Introduction

The increasing demand for rechargeable energy sources to power electronics, electric vehicles, and large-scale grid energy storage has driven extensive research of energy-dense lithium-based batteries1,2,3. To meet such demand, high energy density batteries other than state-of-the-art lithium-ion batteries (LIBs) with typical specific energies above 300 Wh kg−1 must be developed4,5. The metallic lithium negative electrode has a high theoretical specific capacity (3857 mAh g−1) and a low reduction potential (−3.04 V vs standard hydrogen electrode), making it the ultimate choice of negative electrode material for high energy Li-based rechargeable batteries1,6,7,8. Although Li metal (Li0) negative electrodes potentially enable batteries with high energy density, they tend to form dangerous Li0 morphologies (dendritic and the subsequent mossy Li0) and induce sustained electrolyte decomposition, which eventually leads to poor reversibility as indicated by low Coulombic efficiency (CE) and even safety hazards9. Considerable efforts have been devoted to addressing the challenges of Li0 negative electrodes, including interfacial engineering8,10,11,12,13,14,15, electrolyte engineering5,16,17,18,19,20,21, minimizing volume change by architecting stable hosts22,23,24,25, and preventing dendrite propagation with modified separators26,27,28.

Partial physical suppression of dendrite growth has been well achieved in the previous work, but to fully eliminate the unstable interface between the Li0 electrode and electrolyte, one must understand the fundamental mechanism of dendrite growth. The formation of Li0 dendrites is induced by the chemical and morphological inhomogeneity of the in-situ SEI on the Li0 surface, which leads to uneven local current density. Efforts have been made to relate dendritic Li0 formation to the interfacial kinetics as described by Sand’s equation (Supplementary Note 1)29,30 or to the diffusion-limited aggregation model by Chazalviel31, but the actual factors involved are far more complicated than the diffusion models derived for metal ion deposition from aqueous electrolytes, where the interphase does not exist. Nevertheless, a non-quantitative approximate relation exists between the fractal deposition pattern and the maximum interfacial current, where rapid consumption of Li+ in certain locations does result in concentration polarization, which invites local enrichment of Li+ and subsequent preferential deposition32. Hence, the diffusion energy barrier of Li+ at the interface should play a critical role, and various methods have been proposed to reduce the diffusion barrier33,34,35,36,37,38,39. The Li+ transference number, which quantifies the pure contribution from Li+ to the entire migration flux, has a decisive impact on the manner in which Li+ approaches the interphase-enclosed Li0 surface40. Consequently, the most directive and effective approach to guide Li0 deposition in an even and homogenous manner so that dendrite formation is minimized is to decrease the diffusion energy barrier and increase the Li+ transference number. Most current research toward increasing transference number focuses on single-ion conducting solid polymers, ceramic solid electrolytes, and their composites41,42. The former (polymers) offer both a rigid framework of interconnected nanopores and a high transference number, but are limited by low mechanical strength and poor ionic conductivity41,42, while the latter (ceramics) always encounter poor contact issues of solid-solid interfacing, where the advantage of high Li+ transference number become meaningless unless under high pressure. The brittle nature of the ceramic solid electrolytes further complicates the challenges41,42,43. Hence, it is highly desired to develop a new strategy to engineer the interface between the Li0 electrode and the electrolyte with both near-unity Li+ transference and robust mechanical strength to enable “dual protection” for the stabilization of the Li metal electrode.

Herein, we propose a rational design of an artificial SEI produced by treating Li0 with tetramethylethylenediamine (TEMED), which exhibits a low diffusion energy barrier, high Li+ transference number, and unrivaled mechanical strength to simultaneously overcome diffusion and advection-limited ion transport to achieve dendrite-free Li plating/stripping. Notably, TEMED spontaneously reacts with Li0 upon contact and forms pure α-phase Li3N. Differing from conventional Li3N artificial SEI that is fabricated from the exposure of Li0 in the N2 atmosphere, an artificial SEI achieved in this way offers excellent Li+ conductivity with a lower energy barrier for Li+ migration, directly benefitting ion transport at the interface between electrode and electrolyte. This effectively eliminates the uneven current distribution across the electrode surface, preventing preferential local growth of Li0 seedlings. The high modulus of Li3N ensures excellent mechanical strength that tolerates volume change to enforce a more uniform Li+ ion flux. The TEMED-treated symmetrical cell shows outstanding plating/stripping cycles with reduced overpotential and the full cell exhibits remarkably improved cycling stability and capacity retention as well as capacity utilization at high rates compared to untreated Li0. In this work, we demonstrate phase-pure artificial SEI on Li0 negative electrode that is capable of resolving compounded challenges faced by Li0 electrodes.

Results

Lithium chips were completely immersed into TEMED in a petri dish to ensure complete passivation of Li0 (Fig. 1a). A color change from shiny silver to light black and later to dark black is observed with the reaction time. To obtain the optimum reaction time, lithium chips were kept in the TEMED for 6 h, 12 h, and 18 h, respectively. Figure 1c shows that with a reaction time of 6 h, the film formed from TEMED has not fully covered the Li0 surface yet. With 12 and 18 h, full coverage by the artificial SEI layer is observed. Both visual and scanning electron microscope (SEM) inspection revealed that the artificial TEMED-based SEI layer formed during 6 h of reaction time does not cover the surface completely, which does not prevent the reaction between electrolyte with Li0, leading to consumption of both electrolyte and Li0 resulting in low CE and capacity decay. In comparison, the artificial SEI formed during 12 or 18 h fully covers the Li0 surface based on the SEM images, thereby preventing direct contact of the electrolyte with Li0. Cross-sectional SEM images (Fig. 1j–m) show the average thickness (t) of the artificial SEI layer obtained with different TEMED treatment times to be 5, 10, and 20 µm for 6, 12, and 18 reaction hours, respectively. It is assumed that a thicker SEI will have a higher Li+ barrier energy and higher impedance, resulting in slower Li+ diffusion. Thus, the SEI layer thickness should be optimized in order to prevent direct contact between Li0 and electrolyte while maintaining usefully high Li+ ion conductivity.

Fig. 1: Process of the artificial SEI layer.
figure 1

a Schematic for reaction of TEMED with Li0 to produce lithium nitride. be Photographic images of bare Li0 and TEMED-treated Li0 for different treatment times. fi The corresponding cross-sectional SEM images of bare Li0 and TEMED-treated Li0. jm The corresponding top-view SEM images of bare Li0 and TEMED-treated Li0.

Material characterization, electrochemical spectroscopy, and transference number

Contact angle measurements have been performed to determine the wettability of the electrolyte on untreated Li0 and TEMED-treated Li0 (Fig. 2a, b). To ensure good Li+ ion conductivity, rate capability, and formation of a stable SEI, the electrolyte must be able to significantly wet the electrode44. The contact angle for untreated Li0 is 41° suggesting poor wettability with the electrolyte which could cause lower ionic conductivity. In contrast, TEMED-treated Li0 shows a significantly lower contact angle of 12° with electrolyte, suggesting that the TEMED-originated SEI is better interfaced with the electrolyte, which is beneficial to homogenize ion distributions in the vicinity of the negative electrode, not only supporting efficient Li+ transport but also prevents the unwanted morphologies (dendritic or dead Li0) that are known to be induced by uneven Li+ flux distribution45. The phase purity of TEMED-treated Li0 was characterized by X-ray diffraction (XRD) (Fig. 2c). Distinct Li0 peaks of (110), (200), and (211) were observed at 36°, 52°, and 65°10 and Li3N peaks of (001) and (002) at 22.96° and 46.6°, respectively. The formation of hexagonal α-phase single crystal Li3N from TEMED-treated Li was also indicated by the transmission electron microscope (TEM) characterization as shown in the Supplementary Fig. 1. In comparison, conventional methods to obtain Li3N, where Li0 was treated with nitrogen flow show a polycrystalline structure with (001), (100), (002), (110), and (102) peaks for Li3N (Supplementary Fig. 2). XRD of Li3N after 100 cycles was also performed to examine the change of phase purity (Supplementary Fig. 3), revealing surprising structural stability of Li3N formed by TEMED against charge/discharge cycling, as evidenced by the strong α-phase peaks at 22.9° (001) and 46.6° (002). These α-phase Li3N diffraction peaks also reveal that the Li3N film obtained from TEMED treatment is highly orientated along the direction vertical to the Li0surface, hence offers excellent Li+ conductivity and implies a lower Li+ migration energy barrier46.

Fig. 2: Property comparison between untreated Li0 and TEMED-treated Li0.
figure 2

Contact angle measurements of a untreated Li0 and b TEMED-treated Li0. c XRD spectrum of TEMED-treated Li0. d EIS measurements of untreated Li0 and TEMED-treated Li0. Steady-state current under 10 mV polarization for e Li-Li symmetric cell f TEMED-Li/TEMED-Li symmetric cell. Inset shows EIS measurements before and after polarization.

Electrochemical Impedance Spectroscopy (EIS) measurements were performed to further characterize the electrode interface. Figure 2d and Supplementary Fig. 4 compare EIS for untreated Li0 and for TEMED-treated Li0. The first semicircle in the higher frequency range indicates the interfacial resistance of the artificial SEI or resistance of Li+ flux through an artificial SEI, while the second semicircle in the lower frequency range indicates the charge transfer resistance Rct. The symmetrical cell constructed on untreated Li0 showed a high resistance of ~400 ohms, whereas for the TEMED-treated Li0 symmetric cell, we observed a reduced resistance of ~200 ohms. We attribute the smaller overall impedance of the TEMED-treated cell to the better Li+ transport performance across the Li3N artificial SEI formed at the interface. The Li+ conductivity of the TEMED treated Li was calculated to be ~4.19 × 10−1 mS cm−1 based on the series resistance of the symmetric cell, a sufficiently high value to establish a fast Li+ exchange channel between Li0 metal and electrolyte10.

The transference number of Li+ at the interface between the Li0 electrode and the electrolyte was evaluated using Bruce-Vincent Approach. High cation transference numbers are desirable to avoid concentration gradients in the cell and to delay the nucleation and growth of lithium metal dendrites while charging the cell at a high rate. It should be noted that in conventional carbonate-based electrolytes, the transference number t+ is typically between 0.1 and 0.447. Although higher t+ can be obtained in polymeric, ceramic, or nanoparticle-based electrolytes, in which the anions are immobilized to a stationary or slow-moving support, low ionic conductivity and other compromises in properties such as interfacing or mechanical strength always accompany them. We expect that the lithium nitride layer with a pure α-phase on top of Li0 will address these conflicts. The transference numbers are determined in symmetric cells consisting of untreated Li-Li and treated Li-Li, respectively. The cell was initially conditioned to establish a stable interface by charging and discharging at 0.01 mA cm−2, with 4-h charge, 30-min rest, and 4-h discharge, with the process repeated 6 times. The cell was then polarized at 10 mV for 10 h to ensure a steady state (Fig. 2e, f). EIS spectra before polarization and after the steady state had been reached are shown in inset of Fig. 2e, f. The steady-state cation transference number was then calculated via Eq. (1).

$${t}^{+}=\frac{{I}_{s}(\Delta V-{I}_{o}{R}_{o})}{{I}_{o}(\Delta V-{I}_{s}{R}_{s})}$$
(1)

where t+ is the steady-state cation transference number, \(\Delta V\) is the applied voltage, \({I}_{o}\) is the initial current, \({I}_{S}\) is the steady-state current, \({R}_{o}\) is the initial interfacial resistance, \({R}_{S}\) and is the steady-state interfacial resistance. The calculated result shows that the TEMED-treated Li electrode exhibits a high steady-state cation transference number of t+ = 0.668, untreated in comparison with t+ = 0.37 for untreated Li0. This result strengthens our understanding that the TEMED-treated Li0/electrolyte interphase plays a dominant role in altering Li+ transport behavior. The improvement in the cation transference number with the artificial SEI layer has the potential to suppress dendrite growth by lowering the diffusion energy barrier and regulating the ion concentration at the interface in organic electrolyte rather than solid-state electrolyte.

Phase field simulation

The Arrhenius equation (Supplementary Note 3) indicates that a decrease in the activation energy leads to an increase in the diffusion coefficient. This decrease in activation energy for diffusion lowers Li+ migration energy barriers which increases ion transport at the interface between the electrode and electrolyte. The transference number is directly proportional to its diffusion coefficient. Comparative Arrhenius-plots for TEMED-treated Li0, N2 treated Li0 and untreated Li0 are shown in Supplementary Fig. 5, revealing an activation energy of 0.703 eV for untreated Li0, 0.613 eV for N2-treated Li0, and 0.48 eV. For TEMED-treated Li0, respectively. This successive decrease in activation energy results in a much higher Li+ mobility, which in turn decreases the concentration gradient across the corresponding SEI to provide a more uniform surface for Li+ migration and plating. To test this hypothesis and to further understand the mechanism for suppression of dendritic and dead Li0 growth by the Li3N-based SEI, we further characterized the activation energy of Li+ using integrated phase field simulations to elucidate the fundamental correlation between our novel artificial layer with phase purity and the Li+ transport behavior at the interface, which we then verify with the experimental results. A highly diffusive SEI is introduced on the Li0 surface to mimic the treated Li0 covered under artificial SEI. A small protrude is introduced on the surface of the Li metal to mimic the nucleus of Li0. The diffusivity of Li+ in the electrode (\({D}_{e}\)) and the electrolyte (\({D}_{s}\)) are set to be 4.6 × 10−13 cm2/s and 4.6 × 10−10 cm2/s, respectively, while the diffusivity in the artificial SEI layer (\({D}_{i}\)) is 3 times larger than \({D}_{s}\) for N2-treated Li0 and 10 times for TEMED-treated Li0. These values are calculated based on the activation energy obtained from experiments (Supplementary Fig. 5). Figure 3a–c shows snapshots of the Li0 dendrite structure on untreated Li0, N2-treated Li0, and TEMED-treated Li0 having Li3N as an artificial SEI after 400 s, respectively. For untreated Li0, we observe that an initial Li0 protrude grows into a filament-like dendritic morphology with side branches budding from the primary arm of the dendrite (Fig. 3a). For N2-treated Li0 negative electrode, Li0 dendrite forms and grows at a smaller growth rate, and the side growth of the primary arm of Li0 dendrite is hardly seen (Fig. 3b). Instead, the initial Li0 protrude forms a dome-like morphology with a smooth electrode-electrolyte interface, and its growth rate is significantly reduced. It can thus be inferred that the artificial SEI layer of higher Li+ diffusivity and higher Li+ transference number can indeed significantly suppress the dendritic Li0 growth. To further elucidate our findings, we plotted the 1D evolutions of Li+ concentration along the x direction across the tip of the dendrite, as indicated by the arrows in the 2D inset plots (Fig. 3c, d). The Li+ concentration at the tip of the dendrite increases sharply for untreated Li0 surface, whereas it increases rather gradually for treated Li0.

Fig. 3: Phase-field simulation of Li0 dendrite growth from untreated Li negative electrode, N2-treated Li negative electrode, and TEMED-treated Li negative electrode covered with Li3N SEI.
figure 3

Dendrite morphology grown on a untreated lithium metal, b N2 treated Li negative electrode, and c TEMED treated Li0 negative electrode represented by phase-field variable ξ. df 1D evolutions of Li+ concentration profile along x-axis across the tip of the dendrite for d untreated Li0, e N2 treated Li0, and f TEMED-treated Li0. The images on the inset show the 2D map of the Li+ concentration at t = 400 s. gi 1D evolutions of electric field profile along x-axis across the tip of dendrite for g untreated lithium, h N2-treated Li0, and i TEMED-treated Li0. The images on the inset show the 2D distribution of the local electric field at t = 400 s.

The electric field variation (\({E}_{x}\)) along the dendrite tip at different time steps for untreated Li0, N2-treated, and TEMED-treated Li0 are also compared in Fig. 3g, i. The local electric field remains almost constant in the electrode and electrolyte, but it is maximized at the tip of the dendrite for all the cases. However, for untreated Li0 (Fig. 3g), the maximum \({E}_{x}\) is 2 times higher for that of the N2-treated Li0 and 3 times for TEMED-treated Li0 (Fig. 3h, i), which is due to the sharper tip morphology with larger curvature, leading to a higher Li+ concentration gradient near the tip due to the higher local electric field that further facilitates the growth of the dendrite on an untreated Li0 surface. These results indicate that Li0 dendrite growth is a self-accelerating process, agreeing with previous reports48. We also observe that for the untreated Li0\(\,{E}_{x}\) at the dendrite tip and in the electrolyte solution increases to maximum over time. Whereas, for the treated Li0 negative electrode, the variation in \({E}_{x}\) at different time steps is much less significant, indicating that Li0 dendrite growth is significantly inhibited.

Elemental analysis, topography, and modulus mapping

X-ray photoelectron spectroscopy (XPS) was conducted on both TEMED-treated and untreated Li0 to decipher the chemistry of the artificial SEI as shown in Fig. 4. All the high-resolution spectrums were fitted by the Lorentzian in terms of spin-orbit doublets. Figure 4b shows the high-resolution N 1s spectrum, with peaks at 398.3 eV assigned to Li3N, which is known to be a good Li+ conductor. The absence of any N 1s peak for untreated Li0 (Fig. 4e) confirms that the presence of Li3N arises solely from the reaction between TEMED and Li0. The C 1s (Supplementary Fig. 6) confirms that there is no presence of any other organic moieties, and the excellent performance of the battery is solely by the presence of Li3N. Furthermore, we also detected the presence of N from the energy dispersive spectrum (EDS), which shows the distinct presence of N and a uniform distribution of N over the surface of the TEMED-treated Li0 (Supplementary Fig. 7). Based on the above analysis derived from diversified techniques, we believe that this N-rich SEI stabilizes the Li0/electrolyte interface, leading to uniform Li0 electroplating and increased cycle life.

Fig. 4: Surface analysis of untreated Li0 and TEMED-treated Li0.
figure 4

a Li 1s, b N 1s, and c O 1s XPS spectra of TEMED-treated Li0. d Li 1s, e N 1s, and f O 1s XPS spectra of untreated Li0. g AFM topography and h Young’s modulus mapping of untreated Li0. i AFM topography and j Young’s modulus mapping of TEMED-treated Li0.

Atomic force microscopy (AFM) was employed to visualize surface topography and measure the corresponding Young’s modulus of untreated and TEMED-treated Li0 (Fig. 4g–j). The surface roughness values of the untreated and TEMED-treated Li0 were compared by measuring the average surface root mean square (RMS) via high-resolution AFM, which for untreated Li (Fig. 4g) and TEMED-treated Li0 (Fig. 4i) are 242 and 157 nm, respectively. The higher RMS value for untreated Li0 implies uneven and rough surfaces that can induce high local current at the protuberances and encourage Li0 dendrite on electrode surface49,50. In contrast, the smooth surface of TEMED-treated Li0 provides a route for uniform Li0 plating. The corresponding Young’s modulus mapping values of untreated (Fig. 4h) and TEMED-treated Li (Fig. 4j) exhibit an average Young’s modulus values of 0.32 and 6.85 GPa, respectively. We attribute the 20 times higher Young’s modulus value of TEMED-treated Li0 to the superior structural efficiency and strong mechanical strength of the highly oriented α-phase Li3N. This Young’s modulus value is significantly higher than the threshold value of 6.0 GPa for Li0 growth, indicating that the TEMED-originated SEI is mechanically strong to suppress the dendritic Li0 upon its crystallization11.

Electrochemical performance of TEMED-treated Li

To evaluate the electrochemical performance of the TEMED-based lithiophilic interphase, symmetric cells with pristine and TEMED-treated Li0 were cycled at various current densities (0.5 and 1 mA cm−2) with the platting/stripping capacity of 1 mAh cm−2 in an electrolyte consisting of 1.0 M LiTFSI in 1,3-dioxolane/1,2-dimethoxyethane (DOL:DME = 1:1 by vol). Voltage profile versus cycling time, and voltage hysteresis (estimated by calculating the average difference between the voltage of Li stripping/plating) versus cycle number are shown in Fig. 5.

Fig. 5: Symmetric cell performances of untreated Li0 and TEMED-treated Li0.
figure 5

a, b Galvanostatic cycling and voltage hysteresis at 0.5 mA cm−2/1 mAh cm−2. (insets show (i) short circuit for untreated Li0, (ii) plating/stripping behavior of TEMED-treated Li0 at 1600–1640 h). c, d Galvanostatic cycling and voltage hysteresis at 1 mA cm−2/1 mAh cm−2. (insets show (i) short circuit for untreated Li0, (ii) plating/stripping behavior of TEMED-treated Li0 at 420–440 h). Schematic illustration of e Li0 dendrite growth on untreated Li0 and f uniform deposition on TEMED-treated artificial SEI. SEM images of gi untreated Li0, jl TEMED-treated Li0 at 5th, 20th, and 100th plating, respectively. The thickness of lithium chip is 450 μm. The scale bars are at 20 µm.

The plating/stripping voltage profile of untreated and TEMED-treated Li0 was carried out to investigate the electrochemical stability of the TEMED-originated SEI. Figure 5a, c show the voltage profiles of plating/stripping for symmetrical cells constructed on untreated and TEMED-treated Li0 that achieved a capacity of 1 mAh cm−2 at the current density of 0.5 and 1 mA cm−2, respectively. At a low current density of 0.5 mA cm−2, untreated Li0 exhibited large voltage divergence after 150 cycles, and short circuit after ~600 h. However, TEMED-treated Li0-based symmetric cell showed a stable voltage profile with hysteresis below 20 mV, reflecting the stable plating and stripping process for more than 3500 h. Even at a higher current density of 1 mA cm−2, TEMED-treated Li0 showed stable plating and stripping for more than 500 cycles (1000 h), whereas untreated Li0 failed after ~400 h. TEMED-treated symmetric cells show stable performance even at a higher current density of 2 mA cm−2 (Supplementary Figs. 8 and 9) and 5 mA cm−2 (Supplementary Figs. 10 and 11). 700 h and ~350 h have been achieved for 2 and 5 mA cm−2, respectively, for TEMED-treated Li0, compared to 150 and 50 h for untreated Li0. TEMED-treated Li0 also showed a stable plating and stripping cycle with commercial LiPF6-based electrolyte (Supplementary Figs. 12 and 13). The cell showed a stable plating and stripping for more than 750 h at the current density of 1 mA cm−2 for TEMED-derived SEI whereas untreated Li0 failed after only 270 h (Supplementary Figs. 14 and 15). The voltage hysteresis leads to the same conclusion for untreated Li0, an increase in voltage hysteresis was observed with increasing cycles in Fig. 5b, d. The overpotential increases continuously, leading to early failure of the cell after only 200 cycles (~400 h). This large hysteresis implies the formation of a highly resistive and unstable interphase. The unstable SEI formed during cell operation continues to consume electrolyte to repair new SEI, accompanied by the formation of dendritic and dead Li0, eventually leading to the early failure of the cell10. Symmetrical cell performance of the TEMED-treated Li0 at high current and high capacity with 50-µm-thick Li0 have also been performed to determine the compatibility of TEMED-originated SEI. 900 h and ~600 h have been achieved for the current density of 1 mA cm−2 and 5 mA cm−2 at the capacity of 3 mAh cm−2, respectively (Supplementary Figs. 16 and 17), for TEMED-treated Li0, compared to 300 and 120 h for untreated Li0. Even at the high capacity of 3 mAh cm−2 with 50-µm-thick Li0, TEMED-originated SEI ensures longer cycle life.

To better understand the morphology of Li deposition on untreated and TEMED-treated Li0, we performed SEM examination after 5, 20, and 100 cycle. Figure 5e shows a schematic illustration of the growth of dendritic and dead Li0 on the untreated Li0 after plating and stripping cycles, where the native SEI from the reaction between Li0 and electrolytes are fragile, non-uniform, and unstable. Such SEI can be easily ruptured during electrode volume changes and by uneven plating/stripping.

In contrast, TEMED-treated Li0 (Fig. 5f) prevents side reactions of Li0 with the electrolyte. Figure 5g–l shows the SEM images of untreated Li and TEMED treated Li after 5th, 20th, and 100th plating at 0.5 mA cm−2 with a capacity of 1 mAh cm−2. For untreated Li0 we observed uneven Li0 plating and dendrite growth starting from the 5th cycle (Fig. 5g). The unregulated and unprotected surface of the untreated Li0 creates large protuberances generating a non-uniform electric field. Dendrite growth is also promoted because of the uneven surface and locally concentrated Li+ flux because the sharp end of these dendrites serves as a center at which charges tend to accumulate10,51. This needle-like structure with a sharp end may also penetrate through separator to cause an internal short circuit that results in safety issues52,53. In addition, the high surface area associated with the dendritic morphology and side reactions results in an extremely low CE.

In contrast, the TEMED-treated Li0 leads to a dense and nodule-like morphology in the absence of Li0 dendrites. Cross-sectional SEM (Supplementary Fig. 18) showed ~50 µm thickness Li0 after 100 cycles as compared to ~19 µm for the TEMED-treated Li0. Even after 100 cycles, the surface of TEMED-treated Li0 still maintained a compact surface without discernible dendrites (Fig. 5l). Hence, our structural characterization at a high stripping/plating rate over long cycling times further supports the hypothesis that high transference number of Li+ across TEMED-originated SEI improves Li+ mobility, which in turn decreases the concentration gradient, leading to uniformity of Li0 electrodeposition with suppressed lithium dendrite formation.

Nucleation overpotential is defined as the voltage difference between the beginning voltage dip and the following flat plateau during plating, which is also known as the Li nucleation barrier (Supplementary Fig. 19). Lower nucleation overpotential signifies higher lithiophilicity and is preferred for higher reversibility of Li0 chemistry. Untreated Li0 shows a higher nucleation overpotential of 37 mV indicating a significantly large energy barrier whereas, TEMED-treated Li0 shows higher lithiophilicity with the lowest nucleation overpotential of 15 mV. Electrochemical performance of L−2i0||Cu half cell and TEMED-Li0||Cu cell with the thin Li0 chip of 50 µm thickness have been performed at the cycling current density of 1 mA cm−2 and capacity of 1 and 3 mAh cm−2 (Supplementary Fig. 20 and 21, respectively). The CE of the untreated-Cu cell decreases rapidly after only 75 cycles; however, the CE of the TEMED-Li0||Cu cell maintains a stable cycle during 175 cycles for a capacity 1 mAh cm at current density of 1 mA cm−2. The enhanced CE and lifespan were also significant when the capacity was increased to 3.0 mAh cm−2 (Supplementary Fig. 20).

To evaluate the compatibility of TEMED-treated Li0 as a negative electrode for practical LMBs, we adopted lithium iron phosphate (LFP) and NMC-111 as two positive electrode materials to assemble a full cell LMB. Figure 6a shows the cycling performance of the full cell using untreated or TEMED-treated Li0 as the negative electrode at a constant specific current of 160 mA g−1. We observed linear degradation in capacity for the full cell with untreated Li0 as the negative electrode (Fig. 6a). However, with the TEMED-treated Li0, we obtained a steady and stable capacity for the full cell. In rate capability tests (Fig. 6b), at lower specific currents of 32, 80, and 160 mA g−1 we observed comparable capacity for untreated and TEMED-treated Li0. However, high specific currents of 800 to 80 mA g−1 lead to a large capacity loss for untreated Li (Supplementary Figs. 22 and 23). TEMED-treated Li0, on the other hand, recovers almost 100% of capacity, due to the stable dendrite-free Li0 plating/stripping of the TEMED-treated Li0. LFP-based full cell with a high mass loading (~9.5 mg cm−2) has also shown an improved performance for the TEMED-treated Li0 as compared to the untreated Li0. During the 10th cycle Li0/LFP showed a discharge capacity of ~135 mAh g−1 whereas TEMED-Li0/LFP showed a discharge capacity ~123 mAh g−1 (Supplementary Fig. 24). After the 50th cycle (Supplementary Fig. 25) capacity retention of ~74% has been obtained for TEMED-Li/LFP in comparison with ~50% for untreated Li0/LFP full cell. Similar performance has been obtained for NMC-based cells where a drastic decline in capacity for untreated Li/NMC with capacity retention of ~48% has been obtained as compared to 73% for TEMED-Li/NMC after 200 cycles (Supplementary Figs. 2628) suggesting that the artificial SEI originated from TEMED is more stabilized.

Fig. 6: Electrochemical performances of full cells.
figure 6

a Cycling performance of full cells based on LFP positive electrode and untreated or TEMED-treated Li negative electrode at a specific current of 160 mA g−1. b Rate performances of full cells based on LFP positive electrode and untreated or TEMED-treated Li negative electrode. c, d Charge/discharge voltage profiles at different cycles of full cells based on LFP positive electrode and untreated or TEMED-treated Li negative electrode at 160 mA g−1. The mass loading of LFP is ~2.0 mg cm−2.

Figure 6c, d shows the cycling performance and voltage profile of full cells using both untreated Li and TEMED-treated Li as a negative electrode. Untreated Li0/LFP full cells showed a lower CE in the first cycle, which could be attributed to Li0 consumption and electrolyte decomposition to form the SEI. TEMED-treated Li0 showed ~100% capacity retention from the 10th discharge to the 100th discharge. In contrast, the 10th specific discharge capacity of untreated Li/LFP full cell shows a sharp decrease from ~140 to 102 mAh g−1, retaining 72.8% at the 100th cycle. With increased cycle numbers, the decrease in capacity retention becomes more prominent for untreated Li/LFP. Reduced overpotentials were also observed for the TEMED Li/LFP full cell whereas, bare Li/LFP shows higher overpotential due to the loss of Li and consumption of electrolyte from side reactions and an un-stabilized SEI with dendrites, leading to overpotential and sluggish Li-ion transportation. The stable cycle life and low polarization potential suggest that the TEMED-treated Li negative electrode is capable of working under practical cycling conditions.

Discussion

In this work, we demonstrated a facile and efficient solution processed method to provide phase pure lithium nitride (Li3N) as a protective SEI, which successfully suppresses the dangerous and unstable morphologies of dendritic and dead Li0, due to the low electronic conductivity and the intrinsic electrochemical stability of Li3N. This artificial SEI layer offers excellent Li+ conductivity with lower Li+ migration energy barriers that further benefit ion transport at the interface between the electrode and electrolyte.

As a result, the TEMED-originated SEI ensures long stable plating/stripping cycling up to 3500 h at 0.5 mA cm−2 along with a full cell cycling up to 500 cycles at 160 mAg−1 (1C rate). These dendrite-free TEMED treated Li should facilitate applications of high energy density Li metal batteries.

Methods

Materials and synthesis

Li chips (diameter size = 15.6 mm and thickness = 450 μm) were purchased from MTI Corp. Tetramethylethylenediamine (TEMED) was purchased from Sigma-Aldrich. TEMED was used without any further modifications. Lithium chips were allowed to be completely immersed into the TEMED in the petri dish and were kept overnight. The Li chips were allowed to be dried at 60 °C for half an hour to let the unreacted liquid evaporate away. The dried Li chips were then used for further analysis and cell fabrication.

Electrode fabrication

Lithium iron phosphate (LFP) powders were mixed with Super-P carbon black and polyvinylidene fluoride (PVDF) at a weight ratio of 80:10:10, respectively, in the N-methyl-2-pyrrolidone (NMP) solvent to form a slurry using the mortar and pestle. Similarly, for Lithium nickel manganese cobalt oxides (NMC) based positive electrode, NMC powders were mixed with Super-P carbon black and polyvinylidene fluoride (PVDF) at a weight ratio of 80:10:10, respectively, in the N-methyl-2-pyrrolidone (NMP) solvent to form a slurry using the mortar and pestle. The slurry was coated on an aluminum foil current collector by doctor blading and then dried in the vacuum oven at 80 °C for 12 h. The dried samples were cut into circular disks with a diameter of 12 mm and used as the working electrode. The total areal mass loading of the NMC electrode was ~2.5 mg cm−2 and the areal mass loading of active material LFP was ~2.0 mg cm−2. For the high mass loading full cell, the total mass loading was ~9.5 mg cm−2.

Electrochemical characterization

The CR-2032 Li-ion coin cell was assembled inside an argon-filled glove box (moisture and O2 level <0.1 ppm) for all the electrochemical measurements. Celgard 2500 with a film thickness of 25 μm was used as a separator. The electrolyte was 1 M Lithium bis(trifluoromethanesulfonyl)imide (LiTFSI, Sigma-Aldrich) in 1,3-dioxolane (DOL, Sigma-Aldrich)/1, 2-dimethoxyethane (DME) Sigma-Aldrich) where the volume of both DOL and DME were 1:1 (1:1 volume ratio) with 1 wt% Li nitrate (LiNO3, Alfa Aesar). Commercially available electrolyte 1 M Lithium hexafluorophosphate (LiPf6) in ethylene carbonate (EC)/diethyl carbonate (DEC) (1:1 volume ratio) has also been used for the comparative analysis of the performance of the TEMED treated Li0 in both the electrolytes. For the full cell, LiTFSI in DOL/DME (1% LiNO3) has been used for both LFP and NMC positive electrodes. The amount of electrolyte used was controlled as ~60 μL for each cell. Cells were tested under a different current density of 0.5 and 1 mA cm−2 with a capacity of 1 mAh cm−2 using Land battery analyzers (CT2001A).

Electrochemical impedance spectroscopy (EIS) was carried out by a Biologic VSP potentiostat with 10 mV amplitude AC signal with frequency ranging from 0.1 Hz to 100 K Hz with 6 points per decade. The calculations of diffusion coefficient and activation energy can be found in Supplementary Notes 2 and 3. XRD of the samples was conducted on a Rigaku SmartLab diffractometer with Cu Kα radiation (λ = 1.54178 Å). Topography and Young’s Modulus of untreated Li and graphite–SiO2 Li were measured using an Agilent SPM 5500 atomic force microscope equipped with a MAC III controller using a tip (product RTESPA-525) with the resonance frequency of 75 kHz. Raman spectroscopy was carried out using the Horiba Raman system with a 532 nm laser.

Galvanostatic charge-discharge measurements of the coin cells were carried out using the LAND CT2001A system. Plating/stripping of the symmetric cells was performed at various areal current densities from 0.5 to 5 mA cm−2 to achieve various areal capacities from 1 to 3 mAh cm−2. Full cells were cycled at a constant current density of 160 mAg−1 (1 C) and at various current density rates from 80, 160, 320, and 800 mAg−1 for every 10 cycles and followed back to 80 mAg−1. LFP-based full cells were cycled at the voltage range between 2.5 and 4.2 V at 160 mAg−1 (1C) and NMC-based cells were cycled at the voltage range of 2.7 to 4.2 V. All electrochemical tests were performed at 25 °C in an environmental chamber.

SEM characterization was carried out using a Hitachi S-4300N SEM. TEM characterization was carried out using a JEOL 2100F TEM with 200 kV field emission. The samples utilized for TEM characterization were prepared by plating lithium onto a carbon film-supported copper grid by a coin cell, which served as the TEM sample support. Subsequently, these samples underwent TEMED treatment to replicate our experimental conditions.

Phase-field simulation

The phase field simulations were performed on COMSOL. The Multiphysics software used general PDE and the solver was set as time-dependent. The details are described in Supplementary Note 4. The size of the model is chosen to be 500 × 500 µm2. Dirichlet boundary conditions were selected for the Nernst-Planck equation (Equation 6 in Supplementary Information) and the current continuity equation (Equation 7 in Supplementary Information), while the zero-flux boundary condition is set for the phase-field variable (\(\xi\)). \({C}_{{Li}}\) is fixed at 1.0 mol/L in the electrolyte and 0.0 mol/L in the electrode, while \(\phi\) is fixed at −0.35 V in the electrode and 0.0 V in the electrolyte as the boundary conditions. The initial state is a pure electrolyte, in which \(\xi\) and \(\phi\) are set to be zero, while the initial value for \({C}_{{Li}}\) is set to be 1.0 mol/L. Then, a semi-circle type random noise is added on the surface of the Li negative electrode, which acts as a nucleus for Li dendrite growth. The diffusivity of Li-ion is set to be \({D}^{s}\) = 3.05 × 10−10 m2/s in the electrolyte solution, and \({D}^{e}\) = 3.1 × 10−13 m2/s in the Li electrode, based on the activation energy of untreated Li from experimental measurement. To study the effect of the Li3N protective layer on the Li deposition morphology, we introduced a highly diffusive SEI layer on the surface of the Li negative electrode to mimic the treated Li. The diffusivity of this layer (\({D}^{i}\)) after N2 and TEMED treatment increases by 2 times and 10 times than untreated Li, based on the activation energy of treated Li from experimental measurement.

Density functional theory calculation

All density functional theory (DFT) calculations were carried out by using the Vienna ab initio simulation package (VASP)54. The projector-augmented wave method was used to account for core-valence interactions55, and the generalized gradient approximation (GGA) in the form of the Perdew–Burke–Ernzerhof functional was used for the exchange-correlation interactions56. The electronic wave functions were represented by a plane-wave basis set with a cut-off energy of 500 eV. A 3 × 3 × 3 Li3N supercell structure including 108 atoms was used, and a 2 × 2 × 2 Monkhorst–Pack grid was used for the Brillouin zone integration. The convergence criteria were 1 × 10−5 eV energy differences to solve the electronic wave function, and all atoms were relaxed until the forces were less than 0.01 eV/Å. To determine the Li diffusion pathways and migration barriers, a supercell with one Li vacancy was prepared by removing an individual Li atom. The energy profiles along Li migration were calculated as shown in Supplementary Table 1 and Supplementary Figs. 29 and 30 by the climbing-image nudged elastic band (NEB) method57.