Batteries Advanced Materials and Technologies; CRC Press, Taylor & Francis Group: Boca Raton, FL, USA, 2011. [Google Scholar]
  30. Nitta, N.; Wu, F.; Lee, J.T.; Yushin, G. Li-ion battery materials: Present and future. Mater. Today 2015, 5, 252–264. [Google Scholar] [CrossRef]
  31. Goodenough, J.B.; Kim, Y. Challenges for Rechargeable Li Batteries. Chem. Mater. 2010, 22, 587–603. [Google Scholar] [CrossRef]
  32. Hatchard, T.D.; MacNeil, D.D.; Basu, A.; Dahn, J.R. Thermal Model of Cylindrical and Prismatic Lithium-Ion Cells. J. Electrochem. Soc. 2001, 148, A755. [Google Scholar] [CrossRef]
  33. Abraham, D.P.; Roth, E.P.; Kostecki, R.; McCarthy, K.; MacLaren, S.; Doughty, D.H. Diagnostic examination of thermally abused high-power lithium-ion cells. J. Power Sources 2006, 161, 648–657. [Google Scholar] [CrossRef]
  34. Maleki, H.; Howard, J.N. Role of the cathode and anode in heat generation of Li-ion cells as a function of state of charge. J. Power Sources 2004, 137, 117–127. [Google Scholar] [CrossRef]
  35. Zeng, Y.; Wu, K.; Wang, D.; Wang, Z.; Chen, L. Overcharge investigation of lithium-ion polymer batteries. J. Power Sources 2006, 160, 1302–1307. [Google Scholar] [CrossRef]
  36. Waag, W.; Käbitz, S.; Sauer, D.U. Experimental investigation of the lithium-ion battery impedance characteristic at various conditions and aging states and its influence on the application. Appl. Energy 2013, 102, 885–897. [Google Scholar] [CrossRef]
  37. Roth, E.P.; Doughty, D.H.; Franklin, J. DSC investigation of exothermic reactions occurring at elevated temperatures in lithium-ion anodes containing PVDF-based binders. J. Power Sources 2004, 134, 222–234. [Google Scholar] [CrossRef]
  38. Kawamura, T.; Kimura, A.; Egashira, M.; Okada, S.; Yamaki, J.-I. Thermal stability of alkyl carbonate mixed-solvent electrolytes for lithium ion cells. J. Power Sources 2002, 104, 260–264. [Google Scholar] [CrossRef]
  39. Maleki, H.; Deng, G.; Anani, A.; Howard, J. Thermal Stability Studies of Li-Ion Cells and Components. J. Electrochem. Soc. 1999, 146, 3224–3229. [Google Scholar] [CrossRef]
  40. Forgez, C.; Do, D.V.; Friedrich, G.; Morcrette, M.; Delacourt, C. Thermal modeling of a cylindrical LiFePO4/graphite lithium-ion battery. J. Power Sources 2010, 195, 2961–2968. [Google Scholar] [CrossRef]
  41. Wang, Q.S.; Sun, J.H.; Chen, X.F.; Chu, G.Q.; Chen, C.H. Effects of solvents and salt on the thermal stability of charged LiCoO2. Mater. Res. Bull. 2009, 44, 543–548. [Google Scholar] [CrossRef]
  42. Watanabe, I.; Yamaki, J. Thermalgravimetry–mass spectrometry studies on the thermal stability of graphite anodes with electrolyte in lithium-ion battery. J. Power Sources 2006, 153, 402–404. [Google Scholar] [CrossRef]
  43. Yang, H.; Shen, X.D. Dynamic TGA–FTIR studies on the thermal stability of lithium/graphite with electrolyte in lithium-ion cell. J. Power Sources 2007, 167, 515–519. [Google Scholar] [CrossRef]
  44. Pasquier, A.D.; Disma, F.; Bowmer, T.; Gozdz, A.S.; Amatucci, G.; Tarascon, J.M. Differential Scanning Calorimetry Study of the Reactivity of Carbon Anodes in Plastic Li-Ion Batteries. J. Electrochem. Soc. 1998, 145, 472–477. [Google Scholar] [CrossRef]
  45. Orendorff, C.J.; Roth, E.P.; Nagasubramanian, G. Experimental triggers for internal short circuits in lithium-ion cells. J. Power Sources 2011, 196, 6554–6558. [Google Scholar] [CrossRef]
  46. Leising, R.A.; Palazzo, M.J.; Takeuchi, E.S.; Takeuchi, K.J. Abuse Testing of Lithium-Ion Batteries: Characterization of the Overcharge Reaction of LiCoO2/Graphite Cells. J. Electrochem. Soc. 2001, 148, A838–A844. [Google Scholar] [CrossRef]
  47. Ohsaki, T.; Kishi, T.; Kuboki, T.; Takami, N.; Shimura, N.; Sato, Y.; Sekino, M.; Satoh, A. Overcharge reaction of lithium-ion batteries. J. Power Sources 2005, 146, 97–100. [Google Scholar] [CrossRef]
  48. Guo, G.; Long, B.; Cheng, B.; Zhou, S.; Xu, P.; Cao, B. Three-dimensional thermal finite element modeling of lithium-ion battery in thermal abuse application. Three-dimensional thermal finite element modeling of lithium-ion battery in thermal abuse application. J. Power Sources 2010, 195, 2393–2398. [Google Scholar] [CrossRef]
  49. Chen, S.C.; Wang, Y.Y.; Wan, C.C. Thermal Analysis of Spirally Wound Lithium Batteries. J. Electrochem. Soc. 2006, 153, A637–A648. [Google Scholar] [CrossRef]
  50. Freitas, G.C.S.; Peixoto, F.C.; Vianna, A.S. Simulation of a thermal battery using Phoenics®. J. Power Sources 2008, 179, 424–429. [Google Scholar] [CrossRef]
  51. Zhang, X.W. Thermal analysis of a cylindrical lithium-ion battery. Electrochim. Acta 2011, 56, 1246–1255. [Google Scholar] [CrossRef]
  52. Jeon, D.H.; Baek, S.M. Thermal modeling of cylindrical lithium ion battery during discharge cycle. Energy Convers. Manag. 2011, 52, 2973–2981. [Google Scholar] [CrossRef]
  53. Kim, U.S.; Yi, J.; Shin, C.B.; Han, T.; Park, S. Modelling the thermal behaviour of a lithium-ion battery during charge. J. Power Sources 2011, 196, 5115–5121. [Google Scholar] [CrossRef]
  54. Spotnitz, R.; Franklin, J. Abuse behavior of high-power, lithium-ion cells. J. Power Sources 2003, 113, 81–100. [Google Scholar] [CrossRef]
  55. Bharathy, S.; Parimalam, A.D.; MacIntosh, R.; Brett, L. Lucht, Decomposition Reactions of Anode Solid Electrolyte Interphase (SEI) Components with LiPF6. J. Phys. Chem. C 2017, 121, 22733–22738. [Google Scholar]
  56. Kriston, A.; Adanouj, I.; Ruiz, V.; Pfrang, A. Quantification and simulation of thermal decomposition reactions of Li-ion battery materials by simultaneous thermal analysis coupled with gas analysis. J. Power Sources 2019, 435, 226774. [Google Scholar] [CrossRef]
  57. Wang, Q.S.; Sun, J.H.; Yao, X.L.; Chen, C.H. Thermal Behavior of Lithiated Graphite with Electrolyte in Lithium-Ion Batteries. J. Electrochem. Soc. 2006, 153, A329–A333. [Google Scholar] [CrossRef]
  58. Shigematsu, Y.; Ue, M.; Yamaki, J. Thermal Behavior of Charged Graphite and Lix CoO2 in Electrolytes Containing Alkyl Phosphate for Lithium-Ion Cells. J. Electrochem. Soc. 2009, 156, A176–A180. [Google Scholar] [CrossRef]
  59. Lee, S.Y.; Kim, S.K.; Ahn, S. Performances and thermal stability of LiCoO2 cathodes encapsulated by a new gel polymer electrolyte. J. Power Sources 2007, 174, 480–483. [Google Scholar] [CrossRef]
  60. COMSOL. Batteries & Fuel Cells Module Application Library Manual. © 1998–2016 COMSOL.
  61. COMSOL. Batteries & Fuel Cells Module User’s Guide. © 1998–2018 COMSOL.
  62. Hu, X.; Liu, W.; Lin, X.; Xie, Y. A Comparative Study of Control-Oriented Thermal Models for Cylindrical Li-Ion Batteries. IEEE Trans. Transp. Electrif. 2019, 5, 1237–1253. [Google Scholar] [CrossRef]
  63. Cai, L.; White, R.E. An Efficient Electrochemical–Thermal Model for a Lithium-Ion Cell by Using the Proper Orthogonal Decomposition Method. J. Electrochem. Soc. 2010, 157, A1188–A1195. [Google Scholar] [CrossRef]
  64. Santhanagopalan, S.; Guo, Q.Z.; Ramadass, P.; White, R.E. Review of models for predicting the cycling performance of lithium ion batteries. J. Power Sources 2006, 156, 620–628. [Google Scholar] [CrossRef]
  65. Smith, K.; Wang, C.Y. Power and thermal characterization of a lithium-ion battery pack for hybrid-electric vehicles. J. Power Sources 2006, 160, 662–673. [Google Scholar] [CrossRef]
  66. Moya, A.A.; Castilla, J.; Horno, J. Ionic Transport in Electrochemical Cells Including Electrical Double-Layer Effects. A Network Thermodynamics Approach. J. Phys. Chem. 1995, 99, 1292–1298. [Google Scholar] [CrossRef]
  67. Smith, K.; Kim, G.-H.; Darcy, E.; Pesaran, A. Thermal/electrical modeling for abuse-tolerant design of lithium ion modules. Int. J. Energy Res. 2010, 34, 204–215. [Google Scholar] [CrossRef]
  68. Guo, M.; White, R.E. A distributed thermal model for a Li-ion electrode plate pair. J. Power Sources 2013, 221, 334–344. [Google Scholar] [CrossRef]
  69. Christensen, J.; Cook, D.; Albertus, P. An Efficient Parallelizable 3D Thermoelectrochemical Model of a Li-Ion Cell. J. Electrochem. Soc. 2013, 160, A2258–A2267. [Google Scholar] [CrossRef]
  70. Allu, S.; Kalnaus, S.; Elwasif, W.; Simunovic, S.; Turner, J.A.; Pannala, S. A new open computational framework for highly-resolved coupled three-dimensional multiphysics simulations of Li-ion cells. J. Power Sources 2014, 246, 876–886. [Google Scholar] [CrossRef]
  71. Baba, N.; Yoshida, H.; Nagaoka, M.; Okuda, C.; Kawauchi, S. Numerical simulation of thermal behavior of lithium-ion secondary batteries using the enhanced single particle model. J. Power Sources 2014, 252, 214–228. [Google Scholar] [CrossRef]
  72. Lee, K.-J.; Smith, K.; Pesaran, A.; Kim, G.-H. Three dimensional thermal-, electrical-, and electrochemical-coupled model for cylindrical wound large format lithium-ion batteries. J. Power Sources 2013, 241, 20–32. [Google Scholar] [CrossRef]
  73. Al Hallaj, S.; Maleki, H.; Hong, J.S.; Selman, J.R. Thermal modeling and design considerations of lithium-ion batteries. J. Power Sources 1999, 83, 1–8. [Google Scholar] [CrossRef]
  74. Kim, U.S.; Shin, C.B.; Kim, C.S. Effect of electrode configuration on the thermal behavior of a lithium-polymer battery. J. Power Sources 2008, 180, 909–916. [Google Scholar] [CrossRef]
  75. Melcher, A.; Ziebert, C.; Rohde, M.; Seifert, H.J. Modeling and Simulation of the Thermal Runaway Behavior Li-Ion Cells-Computing of Critical Parameters. Energies 2016, 9, 292. [Google Scholar] [CrossRef] [Green Version]
  76. Melcher, A.; Ziebert, C.; Lei, B.; Zhao, W.; Luo, J.; Rohde, M.; Seifert, H.J. Modeling and Simulation of the Thermal Runaway in Cylindrical 18650 Lithium-Ion Batteries. In Proceedings of the 2016 COMSOL Conference, Munich, Germany, 12–14 October 2016. [Google Scholar]
  77. Lopez, C.F.; Jeevarajan, J.A.; Mukherjee, P.P. Mukherjee, Characterization of Lithium-Ion Battery Thermal Abuse Behavior Using Experimental and Computational Analysis. J. Electrochem. Soc. 2015, 162, A2163–A2173. [Google Scholar] [CrossRef]
  78. Xu, J.; Hendricks, C. A multiphysics Simulation of thermal Runaway in Large-Format Lithium-ion Batteries. In Proceedings of the 18th IEEE ITHERM Conference, Las Vegas, NV, USA, 28–31 May 2019. [Google Scholar]
  79. Frank-Kamentstskij, D.A. Diffusion and Heat Transfer in Chemical Kinetics; Plenum Press: New York, NY, USA, 1969. [Google Scholar]
  80. Roetzel, W.; Spang, B. C3 Typical Values of Overall Heat Transfer Coefficients. In VDI Heat Atlas; Springer: Berlin/Heidelberg, Germany, 2010. [Google Scholar]
  81. Huo, Y.; Rao, Z.; Liu, X.; Zhao, J. Investigation of power battery thermal management by using mini-channel cold plate. Energy Convers. Manag. 2015, 89, 387–395. [Google Scholar] [CrossRef]
  82. Fouchard, D.; Xie, L.; Ebner, W.; Megahed, S.A. Rechargeable Lithium and Lithium Ion (RCT) Batteries; PV, 94-28; Megahed, S.A., Ed.; Electrochemical Society: Miami Beach, FL, USA, 1994. [Google Scholar]
  83. Mendoza-Hernandez, O.S.; Ishikawa, H.; Nishikawa, Y.; Maruyama, Y.; Umeda, M. Cathode material comparison of thermal runaway behavior of Li-ion cells at different state of charges including over charge. J. Power Sources 2015, 280, 499–504. [Google Scholar] [CrossRef]
  84. Jiang, J.; Dahn, J.R. ARC studies of the thermal stability of three different cathode materials: LiCoO2; Li[Ni0.1Co0.8Mn0.1]O2; and LiFePO4, in LiPF6 and LiBoB EC/DEC electrolytes. Electrochem. Commun. 2004, 6, 39–43. [Google Scholar] [CrossRef]
  85. Onda, K.; Ohshima, T.; Nakayama, M.; Fukuda, K.; Araki, T. Thermal behavior of small lithium-ion battery during rapid charge and discharge cycles. J. Power Sources 2006, 158, 535–542. [Google Scholar] [CrossRef]
  86. Vetter, J.; Novák, P.; Wagner, M.R.; Veit, C.; Möller, K.C.; Besenhard, J.O.; Winter, M.; Wohlfahrt-Mehrens, M.; Vogler, C.; Hammouche, A. Ageing mechanisms in lithium-ion batteries. J. Power Sources 2005, 147, 269–281. [Google Scholar] [CrossRef]
  87. Wu, C.; Zhu, C.; Ge, Y.; Zhao, Y. A Review on Fault Mechanism and Diagnosis Approach for Li-Ion Batteries. J. Nanomater. 2015, 2015, 9. [Google Scholar] [CrossRef] [Green Version]
  88. Menale, C.; D’Annibale, F.; Mazzarotta, B.; Bubbico, R. Thermal management of lithium-ion batteries: An experimental investigation. Energy 2019, 182, 57–71. [Google Scholar] [CrossRef]
  89. Wang, Y.; Yan, X.; Bie, X.; Fu, Q.; Du, F.; Chen, G.; Wang, C.; Wei, Y. Effects of Aging in Electrolyte on the Structural and Electrochemical Properties of the Li[Li0.18Ni0.15Co0.15Mn0.52]O2 Cathode Material. Electrochim. Acta 2014, 116, 250–257. [Google Scholar] [CrossRef]
  90. Bodenes, L.; Naturel, R.; Martinez, H.; Dedryvère, R.; Menetrier, M.; Croguennec, L.; Pérès, J.P.; Tessier, C.; Fischer, F. Lithium secondary batteries working at very high temperature: Capacity fade and understanding of aging mechanisms. J. Power Sources 2013, 236, 265–275. [Google Scholar] [CrossRef]
  91. Abada, S.; Marlair, G.; Lecocq, A.; Petit, M.; Sauvant-Moynot, V.; Huet, F. Safety focused modeling of lithium-ion batteries: A review. J. Power Sources 2016, 306, 178–192. [Google Scholar] [CrossRef]
  92. Green, D.V.; Perry, R.H. Perry’s Chemical Engineers’ Handbook, 8th ed.; McGraw-Hill Professional Publishing: Blacklick, OH, USA, 2007. [Google Scholar]
Figure 1. Battery load vs. time.
Figure 1. Battery load vs. time.
Energies 15 04169 g001
Figure 2. Two-dimensional axial-symmetric geometry.
Figure 2. Two-dimensional axial-symmetric geometry.
Energies 15 04169 g002
Figure 3. Thermal runaway ignition time (tTR) as a function of h and ambient temperature for the LCO cathode. Solid lines = literature data [75]; dashed lines = present model results.
Figure 3. Thermal runaway ignition time (tTR) as a function of h and ambient temperature for the LCO cathode. Solid lines = literature data [75]; dashed lines = present model results.
Energies 15 04169 g003
Figure 4. Temperature profile for an LCO cell in the absence of exothermic reactions.
Figure 4. Temperature profile for an LCO cell in the absence of exothermic reactions.
Energies 15 04169 g004
Figure 5. Temperature profile for an LMO cell in the absence of exothermic reactions.
Figure 5. Temperature profile for an LMO cell in the absence of exothermic reactions.
Energies 15 04169 g005
Figure 6. Temperature profile for an NMC cell in the absence of exothermic reactions.
Figure 6. Temperature profile for an NMC cell in the absence of exothermic reactions.
Energies 15 04169 g006
Figure 7. Temperature profile for an LFP cell in the absence of exothermic reactions.
Figure 7. Temperature profile for an LFP cell in the absence of exothermic reactions.
Energies 15 04169 g007
Figure 8. Temperature profiles as a function of time (h = 0 W/m2K).
Figure 8. Temperature profiles as a function of time (h = 0 W/m2K).
Energies 15 04169 g008
Figure 9. Temperature profiles as a function of time (h = 2 W/m2K).
Figure 9. Temperature profiles as a function of time (h = 2 W/m2K).
Energies 15 04169 g009
Figure 10. Temperature profiles as a function of time (h = 4 W/m2K).
Figure 10. Temperature profiles as a function of time (h = 4 W/m2K).
Energies 15 04169 g010
Figure 11. Temperature profiles as a function of time (h = 6 W/m2K).
Figure 11. Temperature profiles as a function of time (h = 6 W/m2K).
Energies 15 04169 g011
Figure 12. Temperature profiles as a function of time (h = 8 W/m2K).
Figure 12. Temperature profiles as a function of time (h = 8 W/m2K).
Energies 15 04169 g012
Figure 13. Temperature profiles as a function of time (h = 10 W/m2K).
Figure 13. Temperature profiles as a function of time (h = 10 W/m2K).
Energies 15 04169 g013
Figure 14. Heating rate as a function of h.
Figure 14. Heating rate as a function of h.
Energies 15 04169 g014
Figure 15. Thermal runaway ignition time for different ambient temperatures and heat exchange conditions (LCO cathode).
Figure 15. Thermal runaway ignition time for different ambient temperatures and heat exchange conditions (LCO cathode).
Energies 15 04169 g015
Figure 16. Thermal distribution inside an LCO cathode cell: (a) after one complete charge–discharge cycle, h = 0 W/m2K; (b) at the thermal runaway ignition time, h = 0 W/m2K; (c) after one complete charge–discharge cycle, h = 10 W/m2K; and (d) at the thermal runaway ignition time, h = 10 W/m2K. Initial T = 313 K.
Figure 16. Thermal distribution inside an LCO cathode cell: (a) after one complete charge–discharge cycle, h = 0 W/m2K; (b) at the thermal runaway ignition time, h = 0 W/m2K; (c) after one complete charge–discharge cycle, h = 10 W/m2K; and (d) at the thermal runaway ignition time, h = 10 W/m2K. Initial T = 313 K.
Energies 15 04169 g016aEnergies 15 04169 g016b
Figure 17. Temperature profile as a function of time for a cell exposed to an external thermal runaway (Tex = 420 K; h = 7.5 W/m2K).
Figure 17. Temperature profile as a function of time for a cell exposed to an external thermal runaway (Tex = 420 K; h = 7.5 W/m2K).
Energies 15 04169 g017
Figure 18. Temperature distribution inside an LCO cell exposed to an external heat flux: (a) at the end of the first charging phase; (b) at the end of the first full charging-discharging cycle; and (c) at the ignition time.
Figure 18. Temperature distribution inside an LCO cell exposed to an external heat flux: (a) at the end of the first charging phase; (b) at the end of the first full charging-discharging cycle; and (c) at the ignition time.
Energies 15 04169 g018
Table 1. Cell’s component size.
Table 1. Cell’s component size.
L_anode55 μmNegative electrode thickness
L_separator30 μmSeparator thickness
L_cathode55 μmPositive electrode thickness
L_cc_negative7 μmNegative collector thickness
L_cc_positive10 μmPositive collector thickness
L_cell 1157 μmCell thickness
1 L_cell = L_anode + L_separator + L_cathode + L_cc_negative + L_cc_positive.
Table 2. Two-dimensional cell component size.
Table 2. Two-dimensional cell component size.
r_cell [m]9 × 10−3Cell radius
h_cell [m]6.5 × 10−2Cell height
r_mandrel [m]2 × 10−3Mandrel radius
r_connector [m]3 × 10−3Connector radius
h_connettor [m]3 × 10−3Connector height
d_can [m]0.25 × 10−3External can thickness
Table 3. Physical properties of cell’s materials.
Table 3. Physical properties of cell’s materials.
Materialk [W/mK]cp [J/kgK]ρ [kg/m3]Source
Copper298.153858933[75]
Aluminum1708752770[75]
Separator0.34419781009[75]
Graphite1.0414371347[75]
LCO1.487002500[77]
LMO1.5812692329[60]
NMC3.410002500[60]
LFP1.4812601500[77]
Table 4. Parameters for kinetic Equation (35) [54,75,77].
Table 4. Parameters for kinetic Equation (35) [54,75,77].
Heat   of   Reaction   q i   [ J kg ] Frequency   Factor A x [ 1 s ] Activation   Energy E A   [ J m o l ] Volume W y   [ k g m 3 ] Nondimensional Concentration
ci,0
SEI decomposition2.57 × 1051.667 × 10151.3508 × 1051.39 × 1030.15
Anode reactions 1.714 × 1062.5 × 10131.3508 × 1051.39 × 1030.75
Cathode reactions 3.14 × 1056.667 × 10131.396 × 1051.3 × 1030.0384
Electrolyte decomposition1.55 × 1055.14 × 10252.74 × 1055 × 1021
Table 5. Ignition time (s) at different h (W/m2K) and ambient temperatures.
Table 5. Ignition time (s) at different h (W/m2K) and ambient temperatures.
h (W/m2K)0246810
Tenv (K)
273.15 K2213266534085374
293.15 K18922172262434766179
313.15 K155017602026244531375379
333.15 K125213781525172720242513
353.15 K111311631278138816271798
373.15 K98310471115122913111474
Table 6. Thermal runaway ignition times for all cathodes at different ambient temperatures and heat exchange conditions.
Table 6. Thermal runaway ignition times for all cathodes at different ambient temperatures and heat exchange conditions.
Cathode MaterialInitial Temperature (K)tR (s)
h = 0 W/m2Kh = 2 W/m2Kh = 4 W/m2Kh = 6 W/m2Kh = 8 W/m2Kh = 10 W/m2K
LCO273.152212266434075373
293.1518922172262434766179
313.15155017602026244531375379
LMO273.152274277337046443
293.1521152579306438447519
313.15157817922083250332545394
NMC273.152313284838907656
293.1523302842378848218055
313.15161218222118258534676475
LFP273.15239430784664
293.152859356547336515
313.1517272015234630674967
Table 7. Physical parameters for air at 420 K [92].
Table 7. Physical parameters for air at 420 K [92].
β [K−1]2.38 × 10−3Volumetric expansion coefficient
L [m]0.065Characteristic dimension
μ [Pa·s]2.4 × 10−5Viscosity
ρ [kg/m3]0.84Air density
cp [J/kgK]827.8Specific heat capacity
k [W/mK]3.45 × 10−2Thermal conductivity
Table 8. Thermal runaway ignition times for different cell materials and conditions.
Table 8. Thermal runaway ignition times for different cell materials and conditions.
Cathode MaterialtR (s)
Text = 420 K,
h = 7.5 W/m2K
tR (s)
Text = 293 K,
h = 0 W/m2K
LCO7901892
LMO10272115
NMC11542330
LFP15222859
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cianciullo, M.; Vilardi, G.; Mazzarotta, B.; Bubbico, R. Simulation of the Thermal Runaway Onset in Li-Ion Cells—Influence of Cathode Materials and Operating Conditions. Energies 2022, 15, 4169. https://doi.org/10.3390/en15114169

AMA Style

Cianciullo M, Vilardi G, Mazzarotta B, Bubbico R. Simulation of the Thermal Runaway Onset in Li-Ion Cells—Influence of Cathode Materials and Operating Conditions. Energies. 2022; 15(11):4169. https://doi.org/10.3390/en15114169

Chicago/Turabian Style

Cianciullo, Martina, Giorgio Vilardi, Barbara Mazzarotta, and Roberto Bubbico. 2022. "Simulation of the Thermal Runaway Onset in Li-Ion Cells—Influence of Cathode Materials and Operating Conditions" Energies 15, no. 11: 4169. https://doi.org/10.3390/en15114169

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop