Hierarchical structure design of AOB binder
To mimick the hierarchical structure of spidroin, it starts with designing the binder topological structure with dynamic ionic bonding. To accomplish this, polymer design based on acid-base interaction is a viable route. As mentioned above, traditional water-soluble binder PAA usually suffers from large elasticity modulus yet low elongation at break at ambient temperature, leading to irreversible electrode film fracture and thus fast degradation of cycling performance.23,24 To improve the elongation at break of PAA, it is hopeful to introduce other polymers that can construct dynamic ionic bonding with it.25,26 As a typical nitrogen-rich heterocyclic backbone, tetrazole group is an excellent proton acceptor with multiple coordination sites to form acid-base interactions.27,28 Another benefit is that the conjugated tetrazole structure can form reversible lithiation sites, beneficial to transport lithium ions and thus to improve rate performance of lithium batteries.29,30 As a consequence, a robust binder system with a strong dynamic ionic bonding network can be constructed by mixing PAA with a tetrazole groups-containing polymer.
Therefore, a hydrophobic, tetrazole groups-containing polymer PPB consisting of PAN and PEG segments was designed, both of which help improve elasticity and meanwhile conduct lithium ions. A synthesis schematic diagram of PPB is shown in Fig. 2a. Via a simple "click" reaction between − CN and − N3 groups, PPB was obtained via thermally induced cycloaddition of N3 − PEO − N3 and PAN under 140 oC.31 This reaction occurred smoothly, as evidenced by the peak change of − CN and − N3 groups in 1H-NMR, 13C-NMR and FTIR spectra (Figure S1-3). Meanwhile, the N 1 s XPS spectrum further supports this result; four peaks generate corresponding to N = C at 398.7 eV, N − C at 400.4 eV and +HN − C bonding at 401.5 eV on tetrazole units (Figure S4).32,33
In the AOB binder (Fig. 2a), the strong ionic bonds between tetrazole with carboxyl motifs was confirmed through FTIR spectra (Fig. 2b). For PAA, the stretching vibrational absorption peaks of − OH, C = O, C − O are located at 2500 ~ 3000 cm− 1, 1708 cm− 1 and 1242 cm− 1, respectively. Tetrazoles in PPB contain four nitrogen atoms as coordination sites, which can all act as carboxylic proton acceptors to form hydrogen bonding. After mixing with PPB, the weakened absorption peak of C = O in PAA accompanied by an obvious redshift (1708→1677 cm− 1) occurs, indicating that the carboxylic acid in PAA coordinates with nitrogen atoms of tetrazoles in PPB forming tight ionic bonding NH+∙∙∙–O − C = O. Meanwhile, the characteristic peak intensity of C − O in carboxylic acid of PAA and of C − N in tetrazoles of PPB enhances, accompanied by apparent blueshifts (1242→1265 cm− 1, 1163→1172 cm− 1, respectively). Furthermore, the peak intensity of –OH in PAA becomes weak and a new peak (2700 cm− 1) generates due to the generation of − NH+ on tetrazoles. Temperature-dependent real-time IR analysis was employed to further demonstrate the presence of such ionic bonding in AOB binder. As shown in Fig. 2c, with the increase of temperature, the peaks ascribed to C − N, C = O and C–O stretching vibration gradually shift from ~ 1172 to ~ 1164 cm− 1, from ~ 1670 to ~ 1700 cm− 1 and from ~ 1265 to ~ 1225 cm− 1, respectively (Fig. 2c), suggesting that the breakage of NH+∙∙∙–O − C = O and hydrogen bonding involving carboxyl, and the generation of a mass of free carboxylic and tetrazole groups. Evidently, the dynamically reversible ionic bond network within AOB binder helps dissipate stress and thus can accommodate enormous electrode volume deformation.34
Further theoretical calculation shows that the ionic bond energy between every N of tetrazole with carboxyl group is similar (ranging from − 11 to − 14 kcal mol− 1), but much larger than the hydrogen bonding formed between carboxyl groups themselves (–4.7 kcal mol− 1) (Fig. 2d). It has been extensively established that such ionic bonding can reconstruct at shattered interfaces, restoring the mechanical strength and original shape of materials.35 Thus, the dynamic ionic bonding allows AOB binder to deliver a good self-healing ability of as-prepared electrode.
Given the hierarchical structure of spidroin is highly linked with the assembly of corresponding polyamino acid segments in aqueous solutions, here, a facile aqueous-oil binary mixing process was conducted to regulate the condensed structure of AOB binder. As fair comparison, the PAA/PPB blend binders in sole water or sole NMP (denoted as PAB-W and PAB-N binder, respectively) were also prepared. To begin with, a PAA/PPB weight ratio of 3/1 was selected in the condensed structure regulation research. As depicted in Figure S5a, the PAB-N film shows a uniformly dispersed fiber morphology. After addition of water, the sample surface shows many uniformly dispersed or unequally distributed micron-sized or submicron-level irregular rod- or sphere-like clusters (Figure S5b-d). By adjusting the volume ratio of water/NMP, the condensed structure of AOB binder can be regulated. It is evident that AOB binder films prepared with a water/NMP ratio of 1/5 exhibit the medium uniformly distributed clusters; submicron-sized irregular spherical domains are observed via higher resolution SEM and top-viewed AFM height sensor imaging analyses (Figure S6b and S7). Notably, SEM mapping demonstrates that there are much more N elements distributed in the cluster domains than the rest of the film, while C elements are relatively uniform in the AOB binder (Figure S6c-d). For further microstructure analysis of the AOB binder film, confocal laser scanning microscope (CLSM) and atomic force microscope (AFM) were performed. Apparently, distinct distribution of PPB and PAA in the AOB binder film can be observed (Figure S8). Consistent with the SEM element mapping results (Figure S6c-d), PAA is uniformly distributed in the whole film (Figure S8c), while the PPB exists more in the submicron-sized irregular spherical domains (Figure S8b); these results are highly correlated with the ionic bonding-induced co-assembly of PAA and PPB. X-ray diffraction demonstrates that there is a strong crystalline peak at 16.78° with d spacing of 5.28 Å in the AOB binder film, in sharp contrast to that of PAB-N film (Figure S9a-b). This result clearly proves that the aqueous-oil binary mixing process renders the generation of crystalline zones in the AOB binder film. Given that PAA is amorphous (Figure S9c), it can be concluded that the crystallization domains belong to the ordered assembly of PAN segments in PPB. Furthermore, considering the oleophilic PPB distribution, the crystallization domain is prone to existing in the submicron-sized irregular spherical clusters of AOB binder films. This indicates that submicron-sized irregular spherical clusters formed by PPB within the AOB binder are, in part, stacked in order. More importantly, the crystallization peak is also present in the Si electrode with AOB binder (Figure S9d-e).
Figure 2e vividly illustrates the condensed structure formation of AOB binder via an aqueous-oil binary mixing process. In the cosolvent, by virtue of strong ionic bonding interactions between PAA and PPB, submicron-sized irregular spherical clusters can generate through the co-assembly of major PPB with minor PAA, and other PPB and PAA synergistically constitute long polymer chains that undergo physical/ionic crosslinking themselves forming the uniform domain in the AOB binder film. During this process, part of PAN segments on PPB can crystallize within the submicron-sized irregular spherical clusters. Eventually, the spidroin-like hierarchical structure binder with amorphous and crystallization domains is constructed. While the PAB-W film shows clusters with irregular shape, size and distribution area, which have negative impacts on the electrode particle dispersion (mentioned below).
Physical property evaluations of AOB binder
Via stress-strain curves strength analyses, the water/NMP ratio of 1/5 in the aqueous-oil binary mixing process of AOB binder samples, which renders superior mechanical properties, was chosen as the optimum one for following research (Figure S10). Additionally, it is worth mentioning that compared with PAB-N and PAB-W binder, the AOB binder renders enhanced mechanical and adhesion properties (Figure S10-11), clearly proving the significance of binder hierarchical structure design. Compared with the traditional PAA binder film, AOB binder achieves evidently enhanced tensile strength (68 vs. 21 MPa) and higher elongation at break (22.1% vs. 5.5%) (Fig. 3a). To reflect the practical mechanical properties of binders within cells, the AOB binder films were soaked in electrolyte for 24 hours. Figure 3b shows the sequential loading-unloading curves of the as-prepared electrolyte soaked AOB binder film at a strain limit of 30%. The stress and strain values remain approximately constant for four cycles, indicative of the excellent recoverable behavior of AOB binder films. Afterwards, AOB binder films can achieve a tensile strength of 55 MPa and an elongation at the break of 58%. Superior mechanical properties of AOB binder films are also reflected in the nanoindentation test. At a given nanoindentation force (a maximum load of 500 µN), AOB binder films exhibit an evidently smaller indentation depth (479 vs. 738 nm) than PAA binder (Figure S12). Moreover, the Si electrode with AOB binder shows much higher reduced modulus (1.08 vs. 0.6 GPa) and hardness (0.08 vs. 0.03 GPa) than those of PAA binder (Fig. 3c). Enhanced mechanical properties of the as-developed binder are more favorable to decrease Si-based electrode volume expansion.36
In a 180° peeling test for Si electrodes, the average peel strengths of the AOB and PAA binders based Si electrodes are 2.56 N and 0.64 N, respectively (Fig. 3d), showing the enhanced adhesion to Cu current collector. Additionally, the wettability test on Si surface was conducted to provide good insight into the affinity of binder to Si. Polymer solutions containing 1 wt% of AOB or PAA binder were dropped onto the surface of the monocrystalline silicon wafer to conduct contact angle measurements. After standing for 2 min, the contact angles of the AOB and PAA binder solutions are 39.1° and 50.6°, respectively (Figure S13), indicative of the improved wettability of the AOB binder solution. Enhanced wettability can prevent stress concentration at the interface between binder and Si anode from forming a large number of defects; this effect is effective to stabilize the Si-based electrode structure.37 This observation can be explained by the higher affinity of AOB binder to the Si surface due to abundant polar groups such as carboxyl and tetrazole. Furthermore, the AOB binder film exhibits a lower swelling ratio (4% vs. 6%) of electrolytes than the PAA counterpart (Figure S14), which helps retain mechanical strength and high adhesion during battery cycling. More impressive is that AOB binder films exhibit an excellent self-healing ability. It is demonstrated that two separated AOB binder films can recover within 12 hours under room temperature (Figure S15), mainly attributed to its dynamically reversible ionic bonding network. Superior mechanical, adhesive and self-healing abilities of AOB binder are anticipated to better withstand the huge volume change of Si particles and suppress excessive volume expansion of Si-based electrodes.
Evolution of electrodes during cycling
SEM imaging was used to explore the surface morphology evolution of Si electrodes with varied binders before and after 50 cycles. As shown in Fig. 4a-b, for the pristine Si electrode, AOB binder renders more uniformly dispersed electrode particles than PAA binder, mainly due to the superior wettability of AOB binder to Si. In sharp contrast with AOB binder, PAB-W binder incurs large cluster and obvious cracks on the pristine Si electrode surface (Figure S16), which is highly correlated with its clusters with irregular shape, size and distribution area. After 50 cycles, surface SEM images show more cracks on PAA based Si electrodes, compared with the AOB binder based one (Fig. 4c-d). Furthermore, cross-sectional ion milling-scanning electron microscopy (IM-SEM) analyses present the thickness variation of Si electrodes with different binders before and after 50 cycles (Fig. 4e-h). The AOB binder-based Si electrode shows a thickness enhancement of ~ 22% (18→22 µm), much smaller than that (~ 50%, 20→30 µm) of PAA counterparts. Notably, there were several cracks in the cycled PAA-based electrode, in sharp contrast to the well integrity of the one with AOB binder. In addition, the thickness change of electrode during the first lithiation and delithiation process was observed through the in-situ optical microscope (OM) in real time. It is evident that the PAA-based Si electrode thickens severely after first lithiation (Movie S1). In sharp contrast, AOB binder can better suppress electrode expansion (Movie S2). For understanding the effect of varied binders on electrode evolution during cycling, finite element simulation was carried out to investigate the contact stress on the surface of Si particles at the phase of Li15Si4 (Fig. 4i-k and S17). It can be seen that the Von Mises stress between adjacent Si particles in the presence of PAA binder is much higher than that (~ 2000 vs. ~800 MPa) of AOB binder. These results demonstrate that the AOB binder can better buffer the volume stress and thus maintain the integrity of Si anode. This is mainly ascribed to its superior mechanical properties benefited by the rational hierarchical structure design.
Decreased Si electrode expansion in the presence of AOB binder helps to maintain solid electrolyte interphase (SEI) stability. This can be reflected by surface chemical components of cycled Si electrodes characterized through X-ray photoelectron spectroscopy (XPS) measurements. In the XPS C1s spectra, one of the most noticeable differences is the intensity of the peak at 289.6 eV corresponding to electrolyte decomposition side products ROCO2Li /Li2CO3 (Figure S18a-b);38 the AOB binder based Si electrode shows a weaker signal intensity ratio than that of the one with PAA binder, demonstrating that less electrolyte decomposition occurs at the interface of the cycled AOB binder based Si electrode. Moreover, in the F 1s spectra, there were three typical peaks, LiF (685 eV), LixPOyFz (687 eV) and LixPFy (689 eV) (Figure S18c-d).38,39 Apparently, the Si electrode with AOB binder shows higher LiF, which is a crucial component of SEI offering preferable mechanical stability. The N 1s spectrum of the AOB based anode surface after 30 cycles shows a strong N–Li peak (Figure S19), corresponding to lithiation of tetrazole motifs, which is beneficial to transport lithium ions. This finding proves that AOB binder tends to participate in SEI formation. To elucidate the mechanism behind this, density functional theory (DFT) calculations were performed. As shown in Figure S20, AOB binder has the lower-lying LUMO level (− 0.97 eV) than those of PAA, PPB and carbonate solvents (i.e. DMC and EC). This finding demonstrates that AOB binder would be subjected to be reduced prior to the electrolyte solvents and participate in the formation of polymer-reinforced SEI; such a function of AOB binder toughens SEI and effectively decreases electrolyte decomposition, beneficial to electrochemical performance of Si-based anodes.
Si-based electrode performance evaluations
As plotted in Figure S21, the electrochemical stability of Si/Li half-cells with AOB binder was measured via cyclic voltammetry (CV) analysis at a scan rate of 0.5 mV/s. Two broad peaks around ~ 0.37 and ~ 0.51 V correspond to the delithiation process of Li-Si phases.40 Moreover, a reduction peak at about 0.19 V appears, assigned to the reversible lithiation process of amorphous Si domains to Li-Si phase. The intensity of these peaks increases with cycling, attributed to the formation of ion-conducting SEI, and more electrolyte permeation into electrode facilitating ionic transport (deduced via battery cycling test below). These results confirm the high electrochemical stability of AOB binder. To verify the utility of AOB binder, Si-based electrode assembled cell performance was evaluated. By measuring Si/Li battery performance, the AOB binder with a PPB/PAA weight ratio of 3:1 was selected as the optimal component ratio for further cell performance evaluations (Figure S22). Noting that compared with PAB-N and PAB-W, AOB binder shows the improved battery cycle performance attributable to its enhanced mechanical and adhesion abilities; this clearly proves hierarchical structure design significance of AOB binder.
In coin-type half-cells, AOB binder renders much superior electrochemical performance to those of the PAA binder (Fig. 5a-d). Specifically, the AOB binder based nano-Si electrodes can retain a delithiation capacity of 1617 mAh/g in half-cells after 1000 cycles at 0.4 C under 0.005 − 1.5 V (Fig. 5a). The average Coulombic efficiency (CEs) of the AOB and PAA binders in this half-cell are 99.8% and 98.9%, respectively. Higher CEs represent fewer electrolyte decomposition side reactions in the presence of AOB binder, mainly ascribed to that its superior mechanical and adhesive properties can maintain electrode interface stability. Additionally, the AOB binder-based nano-Si electrode shows excellent rate performance (Fig. 5b), achieving a delithiation capacity of 1350 mAh/g at 2 C, evidently exceeding the PAA based one (980 mAh/g). Application superiority of AOB binder was further evidenced by stable cycling of commercial-level S600 electrode. As shown in Fig. 5c, S600 electrodes based on AOB binder deliver an initial delithiation capacity of 650 mAh/g at 0.1 C and of 539 mAh/g at 0.5 C, and then stabilize at 464 mAh/g in half-cells after 300 cycles at 0.5 C, much superior to that (230 mAh/g after 300 cycles at 0.5 C) of the PAA counterpart. At varied C-rates from 0.1 C to 2 C, AOB binder based S600 electrodes also show apparently enhanced delithiation capacity when compared with PAA (Fig. 5d). Moreover, typical galvanostatic voltage profiles of Si and S600 electrodes in half-cells present that AOB binder endows slighter battery polarization than PAA binder upon long cycling (Figure S23 and S24); this can be supported by lower impedance evolution (Figure S25 and Table S1), which is highly correlated with stable SEI enabled by AOB binder.
More impressively, in commercial-level LiNi0.8Co0.1Mn0.1O2 (NCM811) full cells, the as-developed binder enables excellent cyclabilities (Fig. 5e-j). The AOB binder based NCM811/S600 coin-type full cell (a NCM811 mass loading of 17.6 mg/cm2) displays an areal capacity of 2.85 mAh/cm2 at the initial 3 cycles of 0.1 C followed by stable cycling at 0.2 C in the voltage range of 3.0-4.2 V (Fig. 5e and S26). After 100 cycles at 0.2 C, the full-cell can still maintain an areal capacity of 1.48 mAh/cm2. Then, by extending the voltage range from 2.7 to 4.25 V, the battery can still stably cycle in following cycling. As for NCM811/nano-Si coin-type full cells (a NCM811 mass loading of 18.5 mg/cm2), a capacity retention of 80% after 200 cycles is achieved (Fig. 5f and S27). In order to further prove the practicability of AOB binder, the 3.3Ah soft package cell (Fig. 5g) was prepared by pairing the AOB binder based S600 anode with a commercial-level NCM811 cathode (see detailed information in Table S2). The full cells were activated for 3 cycles at 0.1 C followed by charging and discharging at 1C (1C = 600 mA/g) between 2.8 and 4.25 V. When cycling at 1 C, the NCM811/S600 full cell delivers an initial capacity of 3.22 Ah and still preserves a capacity of 2.92 Ah after 700 cycles (capacity loss: 0.013%/cycle, Fig. 5g and S28). Undoubtedly, electrochemical performance evaluations fully confirm that AOB binder is very valuable in practical implementation of Si-based batteries.
Furthermore, we compared the Si-based electrode cycle performance of previously reported typical binders with AOB binder in half-cells and soft package full cells (Fig. 5h-i and Table S3). Obviously, AOB binder has the leading cycling performance of Si-based electrodes, to the best of our knowledge. These results clearly verify the hierarchical structure design rationality of AOB binder.
In summary, we develop a spidroin-inspired hierarchical structure binder to unlock the stiff challenges faced by Si-based electrode. In the structure of AOB binder, hydrophobic PPB polymer condensates with a minor part of PAA in the mixture solution to form crystalline regions within submicron-sized irregular spherical domains, functioning as the rigid node of binder mimicking β-sheet of spidroin; while amorphous PAA imitates the α-helix structure of spidroin, and builds an ionic bonding network structure with PPB, similar to the interactions between NTD and CTD in the primary structure of spidroin. Benefited by such bionics design, the as-developed AOB binder has the two following competitive edges: (1) Superior mechanical and adhesive capabilities were achieved via the hierarchical structure design including the dynamically reversible ionic bonding network. These factors help achieve energy dissipation and thus can accommodate enormous electrode volume deformation; (2) Low-lying LUMO leads to preferentially reduction, which helps form a polymer-reinforced SEI layer. As a result, AOB binder endows Si and S600 electrodes with superior electrochemical performance to traditional PAA binder, showing good potential for practical implementation of Si-based electrodes that meet the requirement of commercial-level high energy density LIBs. This work marks a milestone in achieving advanced silicon-based electrodes towards high energy lithium battery applications.