1 Introduction

Lithium-ion batteries (LIBs) are widely used in various portable electronics, electric tools, and electric vehicles due to their high energy density, long cycle life, and environmental friendly [1, 2]. Nevertheless, the conventional graphite anode of LIBs with a specific capacity of 372 mAh g−1 can hardly meet the rapidly increasing demand of high energy density. Great efforts have been made to develop promising anode materials with high capacity, such as transition metal oxides [3, 4], alloys [5, 6], metal oxides/sulfates [7,8,9], and phosphorous [10,11,12]. Among various metal oxides, SnO2 has attracted a lot of attention due to its relatively high theoretical capacity (790 mAh g−1, which is twice than the currently used graphite), low average working potential (~ 0.6 V vs. Li+/Li), natural abundance and low price [13,14,15,16]. Unfortunately, its practical application as anode material in LIBs is seriously limited by the poor cycle stability resulting from the severe volume change (> 300 vol%) during the charge/discharge process. Meanwhile, SnO2 also suffers from low electrical conductivity, resulting in poor rate capability. To improve the electrochemical performances of SnO2, several strategies have been proposed to overcome these issues. One of the effective way is to fabricate nanostructured SnO2, such as SnO2 quantum dots (QDs) [17], SnO2 hollow spheres [18], or SnO2 nanowires [19], which could restrain the structure changes during lithium alloying and shorten the ion diffusion lengths [20]. The other way is to combine SnO2 with various carbon materials possessing high conductivity such as carbon nanotubes, carbon fiber or graphene [21,22,23,24,25,26]. The conductive carbon can not only improve the overall electrical conductivity of the composites but also act as a buffer to slow down the structure collapse of the electrode.

MXenes are a newly emerging family of 2D materials with a general formula of Mn+1XnTx, where M represents early transition metal (M = Ti, Sr, V, Cr, Ta, Nb, Zr, Mo, and Hf), X is carbon and/or nitrogen, and T stands for the surface termination groups (–OH, –F=O) [27,28,29,30,31]. Due to their excellent electrical conductivity, tailorable surface chemistries, and mechanical properties, MXenes have recently attracted great interests in energy storage devices such as LIBs, sodium-ion batteries, and supercapacitors [32,33,34,35]. Particularly, as one of the most widely studied MXenes, Ti3C2Tx is a potential anode material for LIBs, which can deliver excellent cycle stability and superior rate performances. However, the theoretical reversible capacity of Ti3C2Tx is only 320 mAh g−1 [36]. On the other hand, MXene can also be used as a substrate to fabricate hybrids with other active materials, including metal oxides, which have higher theoretical capacity but poor conductivity and are prone to large volume change during charge/discharge. In the hybrids, MXene can improve the conductivity and buffer the volume changes of the metal oxide, while the metal oxide provides high capacity. Thus, the MXene/metal oxide hybrid could achieve high capacity, stable cycle and good rate performance [37,38,39,40,41]. For example, Ti3C2Tx/Fe2O3 nanocomposite was prepared by confining Fe2O3 nanoparticles into Ti3C2Tx nanosheets through ball-milling method, which showed reversible capacities of ~ 203 mAh g−1 at 1 C and 100 mAh g−1 at 10 C [37]. MXene/NiCo2O4 composite with a capacity reaching up to 1330 mAh g−1 was synthesized by spray coating method [38]. Some recent works indicate that MXene is also a good substrate to improve the lithium storage performance of SnO2. The hybrids of SnO2 with MXene prepared by wet hydrothermal approach has shown a high capacity of 1021 mAh g−1 and improved cycle stability [40]. However, a few reports indicate that MXene is unavoidable to be oxidized to titanium dioxide (TiO2) as a by-product during these processes, thereby, influencing the electrochemical performance of the composites. Recently, our group proposed a general route to self-assemble transition metal oxide nanostructures on Ti3C2Tx MXene nanosheets through van der Waals interaction. The proposed method allowed fabrication of hybrids without getting MXene to be oxidized and achieved enhanced cycle and rate performance [15].

In this work, a simple method is proposed to prepare 0D–2D SnO2 QDs/MXene hybrids by electrostatic self-assembling SnO2 QDs on the surface of 2D Ti3C2Tx MXene nanosheets under ultrasonication. Since, the synthesis is accomplished in a very mild condition, i.e., ultrasonication treatment of the mixture of negatively charged MXene and positively charged SnO2 QDs at room temperature, the MXene oxidation is effectively avoided. In the hybrids, 2D nanosheet of MXene acts as a substrate to support 0D SnO2 QDs, which can not only provide large electrode/electrolyte interface area for fast reversible transport of electrons and ions, but could also inhibit the aggregation of SnO2 QDs and buffer the volume changes during charge/discharge process. The SnO2 QDs with ultra-small particle size can effectively maximize the activity and specific capacity and minimize the volume change while inhibiting the structure collapse during charge/discharge process and shortening the lithium diffusion pathways. In addition, the SnO2 QDs acts as a “spacer” to prevent the MXene nanosheets from restacking and thus protecting the Li+ migration channels and active sites. These unique features endow the 0D–2D SnO2 QDs/MXene hybrids with high lithium storage capacity, excellent cycle stability and superior rate performance, indicating a promising anode for LIBs.

2 Experimental

2.1 Synthesis of Ti3C2Tx MXene

The Ti3C2Tx MXene was synthesized by etching Ti3AlC2 (400 mesh, purchased from Yiyi Technology Co., Ltd.) with LiF + HCl solution as reported previously [41]. Typically, 1.0 g Ti3AlC2 powder was subjected to a mixture containing LiF (1.0 g) and hydrochloric acid (12 M, 10 mL) under stirring conditions for 24 h at 35 °C. After several times of centrifugation-washing with deionized (DI) water, the product was then dispersed in 50 mL DI water, stored under ultrasound for 30 min. The dark green supernatant was collected by centrifugation at 3500 rpm for 1 h. Finally, the black MXene liquid was obtained and sealed for future use.

2.2 Synthesis of SnO2 QDs

In order to obtain SnO2 QDs, 4 mmol SnCl2·2H2O and 4 mmol thiourea (CH4N2S) were added to 30 mL DI water and magnetically stirred at room temperature to form a milky suspension. After stirring for 24 h, a clear yellow aqueous solution containing SnO2 QDs was obtained.

2.3 Synthesis of 0D–2D SnO2 QDs/MXene Hybrids

In a typical synthesis, 5 mL of as-prepared SnO2 QDs solution was added to 40 mg 0.1 mg mL−1 Ti3C2Tx MXene solution under ultrasonication for 6 h. The SnO2 QDs were deposited on the Ti3C2Tx layers during this process and solid–liquid separation was observed. Finally, the black precipitate of SnO2 QDs/MXene hybrids were collected by vacuum filtration, washed with water and dried in a vacuum oven at 80 °C for 6 h. The above hybrid was denoted as SnO2 QDs/MXene-52. For comparison, we also prepared another SnO2 QDs/Ti3C2Tx hybrids (SnO2 QDs/MXene-51) under the same conditions, but the addition amount of MXene in this solution was 20 mg.

2.4 Materials Characterization

Scanning electron microscopy (SEM) characterization was conducted using a Gemini SEM 500. Transmission electron microscopy (TEM) characterization was conducted on a JEOL JEM-F200 (HR) equipped with selected area electron diffraction (SAED). X-ray diffraction (XRD) patterns were recorded on a Bruker D8 ADVANCE with monochromatic Cu Kα radiation (λ = 1.54060 Å). Raman spectra were obtained with a LabRAM HR Evolution Raman spectrometer (633 nm). X-ray photoelectron spectroscopy (XPS) analysis was performed using a Thermo Fisher ESCALAB Xi+ to analyze the chemical compositions of the samples.

2.5 Electrochemical Measurements

The working electrodes were prepared by mixing 80 wt% active material, 10 wt% Super P and 10 wt% carboxymethylated cellulose (CMC) in DI water. After coating the slurry on the copper foil, the electrodes were dried at 60 °C in vacuum oven for 8 h to remove the solvent. To test electrochemical performances, CR2025-type coin cells were assembled in an argon-filled glove box (Mikrouna, H2O, O2 < 0.1 ppm) using Li foil as half-cell counter electrode, and microporous membrane (Celgard 2400) as separator. The electrolyte was 1 M LiPF6 in a mixture of ethylene carbonate (EC)/diethyl carbonate (DEC)/dimethyl carbonate (DMC) with a volume ratio of 1:1:1. The charge/discharge (GCD) tests were performed using a LAND BT2000 battery tester. The potential window was 0.01–2.5 V versus Li/Li+. The CV traces were recorded on a CS350 electrochemical workstation from 0.01 to 2.5 V. Electrochemical impedance spectroscopy (EIS) measurements were performed on the VSP Bio-Logic SAS at frequencies ranging from 10 mHz to 100 kHz with an applied AC signal amplitude of 10 mV. The capacities of the samples were calculated based on all the components in the active materials, i.e., the mass of SnO2 QDs/Ti3C2Tx hybrids.

3 Results and Discussion

The synthesis route for 0D–2D SnO2 QDs/MXene by electrostatic self-assembly is illustrated in Fig. 1. Firstly, transparent yellow aqueous solution of positively charged SnO2 QDs is prepared by hydrolysis, dehydration and oxidation of SnCl2·2H2O in DI water, where thiourea is added as a promoter and stabilizer (Fig. S1a). The mercaptan group can easily bound to Sn2+, and the SnO2 QDs whose zeta potential is +99.0 mV (Fig. S1b) is surrounded by a positively charged protic amino group (–NH3+), which makes it highly stable and electropositive [42]. The Ti3C2Tx MXene solution with its negatively charged groups (–F, –OH) [43], and zeta potential of − 36.2 mV, is very stable in water (Fig. S1a) due to the hydrophilicity and electrostatic repulsion between neighboring nanosheets. When positively charged SnO2 QDs are added into the negatively charged Ti3C2Tx colloidal solution, the SnO2 QDs can easily load onto the surface of the MXene nanosheets. After the continuous ultrasonication of 6 h, the supernatant becomes clear and colorless, and black precipitates are obtained. The positively charged SnO2 QDs are captured by the negatively charged MXene nanosheets only by ultrasonic treatment without other components and additional treatments, implying an electrostatic self-assembly mechanism for the designed 0D–2D SnO2 QDs/MXene hybrids. The Ti3C2Tx MXene with 2D layered structure as a conductive matrix facilitates the charge transfer and accommodates the volume change of SnO2 QDs; the SnO2 QDs thus prevents the MXene nanosheets from restacking by working as “spacer” and providing channels for fast transfer of Li ion with promising electrochemical performance.

Fig. 1
figure 1

Schematic illustration for the preparation of 0D–2D SnO2 QDs/MXene hybrids

SEM and TEM were employed to characterize the morphology and structure of the as-prepared samples. The pure Ti3C2Tx (Fig. S1c) displays a compact 2D layered structure. After decoration with SnO2 QDs, the SnO2 QDs/MXene hybrids exhibit 0D–2D structure, as shown in Fig. 2a. From a closer view in Fig. 2b, it can be seen that the spatially dispersed SnO2 QDs are evenly distributed over the surface of the MXene nanosheets. The particle-size distribution analysis (Fig. S1d) indicates the SnO2 QDs have an average particle size of ~ 4.5 nm, favoring the electrode/electrolyte interactions. Since the diffusion of Li ions is strongly dependent on the transport length and the active sites, the ultra-small particle size of SnO2 QDs can not only expose a large number of electrochemical active sites shuttling Li ions in and out, but also shorten the diffusion path of Li ion transport, which are beneficial for improving the capacity and rate performance. Figure 2c shows clear lattice fringes of the SnO2 QDs, indicating a high degree of crystallinity. The crystal lattice with a spacing of 0.33 nm is consistent with the d spacing of (110) planes of the SnO2 tetragonal phase. The selected area electron diffraction (SAED) pattern (Fig. 2d) shows clear diffraction rings, implying the polycrystalline nature of SnO2. These diffraction rings can be well indexed to the (110), (101), and (211) planes of SnO2. Moreover, the typical scanning TEM (STEM) and elemental mapping images of SnO2 QDs/MXene hybrid (Fig. 2e) show homogeneously distributed Ti, C, Sn, and O elements, demonstrating the uniform loading of SnO2 QDs on the surface of MXene sheets.

Fig. 2
figure 2

a SEM image, b TEM image, c HRTEM image of 0D–2D SnO2 QDs/MXene hybrid. d SAED patterns of SnO2 QDs. e STEM image and corresponding elemental mapping images of Ti, C, Sn, and O

The phase structures of the as-prepared SnO2 QDs/MXene hybrids were verified by XRD as shown in Fig. 3a. Ti3C2Tx exhibits the major peaks, such as the (002), (006), (008), (0010), and (0012) [44]. Typically, the (002) plane of the Ti3C2Tx MXene is located at 2θ = 6.4°, corresponding to an interplanar spacing of 13.7 Å [45]. In addition, the pure tetragonal phase of crystalline SnO2 (JCPDS No. 41-1445) is located at 26.5°, 34.1°, and 52.4°, corresponding to (110), (101), and (211). The XRD patterns of the SnO2 QDs/MXene hybrids consist of Ti3C2Tx and tetragonal phase SnO2, indicating that SnO2 QDs/MXene hybrids have successfully formed. No any extra crystalline phase peaks such as TiO2 are observed, indicating the electrostatic self-assembly process is mild with no oxidation of MXene. In addition, the (002) peak of the SnO2–MXene composite (Fig. S1e) was downshifted compared to pure MXene, indicating increased interlayer spacing of the MXene in the hybrids.

Fig. 3
figure 3

a XRD patterns, b Raman spectra of MXene, SnO2, and SnO2 QDs/MXene hybrid. XPS spectra of c MXene for high-resolution C 1s, d Ti 2p, e SnO2 QDs/MXene-52 for high-resolution C 1s, f Ti 2p

Raman spectra further confirm the phase characteristics of the SnO2 QDs/MXene-52 hybrid. As shown in Figs. 3b and S1f, the peaks of pure SnO2 at 271, 465, and 698 cm−1 can be assigned to the Eg and A1g active mode of SnO2, respectively [46]. Moreover, the pure MXene exhibits a peak at 199 cm−1 which corresponds to the A1g symmetry out-of-plane vibrations of Ti atoms, whereas the peaks at 381 and 622 cm−1 are related to the Eg group vibrations, including in-plane (shear) modes of Ti, C, and surface functional group atoms [47]. For the SnO2 QDs/MXene-52, the spectrum manifests six cognizable Raman-active modes, combining the characteristic Raman peaks of Ti3C2Tx and SnO2. In comparison with the pure SnO2, the peak intensity of SnO2 in the hybrid has significantly declined. This suggests that the SnO2 QDs are well separated by the Ti3C2Tx nanosheets, which weakens the Raman signal from the SnO2. It’s worth noting that no strong peak for TiO2 was detected at 144 cm−1, confirming that MXene has not oxidized during the hybrid formation.

The elemental composition of SnO2 QDs/MXene hybrid was further analyzed by XPS. For pristine Ti3C2Tx MXene, there are only four main elements, Ti, C, F, and O, whereas for SnO2 QDs/MXene-52, the Sn element is also detected, indicating the presence of SnO2 in the hybrid (Fig. S2a). In the high-resolution C 1s spectrum (Fig. 3c), the characteristic peaks correspond to C–Ti (284.2 eV), C–C (288.7 eV), C–O (288.1 eV), C=O (289.5 eV), and O–C=O (291.6 eV) [48]. The particular peak located at 282.1 eV can be assigned to the C–Ti bond in Ti3C2Tx sheets. The XPS spectrum of the Ti 2p (Fig. 3d) reveals that peaks of Ti bound to C, Ti(II), and Ti(III). The Ti–C 2p3/2, Ti(II) 2p3/2, Ti(III) 2p3/2, T–C 2p1/2, Ti(II) 2p1/2, and Ti(III) 2p1/2 peaks are detected at binding energies of 455.1, 455.6, 456.8, 458.8, 461.3, and 462.3 eV, respectively. After loading the SnO2 QDs, no Ti(IV) peak is observed, suggesting no oxidation of MXene in the SnO2 QDs/MXene-52 (Fig. 3f). Meanwhile, the XPS in Sn 3d region has two peaks at 487.2 and 495.6 eV, which are attributed to Sn 3d5/2 and Sn 3d3/2 of SnO2, confirming the formation of SnO2 (Fig. S2b). Besides this, the relative intensities of Ti–C, Ti(II), and Ti(III) peaks in SnO2 QDs/MXene-52 are relatively weaker than pure Ti3C2Tx, indicating decreasing Ti3C2Tx signal for the hybrid. Furthermore, the slight shift of the Ti–C component to lower binding energy indicates that the synthesis of SnO2 QDs/MXene hybrid by electrostatic attraction might have caused shift in the electron density (Fig. 3e). All of these observations demonstrate that SnO2 QDs have been successfully deposited on the surface of the Ti3C2Tx sheets. In addition, the introduction of SnO2 QDs on the Ti3C2Tx surface increases the accessible surface area [49], which is evident from the nitrogen adsorption measurements (Fig. S2c). The SnO2 QDs/MXene-52 shows a typical type I adsorption/desorption isotherms with a BET specific surface area of 184 m2 g−1, much higher than that of the pure MXene (19 m2 g−1). The large surface area is beneficial for accelerating electrolyte diffusion and accommodating volume change of the SnO2 QDs during charge/discharge. Thus, much improved lithium storage performances could be expected.

To evaluate the electrochemical performances of the SnO2 QDs/MXene hybrids as anode materials in LIBs, coin-type half cells were assembled using lithium as counter electrode. CV curves of Ti3C2Tx MXene and SnO2 QDs/MXene hybrids were scanned between 0.01 and 2.5 V. For the pure Ti3C2Tx MXene, the broad irreversible reduction peak at around 0.7 V is observed in the first lithiation process, which is attributed to the formation of a solid electrolyte interphase (SEI) generated from the reaction of Ti3C2Tx with Li ion [17]. In the subsequent cycles, the reversible peaks near 0.77 and 1.5 V may be the consequence of the reaction between Li+ and the titanium-based compounds (Fig. S3a) [50]. For the pure SnO2 QDs (Fig. S3b), the SnO2 is considered to transform into LixSn (x ≤ 4.4) and Li2O during the initial discharge process (lithium insertion). The cathodic peak at 0.05 V result from the alloying process of Sn to LixSn (0 ≤ x ≤ 4.4), and the strong anode peak at about 0.5 V is caused by the de-alloying process of LixSn (0 ≤ x ≤ 4.4). The anode peak at 1.2 V is caused by the partial reversible transformation of Sn to SnO2 owing to its quantum size [51, 52]. However, in the second and third cycles of the pure SnO2, the intensities of the cathodic and anodic peaks show continuous and significant decline, indicating rapid capacity degradation because of the structural variation of SnO2 during lithiation/delithiation processes (Fig. S3b). The SnO2 QDs/MXene-52 (Fig. 4a) shows CV curves with similar characteristic peaks, but the curves in the 2nd and 3rd cycles almost overlap, indicating good cycle performance. In comparison with the SnO2 QDs/MXene-51 (Fig. 4b), the SnO2 QDs/MXene-52 display better overlapping CV curves with the same peak positions, implying excellent reversibility in its conversion reaction.

Fig. 4
figure 4

Electrochemical performances as anode in LIBs. a, b CV curves of SnO2 QDs/MXene at a scan rate of 0.1 mV s−1 in 0.01–2.5 V. c, d Charge/discharge curves of SnO2 QDs/MXene at 50 mA g−1

Furthermore, the GCD profiles (Fig. 4c, d) of the SnO2 QDs/MXene at 50 mA g−1 coincides well with the CV curves. The GCD profile of the Ti3C2Tx and SnO2 QDs electrodes were also tested at the same conditions (Fig. S3c, d). In the initial discharge profile of the SnO2 QDs/MXene hybrids, the plateaus at 0.7 and 0.05 V correspond to the SEI formation and lithium alloying; while during the initial charge, the plateaus at 0.5 and 1.2 V relate to lithium de-alloying and partial reversible SnO2 formation, respectively. The reversible capacity of the SnO2 QDs/MXene-52 is 887.4 mAh g−1 in the first cycle, with initial Coulombic efficiency (CE) of about 51.2%. The low CE might have resulted from the SEI and Li2O formation as well as the electrolyte decomposition. In the 5th cycle, the capacity was determined to be about 847.6 mAh g−1 and the corresponding CE reached about 100%. With the SEI film protection, the capacity reaches a stable state. In contrast, the SnO2 QDs/MXene-51 delivers an initial reversible capacity of 897.5 mAh g−1 with a CE of 43.0% (Fig. 4d). The results indicate that the introduction of more SnO2 QDs can increase available active sites and the capacity of the hybrid material, but at the same time cause more side reactions and reduce the initial CE, due to its ultra-small particle size and high surface area. Therefore, only appropriate ratio of SnO2 QDs and MXene sheets could reach an optimum electrochemical performance.

Figure 5a shows the comparison of the charge/discharge cycle performance among Ti3C2Tx MXene, SnO2 QDs, and SnO2 QDs/MXene at 100 mA g−1. The bare Ti3C2Tx MXene has a low initial capacity of 79.2 mAh g−1 with a CE of 34.2%, which is attributed to the restacking of MXene nanosheets [36]. The initial capacity of the pure SnO2 QDs can reach 835.9 mAh g−1, but it fades very rapidly due to the severe pulverization of SnO2 during charge/discharge. After 20 cycles, the capacity remains only 144.6 mAh g−1 with a capacity retention ratio of about 17% of the initial capacity. In contrast, by electrostatic self-assembly of the SnO2 QDs on the 2D Ti3C2Tx MXene nanosheets, the SnO2 QDs/MXene hybrids exhibit much enhanced cycle stability. The capacity of SnO2 QDs/MXene-52 reaches to 659.8 mAh g−1 with 91% retention of the initial capacity after 100 cycles, which is much higher than that of the pure SnO2 QDs and Ti3C2Tx MXene. In a comparative evaluation of SnO2 QDs/MXene hybrids with different SnO2/MXene ratio, it is found that the SnO2 QDs/MXene-52 with abundant MXene shows better cycle stability than the SnO2 QDs/MXene-51, indicating the importance of adequate MXene substrate in achieving optimum performance [53].

Fig. 5
figure 5

a Cycle stability, b rate performance of all the samples as LIB electrodes. c Comparison of rate capacity between the SnO2 QDs/MXene-52 and other MXene-based electrode materials reported for LIBs. d Nyquist plots of the SnO2 QDs/MXene-52 and the pure SnO2 electrodes; e energy storage mechanism of the 0D–2D SnO2 QDs/MXene hybrids

Another attractive feature of the SnO2 QDs/MXene electrodes is the excellent rate performance. As shown in Fig. 5b, the SnO2 QDs/MXene electrode exhibit much enhanced rate capability compared to pure SnO2 QDs. Especially for SnO2 QDs/MXene-52, as the current density increases from 50 to 100, 200, 500, 1000, 2000, and 3000 mA g−1, its reversible capacity remains 887.4, 655.2, 584.5, 522.0, 454.7, 409.7, and 364.0 mAh g−1, respectively, demonstrating superior rate performance. When the current density returns back to 50 mA g−1 again, the capacity recovers to 688.1 mAh g−1. These results indicate that the unique 0D–2D architecture of the SnO2 QDs/MXene hybrids facilitates the Li ion diffusion and the electron transfer, thereby enhancing the reaction kinetics and the rate capability. The comparison of the rate performance for QDs/MXene-52 with previously reported SnO2-based anodes and other MXene-based anodes is plotted in Fig. 5c. The SnO2 QDs/MXene-52 demonstrate superior rate capability compared to other anode materials based on MXene matrix, such as SnO2 QDs/MXene backbone [40], SnOx nanosheets/MXene [54], Fe2O3/Mxene [37], Fe3O4 nanoparticles@MXene [50], TiO2 nanowire/Ti3C2 [55], CoO nanoparticles/Ti3C2 [44], MoS2/p-Ti3C2 [56], and Si@Ti3C2 [57] etc.

Electrochemical impedance spectroscopy was employed to compare the reaction kinetics of the SnO2 QDs/MXene-52 and the bare SnO2 QDs. The typical Nyquist plots of the two electrodes (Fig. 5d) consist of a compressed semicircle in the intermediate frequency region and a diagonal line in the low frequency range. The semicircle is related to the charge transfer resistance (Rct), and the oblique line is related to Warburg impedance, suggesting the diffusion of Li ion in the active materials [58]. The Rct values of the SnO2 QDs/MXene-52 and SnO2 QDs electrode were calculated to be 75.50 and 100.20 Ω, respectively. Obviously, SnO2 QDs/MXene-52 possesses much lower Rct value compared to bare SnO2 QDs. This can be attributed to the high electrical conductivity of the MXene and the fast charge diffusion reaction due to its unique 0D–2D structure. Moreover, the SnO2 QDs/MXene-52 shows a relatively steep low-frequency tail, indicative of high Li ion diffusibility, which results from the efficient ion transfer pathways constructed by MXene.

The above excellent electrochemical performances are because of the synergic effect between MXene nanosheets and SnO2 QDs, and the mechanism is illustrated in Fig. 5e. In the SnO2 QDs/MXene hybrids, the MXene nanosheets act as 2D substrates for uniform anchoring of SnO2 QDs. The MXene nanosheets prevent the aggregation of SnO2 QDs and work as an elastic buffer space to adapt to the volume expansion/contraction of SnO2 QDs during charging and discharging, thus leading to good cycle stability. Furthermore, MXene nanosheets with good electric conductivity construct effective conductive channels for SnO2 QDs, facilitating fast charge transport and improving the rate performance. Moreover, the unique 0D–2D structure offers massive electrochemically active sites for high specific capacity, which contribute in improving the electrochemical performance of electrode materials, thus excellent lithium-ion storage performances are obtained for SnO2 QDs/MXene hybrids.

4 Conclusions

The 0D–2D SnO2 QDs/MXene hybrids have been successfully synthesized by an efficient electrostatic self-assembly strategy. The 0D SnO2 QDs with an average size of 4.7 nm are uniformly distributed on the 2D MXene nanosheets with strong adhesion, acting as structurally stable host for lithium storage. The 2D MXene nanosheets buffer the volume change of SnO2 QDs during charge/discharge and construct effective channels for charge transport. Besides this, the 0D–2D structure creates additional active sites. The 2D conductive Ti3C2Tx MXene, ultra-small SnO2 QDs, and unique 0D–2D nanoarchitecture are synergistically responsible for the outstanding electrochemical performances of the SnO2 QDs/MXene hybrids. As an anode for LIBs, it exhibits a high specific capacity of 887.4 mAh g−1 at 50 mA g−1, excellent rate capability (364 mAh g−1 at 3 A g−1), and superior cycle stability (659.8 mAh g−1 after 100 cycles with 91% retention). These results indicate that the 0D–2D SnO2 QDs/MXene is a promising anode material for advanced LIBs. In addition, the electrostatic self-assembly method could be extended to other transition metal oxide/MXene hybrids and will have potential applications in sodium-ion batteries, potassium-ion batteries, and supercapacitors.