Jihun Songa,
Royal C. Ihuaenyia,
Jaejin Limb,
Zihan Wangc,
Wei Lia,
Ruqing Fanga,
Amin Kazem Ghamsaria,
Hongyi Xuc,
Yong Min Lee*b and
Juner Zhu
*a
aDepartment of Mechanical and Industrial Engineering, Northeastern University, Boston, MA 02115, USA. E-mail: j.zhu@northeastern.edu
bDepartment of Chemical and Biomolecular Engineering, Yonsei University, Seoul 03722, Republic of Korea. E-mail: yongmin@yonsei.ac.kr
cDepartment of Mechanical Engineering, University of Connecticut, Storrs, CT 06269, USA
First published on 20th February 2025
Cell-level battery models, most of which rely on the successful porous electrode theories, effectively estimate cell performance. However, pinpointing the contributions of individual components of an electrode remains challenging. In contrast, particle-level models based on real microstructures describe active material characteristics but do not accurately reflect performance under cell-level operating conditions. To bridge this modeling gap, we propose a microelectrode modeling framework that considers each component of a composite electrode. This framework enables us to analyze the complex electrochemo-mechanical relationships within the composite electrode. The realistic 3D microstructure of the LiNi0.7Mn0.15Co0.15O2 composite electrode is reconstructed from focused ion beam-scanning electron microscopy images. By applying the intrinsic properties of every component, the composite microelectrode model achieves more than 98% accuracy in terms of the voltage profile compared to the measurement on coin cells. This model allows us to identify three important mechanisms that contribute to the discrepancy between cell and particle levels, i.e., reduced reaction area, increased diffusion length, and insufficient amount of electrolyte. Simulations under excessive electrolyte conditions reveal a significant improvement in rate capability with 94% capacity retention at 4C. In addition, the model considers the role of conductive materials and binders as well as the viscoplasticity of the polymeric binder, enabling the study of degradation mechanisms involving the stability of the binder-particle connection.
Broader contextLithium-ion batteries (LIBs) have been pivotal in advancing electric transportation but still face significant challenges in achieving higher energy densities, faster charging times, and longer cycle life. Addressing these challenges requires a comprehensive analysis of the factors affecting battery performance. To this end, a microelectrode modeling framework is proposed. This framework enables a thorough analysis of the complex electrochemical and mechanical interactions by considering each component within the composite electrode. The model identifies three primary mechanisms that impact battery performance: reduced reaction area, increased diffusion length, and insufficient electrolyte volume. To mitigate these issues, a design incorporating excessive electrolyte is proposed. Simulations indicate that this design not only significantly enhances rate capability but also exhibits excellent mechanical properties. Furthermore, the model provides valuable insights into degradation mechanisms by examining the roles of conductive and binder materials and the viscoplasticity of the polymer binder. |
From this perspective, extensive research has focused on analyzing the electrochemo-mechanical properties of active materials at the particle level. Techniques such as single particle measurement, scanning transmission X-ray microscopy (STXM), Bragg coherent diffractive imaging (BCDI), and nanoindentation have been employed for these investigations.4–7 Moreover, to overcome experimental constraints, modeling approaches for operando electrochemo-mechanical analysis have been developed, including phase field modeling and 3D reconstruction-based modeling that implement a realistic particle.8–11 These models assess the feasibility of fast charging and address critical factors such as cracks, critical factors affecting the degradation of high-nickel particles. However, particle-level studies often overlook the structural attributes of the composite electrode beyond the active material, complicating accurate predictions of overall cell behavior.
In contrast, cell-level modeling can directly measure cell performance but is too large in scale to accurately reflect the characteristics of all components. The battery electrodes are composed of millions to hundreds of billions of particles: a coin cell (∼1 A h) includes millions, a cylindrical cell (∼5 A h) contains tens of billions, and an EV pouch cell (∼100 A h) contains hundreds of billions. Thus, the preferred approach involves pseudo-2-dimensional (P2D) modeling based on the Doyle–Fuller–Newman model,12,13 which visually represents particles. To estimate the temperature effects and behavior of the cell, a P2D model is integrated with a 3D heat transfer modeling framework to develop a thermal-electrochemical model. This model calculates heat generation within the P2D framework and simulates heat transfer throughout the 3D framework.14,15 While these multiphysics models effectively predict cell performance under various scenarios through parameter adjustments, their virtual construction that ignores the microstructure of the electrodes limits the ability to analyze the specific characteristics of each battery component individually.
Binders are a component that is often overlooked in cell-level modeling. It binds the active materials, preventing delamination, and enhances the electrical conductivity of the composite electrode by combining it with conductive materials. During battery operation, binders undergo cyclic mechanical loads due to the continuous volume expansion and contraction of the active materials. These cyclic loads can potentially cause fatigue of the binder materials and the failure of the binder-particle interface, greatly shortening the battery lifespan.
Microelectrode-level modeling allows for the individual implementation of tens to hundreds of particles, conductive and binder materials (CBM), pores, and current collectors, bridging the gap between particle-level and cell-level modeling. Integration of advanced imaging technologies such as focused ion beam-scanning electron microscopy (FIB-SEM), X-ray microscopy (XRM), or nano-computed tomography (nano-CT) enables realistic representations of active materials, CBM, pores, and current collectors. This approach facilitates the analysis of structure-influenced electrochemical properties, including changes in effective exchange current density due to alterations in active surface area and effective diffusivity influenced by particle deformation during cycling.16–18 One of the fundamental challenges of furthering microelectrode-revolved models lies in the complex coupling between electrochemistry and mechanics. Most of the existing successful models are developed for small deformation, limited in simulating the physical contraction and expansion of active materials during cycling. This significantly affects both electrochemical and mechanical properties and reduces prediction accuracy. Also, these models only consider elastic deformation when analyzing CBM mechanical characteristics, neglecting the plastic deformation observed in real cells.
To address these challenges, we acquired hundreds of FIB/SEM images of the LiNi0.7Mn0.15Co0.15O2 (NMC711) composite electrode, capturing the detailed structure of active materials, CBM, pores, and current collectors. Leveraging these reconstructions, we developed an electrochemo-mechanical model at the microelectrode level capable of simulating structural changes within the composite electrode during cycling. Our investigation meticulously examines how these changes impact both electrochemical performance and mechanical characteristics. Furthermore, our model simulates the inelastic deformation within composite electrodes and predicts mechanical degradation in CBM based on viscoplastic deformation.
One of the effective ways to measure the properties of active materials is through particle-level experiments. In this study, we used the reported electrochemical and mechanical parameters of single NMC711 particles, measured through single particle measurements and nanoindentation (Fig. 1A), and applied these properties to the active material structures.9 On the other hand, we manufacture coin cells and directly investigate the electrochemical properties of the whole cell. These measurements can be compared with the model predictions to verify the parameters of each component of the composite electrode. To this end, we fabricated a cathode composite electrode with a weight ratio of NMC711:
PVdF
:
Super-P = 96
:
2
:
2 and conducted rate capability evaluations at 1C, 2C, 4C, and 8C rates using a half cell (Fig. 1B). By comparing the rate capabilities at both the particle and cell levels, we aim to understand the factors influencing the performance of the composite electrode.
A critical step in developing a composite microelectrode model is accurately representing the structure of the components. This involves generating the structures of all components, including the active materials, conductive additives, binders, and current collectors, and appropriately applying their electrochemical and mechanical properties. We reconstructed a 3D structure using 540 tomography images taken by FIB/SEM with a resolution of 43 nm (Fig. 1C). In the SEM images, the active materials, current collector, and pores are distinguishable. After image preprocessing, we set a threshold for grey values to form these structures. Nevertheless, due to the limitations of equipment precision and computational resources, it is almost impossible to accurately capture the exact shapes and distribution of conductive materials that range from sub-nano to tens of nanometers. Consequently, we combined the conductive materials and binders in the model and applied electrochemical and mechanical parameters to the combined CBM domain.
The mechanical properties of binder film have only been reported by a few experimental studies.24,25 Our study utilized the reported stress–strain curve at different strain rates (0.00003 s−1–0.003 s−1) to derive Young's modulus of 1.05 GPa and yield strength of 19.36 MPa (Fig. 1D). It is worth noting that these values are measured using dried binder film, so they do not accurately reflect the change in mechanical properties influenced by the electrolyte. Among various electrolyte effects, the most important is probably the binder swelling induced by electrolyte impregnation. We observed this phenomenon during the first tens of hours after cell manufacturing. The essential electrochemical property of CBM for model development is electronic conductivity, as there is no electrochemical reaction but current flow through the CBM. Although some studies have measured the electronic conductivity of CBM, the swelling caused by electrolyte impregnation and structural deformation during cycling alter the electronic conductivity. Due to the lack of additional physical measurements, this effect has not been included in our model. We applied an average measured value for CBM's electronic conductivity.26
Compared with cell-level models, the microstructure-resolved model enhances the accuracy by directly defining the governing equations on each component (Fig. 1E), avoiding homogenization and approximations. Conventional cell-level models commonly approximate porosity, tortuosity, and active surface area using effective values, which hinders the accurate prediction of electrochemo-mechanical properties during cycling. Our model fundamentally eliminates the necessity for these approximations and can account for structural changes that occur during cycling. More detailed information is available in the Modeling methodology section.
When the C-rates were increased to 2C (30 minutes), 4C (15 minutes), and 8C (7.5 minutes), the particle showed overpotentials of 0.0142 V, 0.0244 V, and 0.0405 V, respectively, at x = 0.24 in LixNi0.7Mn0.15Co0.15O2, while the coin cell exhibited 18–20 times higher overpotentials of 0.2538 V, 0.4505 V, and 0.8234 V, respectively. The capacities of the coin cell were decreased by 23.89%, 89.23%, and 96.95% compared to the equilibrium potential at 2C, 4C, and 8C, respectively. This demonstrates the difficulty in fully charging within 15 minutes at the cell level, although the particle results indicate the possibility of fast charging. Therefore, to achieve fast charging at the cell level, we need to closely analyze and bridge the gap between particle and cell performance. In coin cells, the overpotentials are significantly affected by both the anode and the cathode. In this study, since lithium metal is used as the anode and its overvoltage is assumed to be constant, the analysis is focused on the cathode.
In this regard, we used microelectrode modeling of the cathode to determine the factors affecting fast charging and conducted an in-depth analysis. From a structural perspective, the higher initial overpotential of the coin cell compared to the particle can be attributed to differences in the electron pathways of the active material and the active surface area which is the contact area between the electrolyte and the active material. The specific active surface areas (surface area/volume) for the particles and the microelectrode are 2476
784 m2 m−3 and 1
720
752 m2 m−3, respectively, a difference of 30.52%.
For the composite electrode of the coin cell, both the CBM and current collector must be considered. Given that CBM contains 12% electrolyte, we assumed that 12% of the interfaces between active materials and CBM are active surface area and the remaining 88% of the interfaces (8.30 × 10−9 m2) are non-reactive. Also, the interfaces of active materials and the current collector (7.71 × 10−10 m2) are non-reactive. By subtracting these non-reactive surface areas from the total area of active materials (7.62 × 10−8 m2), the specific active surface area of the microelectrode is 1538
340 m2 m−3, which is only 62.11% of the particle's specific surface area. Additionally, since the active material surfaces positioned at the sides of the composite electrode are not in contact with the electrolyte, the specific active surface area of the microelectrode is 1
529
744 m2 m−3, which is 61.76% of the particle's specific active surface area.
Since the experiment was conducted under constant current conditions, assuming that the ohmic overpotential remains constant, the increase in overpotential as the discharge progresses can be attributed to concentration overpotential. Although some literature reports that concentration within a particle can be analyzed using experimental methods, these methods have limitations.30 In the following of the paper, we will use our model to investigate this aspect.
At the particle level, all particle surfaces are surrounded by electrolytes, ensuring sufficient reaction sites. However, at the cell level, contact areas between active materials, between active materials and CBM, and between active materials and the current collector reduce the effective active surface area. Insufficient electrolyte further exacerbates the issue, creating a gradient of lithium-ion concentration within the electrolyte, leading to significant overpotential.31 From this perspective, we compared the rate capabilities and overpotentials at 1C, 2C, and 4C under realistic electrode conditions and excessive electrolyte conditions (Fig. 3A and B). The realistic electrode condition is when the electrolyte does not cover the electrode side, and the excessive electrolyte condition is when the electrolyte surrounds the entire electrode as shown in Fig. S1A and D (ESI†). So, the difference between the realistic electrode condition and the excessive electrolyte condition is whether the electrolyte is in contact with the sides. Under excessive electrolyte conditions, capacities were maintained at 100%, 96.72%, and 94.01% at 1C, 2C, and 4C, respectively, despite the relatively higher overpotential at the initial stage of charge compared to the overpotential at the late stage of charge. This was attributed to sufficient reaction sites and electrolytes mitigating the initial overpotential. In contrast, under the realistic electrode condition where sufficient reaction area and electrolyte volume were not ensured, charging capacities dropped significantly to 90.5%, 81.58%, and 26.5% at 1C, 2C, and 4C, respectively, much lower than those under excessive electrolyte conditions. In these realistic electrode conditions, smooth delithiation did not occur due to inadequate reaction area and electrolyte, leading to concentration gradients of lithium ions in both the electrolyte and electrode during charging.
These findings highlight the challenge of achieving high-rate charging above 2C (30 minutes) with currently commercialized batteries, influenced by both cathode and anode rate capabilities. To gain insight into fast charging, we employed our model's advanced analysis technique, 3D operando analysis, which was conducted at 25%, 50%, 75% charge, and cutoff voltages at 1C, 2C, and 4C under both realistic electrode and excessive electrolyte conditions. This analysis included evaluation of lithium-ion concentration in the active materials (Fig. 3C and G, and Fig. S4–S6, ESI†), CBM (Fig. 3D and H, and Fig. S7–S9, ESI†), electrolyte (Fig. 3E and I, and Fig. S10–S12, ESI†), and overpotential of the active material (Fig. 3F and J, and Fig. S13–S15, ESI†).
In the realistic electrode condition at 25% charge at 4C, despite some electrolyte penetrating the electrode and facilitating simultaneous reactions on the electrode surface and within, a larger quantity of lithium ions is deintercalated near the electrode surface compared to near the interface of the active materials and the current collector, creating a significant concentration gradient of lithium ions (Fig. 3C). Analysis of the lithium-ion concentration in the CBM and electrolyte reveals a gradient where concentrations are higher near the current collector and lower near the electrode surface, as lithium ions struggle to diffuse sufficiently into the bulk electrolyte (Fig. 3D and E). Furthermore, due to the migration of lithium ions towards lithium metal, the lithium-ion concentration in the bulk electrolyte is much lower than in the electrolyte impregnated within the electrode pores and CBM, resulting in a pronounced concentration gradient of lithium ions and consequently a high overpotential near the current collector (Fig. 3F).
On the other hand, under the excessive electrolyte condition at 25% charge at 4C, lithium-ion concentration near the electrode surface and current collector remains uniform (Fig. 3G). The surplus electrolyte ensures sufficient lithium ions, resulting in even distribution within the CBM and electrolyte, leading to uniform lithium-ion concentrations (Fig. 3H and I) and low overpotential throughout the electrode (Fig. 3J). These findings suggest the potential for enhanced performance through various patterns etched onto composite electrodes, which increase specific active surface area and promote rapid lithium-ion diffusion from active materials to the reaction surface near the current collector.32–34 Moreover, they indicate the feasibility of supporting charging rates of 4C or higher in a 70 μm-thick composite electrode, albeit with a trade-off between improving rate capability and energy density through patterning.
Particle size is another factor influencing fast charging. Larger particles exhibit slower diffusion between the center and surface, resulting in reduced rate capability due to increased diffusion lengths within the particle. Our results also demonstrate less delithiation in larger particles under both excessive electrolyte and realistic electrode conditions (Fig. 3C and G). Previous literature investigating lithium-ion diffusion in active materials during charging and discharging confirms that particle size significantly impacts rate capability, as effective lithium diffusion across the composite electrode depends on particle size optimization.35–37 However, particle size must balance with mechanical robustness, as smaller particles are more prone to cracking.38 Therefore, achieving optimal rate capability hinges on effectively managing the trade-offs between energy density and mechanical durability.
The volume changes in active materials during cycling can lead to deformation of CBM and potential detachment of the composite electrode from the current collector. However, to achieve high energy density, composite electrodes minimize the fraction of CBM and emphasize uniform distribution for mechanical robustness.41 Less discussed in literature is CBM's impact not only on mechanical strength but also on electronic conductivity within composite electrodes. Specifically, CBM greatly enhances effective electrical conductivity due to its electronic conductivity (3.75 × 106 S cm−1), which is billions of times higher than that of the active material (10−7–10−2 S cm−1).26,42 Therefore, CBM deformation by active material changes can significantly alter current density within the electrode.
We analyzed current densities inside the electrode, revealing averages of 16.29 A m−2, 35.20 A m−2, and 96.60 A m−2 in active materials, and 222.10 A m−2, 431.01 A m−2, and 1012.92 A m−2 in CBM at 1C, 2C, and 4C rates, respectively (Fig. 4C and D). CBM exhibited over 1000% higher current density than active materials, resulting in effective current densities (averages of active material and CBM) of 38.78 A m−2, 78.39 A m−2, and 196.40 A m−2, approximately 200% higher than that of the active material, despite CBM comprising only 4 wt% (Fig. 4C). This indicates that current flow from the current collector to the electrode surface predominantly occurs through CBM rather than active materials. However, the volume change and specific surface area do not proportionally correspond to the current density, as the overpotential increases sharply. Consequently, the current density decreases, as shown in Fig. S15–S17 (ESI†), after the gray circle in Fig. 4. This finding was corroborated by operando analysis, which revealed high current density areas predominantly within CBM dispersed throughout the electrode at 25% charge by 4C in both 3D and 2D views (Fig. 4E–J, and Fig. S16–S18, ESI†).
High current density areas are sporadically observed within CBM, while active materials show high current density only at surfaces in contact with CBM, diminishing towards the particle center (Fig. 4F and I). Within CBM, the current density is high from the current collector to two-thirds of the composite electrode but low near the electrode surface. This is due to binder boiling during electrode drying at 130 °C after slurry production, which creates voids and reduces CBM connectivity near the electrode surface (Fig. 4G and J). Consequently, CBM near the electrode surface exhibits current densities similar to those of active materials, primarily facilitating current flow between particles rather than from the current collector.
Using these assumptions, the stress evolution in the electrode was predicted by applying hygroscopic swelling theory for the active materials and large-deformation elasto-viscoplasticity for the CBM. Mechanical properties were evaluated under excessive electrolyte conditions, demonstrating superior characteristics despite higher stress levels (Fig. 5, Fig. S19–S30, and Videos S3 and S4, ESI†). Furthermore, overpotential was lower under excessive electrolyte conditions compared to the realistic electrode condition (Fig. 3F and J), enabling a broader x range in LixNi0.7Mn0.15Co0.15O2 utilization, which resulted in higher average and maximum stresses at the end of charge (Fig. 5A and B). Consequently, superior mechanical integrity was observed under excessive electrolyte conditions when compared at similar x in LixNi0.7Mn0.15Co0.15O2. 3D operando analyses of stress and strain were conducted under both excessive electrolyte and realistic electrode conditions. Under the realistic electrode condition at 25% charge by 4C, stress concentrated near the electrode surface with significant strain in active materials (Fig. 5C and D). Conversely, the excessive electrolyte condition at 25% charge by 4C exhibited relatively uniform deformation throughout the electrode, resulting in uniform stress distribution (Fig. 5G and H). Stress and deformation in CBM were concentrated near the surface in the realistic electrode condition (Fig. 5E and F), while they were evenly distributed in the excessive electrolyte condition (Fig. 5I and J), similar to the behavior observed in active materials. However, in both excessive electrolyte conditions and realistic electrode conditions, significant stress was observed in CBM near the current collector. Upon completion of charging (94% charge) in the excessive electrolyte condition (Fig. 5K–N), the maximum stress near the current collector increased to 314 MPa (Fig. 5B). This high stress in active materials translated to high stress and strain in CBM (Fig. 5M and N), eventually leading to plastic deformation in CBM (Fig. S31, ESI†). Although plastic deformation in CBM may not cause immediate mechanical degradation in composite electrodes within a single cycle, hardening evolution over repeated cycles can lead to mechanical degradation in CBM and potentially result in electrode delamination from the current collector.
Using these well-fitted parameters, mechanical degradation was predicted under the assumption that the volume change of the active materials remains constant over cycles. Specifically, the model was run for 5 cycles at 1C in the realistic electrode condition, with the strain evolution quantified at each cycle. The average strain over 5 cycles ranged from 0 to 0.00316, remaining nearly constant within each cycle (Fig. 6C). The average yield strength increased by 0.01% after 5 cycles (Fig. 6D), indicating minimal hardening of CBM. Therefore, the overall strain in CBM is insufficient to induce significant plastic deformation during cycling in the NMC composite electrode containing 2 wt% PVdF. Nevertheless, analyzing the maximum strain is crucial as partial degradation of CBM can lead to a significant decline in cell performance. Literature reports indicate that after hundreds of cycles, particles may detach from the surface of the composite electrode, with these separated particles observed on the separator surface in post-mortem analyses.14,46,47 Furthermore, if the CBM bonded to the current collector degrades, the entire composite electrode may separate, resulting in dramatic deterioration in cell performance.48
To further elucidate the mechanical behavior, we analyzed the maximum strain occurring near the current collector in our microelectrode model (Fig. 6E). Unlike the average strain behavior, the maximum strain decreases with each cycle, indicating plastic deformation in the CBM. This plastic strain drives the increase in the yield stress, reflected in the hardening of the CBM. However, since the maximum strain is insufficient to cause fracture or failure, the increase in yield stress diminishes over cycles and stabilizes at 42 MPa, as calculated by linear extrapolation (Fig. 6F). Given that the fracture strength of PVdF is 45 MPa at a slow loading rate of 0.00003 s−1 (Fig. 6B), it is evident that the PVdF material in the NMC active material composite structure ensures mechanical integrity. This finding underscores the widespread use of PVdF in batteries.
Despite the development of an advanced CBM model, its limitations are evident. This model predicts the mechanical degradation of CBM solely based on the hardening due to plastic deformation. To achieve more accurate simulations of CBM degradation, such as electrode separation from the current collector or particle peeling, a sophisticated fracture model incorporating precise CBM mechanical properties is required.
To thoroughly analyze and understand the experimental findings, we developed an electrochemo-mechanical model at the microelectrode level using 3D reconstruction that accurately reflects the active materials, pores, and CBM in a domain of tens of micrometers. Our results indicate that the amount of electrolyte, specific surface area, and electrode thickness significantly influence both electrochemical and mechanical characteristics by generating lithium-ion concentration gradients in the active materials, CBM, and electrolyte. Additionally, we analyzed the impact of CBM, which is not considered at the particle level. CBM, being a material through which current predominantly flows inside a composite electrode, significantly affects electrochemical performance by increasing effective electrical conductivity. Uneven binder distribution in the composite electrode interferes with current flow, causing overpotential, and this uneven distribution becomes more frequent as the electrode thickens.
From a mechanical perspective, the degradation of the composite electrode is closely related to the plastic deformation of the CBM. We applied an elastic-viscoplastic model to estimate the large deformation of CBM over cycles. Our results show that PVdF can withstand the stress generated by the deformation from nickel-based active materials during cycling, as the yield stress saturates before reaching the fracture stress. Therefore, PVdF is one of the most stable materials that has been widely used for NMC composite electrodes. At the same time, it is worth noting that due to computational limitations, this model has not considered the long-term degradation effects such as particle fatigue crack, SEI growth, Li plating, and gas generation. In reality, as the cycle number increases, these effects will become prominent, and the plastic deformation of PVdF binder is expected to accumulate.
Despite the development of this advanced analytics platform, several challenges remain to be addressed: (1) during the battery manufacturing stage, it is crucial to account for the expansion of the composite electrode due to electrolyte impregnation, as CBM impregnated with electrolyte exhibits poorer mechanical properties than dried CBM. Therefore, the properties of CBM impregnated with electrolyte must be accurately measured and applied to the model. Although several studies have attempted to analyze the mechanical properties of binders, ensuring their reliability is challenging due to the inconsistency in the reported properties of PVdF.24,25,49 (2) For a more accurate simulation of current flow, the conductive materials and binder materials should be separately reconstructed. However, this presents a significant challenge due to the disparity in scale: conductive materials are only tens of nanometers in size, while particles are several to tens of micrometers. Focusing on the conductive materials and performing 3D reconstruction at high resolution would result in a very heavy simulation, as the entire electrode would need to include tens of billions of voxels. (3) For long cycle estimation, it is crucial to consider electrochemo-mechanical degradation phenomena such as particle cracking, nickel dissolution, and SEI cracking/formation. Additionally, to accurately simulate the separation of electrodes or particles, the general model framework should incorporate a fracture model. (4) Since our model used a separator with a significantly high porosity, the effect of the separator was not considered. However, the influence of the separator cannot be ignored in general batteries. Since the technology for 3D reconstruction of separators using Cryo-FIB-SEM has been developed, it is possible to develop a microstructure model of the separators, reflecting this will enable the development of a more sophisticated model.
True structures | Synthesized structures | |
---|---|---|
Domain: W × T × D (μm) | 30 × 71 × 30 | 30 × 70.8 × 30 |
NCM711 fraction in domain | 0.68355 | 0.68351 |
CBD fraction in domain | 0.08062 | 0.08067 |
Porosity | 0.23583 | 0.23582 |
Voxel length (nm) | 43.78 | 600 |
Electronic conductivity (S m−1) | 3.58 × 107 |
Density (g cm−3) | 2.70 |
Poisson's ratio | 0.3314 |
Young's modulus (GPa) | 6.88 |
Yield strength (MPa) | 276 |
Isotropic tangent modulus (GPa) | 0.562 |
Resistivity (Ω m) | 2.79 × 10−8 |
Young's modulus (GPa) | 2.611 |
![]() | (1) |
Boundary condition: , and cs|center = cs,init.
Ds,eff (m2 s−1) is effective diffusion coefficient in the composite electrode, which determines rate of lithium-ion intercalation or deintercalation. In our model, Ds,eff = Ds can be applied as intrinsic properties that does not changed with cell design because active materials are replicated completely by 3D reconstruction. So, eqn (1) is simplified as follows:
![]() | (2) |
In this way, the realistic structures of each component can contribute to the development of more accurate models by fundamentally removing structural parameters and preventing overfitting. Therefore, we were able to apply the Ds in a way that depends on the lithium-ion concentration as follows:
![]() | (3) |
As for the electrolyte, mass conservation equation is as follows:
![]() | (4) |
Boundary condition: ce|Electrolyte|CCsurface = ce,init.
In the electrolyte, we can apply εe = 1, De,eff = De, and σe,eff = σe because realistic structures include all electrolyte structure in the composite electrode. So, eqn (4) is changed as follows:
![]() | (5) |
However, in the CBM, εe, De,eff, and σe,eff are determined by the amount of electrolyte that penetrates the CBM. In this paper, PVdF was used for binders, so the εe is 0.16, which is the volume expansion rate of PVdF due to electrolyte penetration24 and with the Bruggeman relation, De,eff = εe1.5De and σe,eff = εe1.5σe are used as effective value. So, eqn (4) can be written for CBM as follows:
![]() | (6) |
∇·(σs,eff∇φs) = asj, | (7) |
Boundary condition: , and ∇·(σs,eff∇φs)|AM
surface = 0.
In the active materials, the σs,eff = σs was applied thanks to the realistic structures (Fig. 1). Eqn (7) can be written as follows:
∇·(σs∇φs) = asj. | (8) |
The charge conservation equation in liquid phases of electrolyte and CBM is as follows:
![]() | (9) |
Likewise, in the electrolyte, we can apply σe,eff = σe so, eqn (9) is changed as follows:
![]() | (10) |
Boundary condition: φe|Electrolyte|CCsurface = 0.
However, eqn (10) cannot be applied for CBM because electrolyte is combined with CBM. So σs,eff can be determined by Bruggeman relation as follows:
![]() | (11) |
![]() | (12) |
![]() | (13) |
The current density j in eqn (12) is driven by overpotential defined as follows:
η = φs − φe − Ueq. | (14) |
![]() | (15) |
Here, ε and u denote the strain and displacement tensor fields within the domain Ω respectively. Also, a Dirichlet boundary condition is imposed on the displacement field over the domain, expressed as:
u = û, on Ωu, | (16) |
The conservation of linear momentum gives us the force equilibrium condition:
∇·σ + b = 0, in Ω, | (17) |
σ·n = ![]() | (18) |
σ = σ(ε, ![]() | (19) |
ε = εe + εvp + εs | (20) |
The Cauchy stress is determined by:
σ = ![]() | (21) |
NMC711 active material has been well documented to show brittle mechanical behavior.51,52 Hence its strain tensor formulation excludes consideration of viscoplastic strains and is limited to elastic and hygroscopic strains.
ε = εe + εs | (22) |
An isotropic volumetric change is assumed to occur based on the lithium-ion concentration, allowing for the application of hygroscopic theory to simulate the lithium-induced strain as outlined in previous studies:53,54
εs = βLMm(cs − cs,init), | (23) |
The lithium-induced strain, εs dependent on the lithium-ion concentration are reported in ref. 55 and 56, the molar mass (Mm), and lithium-ion concentration change (cs − cs,init) are calculated (Fig. S32A–C, ESI†). So, we reversely derived the lithium-induced strain coefficient, βL (Fig. S32D, ESI†).
In the CBM, both elastic and viscoplastic deformations are observed, typical of such polymeric materials. These deformations are induced by the deformation of active materials. Thus, eqn (20) reduces to the following formulation:
ε = εe + εvp. | (24) |
The Perzyna model is used to describe the viscoplastic strain evolution.57 This model, originally formulated for the viscoplastic response of metals at high temperatures has been successfully applied to model the inelastic rate-dependent response of polymeric materials.58,59 The model equation for the viscoplastic strain rate is:
![]() | (25) |
f(σ,![]() ![]() | (26) |
![]() | (27) |
σy = σy0 + k(![]() | (28) |
![]() | (29) |
![]() | (30) |
A is the viscoplastic rate coefficient, b is the stress exponent typically set to be unity for Perzyna type models and 〈·〉 is the Macauley bracket,
![]() | (31) |
All parameters for the model are described in Tables 2–5 and their descriptions are provided in Table 6.
Ds,eff | Effective diffusion coefficient in active materials, [m2 s−1] |
cs | Lithium-ion concentration in active materials, [mol m−3] |
cs,init | Initial lithium-ion concentration in active materials, [mol m−3] |
j | Current density on the electrode, [A m−2] |
F | Faraday constant, [96![]() |
Ds | Intrinsic diffusion coefficient in active materials, [m2 s−1] |
Ds,init | Initial diffusion coefficient in active materials, [m2 s−1] |
cs,max | Maximum lithium-ion concentration in active materials, [mol m−3] |
εe | Electrolyte volume fraction in porous materials |
ce | Lithium-ion concentration in the liquid phase, [mol m−3] |
De,eff | Effective diffusion coefficient in electrolyte, [m2 s−1] |
σe,eff | Effective electronic conductivity of electrolyte [S m−1] |
R | Gas constant, [8.3143 J mol−1 K−1] |
T | Temperature, [303.15 K] |
t+ | Transport number |
De | Intrinsic diffusion coefficient in electrolyte, [m2 s−1] |
σe | Intrinsic electronic conductivity of electrolyte [S m−1] |
σs,eff | Effective electronic conductivity of active materials [S m−1] |
σs | Intrinsic electronic conductivity of active materials [S m−1] |
I | Applied current |
φs | Potential in the active materials, [V] |
as | Specific surface area, [m2 m−3] |
φe | Potential in the electrolyte, [V] |
i0 | Exchange current density [A m−2] |
aa | Anodic transfer coefficient |
ac | Cathodic transfer coefficient |
i0,init | Initial exchange current density [A m−2] |
η | Overpotential [V] |
Ueq | Equilibrium potential [V] |
ρ | Density, [kg m−3] |
ε | Strain tensor |
u | Displacement field, [m] |
Ω | Domain |
û | Prescribed displacement |
Ωu | Dirichlet boundary condition |
σ | Second order stress tensor [Pa] |
b | Body force per unit volume [N m−3] |
![]() |
Prescribed traction vector |
![]() |
Strain rate tensor |
εe | Elastic strain |
εvp | Viscoplastic strain |
![]() |
Fourth-order elastic modulus tensor |
Qp | Plastic potential [Pa] |
εs | Hygroscopic strain |
βL | Lithium-induced strain coefficient [m3 kg−1] |
Mm | Molar mass [kg mol−1] |
![]() |
Viscoplastic strain evolution |
f | Yield function of the material |
![]() |
Equivalent stress |
s | Deviatoric stress tensor |
σy | Yield stress [Pa] |
σy0 | Initial yield stress [Pa] |
k | Strength coefficient |
n | Hardening exponent |
A | Viscoplastic rate coefficient |
b | Stress exponent |
The numerical calculations were performed on a workstation featuring an AMD Ryzen Threadripper PRO 3995WX processor (64 cores, 2.70–4.20 GHz) and 1 TB of SK Hynix PC4-3200AA-L 3200 MHz ECC server RAM (4 × 256 GB). The simulation required approximately two weeks to compute a single discharge curve.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ee04856c |
This journal is © The Royal Society of Chemistry 2025 |