1. Introduction
Sodium-ion batteries (SIBs) are increasingly recognized as a promising technology for large-scale energy storage applications, primarily driven by the abundance of sodium resources and their inherent cost advantages [
1,
2,
3,
4]. However, the practical application of SIBs faces significant challenges. The larger ionic radius (~1.06 Å, 1 Å = 10
−10 m) and higher molar mass of Na
+ compared to Li
+ seriously induces substantial electrode volume changes, including expansion and potential pulverization, resulting in sluggish diffusion kinetics [
5,
6]. Consequently, identifying suitable electrode materials, particularly for the key anode materials, is crucial. While graphite serves as the commercial anode in lithium-ion batteries, it exhibits extremely low capacity for sodium storage, rendering it unsuitable for SIBs [
7,
8]. Similarly, commercially available hard carbon anodes, despite their usefulness, struggle to meet the escalating demands for high energy density SIBs due to their limited capacities [
9,
10]. This context underscores the imperative to develop novel high-performance anode materials tailored for advanced SIBs.
Molybdenum disulfide (MoS
2) has emerged as a promising anode candidate for SIBs, largely attributed to its distinctive layered architecture. The relatively large interlayer spacing (~0.62 nm), combined by weak van der Waals (vdW) forces, facilitates rapid sodium ion (de)intercalation and underpins a high theoretical capacity of 670 mAh g
−1 [
11,
12,
13,
14]. MoS
2 primarily exists in two distinct phases, 2H and 1T, dictated by variations in Mo-S atomic coordination and stacking sequences [
15]. The thermodynamically stable 2H-MoS
2 phase adopts a trigonal prismatic coordination, exhibiting semiconducting behavior characterized by a bandgap of ~1.9 eV [
16]. However, its application as a SIB anode is still hampered by inherent limitations: sluggish Na
+ diffusion kinetics, poor reversibility in Mo-S bonding, and significant volume expansion during cycling. These factors collectively lead to rapid capacity degradation [
17,
18]. While extensive strategies, such as nanostructuring [
19,
20] and carbon compositing [
21,
22] have been employed to enhance sodium storage performance of 2H-MoS
2, its intrinsic properties pose fundamental challenges, resulting in sodium storage capabilities that fall short of practical requirements. In contrast, the 1T-MoS
2 features an octahedral coordination structure and exists in a thermodynamically metastable state. Crucially, it exhibits metallic conductivity, surpassing that of the 2H phase by 10
5 to 10
7 times [
23,
24]. This exceptional electrical conductivity, coupled with a larger interlayer spacing (~1 nm), makes 1T-MoS
2 particularly advantageous for facilitating rapid Na
+ and electron transport, positioning it as an exceptionally suitable anode material for SIBs [
25,
26]. Nevertheless, the inherent metastability of the 1T phase presents a significant synthesis challenge, as it tends to degradation toward 2H-MoS
2 under ambient conditions, thereby severely limiting its practical utility.
To achieve stable 1T-MoS
2 synthesis for enhanced sodium storage performance, diverse strategies have been explored. These include alkali metal intercalation [
27], organic solvent-assisted exfoliation [
28], and heteroatom doping [
29]. Yet despite the effectiveness of these proposed strategies in enhancing the sodium storage performance of MoS
2, intrinsic challenges still exist in the resulting 1T-MoS
2 anodes. These typically involve complex and time-consuming synthesis procedures, or manifest as lower 1T phase purity, long-term stability, and production yields. Such limitations hinder their large-scale application [
30,
31]. Consequently, developing an efficient and scalable approach to obtain high-purity and highly stable 1T-MoS
2 for SIBs anodes remains a significant challenge. Recently, the electron injection strategy could be promising as an effective method to obtain high-purity and highly stable 1T-MoS
2 [
32]. For instance, Zhang et al. [
33] developed an electron-modulated and phosphate radical-stabilized approach to produce stable 1T-rich MoS
2, accelerating Na
+ insertion/extraction kinetics and yielding ultrahigh rate capability. Similarly, He et al. [
34] utilized facile electron injection engineering to fabricate a TiO
2-1T MoS
2 nanoflower composite, which reduced the Na
+ diffusion barrier and suppressed sulfur dissolution, leading to outstanding rate performance and cycling stability. Sun et al. [
35] synthesized a defective, 1T-rich MoS
2/m-C nanoflower composite, which facilitated rapid charge transport, resulting in high sodium storage capacity, superior rate capability, and excellent long-term cycling stability. The above-listed studies employed some smaller organic molecules or cations/anions, such as ethylene glycol, ascorbic acid, or Ni
2+/PO
43−, that were introduced into the hydrothermal synthesis of MoS
2, inserting the interlayer of MoS
2 and inducing the formation of 1T-MoS
2. However, the organic molecules derived carbon or inactive cations/anions could influence full delivery of specific capacity of the obtained 1T-MoS
2 electrode. How to achieve the formation of the 1T-MoS
2 through the induction of an external medium without affecting its full performance potential might be of great significance. Thus, the ethanol molecule could be appropriately considered due to the fact that the smaller ethanol molecule not only easily entered the interlayer, but also did not remain in the interlayer to derive the carbon layer due to its extremely low boiling point, which is promising for obtaining high-purity 1T-MoS
2 to realize full capacity release. Furthermore, ethanol is cleaner and more environmentally friendly compared to other small alcohol molecules (e.g., methanol, isopropanol), which could be more suitable for large-scale sample synthesis and potential battery application.
Herein, we developed an ethanol molecule assisted solvothermal method and realized a stable E-1T MoS2 via a synergetic ethanol molecule intercalation and electron injection engineering. The intercalation of the ethanol molecule and its -OH group as an electron donor facilitates the formation and stabilization of 1T-MoS2. The obtained E-1T MoS2 exhibits the sphere-like nanoflower structure composed of the regularly arranged few-layered ultrasmall 1T-MoS2 nanosheets. Benefiting from the synergetic effect of the high electrical conductivity, expanded interlayer spacing and regularly arranged nanosheets structure, the E-1T MoS2 electrode displays an excellent sodium storage performance with the high reversible capacity of 459.7 mAh g−1 after 200 cycles at 1 A g−1 and remarkable rate capability of 312.2 mAh g−1 at 10 A g−1.
2. Results and Discussion
The synthesis processes of E-1T MoS
2 and 2H MoS
2 samples are schematically illustrated in
Figure 1a. First, the ammonium molybdate (AMT) and thiourea were mixed into the deionized (DI) water via magnetic stirring, and then underwent a typical hydrothermal process to obtain a pristine 2H MoS
2 sample with an irregular nanoflower-like structure. When the DI water solvent was replaced by the ethanol solvent, the similar solvothermal process was performed to achieve an extraordinary E-1T MoS
2 sample with a regular sphere-like interlaced nanosheets structure. The transformation mechanism from 2H to 1T phase could be attributed to the ethanol molecules intercalating into the interlayer of MoS
2 grain in the solvothermal process, and its -OH functional group serving as an electron donor to induce the formation of 1T phase with expanded interlayer spacing, which could be universal for the general alcohol-based small molecules [
32,
35]. Moreover, the low-boiling point ethanol molecule did not remain in the interlayer, which is beneficial to full delivery of high sodium storage capacity discussed below. Meanwhile, the rich -OH functional groups brought by the ethanol molecules could also act as nucleation sites to confine the growth of MoS
2 nanosheets and prevent their irregular stack, forming the regular nanosphere-like structure assembled by ultrasmall nanosheets
Figure 1b and
Figure S1 clearly show the different magnification SEM images of the 2H MoS
2 sample exhibiting the nanoflower-like morphology with a diameter of about 1–2 μm, composed of many irregularly aggregated nanosheets with length sizes of around 400–500 nm. As shown in
Figure 1c,d, the E-1T MoS
2 sample presents the smaller micron-sized nanosphere-like shape due to the smaller nanosheet unit with a lateral size of about 50–100 nm.
Figure 1e also exhibits the TEM image of E-1T MoS
2 sample, further demonstrating a sphere-like morphology consisting of many regularly interwoven nanosheets with lateral sizes of around 50 nm and thicknesses of about 2–3 nm. The HRTEM image of E-1T MoS
2 sample (
Figure 1f) displays the few-layer nanosheet character, with about three layers, accompanied by a larger interlayer spacing of around 0.97 nm. As shown in the in-plane HRTEM image of the E-1T MoS
2 sample (
Figure S2), there exists hexagonal and trigonal lattice areas representing 2H and 1T phase structure characteristics, respectively, suggesting the co-existence of 1T and 2H phases in the E-1T MoS
2 sample. Meanwhile, the 2H MoS
2 sample exhibits the thicker multi-layer nanosheet structure with smaller interlayer spacing of about 0.67 nm (
Figure S3), proving the intercalation and electron transfer of the ethanol molecule induced the formation of 1T phase with the larger interlayer spacing, which is beneficial to Na
+ transport kinetics as discussed below.
Figure 1g and
Figure S4 show the selected area electron diffraction (SAED) patterns of the E-1T MoS
2 and 2H MoS
2 samples, respectively, all exhibiting the concentric ring characteristic of polycrystalline structures. Compared to the obvious concentric rings in 2H MoS
2 sample (
Figure S4), the SAED pattern of E-1T MoS
2 became weaker and its (002) diffraction ring almost disappeared, indicating the significantly reduced nanosheet size and few-layer nanosheet structure, which is also consistent with the XRD and TEM results. The EDS mapping images display the homogeneous distribution of Mo and S elements, indicating the reliable MoS
2 structure.
Figure 2a shows the XRD patterns of 2H MoS
2 and E-1T MoS
2 samples. The pristine 2H MoS
2 sample exhibits the four obvious diffraction peaks, which can be well indexed to the standard JCPDS Card No. 75-1539, indicating a pure hexagonal phase structure [
36]. Compared to 2H MoS
2 sample, the XRD diffraction peaks of the E-1T MoS
2 sample became significantly weak, and a new (002) peak at about 8.8° appeared, suggesting the formation of the ultrasmall few-layer 1T-MoS
2 nanosheets structure with low crystallinity. For the obvious left shift of main (002) peak, we also calculated the corresponding interlayer spacing (d) based on Bragg’s law: 2dsinθ = λ (for detailed calculation, see
Note S2), increasing it from 0.63 to 1 nm. Thus, we speculated that the expanded interlayer spacing and reduced nanosheet size could be attribution to the intercalation and confined growth of the ethanol molecules, respectively, which is aligned with the TEM and HRTEM results. To further understand the interesting formation mechanism of 1T-MoS
2, the XRD patterns comparison of E-1T MoS
2 samples with the different ethanol concentrations were also provided (
Figure S5). The results indicate that the full ethanol solvent environment could induce the ultrasmall few-layer 1T-MoS
2 nanosheets structure. Furthermore, for 1T-MoS
2, 1T phase stability over time is critical. Thus, we have provided the XRD patterns comparison of E-1T MoS
2 sample at fresh synthesis state and after nine months, as shown in
Figure S6. The results indicate there are no obvious changes for the XRD patterns of the sample from the two time periods, indicating the excellent 1T phase ability of the obtained E-1T MoS
2 sample. The Raman spectra of both samples are shown in
Figure 2b. For the 2H MoS
2 sample, the two obvious vibration peaks are located at about 380.5 and 406.2 cm
−1, which can be attributed to
and A
1g modes of 2H phase MoS
2 [
37]. Meanwhile, for the E-1T MoS
2 sample, the new evident characteristic peaks of the main E
1g mode (~284.9 cm
−1), and three additional J
1 (~148.5 cm
−1), J
2 (~238.1 cm
−1), and J
3 (~338.1 cm
−1) modes can be observed, which correspond to vibration modes of 1T-MoS
2 [
38]. Furthermore, an XPS test was conducted to explore the compositions and phase structures.
Figure S7 displays the survey XPS spectrum of the E-1T MoS
2 sample, which is mainly composed of Mo and S elements.
Figure 2c and
Figure S8 display the high-resolution Mo 3d spectra of E-1T MoS
2 and 2H MoS
2 samples, respectively. Compared to Mo 3d
3/2 (232.9 eV) and Mo 3d
5/2 (229.7 eV) characteristic peaks of 2H MoS
2 sample (
Figure S8), the Mo3d spectrum of the E-1T MoS
2 sample can be deconvoluted into the four main characteristic peaks at 231.4 and 228.2 eV (indexed to 1T-MoS
2), and 232.9 and 229.7 eV (assigned to 2H-MoS
2) [
39]. The peak positions of Mo 3d in E-1T MoS
2 sample shifted toward the lower binding energies, indicating enhanced Mo charge density due to the introduction of the ethanol molecules [
40]. Moreover, based on the fitting results of Mo 3d spectrum in the E-1T MoS
2 sample, the 1T phase proportion can be calculated to be as high as 70%. Furthermore, we have provided the comparison of the efficiency and phase purity of E-1T MoS
2 in previous studies [
33,
34], as shown in
Table S1. The results indicate that the synthesis efficiency of our E-1T MoS
2 sample is much higher just via a one-step solvothermal process, compared to the two previous studies (two-step or multi-step synthesis process). Although 1T phase purity is only 70% in the obtained E-1T MoS
2 sample, higher than that of this reported work [
33] and lower than that of another work [
34], it still demonstrated the highest synthesis efficiency and excellent sodium storage performance discussed below, indicating its high cost-effectiveness for SIBs anode toward practical application. Besides, the remarkable phase stability after nine months and high yield via solvothermal method further emphasize its practical value. The additional peaks at about 225.5 and 234.8 eV correspond to S 2s and Mo–O bonding, respectively. Compared to the 2H MoS
2 sample, the formation of the obvious Mo–O bonding suggests that -OH functional group in the ethanol molecule bridged MoS
2 counterpart and inserted ethanol molecule, realizing the electron transfer to induce the formation of 1T phase [
41]. In addition, the S 2p spectrum of the E-1T MoS
2 sample can be fitted into the two obvious characteristic peaks at 162.5 and 161.1 eV, as shown in
Figure 2d, which are attributed to S 2p
1/2 and S 2p
3/2 of S
2−, respectively.
Figure 2e and f show the isothermal N
2 adsorption/desorption curves and corresponding pore size distributions of 2H MoS
2 and E-1T MoS
2 samples, respectively. As shown in
Figure 2e, the E-1T MoS
2 sample displays a more evident H3-type hysteresis loop compared to the 2H MoS
2 sample, indicating a richer mesoporous structure. Furthermore, the Bruauer–Emmett–Teller (BET) specific areas of both samples can be calculated to be about 31.3 and 7.0 m
2/g for the E-1T MoS
2 and 2H MoS
2 samples, respectively. The increased specific area for E-1T MoS
2 could be related to the regularly arranged nanosheets structure and corresponding reduced nanosheets size as demonstrated in SEM and TEM analyses. Moreover, the E-1T MoS
2 sample has more abundant pore diameter distributions and larger pore volume than the 2H MoS
2 sample based the Barret–Joyner–Halenda (BJH) method (
Figure 2f), which will promote the Na
+ diffusion ability as discussed below.
The sodium storage performance of the E-1T MoS
2 electrode was evaluated via assembling a 2032 coin-type half-cell with Na foil as a counter electrode.
Figure 3a and
Figure S8 show the first three cycles CV curves of E-1T MoS
2 and 2H MoS
2 electrodes at 0.1 mV s
−1, respectively. Compared to the CV curves of the 2H MoS
2 electrode with the typical reversible conversion reaction process from MoS
2 to Mo and Na
2S (
Figure S9), the redox peaks of the CV curves in the E-1T MoS
2 electrode became less obvious, indicating the enhanced pseudocapacitive behavior, discussed below, which could be originated from the regularly arranged ultrasmall nanosheets structure with the larger specific area. Meanwhile, in the first cathodic scanning, the weak reduction peak at 0.21 V is attributed to the conversion reaction toward metallic Mo and Na
2S, accompanied with the formation of the solid electrolyte interphase (SEI) film [
42]. In the subsequent cathodic scanning, the reduction peaks shifted upward to about 1.51 and 0.65 V, which correspond to the insertion of Na
+ and follow conversion reaction processes, respectively. In the corresponding anodic scanning, the oxidation peaks of about 1.84 and 2.18 V can be attributed to the gradual Na
+ desertion and reversible conversion reaction processes. Moreover, the subsequent CV curves overlapped well, indicating the excellent electrochemical reaction reversibility.
Figure 3b and
Figure S10 display the GCD profiles for first three cycles of the E-1T MoS
2 and 2H MoS
2 electrodes at 0.1 A g
−1. The E-1T MoS
2 electrode delivered the high initial discharge and charge specific capacities of 972.0 and 716.3 mAh g
−1, much higher than those of the 2H MoS
2 electrode (462.0 and 381.2 mAh g
−1). Moreover, the subsequent GCD curves almost coincide, further suggesting the remarkable reaction reversibility of E-1T MoS
2 electrode, which agrees well with the above CV analysis. Furthermore, it can be seen that there only appear some weak inclined plateaus in the GCD curves in the E-1T MoS
2 electrode, which correspond to the extremely weak redox peaks in the CV curves, further indicating the strong pseudocapacitive behavior.
Figure 3c displays a comparison of the cycling performance of the E-1T MoS
2 and 2H MoS
2 electrodes at 1 A g
−1 (red arrow corresponds to the left coordinate axis, and blue arrow corresponds to the right coordinate axis). The E-1T MoS
2 electrode exhibits an outstanding reversible capacity of 459.7 mAh g
−1 after 200 cycles at 1 A g
−1, accompanied with a high capacity retention of 86.8%. Meanwhile, the specific capacities of the E-1T MoS
2 electrode exhibited obvious fluctuations, and demonstrated an evident capacity decay after 200 cycles, which decreased to 375.2 mAh g
−1 after 250 cycles, with a lower capacity retention of 70.8% (
Figure S11). Furthermore, to emphasize potential practical value of E-1T MoS
2 electrode, its cycling performance demonstrated a higher loading mass of about 4 mg cm
−2, also shown in
Figure S12. The results indicate that the E-1T MoS
2 electrode with the higher area loading can still deliver a high specific capacity of about 400 mAh g
−1 at 5 A g
−1 after 100 cycles. However, its cycling performance is still unsatisfactory and obvious capacity decay is apparent after 120 cycles, which is consistent with that of the above E-1T MoS
2 electrode with the low area loading. The capacity decay could be attributed to the conversion reaction mechanism between Mo/Na
2S and MoS
2 accompanied by the larger volume variation, resulting in a cycling performance that is worthy of further improvement. However, the pristine 2H MoS
2 electrode displays more inferior cycling performance, which maintains a stable capacity of 329.5 mAh g
−1 in the former 40 cycles and then exhibits a significant capacity fading in the subsequent cycles. It retains an unsatisfactory capacity of 43.6 mAh g
−1 after 200 cycles at 1 A g
−1, with an extremely low cycling retention of about 13.2%. Furthermore, the comparison of the coulombic efficiencies (CE) of both electrodes during long-term cycling were also provided in
Figure S13 (for CE calculation, see
Note S3). The results indicate the E-1T MoS
2 electrode exhibits relatively more stable CE values than the pristine 2H MoS
2 electrode, further indicating the superior electrochemical reversibility. The comparison of the rate capability for E-1T MoS
2 and 2H MoS
2 electrodes is displayed in
Figure 3d. The E-1T MoS
2 electrode shows excellent rate performance with the reversible capacities of 662.0, 596.0, 537.1, 487.1, 463.5, and 390.1 mAh g
−1 at the respective current densities of 0.1, 0.2, 0.5, 1, 2, and 5 A g
−1. Even at a high current density of 10 A g
−1, a prominent capacity of 312.2 mAh g
−1 can still be obtained. Nevertheless, the 2H MoS
2 electrode shows a poor rate capability of 119.7 mAh g
−1 at 10 A g
−1. To intuitively demonstrate the enhanced rate capability for E-1T MoS
2 electrode, its GCD curves at the various current densities were plotted in
Figure 3e. It can be seen that the GCD curves overlap well with a smaller voltage polarization from 0.1 to 10 A g
−1, accompanied with a high rate capacity retention of about 47.2%, magnifying a rapid Na
+ diffusion kinetics, which is consistent with the rate performance analysis. Impressively, the rate capability of the E-1T MoS
2 electrode outperforms those of many previously reported MoS
2-based anodes for SIBs, as shown in
Figure 3f. To emphasize the practicality of E-1T MoS
2 anode for SIBs, the corresponding full cell was also assembled via E-1T MoS
2 as anode and Na
3V
2(PO
4)
3 as cathode.
Figure S14a shows the GCD curves for first three cycles of the E-1T MoS
2 ‖ NVP full cell within 0.5–3.5 V at 0.1 A g
−1, displaying normal charge/discharge behavior within voltage range and outputting the sloping charge and discharge plateaus at about 2.3 and 1.7 V, respectively, in the subsequent cycles. The cycling performance of the full cell at 0.1 A g
−1 is displayed in
Figure S14b. It displayed a high initial charge capacity beyond 400 mAh g
−1, which could be attributed to the formation of SEI film on the anode side contributing to the higher irreversible capacity. It also delivered an initial discharge capacity of about 223.6 mAh g
−1, which slightly decreased to 105.8 mAh g
−1 after 100 cycles, accompanied by a capacity retention of about 47.3%. The lower specific capacity of the full cell compared to that of the half cell could be closely related to the operation voltage range of the anode in the full cell. Based on the charge and discharge output voltages of about 2.3 and 1.7 V, respectively, it can be determined that only the capacity within 0–1.5 V or even a narrower voltage range can be contributed to the full cell, leading to the obtained specific capacities lower than 400 mAh g
−1. Furthermore, the present specific capacity and retention remain limited, yet they clearly demonstrate the feasibility of applying E-1T MoS
2 in full-cell devices and highlight directions for further optimization, including capacity matching between anode and cathode, voltage window, and anode performance.
To elucidate the enhanced Na
+ reaction kinetics of the E-1T MoS
2 electrode, CV measurements were conducted at various scan rates ranging from 0.2 to 1.2 mV s
−1 (
Figure 4a). Despite increasing scan rates, the CV profiles retain similar shapes. However, they exhibit progressively larger peak currents and minimal peak shift, indicating reduced polarization. Furthermore, the relationship between the peak current (
i) and scan rate (
v) follows a power-law behavior described by Equation (1): [
43]
where a represents dimensionless coefficient, b = 1.0 signifies pseudocapacitive charge storage, b = 0.5 indicates diffusion-limited behavior, and 0.5 < b < 1.0 denotes a mixed charge storage mechanism. By analyzing the anodic and cathodic peaks across multiple scan rates, the linear relationship between log(
i) and log(
v) was plotted and fitted as presented in
Figure 4b. The calculated b values are about 0.90 and 0.91 for the cathodic and anodic peaks, respectively, thus confirming a surface pseudocapacitance-dominated charge storage mechanism in the E-1T MoS
2 electrode. To further quantify the pseudocapacitive contribution, the current response (
i) can be deconvoluted as a function of scan rate (
v) using Equation (2) [
44]
where k
1v corresponds to the pseudocapacitive contribution, while k
2v1/2 represents the diffusion-controlled contribution.
Figure 4c illustrates the calculated pseudocapacitive contributions for the E-1T MoS
2 electrode across varying scan rates. Notably, this contribution increases progressively with increasing scan rates, reaching ratios of 73.4%, 74.9%, 78.2%, 82.1%, 88.4%, and 95.4% at 0.2, 0.4, 0.6, 0.8, 1.0, and 1.2 mV s
−1, respectively. Specifically, at 1.2 mV s
−1, a representative CV profile was fitted to intuitively demonstrate this behavior (
Figure 4d), revealing a dominant surface pseudocapacitance contribution of 95.4% (for detailed fitting process, see
Note S4). Such a high surface contribution even at a lower scan rate of about 1 mV s
−1 could be a common phenomenon in the modified electrode materials for batteries [
29,
33,
35]. Furthermore, the CV curves within ultrahigh scanning rates of 10–100 mV s
−1 (
Figure S15) no longer represent charge storage behavior of the electrode for batteries, but rather the charge storage behavior similar to that of in supercapacitors, accompanied by significant variation of shapes and currents of CV curves compared to those of 0.2–1.2 mV s
−1, leading to the failed simultaneous pseudocapacitive fitting with those of within 0.2–1.2 mV s
−1, which is originated from the failed linear fitting for obtaining k
1 values via the fitting process shown in
Note S4, thus failing to obtain continuous pseudocapacitive contribution variations from 0.2 to 100 mV s
−1. While the pseudocapacitive contributions at 10–100 mV s
−1 can still be obtained, the fitting process is only carried out within this scanning rate range, obtaining a pseudocapacitive contribution that is no more than 70% at most, as also demonstrated by the previous studies on most MoS
2-based electrodes materials for supercapacitors [
45], which seems to be contradictory with the pseudocapacitive fitting results obtained at 0.2–1.2 mV s
−1. It could be that, at ultrahigh scan rates, the ion diffusion process cannot be completed within such an extremely short timescale, leading to surface-limited charge accumulation that resembles supercapacitor-type behavior rather than true battery-type pseudocapacitance. The above results further indicate that overhigh scanning rates could be unsuitable for studying pseudocapacitive contribution of the present E-1T MoS
2 electrode for SIBs. Meanwhile, the manifestation of pseudocapacitance could be different between battery-type and supercapacitor-type systems even on the same MoS
2-based electrode materials, deserving further deep clarification in the future. Inspiringly, this high surface pseudocapacitance contribution of E-1T MoS
2 for SIBs significantly enhances rapid Na
+ transport kinetics, particularly at high current densities, underpinning the superior rate capability demonstrated earlier. The origin of this high surface pseudocapacitance primarily stems from the regularly arranged sphere-like ultrasmall nanosheet structure and former nanoclusters during cycling, as discussed below, facilitating rapid surface sodium adsorption/desorption processes, which is fully consistent with the CV analysis. To further understand the elevated Na
+ diffusion kinetics of the E-1T MoS
2 electrode with the outstanding rate capability, galvanostatic intermittent titration technique (GITT) measurements of E-1T MoS
2 and 2H MoS
2 electrodes were conducted, and are displayed in
Figure 4e. The E-1T MoS
2 electrode exhibits the longer discharge/charge time and smaller voltage drop compared to the 2H MoS
2 electrode. Furthermore, the diffusion coefficient (
DNa+) can be calculated based on Equation (3): [
46]
where
mB,
Vm,
MB, and
S are mass, molar volume, molar mass and contact area of electrode material, respectively.
τ, Δ
ES, and Δ
Eτ are pulse current time, voltage variations of steady state, and pulse process, respectively. As shown in
Figure 4f (blue arrow corresponds to the left coordinate axis, and green arrow corresponds to the right coordinate axis), the
DNa+ values of E-1T MoS
2 electrode during discharging and charging processes were calculated to be about 10
−9~10
−10 and 10
−8~10
−10 cm
2 s
−1, respectively, evidently higher than those of 2H MoS
2 electrode (10
−10~10
−11 and 10
−9~10
−11 cm
2 s
−1), indicating the significantly enhanced Na
+ diffusion kinetics of E-1T MoS
2 electrode, which could be attributed to the synergetic effect between the metallic 1T phase and regularly arranged ultrasmall nanosheet structure.
To further reveal the sodium storage mechanism of the optimal E-1T MoS
2 electrode during the sodiation/desodiation processes, ex-situ XRD measurements at the different discharging/charging states were performed to explore its phase structure evolution, as shown in
Figure S16. At the open circuit voltage (OCV) state, The E-1T MoS
2 electrode still displays the similar weak XRD pattern with its powder sample. When discharged from OCV to 0.01 V, several obvious peaks representing Na
2S species appeared, indicating the phase transformation from MoS
2 to Mo and Na
2S during the sodiation process. When charged to 3 V, the XRD peaks of Na
2S disappeared, while there appeared no obvious peaks of MoS
2, indicating the amorphous character of the reconstructed MoS
2 structure, and possible conversion reaction from Mo and Na
2S to MoS
2 nanoclusters, which will be further verified below. To further explore the phase structure evolution in E-1T MoS
2 electrode in-depth, ex-situ TEM tests were also conducted, as shown in
Figure 5a–c. At OCV state, the HRTEM image of E-1T MoS
2 electrode still exhibits the few-layer ultrasmall 1T-MoS
2 nanosheets structure with the larger interlayer spacing of about 0.97 nm, which is consistent with its powder sample (
Figure 5a). At the fully discharged state of 0.01 V, the typical lattice fringes of (002) planes in the HRTEM image of E-1T MoS
2 electrode disappeared, while two sets of new lattice fringes appeared, consisting of the smaller particle fringes of about 0.22 nm and larger continuous fringes of about 0.23 nm (
Figure 5b), indicating the formed metallic Mo particles became embedded into the Na
2S matrix. Moreover, the EDS-mapping images at the fully discharged states (
Figure 5d) exhibit the uniform distribution of Mo, S, and Na elements, further suggesting the conversion reaction mechanism during the sodiation process. Moreover, we have provided the ex-situ TEM measurements of the different regions in the E-1T MoS
2 electrode at the fully discharged state as shown in
Figure S17. The results indicate that there exist obvious SEI film layers coated on the surface of the E-1T MoS
2 electrode after the sodiation reaction process, which could be beneficial to the cycling stability of the electrode, further proving the above conversion reaction process accompanied by the formation of SEI film. Furthermore, at the fully charged state of 3 V, the ex-situ HREM image in
Figure 5c displays there appeared many ultrasmall 1–2 layer nanosheets stripes with size of about 2 nm (yellow dotted circles) in the electrode, which are short-term ordered but long-term disordered. Thus, we vividly referred to it as “MoS
2 nanoclusters”. The above results indicate the reversible transformation of Mo and Na
2S species toward MoS
2, as confirmed by the EDS-mapping images in
Figure S18, which could contribute the more rapid pseudocapacitive behavior to sodium storage capacity as discussed above. The ex-situ EDS mapping images at the fully charged state exhibit the same distribution of Mo and S elements, while the emerged dispersive distribution of Na element could be attributed to the formed SEI film on the electrode, further suggesting the reversible conversion reaction process accompanied by SEI film formation. Moreover, the ex-situ SAED pattern at the fully de-sodiation state (
Figure S19) presents a fuzzy central spot, further confirming the formed MoS
2 nanoclusters with the amorphous character during the charging process. To intuitively reflect the sodium storage mechanism of the E-1T MoS
2 electrode, a schematic illustration is presented in
Figure 5e. During the first sodiation process, E-1T MoS
2 converted to the metallic Mo and embedded into Na
2S. The Mo and Na
2S species performed the reversible conversion reaction toward the MoS
2 nanoclusters during the subsequent sodiation/desodiation processes. Based on the above sodium storage mechanism, the excellent sodium storage performance of E-1T MoS
2 electrode could be assigned to the following aspects: (1) the induced metallic 1T phase structure with the large interlayer spacing and high electrical conductivity is beneficial to the rapid transfer of Na
+ or e
-, promoting sodium reaction kinetics; (2) the regular sphere-like ultrasmall nanosheets structure provides more sodium storage active sites, shortens the diffusion path of Na
+ and accommodates the larger volume expansion during the cycling process; (3) the formed MoS
2 nanoclusters display the stronger surface pseudocapacitance behavior, leading to the enhanced Na
+ reaction kinetics.