Jie
Ren
a,
Xiang
Wang
a,
Jihao
Li
b,
Qianzi
Sun
a,
Shaozhou
Li
a,
Ling
Bai
a,
Xianming
Liu
c,
Guilong
Liu
c,
Ziquan
Li
*a,
Haijiao
Zhang
*b and
Zhen-Dong
Huang
*a
aState Key Laboratory for Organic Electronics and Information Displays & Jiangsu Key Laboratory for Biosensors, Institute of Advanced Materials, Jiangsu National Synergetic Innovation Center for Advanced Materials, Nanjing, 210023, P. R. China. E-mail: iamzdhuang@njupt.edu.cn; lizq@njupt.edu.cn
bInstitute of Nanochemistry and Nanobiology, School of Environmental and Chemical Engineering, Shanghai University, Shanghai, 200444, P. R. China. E-mail: hjzhang128@shu.edu.cn
cCollege of Chemistry and Chemical Engineering, Luoyang Normal University, Luoyang, Henan 471934, P. R. China
First published on 31st January 2025
Hard carbon and antimony (Sb) are two promising anode candidates for future potassium-ion batteries. Herein, we successfully solve the low-capacity problem of highly conductive carbon and poor cycling stability of high-capacity Sb through uniformly dispersing and embedding sub-nano and nanoscale Sb particles (∼36.4 wt%) inside nitrogen-doped two-dimensional hard carbon nanosheets to form a multi-scale carbon@Sb mesoporous composite, denoted as Sb3@HCNS. The electrochemical results show that the optimized Sb3@HCNS anode exhibits an exceptional potassium-ion storage performance, delivering a reversible capacity of 580.8, 413.0, and 215.5 mA h g−1 at the current density of 0.1, 1, and 4 A g−1, respectively. Furthermore, it still maintains a high capacity of 382 mA h g−1 at a high current density of 2 A g−1 after 1000 cycles. The characterization results further manifest that the in situ localized electrochemical pulverization activation of Sb during the (de)alloying process and the pseudo-capacitive effect of good electronic conductive hard carbon nanosheets are mainly responsible for the exceptional properties of Sb3@HCNS. Together with its controllable preparation strategy, the newly-developed Sb3@HCNS composite is expected to be a promising anode material for high-performance potassium-ion batteries.
New conceptsHerein, we successfully solve the low-capacity problem of highly conductive carbon and poor cyclic stability of high-capacity Sb by developing a multi-scale carbon@Sb mesoporous composite (Sb@HCNS). The sub-nano and nanoscale Sb particles (∼36.4 wt%) are dispersed and embedded inside N-doped 2D hard carbon nanosheets (HCNS). The in situ electrochemical pulverization activation of Sb nanoparticles localized within the good electronic conductive HCNS endow Sb@HCNS with exceptional properties, including an exceptional potassium-ion storage performance. In particular, it can maintain a high capacity of 382 mA h g−1 at a high current density of 2 A g−1 after 1000 cycles. |
Among the widely studied anode materials for PIBs, hard carbon has been considered one of the most promising anode materials. Unlike graphite with small interlayer space and limited potassium storage capacity,21,22 hard carbon has a larger interlayer space than 0.34 nm and a relatively disordered structure, which can effectively suppress the volume change caused by the insertion or reaction of K+ during the charge/discharge process.23,24 Moreover, the theoretical capacity of the graphitic component of hard carbon to form K8C generally doesn’t exceed 300 mA h g−1.25 This low-capacity anode will significantly hamper the development and practical applications of PIBs.26,27
Compared to carbonaceous anode materials, alloying-type antimony (Sb) has a high theoretical capacity (660 mA h g−1) to form K3Sb with potassium. However, the abnormal volume expansion makes the Sb electrode powdery, resulting in poor cycling performance.28,29 Alloy-based nanomaterials and composites with two-dimensional (2D) conductive substrates, along with nanotechnology engineering, are widely used strategies to suppress excessive volume changes for accommodating the volume changes and improving the reaction kinetics of Sb-based anode materials.30–35 As a result, various carbon/Sb composites have been developed in recent years. For example, a nitrogen-doped porous carbon framework encapsulating antimony nanoparticles (Sb@NPC) was developed using polyvinyl pyrrolidone (PVP) as the carbon source.23 Unfortunately, the resulting electrode materials suffered from a significant capacity fade after 500 cycles.36 Highly conductive reduced graphene oxide (rGO) and MXene have also been applied to help Sb nanoparticles maintain a high capacity-retention rate and rate capability.37 Additionally, tri-mesic acid has been used as the carbon source to prepare N-doped carbon to alleviate the significant volume changes of antimony.38 To effectively enhance the utility and activity of Sb, atomic level or ultra-small Sb has been synthesized and impregnated into a nanostructured carbon matrix.39–42 Although much progress has been made, how to develop high-performance PIB anode materials remains a high-concern issue. The low-capacity and low-density problems of carbon and the poor cycling stability problem caused by the vast volume expansion of Sb particles still need to be solved more effectively.
Herein, we successfully solve the low-capacity problem of highly conductive carbon and poor cycling stability of high-capacity Sb through uniformly dispersing and embedding the sub-nano and nanoscale Sb particles (∼36.4 wt%) inside nitrogen-doped 2D hard carbon nanosheets to form a multi-scale carbon@Sb mesoporous composite, marked as Sb3@HCNS. Benefiting from the multiscale nanostructure and the activation by the in situ localized electrochemical pulverization of Sb nanoparticles, the as-prepared Sb3@HCNS composite anode shows an exceptional potassium-ion storage performance, delivering a high reversible capacity of 580.8, 413.0, and 215.5 mA h g−1 at the current density of 0.1, 1, and 4 A g−1, respectively. Even being cycled at 2.0 A g−1 for 1000 cycles, Sb3@HCNS still can maintain a high capacity of 361 mA h g−1 with a capacity retention ratio of 94%, which is superior to the reported counterparts. Furthermore, the potassium-ion storage mechanism of Sb@HCNSs is also revealed via different characterization techniques.
After sintering under a N2 atmosphere at 800 °C for 2 h in a tube furnace, the DCDA/PACP-derived products are hard carbon nanosheets (HCNS), as shown in Fig. 1b and c. Interestingly, the thickness and in-plane size of the obtained HCNSs decrease with more SbCl3, as observed from the SEM images (Fig. 2). Moreover, the multiscale Sb are uniformly dispersed and encapsulated within the crumpled HCNS, as shown in Fig. 1d, e and Fig. S2 (ESI†). The sizes of the Sb nanoparticles are around 40 to 70 nm. Sub-nano Sb particles or clusters can also be observed from the high-resolution element map in Fig. S2 (ESI†). The HRTEM analysis results of the inset of Fig. 1e indicate that the lattice fringes have a lamellar thickness of 3.15 Å, which can be assigned to the interlayer space of Sb's crystal plane (012). The analysis results of the HRTEM image in Fig. 1f indicate that the hard carbon nanosheets of as-prepared Sb3@HCNS have a thickness of 6.3 nm. The long-range disordered graphite domains of Sb3@HCNS have an interlayer distance of ∼0.44 nm. The interlayer spacing of this material is greater than the interlayer spacing of the (002) plane of the reported hard carbon (0.37–0.4 nm).43
![]() | ||
Fig. 2 SEM images in different magnifications: (a)–(c) Sb2@HCNS, (d)–(f) Sb3@HCNS, and (g)–(i) Sb4@HCNS. |
Fig. 1g–m show the high-angle annular dark field (HAADF) image and the elemental distribution maps corresponding to C, N, O, and Sb within the as-prepared Sb3@HCNS. The results shown in Fig. 1i and m further confirm the existence of metal Sb nanoparticles, as observed in Fig. 1d. Moreover, C, N, and O elements are uniformly distributed in Sb3@HCNS, as displayed in Fig. 1j–l.
Fig. 3a and b present the X-ray diffraction (XRD) patterns and Raman spectra of the Sb2@HCNS, Sb3@HCNS, and Sb4@HCNS composites, respectively. It is found that both the intensity of the XRD peaks indexed to metallic Sb and the HCNS component's graphitic degree increased with Sb content. As exhibited in Fig. 3b, the peak intensity ratios between the D and G band (ID/IG) of the Sb2@HCNS, Sb3@HCNS, and Sb4@HCNS composites decreased from 2.44 to 2.15 with the content of Sb. The results suggest that Sb3@HCNS has a relatively better balance on the electric conductivity and active defect sites, which is suitable for potassium-ion storage. N2 adsorption/desorption isotherms and the corresponding pore size distribution shown in Fig. 3c manifest that Sb3@HCNS has a mesoporous structure with a pore size centered at around 3.7 nm and a specific surface area of 71.6 m2 g−1.
Fig. S3a (ESI†) provides the XPS full spectrum of Sb3@HCNS. The results confirm the observation of element mapping shown in Fig. 1. The compositions listed in Tables S1 and S2 (ESI†) obtained by element mapping and XPS measurements are similar. The contents of dopants of N and O listed in Table S2 (ESI†) are higher than those shown in Table S1 (ESI†). The different scales of surface area should cause the differences in content. Due to most of the Sb nanoparticles and clusters being encapsulated within HCNS, the content of Sb observed by element mapping and XPS is much smaller than 12.0 wt%, 36.4 wt%, and 62.2 wt% of Sb2@HCNS, Sb3@HCNS, and Sb4@HCNS that was measured using the thermogravimetry (TG) method, seen in Fig. S4 (ESI†). The core XPS spectra corresponding to Sb 3d + O 1s, C 1s, and N 1s of Sb3@HCNS are shown in Fig. 3d, e and f, respectively. The Sb 3d + O 1s of Sb3@HCNS shown in Fig. 3d can be deconvoluted into two prominent peaks at 539.9 (Sb0 3d3/2) and 530.6(Sb0 3d5/2), and two corresponding satellite peaks at 537.4 and 528.2 eV, with one O 1s peak at 532.5 eV, respectively.37 The C 1s of Sb3@HCNS shown in Fig. 3e can be deconvoluted into three main peaks at 284.5, 285.3, and 287.5 eV, corresponding to C–C/CC, C–N/C–O, and O
C–O, respectively.44,45 The N 1s spectrum in Fig. 3f confirms the presence of pyridinic N (398.2 eV), pyrrolic N (399.6 eV), pyridinic N–H (400.7 eV), graphitic N (401.5 eV), and oxidized N (403.5 eV).45–47 As summarized in Fig. S3b (ESI†), the content of electrochemical active pyridinic N (36.8%), pyrrolic N (14.7%), pyridinic N–H (28.1%), and oxidized N (112.5%) is more than 90% that of doped N in Sb3@HCNS.
Fig. 4a shows the rate capability of potassium-ion half batteries using the as-prepared Sb2@HCNS, Sb3@HCNS, and Sb4@HCNS samples as anodes. As expected, the Sb3@HCNS composite shows the highest capacity and the best rate capability. At the same time, Sb4@HCNS has a slightly high initial capacity at 0.1 A g−1 but can’t work at the current density of 4 A g−1 due to the relatively higher resistance and polarization. Interestingly, a capacity enhancement can be clearly observed from the second testing cycle of the rate performance of both Sb2@HCNS and Sb3@HCNS. As demonstrated in Fig. S5 (ESI†), the plateau capacity below 0.6 V rises from 41 to 155 mA h g−1 from the 2nd cycle to the 68th cycle at 0.1 A g−1 due to the gradual activation of Sb nanoparticles. Meanwhile, the polarization of the Sb3@HCNS electrode is reduced by 0.18 V. During the second testing cycle on rate performance, Sb3@HCNS delivers a specific capacity of 580.8, 540.6, 477.5, 413.0, 325.0, and 215.5 mA h g−1 at the current density of 0.1, 0.2, 0.5, 1, 2, and 4 A g−1, respectively. The corresponding discharge/charge profiles of Sb3@HCNS at the 38th, 43rd, 48th, 53rd, 58th, and 63rd cycles are shown in Fig. 4b. Fig. S6 (ESI†) also provides the (dis)charge profiles at different rates of all Sb@HCNSs. As listed in Fig. 4c and Table S3 (ESI†), the potassium-ion storage properties achieved by Sb3@HCNS are superior to most reported C composites with Sb,48–60 even the atomically dispersed Sb.37,39,41,48
![]() | ||
Fig. 4 Electrochemical performance of Sb@HCNS as anodes for PIBs: (a) rate performances of all Sb@HCNS, (b) the typical charge/discharge profiles of Sb3@HCNS at various current densities, (c) potassium-ion storage performances of the as-prepared Sb@HCNS compared with other reported Sb/C-based counterparts listed in Table S3 (ESI†). (d) Long-term cycling performance of Sb@HCNS at a constant current density of 2 A g−1. (e) dQ/dV profiles of Sb3@HCNS at 2 A g−1. (f) Schematic illustrating the K-ion storage mechanism of Sb3@HCNS. |
Fig. 4d exhibits their long-term cycling performances as anode materials for PIBs. Ten discharge/charge cycles at 0.1 A g−1 are applied to pre-activate Sb2@HCNS, Sb3@HCNS, and Sb4@HCNS electrodes. The corresponding initial three cycles’ discharge/charge profiles at 0.1 A g−1 are provided in Fig. S7 (ESI†). The initial coulombic efficiencies of the Sb2@HCNS, Sb3@HCNS, and Sb4@HCNS anode materials are 52.8%, 58.7%, and 61.8%, respectively, rising with the increment of Sb content. However, introducing too much Sb into HCNS can’t continuously enhance the potassium ion storage properties of HCNS. As shown in Fig. 4d, although Sb4@HCNS can deliver a higher capacity than Sb2@HCNS and Sb3@HCNS, the cycling stability of Sb4@HCNS is not as good as that of Sb3@HCNS. Because Sb4@HCNS has much higher Sb content than Sb2@HCNS and Sb3@HCNS, the apparent fluctuation of potassium-ion storage performance observed during the long-term cycling performance measurements may be caused by the stress-concentration-induced deterioration of the Sb4@HCNS electrode or the pulse change in the test room temperature. After cycling at 2 A g−1 for 1000 cycles, the Sb3@HCNS electrode still maintains 382 mA h g−1 with a capacity loss of 0.05% per cycle. Moreover, a slight capacity enhancement can also be observed from the cyclic testing results of three Sb@HCNS electrodes. As demonstrated in Fig. S8 (ESI†), the charge plateau below 1.0 V gradually appears. Meanwhile, the plateau capacities of the Sb@HCNS electrodes increase obviously, and the polarizations of the Sb@HCNS electrodes are reduced, extending the discharge/charge cycle to about 400 cycles. According to the corresponding dQ/dV curves of Sb@HCNS at the 30th, 50th, and 200th cycles provided in Fig. 4e and Fig. S9 (ESI†), the redox reaction peaks are broad and weak, indicating that the primary potassium-ion storage mechanism is dominated by surficial redox and absorption/desorption behaviors during the first 200 cycles. It should be pointed out that the (de-)alloying reaction at the surface layer of Sb nanoparticles is enhanced, as the oxidation peak at 0.6 V can be observed clearly, although it is still broad and weak. With the extension of the discharge/charge cycle, the activation degree of Sb nanoparticles also gradually deepens. As schematically illustrated in Fig. 4f, during the continuous discharge/charge process, due to the volume expansion and contraction caused by (de)potassiation reactions during continuous rapid charge and discharge, the surface layer of the Sb nanoparticles is continuously pulverized and peeled off, forming a large amount of finer antimony nanoparticles, while exposing more new surfaces. As a result, the utilization rate of Sb active materials increases continuously. Therefore, the peak currents of the oxidation peak centered at 0.6 and reduction peak centered at 0.15 V increase obviously from the 200th to 400th cycle. Meanwhile, the redox peak gap decreases. After the complete activation, the charge/discharge capacities of three Sb@HCNS electrodes become stable until 1000 cycles, which further confirms that the optimized Sb3@HCNS electrode has excellent stability and rate capability.
Fig. 5a and b display the HRTEM image and the corresponding selected area electron diffraction (SAED) pattern of the Sb3@HCNS electrode being discharged to 0.01 V at the current density of 0.1 A g−1. The SAED pattern in Fig. 5b indicates the formation of K3Sb during the discharge process. And the sizes of the obtained K3Sb nanoparticles are smaller than 10 nm, and the pristine Sb nanoparticles (40–70 nm), as seen in Fig. 5a. Furthermore, after 1000 cycles, the particle size of the as-obtained KSb partially recovered during the charging process to 3 V is also much smaller than that of the pristine Sb particles, as manifested in Fig. 5c and d. Fig. S10 (ESI†) presents the SEM images of the pristine Sb3@HCNS electrode and Sb3@HCNS electrode being cycled for 1000 cycles at 2 A g−1. The porous structure remains relatively intact. The HCNSs can still be observed clearly, only covered by a solid electrolyte interphase layer. These observations further verify the enhanced mechanism of potassium-ion storage performance proposed in Fig. 4f for as-prepared Sb@HCNS anode materials. Namely, during the successive discharge/charge processes, the surface layer of carbon-wrapped Sb nanoparticles is gradually pulverized, resulting from the volume change caused by the potassiation/depotassiation, until the whole Sb nanoparticles can be reversibly potassiated/depotassiated at the high current density of 2 A g−1. Since the newly formed ultrafine Sb nanoparticles are still encapsulated within the HCNS, they are still highly active for potassium-ion storage. Taking advantage of the in situ localized electrochemical pulverization activation discussed above and the creative concept of multi-scale carbon@Sb mesoporous composites, the newly developed Sb3@HCNS delivers a higher capacity than single-Sb-atom-enforced carbon materials,37,39,41,48 and achieves a much better cycling stability and rate capability than common Sb nanoparticle-modified carbon materials.49–60
Fig. 5e provides the ex situ Raman spectra of the Sb3@HCNS electrode in different discharge/charge states. The corresponding evolution of the intensity ratio between the D and G band (ID/IG) of carbon is present in Fig. 5f. During the discharge process, the value of ID/IG decreases from 2.32 of pristine to 2.12 of 0.8 V, and to 2.05 of 0.01 V. Inversely, the value of ID/IG increases from 2.05 of 0.01 V to 2.1 of 0.8 V, and to 2.3 of 3 V, during the recharge process. This observation indicates that the redox reaction of the active defects within the N-doped mesoporous HCNS are one of the important potassium ion storage mechanisms, especially within the voltage window of 0.8 to 3 V.
To elucidate the potassium-ion storage kinetics of the newly developed Sb@HCNS composites, cyclic voltammetry (CV) measurements were conducted at scan rates ranging from 0.1 to 1 mV s−1. The results are shown in Fig. 6a and Fig. S11a and S12a (ESI†). It can be seen from the CV curves that there is no significant alloying reaction during the reduction and oxidation processes because the Sb3@HCNS required a relatively long-term activation process to activate the potassium-ion storage activity of Sb. As shown in Fig. 6b and Fig. S11b and S12b (ESI†), the obtained b values corresponding to all three redox peaks of as-developed Sb2@HCNS, Sb3@HCNS, and Sb4@HCNS are much larger than 0.6, close to 1. The results imply that the surface pseudo-capacitive effect is prominent in electrochemical potassium-ion storage.61 As summarized in Fig. 6c and Fig. S11c and S12c (ESI†), at high current sweep rates, the charge–discharge capacities of Sb2@HCNS, Sb3@HCNS, and Sb4@HCNS are mainly from pseudo-capacitance. This observation is consistent with the analysis results of dQ/dV profiles shown in Fig. 4e and Fig. S5 (ESI†). The diffusion kinetics of K+ within the Sb2@HCNS, Sb3@HCNS, and Sb4@HCNS electrodes were further studied using the galvanostatic intermittent titration technique (GITT). The diffusion coefficient was calculated based on Fick’s second law with the following equation:62
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4nh00621f |
This journal is © The Royal Society of Chemistry 2025 |