The following article is Open access

3D-Printed Testing Plate for the Optimization of High C-Rates Cycling Performance of Lithium-Ion Cells

, , , , , and

Published 3 May 2021 © 2021 The Author(s). Published on behalf of The Electrochemical Society by IOP Publishing Limited
, , Citation Gilberto Carbonari et al 2021 J. Electrochem. Soc. 168 050508 DOI 10.1149/1945-7111/abfab9

1945-7111/168/5/050508

Abstract

Nowadays, long charging times have become one of the main limitations to a greater worldwide spread of electric vehicles (EV). Enabling high C-rates charging is a promising approach to eliminate this problem and alleviate range anxiety. When a battery is charged at high currents, several factors have to be taken into account. Temperature is certainly a key parameter because when it is too high this can lead to degradation of components (binder, electrolyte, active material, etc), however, when it is too low intercalation kinetics becomes sluggish. Using 3D-printed testing plates (PP3D plates) with Li-reference electrode, we developed a tool for electrochemical investigations of pouch cells. These plates enabled to build a new well-designed 3-electrode pouch cell. This setup allows the identification of the best high C-rate cycling procedure to improve the performance and cycling life of the lithium ion cells. We explored the electrochemical behavior of NMC811 cathodes and graphite anodes, during high discharge C-rates test up to 7 C and charge C-rates up to 2 C. Moreover, the temperature influence on charging performance and longtime cycling stability is investigated. The cells cycled at 25 °C using optimized procedures reached an 80% state of health after more than 1000 cycles.

Export citation and abstract BibTeX RIS

This is an open access article distributed under the terms of the Creative Commons Attribution Non-Commercial No Derivatives 4.0 License (CC BY-NC-ND, http://creativecommons.org/licenses/by-nc-nd/4.0/), which permits non-commercial reuse, distribution, and reproduction in any medium, provided the original work is not changed in any way and is properly cited. For permission for commercial reuse, please email: permissions@ioppublishing.org.

The last few years have seen a dizzying increase in the worldwide demand for electric vehicles (EV) and portable electronics. 1 Both of them share a common energy source: rechargeable lithium-ion batteries (LIBs). Their development follows the market demands for higher energy density as well as safer, cheaper, and more environmentally friendly batteries. 2 These tasks can be accomplished by pursuing different approaches but several of them could be fulfilled simultaneously. The traditional strategy to improve the battery-specific energy suggests increasing the average cell potential 2 and/or improving the specific capacity of the material. This can be easily estimated by applying Eq. 1.

Equation (1)

One solution to these challenges is the use of layered lithium oxides also known as NMCs. Their general formula is Li1+x (NiyMnzCo1−y−z)1−xO2 and their names depend on the relative ratio of the three different transition metals (TM). NMC111 with a relatively low nickel content has been and still is used in LIBs because of its cycle life, stability, and capacity. On the other hand, Ni-rich materials such as NMC532, NMC622, and NMC811 seem to be reasonable choices for the future of the battery producers. 3 In particular, NMC811 provides high capacities of ∼200 mAh g−1 and an average discharge potential of ∼3.80 V together with a substantial decrease of the Co content. 4,5 Cobalt amount reduction is important in cutting down the costs and reducing the dependence on suppliers with a questionable supply chain. 6 These advantages come at the expense of its cycling stability and safety, both mainly originating from a high nickel content. 7 Several solutions have been taken into consideration such as the tailored coating on the active material surface 8,9 and electrolyte additives. 1012 In terms of cost reduction and environmental sustainability, one further challenge for these materials is the implementation of aqueous processing. In fact, NMCs are sensitive to water and air exposure during storage and processing. 13 Nevertheless, several research groups have already obtained promising results from water-based cathode formulations. 14,15

In order to minimize inactive material thus increasing battery energy density, an optimal capacity balancing (Qnegative electrode/Qpositive electrode = N/P ratio) of the electrodes used in the cell must be defined. 16,17 A N/P ratio < 1 is not beneficial for cell-specific capacity and leads to anode overcharge and metallic lithium deposition on the anode side, whereas an N/P ratio ≫ 1 leads to cathode over-discharge and possibly cathodic material degradation. Ideally, an N/P ratio equal to 1 should be chosen to minimize these effects, but due to intrinsic defects in the electrodes and to avoid edge effect phenomena, N/P ratios slightly over 1 are used (e.g. 1.1–1.2). It should be noted that in the case of cells for high-energy applications, the ratio could be closer to 1, but in the case of quick charge application, the adopted N/P ratio is usually higher.

In a classical laboratory approach, the N/P ratio is usually determined by testing separately anode and cathode in coin half-cells. These capacity values are often obtained at low C-rates (e.g. C/10) and they do not reflect the ratio that can be obtained at relatively high C-rates. 18 On the other hand, when high currents are applied in coin half-cells, polarization phenomena can affect the results. The low lithium surface area leads to a large electrode polarization affecting its potential and capacity. Moreover, if we consider relatively high currents, every single coating defect (e.g. active material A.M. agglomerations, edge effect, edge cutting defects, etc.) starts to play a role, and often non-homogeneous coatings heavily affect the electrochemical results of small electrodes. For these reasons, pouch full-cells were used in this work to study especially the crosstalk of the paired electrodes, which gives more useful information about their electrochemical performances in real working conditions.

The development of additional strategies to overcome these issues can be facilitated by an accurate analysis of the processes leading to cell aging. Parameters such as anode and cathode potentials, cycling temperature, and stack pressure could be extremely helpful for a deep understanding of the cell behavior. Particularly, the usefulness of a reference electrode (RE) has never been questioned. An ideal RE should be non-polarizable and should exhibit a stable and reproducible potential during the electrochemical test. Since in full-cells voltage is given by the difference of the cathode and anode potential (two-electrode setup), the introduction of a RE allows measuring the individual electrode contribution, which is a key parameter for an in-depth understanding of cell aging and degradation mechanisms. 19

A significant work 2022 was done in the case of small-sized electrodes on the lab scale. In case of larger electrode area and full-cells, most of the published works insert the RE inside the jelly roll or the electrode stack. 23,24 In this case, the RE surely is in the best position taking into account the electrical field and equipotential lines, 25 however, its presence in the stack will influence the cell behavior. 26 Different groups tried to minimize its dimension to avoid any side effects. 23,27 In commercial batteries, a certain stack pressure is always present and the minimal defect, as in the case of a RE, quickly results in lithium plating 28 even at moderate C-rates. For the sake of clarity, we would like to point out that in commercial batteries thin polymeric and ceramic coated separators are used. In this case, the presence of an internal RE will result in an even greater problem since the buffer ability of the glass fiber separator (e.g. Whatman GF/A) is lost.

Here is proposed a simple solution that involves a 3D printed plate (Fig. 1a). This device has proven to be cheap, easy to print, and non-invasive. Besides, it can be easily up-scaled and used to monitor larger electrode stacks. Its versatility allows the introduction of even more sensors (temperature, strain, etc) to gain a more complete understanding of cell aging mechanisms. Custom models can be realized thanks to the ease of the 3D design. In particular, the implementation of other sensors would help in a more complete mapping of cell aging.

Figure 1.

Figure 1. PP3D plate design with a magnified vision of the reference electrode area (II), channels for the electrodes' tabs and the surface where the stack is placed (III) (a); b) Pouch cell holder design (b); Assembled PP3D cells (c).

Standard image High-resolution image

In this paper, graphite∣∣NMC811 single-layer pouch cells were tested to assess their discharge capabilities. In order to improve their charging step, 3D printed plates with an integrated lithium reference electrode (Li-RE) were used to follow anode and cathode potentials. Setup validation was investigated electrochemically by the means of pouch cells without 3D plate as well as optical visualization and Glow Discharge—Optical Emission Spectroscopy (GD-OES) depth profiling. Remarkable results in terms of performance, reproducibility, and reliability were obtained using this setup. The purpose of this work is to develop a tool able to predict and find the optimal cycling operation of a cell in terms of customized cycling procedures for extending the cell's lifetime.

Experimental

3D plate design

The 3D plate was designed with free CAD software (Blender) and printed with a Fused Deposition Modeling (FDM) 3D printer (Ultimaker3, Ultimaker). Aside from a good extrusion accuracy, the choice of the material is fundamental since it needs to be electrolyte resistant. The most used in commercially available test cells is the expensive polyether ether ketone (PEEK) thanks to its superior mechanical and chemical properties. Since the extrusion process of this polymer requires high temperatures and high-end 3D printers, we have opted for a cheaper and easy to print polymer such as polypropylene (PP). Figure 1a shows its simple design. The electrodes´ tabs are fixed in two guides (see Fig. 1a—I) allowing the electrodes stack to stay in the desired position. The reference electrode (RE) position is well defined by design: to avoid short circuits and limit measurement's artifacts, the Li-RE (0.2 mm2 surface) is placed at a horizontal distance of 300 μm from the stack (see Fig. S1). Single-layer pouch cells assembled in the 3D plate setup will be named PP3D cells (Fig. 1c).

Coatings and electrodes preparation

All materials were used as received. Anode and cathode slurries were prepared in a planetary mixer where active materials portions were kept as high as possible to match industrial targets and good coating performance in terms of mechanical and electrochemical stability. The used materials were Graphite (Hitachi Chemical), LiNi0.8Mn0.1Co0.1O2 (NMC 811, Targray), Na-CMC (Sunrose), PVdF (Solvay Solexis), SBR emulsion (Zeon), Super C65 (SC65, Imerys) and SFG6L graphite (Imerys). Distilled water and N-methylpyrrolidone (NMP, Sigma Aldrich) were used as solvents respectively for anode and cathodes. The slurries were then cast in a pilot line (LACOM Gmbh, Germany) equipped with a comma bar system, four drying ovens (total length: 8 m), and a Kr online mass scanner to ensure a uniform loading throughout the coatings. All electrodes were prepared by single-side coating and calendering to a specific density. The most important parameters of the electrodes are shown in Table I.

Table I. Anode and cathode characteristics.

Anode94 wt% SMG-HT1(Hitachi Chemical)
 2 wt% Super C65 (Imerys)
 2 wt% Na-CMC (Sunrose)
 2 wt% SBR (Zeon)
 Cu current collector: 10 μm
 Coating Thickness: 50 μm
 Density: 1.4 g cm−3
 AM loading: 8 mg cm−2
 Capacity (@ C/10): 2.9 mAh cm−2
 2 wt% Super C65 (Imerys)
Cathode93 wt% NMC811 (Targray)
 2 wt% Super C65 (Imerys)
 2 wt% SFG6L graphite (Imerys)
 3 wt% PVdF (Solvay Solexis)
 Al current collector: 14 μm
 Coating thickness: 44 μm
 Density: 3.2 g cm−3
 AM loading: 13.3 mg cm−2
 Capacity (@ C/10): 2.7 mAh cm−2
SeparatorCelgard 2325
 Thickness: 25 μm
 Porosity: 39%
Electrolyte1.4 M LiPF6 in EC:DMC
 (3:7 wt%) + 2% VC (LP40)
 Conductivity @25 °C ≈ 12 mS cm−1
 Conductivity @40 °C ≈ 15 mS cm−1

Cells assembly

Where not specified, all the cells were assembled with the materials and components listed in Table I. For single-layer pouch cells (SLPCs), electrodes were punched with a customized mechanical puncher (anode area: 26 cm2; cathode area: 23.94 cm2), dried overnight at 130 °C under vacuum and stored in a dry cabinet. A tri-layered polymeric separator (Celgard 2325) was used and Al and Ni tab collectors were ultrasound-welded respectively to cathode and anode. Before electrolyte filling, the cell was furthermore dried at 80 °C overnight under vacuum.

Unlike regular single-layer pouch cells, when the PP3D plate is used, Celgard 2325 separator (26 cm2) was stacked between anode and cathode and not rolled around them as in the SLPCs. A small piece of freshly cleaned metallic lithium (PI-KEM) was filled inside the channel where a thin copper wire was inserted. Anode and cathode were aligned on the reference electrode side and held in position by Kapton tape. Both cell types were assembled in a dry room (dew point < −65 °C), then they were vacuum-dried overnight at 80 °C. SLPC and PP3D cells were filled respectively with 0.9 ml and 1.2 ml of electrolyte and vacuum-sealed in a glovebox (MBraun, O2, and H2O levels < 0.5 ppm).

Electrochemical characterization

All the electrochemical procedures listed in Table SI (Supporting Info available online at stacks.iop.org/JES/168/050508/mmedia) were carried out on a VMP-3 (Biologic) system connected to a climate chamber (VB-4004,Vötsch) where the cells were tested at controlled temperatures (25 °C and 40 °C). The resulting capacities are always normalized with respect to the first discharge value after the formation or the cathode active material weight. SLPCs and PP3D cells were cycled using an in-house designed pouch cell holder (Fig. 1b) equipped with pogo pins to test cells in a 4-points configuration. Thanks to the knowledge gained from previous tests on this anode/cathode combination, the bolts of the holder were tightened applying a defined torque momentum resulting in an average force of 100 N on the surface of the pouch cell. The holder assures a homogenous pressure on the cells and it is minimizing the electrical contact resistance between VMP probes and the tabs of the cells.

Post-mortem and glow discharge optical emission spectroscopy (GD-OES)

Aged cells were opened in an Ar-filled glovebox (MBraun, O2, and H2O levels < 0.5 ppm), and pictures of the graphite anodes were taken. The absence or presence of lithium deposition was further confirmed by GD-OES analysis according to the method published earlier. 29 lithium deposition quantification was conducted using reference samples.

Aged anode samples were washed three times in Dimethyl Carbonate (DMC), left to dry, and eventually mounted on airtight sample holders. An additional test was performed on the pristine anode coating to evaluate its binder distribution following the sodium relative concentration (from the Na-CMC binder) throughout the coating thickness. GD-OES analyses were conducted using a GDA750 device (Spectruma). The measurements on the anodes were performed in RF mode at a discharge voltage of 550 V, a gas pressure of 2 hPa, and a pulse rate of 50% at a frequency of 2500 Hz. A mixture of H2 6.0 and Ar 6.0 (1:99 vol. %) was used as a discharge gas. The following emission lines were used for detection: Cu (327.4 nm), C (156.1 nm), O (130.2 nm), Li (670.7 nm), P (178.3 nm), H (121.6 nm), Na (589.0 nm). The sputtered craters in the samples produced by the measurement have a diameter of 2.5 mm.

Results and Discussion

Cell formation and validation

The formation protocol in Table SI (supporting Info) was applied to SLPCs and PP3D cells. Figure 2a shows the three formation cycles for the two different setups. Charge/discharge profiles are overlapped leading to the same capacities. The use of the PP3D plate did not influence the electrochemical performance of the stack tested in a state-of-the-art SLPC. Firstly, both showed an initial irreversible capacity loss (ICL) of 11.5%, a value in perfect agreement with the ICL values found in the literature (10% < ICL < 20%). 30,31

Figure 2.

Figure 2. Galvanostatic profiles during the formation step of the graphite-based anode (a) and NMC811-based cathode (b) in a PP3D cell. Comparison of the first 3 cycles between a SLPC and a PP3D cell (c).

Standard image High-resolution image

The presence of the Li-RE in PP3D cells affords us to separate anode and cathode contributions. Figures 2b–2c report the third formation cycle for the anode and cathode, respectively. In Fig. 2b it is possible to distinguish the lithium intercalation process in the graphite anode. The dQ dE−1 vs. E plot in the inset allows for better visualization of the graphite staging. The redox peaks around 230 (III), 120 (II), and 80 mV (I) correspond to the formation of LiC36, LiC12, and LiC6 graphite intercalation compounds (GICs), respectively. 3235 In Fig. 2c, it is possible to observe the respective dQ×dE−1 vs. E plot for the cathode side. The typical redox peak at 3.76 V is present together with three other oxidation peaks at 3.64 V, 4.02 V, and 4.20 V corresponding to the multiphase transition from hexagonal to monoclinic (H1 → M), monoclinic to hexagonal (M → H2) and hexagonal to hexagonal (H2 → H3). 7,36,37

Graphite is certainly the most used anode material in LIBs and the lithium is inserted in its structure by an intercalation reaction. Because of several factors such as temperature, electrodes structure, electrolyte, charging current, etc, lithium plating can compete with the main Li storage mechanism. 38 This alternative faradaic reaction becomes thermodynamically possible when the anode potential falls below the Li/Li+ couple potential (−3.04 V), where Li+ ions get reduced preferably on the graphitic surface leading to poor performance and safety issues. This phenomenon is exacerbated especially in the case of low temperatures, electrolytes with poor Li transport properties, 39 high C-rates, and not optimized anode structures. Since the N/P ratio changes with the applied current depending on the capacity retention and the different kinetics of anode and cathode, it would be faster to detect single electrode potentials in a full-cell at different C-rates. This is done by introducing a RE that allows monitoring the anode and cathode potentials individually. Its positioning can be internal or external relative to the electrode stack and despite the investigation of both geometries by several groups, 21,22,40,41 it is still hard to select the best setup. 19 To minimize the measurements artifacts, we deliberately used an external reference electrode to avoid shielding errors due to WE surface area coverage by RE 26 and application of inhomogeneous stack pressure. The ohmic drop due to relatively poor conductive organic electrolytes was minimized by placing the RE as close as possible to the stack edge previously aligned to avoid misalignment. Thanks to the near point-like design, the copper wire used as an electric connection to the RE was completely isolated from the electrolyte avoiding mixed surface potentials.

Good cell performance and reproducibility were achieved by the application of an external pressure onto the electrode stack since it can greatly influence the cell performance during formation and cycling steps. 42 Materials undergoing big volume expansions during cycling (e.g. silicon) 43,44 can benefit from the application of specific and tailored pressure. This is also true in the case of traditional materials such as graphite, where a better electrical contact can be obtained. Nevertheless, if excessive mechanical compression is applied especially to anodes and separators, their pores can close 45 leading to a decrease of active area and consequently an increase of the current density on the anode surface. A pouch cell holder was designed and used to ensure homogeneous pressure distribution all over the cell surface providing better and more reproducible results. The holder can squeeze out all the gas bubbles potentially formed during electrolyte degradation in the first cycles. When pressure is not applied, gas bubbles can prevent full utilization of the active material, which often leads to lithium deposition in its surrounding. 28,46 Figure 1b shows its design where a different pressure can be applied simply using a torque wrench to tighten the bolts. In the case of small pouch cells (single- and bi-layer), pressure differences play a bigger role than in the stacked pouch cells. In the latter case, the higher number of layers attenuates the pressure inhomogeneities experienced by single electrodes acting as a buffer. In the following measurements, the pressure distribution is furtherly ensured by soft polymeric pads on both lateral sides of the cells.

Charge/discharge C-rate Tests

The electrode specifications (AM type, AM content, loading, density) were chosen for achieving good performance with high discharge currents (high power). As can be seen from Fig. 3, single-layer pouch cells were discharged up to a rate of 7 C (≈ 19 mAcm−2). The discharge steps were conducted in constant current mode leading to remarkable capacity retentions (see Table II). The final discharge step at C/2 (Fig. 3a) confirmed that the high current discharge did not damage the cells allowing for a full recovery of the initial discharge capacity. Electrodes loadings and charge/discharge conditions greatly affect cell performance. Nevertheless, the obtained results proved to be noteworthy, especially, when taking into consideration full- and half-cells studies with the same electrode materials. 9,12,47

Figure 3.

Figure 3. SLPCs performance at different discharge current rates (a)–(b) ranging from C/2 to 7 C. Charge steps were conducted with C/2 in constant current-constant voltage (CCCV) mode until a current lower than C/20 is reached.

Standard image High-resolution image

Table II. Capacity retention values for several discharge currents. The values are calculated relatively to the 1st discharge capacity value after the formation step. Charge steps were conducted with C/2 in constant current-constant voltage (CCCV) mode until a current lower than C/20 is reached.

Capacity retention at different discharge C-rates
C/21 C2 C5 C7 CC/2
100%97.2%94.2%83.3%64.3%99.2%

To investigate the charging behavior PP3D cells were utilized, where anode and cathode potentials were monitored separately. High-charging current densities can greatly influence battery behavior since different phenomena can occur mainly at the graphite/electrolyte interface. Because of high charging currents and low temperatures, the anode surface potential can drop below 0 V (vs Li+/Li) and lithium plating becomes the thermodynamically favorable process over the intercalation. This can often lead to important safety concerns since lithium dendrites can pierce the separator and cause internal short circuits 38 and deposited lithium can react exothermically with electrolyte. 4850 Using the PP3D plate, we investigated how our cathode/anode couple behaves in different charging conditions. The charge C-rate protocols denoted in Table SI (Supporting Info) were applied.

Figure 4 shows anode and cathode potentials at different charge C-rates as a function of the state of charge (SOC) at 25 °C and 40 °C, respectively. Most battery producers recommend an operating charging temperature ranging from 0 °C to 40 °C. These values depend on several factors such as electrodes' chemistry, electrolyte type, and cell design. Based on literature data and personal experience, it was decided to use a controlled cycling temperature of 40 °C to stress the cells and partially simulate the thermal behavior of bigger cell formats (stacked pouch cells, prismatic cells, cylindrical cells). 51 As can be seen from Figs. 4a–4b, at both temperatures the anode end-of-charge potential is progressively shifted toward zero because of the increasing polarization influencing the N/P ratio. When cells were cycled at 25 °C the anode potential reached the lithium deposition limit only when a 2 C current was applied in the CC charging step. After the application of the CV step, the potential increased to around 70 mV as in the other C-rates. This behavior was present in each 2C-charging cycle in different cells. This leads to the hypothesis that despite the negative potential at the end of the CC step, lithium re-intercalation occurred during the CV step. For the sake of clarity, we cannot exclude that extended cycling in these conditions could lead to heavy a more pronounced lithium deposition. As already mentioned, the temperature during charge and discharge plays a fundamental role in the electrodes kinetics. If the electrolyte and the electrodes are able to withstand and do not degrade at such a temperature, then a beneficial effect on the cell performance should be observed.

Figure 4.

Figure 4. Anode and cathode potentials at different charge current rates and temperatures in a PP3D cell: anode potential vs. SOC % at 25 °C (a) and 40 °C (b); cathode potential vs. SOC % at 25 °C (c) and 40 °C (d).

Standard image High-resolution image

When the same charging protocol was applied at 40 °C, the rate capability improved, and as expected the temperature contribution was particularly relevant at higher C-rates. The increase in Li+ ions mobility due to the electrolyte lower viscosity had a beneficial effect on the cell performance. The SOC reached right after the CC step in charge (Table III) increased from 86.2% to 88% at 1 C (+ 2%) and from 74.6% to 81.3% at 2 C (+ 9%). Potentials shifts were observed at the cathode side as well (Figs. 4c–4d). As it can be seen clearly from Fig. 5, the anode potentials at the end of the CC charge steps at 40 °C (Fig. 5b) are higher than those at 25 °C (Fig. 5a). These more positive potential values are the result of the smaller polarization associated with the temperature increase.

Table III. SOC values reached after the CC step in the charging C-rate protocol (see Table SI).

 SOC after CC step
Charge C-rate T = 25 °C T = 40 °C
C/598.1%97.7%
C/395.5%94.8%
C/292.9%92.8%
1 C86.2%88.0%
2 C74.6%81.3%
Figure 5.

Figure 5. Charging C-rate's influence on the anode potential at the end of lithiation (CC + CV in red) at 25 °C (a) and 40 °C (b). The black squares represent the anode potential at the end of the CC step.

Standard image High-resolution image

Long-term cycling protocols

To confirm the obtained results, we tested single-layer pouch cells. Long-term cycling tests were carried out at different combinations of temperature and charging C-rate. Their results are shown in Fig. 6. As already mentioned above, the cycling temperature is important since it can directly affect the lithiation/delithiation kinetics, and based on the previous result, a symmetrical 1 C/1 C cycling at 25 °C should have not showed a fast capacity decay due to Li plating. 2 C charging at 25 °C showed an anode potential slightly below 0 V, followed by an anode potential increase during the CV step (Fig. 5a). Lithium deposition does not lead always to dendrites and electrically isolated lithium ("dead" lithium) formation. Depending on the conditions, these lithium deposits can re-intercalate in the graphite. 5255 This reversible Li deposition usually does not lead to a large amount of capacity loss through cycling, however, it can be anyway a safety concern 48 and should be avoided when possible. The results obtained in Fig. 4a suggest that perhaps we are dealing mostly with a reversible form of lithium deposition since no plateau is observed below 0 V and in the CV step the potential increases rapidly to ≈70 mV.

Figure 6.

Figure 6. SLPCs cycling performance at different charge current rates and temperatures.

Standard image High-resolution image

At the same time, other cells were cycled at a rate of 2 C at 40 °C. As shown in Figs. 4b and 5b, the anode potential in these pouch cells never dropped below 0 V. If the electrolyte and the NMC811 cathode would not degrade because of the cycling temperature, better results could have been expected compared to the pouch cells cycled at 25 °C. Several studies point out that temperatures higher than 30 °C can accelerate NMC-based cathodes degradation. 7 Moreover, electrolyte decomposition products such as HF can trigger NMC dissolution. 56 The dissolution of transition metals can pose a problem also to the anode since they can be included in the SEI structure worsening cell stability and cycling life.

Post-mortem and GD-OES depth profiling

After long-term cycling, the SLPCs were discharged at 2.7 V and disassembled in a glove box. In Fig. 7, it is possible to observe anodes and separators appearances after they reached 80% state of health (SOH). As predicted by the previous experiments (Figs. 4 and 5), pouch cells cycled with a charge rate of 1 C at 25 °C (Fig. 7a) and with a charge rate of 2 C at 40 °C (Fig. 7c) did not show any presence of dead lithium on their surface. When the temperature is raised to 40 °C better performance was expected due to an improvement of the intercalation kinetics, however the thermal stability of the electrolyte must be taken into consideration as well. Since the temperature was applied continuously throughout the cycling procedure, this can have affected the electrolyte stability resulting in faster cell aging. SLPCs cycled at 40 °C (Fig. 6) reached 80% SOH earlier than the cells cycled at 25 °C and no clear sign of gas evolution was observed. A sign of the electrolyte degradation can be seen in Fig. 7c where the polymeric separator showed local black areas and brownish spots. Taking a closer look at Fig. 7b, it is possible to notice a small "dead" lithium spot near the tabs. During the CV charging step, this region provides the shortest path for charge transport phenomena, therefore is more prone to be overcharged. This behavior was already detected in the PP3D cells where the anode potential was going slightly below 0 V. Nevertheless, the pouch cells cycled in these conditions showed comparable performance to those cycled with a 1 C charge current (see Fig. 6). The negative potential in the first cycles probably led to "dead" lithium that remained in that position without affecting further cycles.

Figure 7.

Figure 7. Post-mortem studies of the SLPCs cycled at different charge-currents and temperatures: anodes and separators appearances (on the left) and respective GD-OES depth profiling results (on the right). The analyzed samples were harvested inside the white dotted circle.

Standard image High-resolution image

In order to assess the presence of deposited lithium by GD-OES depth profiling, it is assumed that the measured amount of lithium is composed of lithium in the SEI and metallic lithium, as defined by Ghanbari et al. 29 The cell has been completely discharged, therefore the amount of intercalated lithium is negligible.

Equation (2)

Furthermore, it is presumed, that the SEI consists only of lithium oxide (Li2O). Li2O has the highest Li-content (46.5 wt %) of all known SEI species, therefore this assumption gives the minimum amount of metallic lithium. The detected oxygen is used to calculate the amount of Li2O in the SEI. If Eq. 2 gives a positive value for Limetallic, Li deposition has been detected. A negative value means that all Li is bound in the SEI. 29 Figure 7 shows all the GD-OES results performed at the center of the anode for each cell (white dotted circle in Fig. 7). Using this calculation, none of the investigated samples have shown lithium deposition. In Fig. 7a, the cell, which has been charged with a 1 C current at 25 °C shows a lithium gradient throughout the depth of the electrodes. This result is consistent with previous findings 57 showing that the SEI is more pronounced at the electrode surface. In Fig. 7b is shown the GD-OES result from the area highlighted by the white dotted circle as in the other two anodes. Nevertheless, since a small lithium deposition was visible, a sample was harvested from the red square and analyzed to confirm the presence of metallic lithium. The GD OES results from this area can be seen in Fig. 8a, where the Li distribution shows a plateau at the anode surface with a higher mass percentage than the other analyzed samples in Fig. 7. Furthermore, Eq. 2 was applied and the calculated Limetallic weight percentages have shown positive values only in this sample. The comparison between the calculated Limetallic weight percentages of these different samples is presented in Fig. S2 (Supporting Info). The cell charged with a 2 C current at 40 °C (Fig. 7c) shows a steady Li content, which could originate from the electrolyte decomposition at elevated temperatures that leads to an enhanced SEI formation. 58,59

Figure 8.

Figure 8. GD-OES depth profiling: a) Li, O, H, P, and C intensity from the red square area in Fig. 7b (SLPC: 2 C charge rate at 25 °C); b) Intensity of the Na line at 588.995 nm in the pristine (black line) and in an aged (red line) anode (left: anode surface, right: Cu current collector). The Na intensity was related to the Ar line at 763.511 nm to compensate for the small variations of the sputtering rate.

Standard image High-resolution image

GD-OES analysis was also used to assess the binder distribution. Slurry preparation, electrodes coating, and the drying step greatly influence the electrode homogeneity. In particular, binder migration toward the electrode surface can occur when high thermal gradients are present during electrode drying. 60 This phenomenon can lead to poor mechanical stability together with poor electrochemical performance of the electrode. By the detection of sodium which is present in the Na-CMC, the binder distribution was monitored. The results can be seen in Fig. 8b. The homogeneous sodium distribution throughout the whole pristine anode coating is shown by the flat course of the Na intensity. For the sake of comparison, an aged anode coating was analyzed as well. The Na distribution resulted to be the same as in the pristine anode. Since it was not detected a binder gradient, these results are the index of a successful anode drying step that resulted in a homogeneous binder distribution. The good performance in the C-rate discharge tests is also supporting these findings.

Conclusions

Graphite and NMC811 electrodes were manufactured and tested in SLPCs and PP3D cells with remarkable results especially at high discharge rates (64.3% capacity retention at 7 C ≈ 19 mA cm−2 discharge rate). The use of a novel 3D printed testing plate allowed us to investigate and understand under which cycling conditions we can obtain the best outcome in terms of cell performance, safety, and cyclic life for this specific anode/cathode combination.

In particular, the following results were obtained:

  • 1)  
    All the cells showed significant power performance and cycling stability, especially due to the optimized production and proper design of electrodes (slurry recipe, loading, density). Results reproducibility was furthermore assured by the use of pouch cell holders (Fig. 1b).
  • 2)  
    The presence of the PP3D plate does not affect stack performance (Fig. 2c) and could be optimal support for the implementation of other sensors.
  • 3)  
    The PP3D plate enables us to record the anode potentials vs. Li reference measured during C-rate tests. Potentials trends are supported by the SLPCs behavior observed during long-term cycling at different charge currents and temperatures.
  • 4)  
    80% SOC is reached after 1000 cycles when the cells are charged at 1 C and 2 C at 25 °C, and around 650 cycles when charged at 2 C at 40 °C.

These findings were furtherly supported by the GD-OES depth profiling. The only SLPCs showing lithium deposition (Fig. 6) were those charged with a 2 C rate at 25 °C. These cycling conditions were the only ones showing negative anode potentials in the PP3D cell results. Also, GD-OES depth profiling can be used as a valid ally to detect coatings homogeneity after electrode production.

Lithium metal is one of the most used reference electrodes in the academic world, however, there are concerns about its stability over time. With few modifications of the plate design, other materials such as Li4Ti5O12 (LTO) 61 or LiFePO4 (LFP) can replace lithium metal as a reference electrode. Because of the 3D printing versatility, several sensors could be easily implemented inside the plate. Implementation of other sensors and further electrochemical studies are currently under investigation and they will be discussed in a further paper.

Acknowledgments

This work was financially supported by project DigiBattPro 4.0 BW (3-4332.62-IPA/69) by the Ministerium für Wirtschaft, Arbeit und Wohnungsbau Baden-Württemberg and by the European Union's H2020 Framework Program under Grant agreement n°814106 (TEESMAT). The authors wish to thank Hitachi Chemical for the SMG-HT1 anode material they provided. The authors would like to thank G. Zettisch (ZSW) for the 3D printing process.

Please wait… references are loading.
10.1149/1945-7111/abfab9