Abstract

The surging demand for Li rechargeable batteries with high energy densities and rapid rate capability has propelled research on materials that can replace the conventional anode materials like graphite. Vanadium pentoxides (V2O5) have emerged as promising anode candidates owing to their excellent rate capability. Specifically, V2O5 allows the electrochemical prelithiation process, in which three Li-ions can be inserted to form Li3V2O5, followed by reversible insertion and extraction of two Li-ions. Nevertheless, the unsatisfactory Li-ion diffusion coefficients and electrical conductivities of these materials remain major drawbacks. Here, we propose a Zn-doped V2O5 anode, fabricated using a two-step sol–gel method, for high-rate Li-ion batteries. Zn was incorporated into V2O5 to enhance the Li-ion transport kinetics through the electrode. The crystal structure (orthorhombic) of Zn-doped V2O5 was identified by X-ray diffraction analysis, and the Zn doping was confirmed by X-ray photoelectron microscopy. The effect of Zn doping was thoroughly examined using various analytical methods, such as cyclic voltammetry and galvanostatic intermittent titration technique. The Zn-doped V2O5 electrode exhibited remarkable cycling durability, enduring for 1000 cycles, while retaining an enhanced capacity, even under a high rate of 2C.

1. Introduction

Li-ion batteries (LIBs) have undergone continuous development over several decades in response to market demand [14]. Carbon-based materials, particularly graphite, are extensively applied as anode materials in LIBs owing to their moderate capacity, strong cycling stability, and affordability [57]. However, their rate performance and capacity remain constrained because graphite can store Li-ions only through an intercalation mechanism [510]. Conversely, anode materials that rely on alloying mechanisms, such as Si and Sn, are also proposed owing to their high capacities and rapid rate characteristics [1118]. However, structural degradation of these active components, caused by substantial volume changes during Li insertion and extraction, results in a subpar cycling performance [1113]. Thus, developing novel materials suitable for extended usage and rapid charging applications is necessary.

Because of their outstanding chemical stability and cost-effectiveness, vanadium pentoxides (V2O5) are considered advantageous as electrode materials for LIBs [1921]. However, the low diffusion coefficient of Li+ ions and poor electrical conductivity of V2O5 limit its use in high-rate applications [2224]. Several strategies such as structural and compositional manipulations, including cation doping, have been suggested to improve its rate performances [2530]. Among these strategies, cation doping, which can enhance the intercalation and deintercalation kinetics of V2O5, has been widely explored [3034]. To date, the feasibility of doping various cations such as Ni2+, Mn2+, Cu2+, and Sn4+ has been investigated [3033]. Li et al. prepared Sn-doped V2O5 microspheres using the solvothermal method [30]. Introducing Sn into V2O5 creates oxygen vacancies that facilitate the diffusion of Li-ions, leading to an enhanced capacity and rate performance. Similarly, Ni-V2O5 microspheres have been synthesized using a nickel-mediated polyol method followed by calcination [31]. Notably, these materials have been developed with a focus on fabricating cathodes for LIBs.

Recently, V2O5 has been repurposed as an anode material through an electrochemical prelithiation process, demonstrating impressive long-term cycling and excellent rate capabilities [35]. The Li3V2O5 material produced through the electrochemical prelithiation process has been further investigated [3638]. Lan et al. exhibited the superior low-temperature performance of the full cell incorporating the Li3V2O5 anode [36]. Ren et al. utilized the Li3V2O5 material in Li-ion capacitors, demonstrating exceptional rate capability and remarkable cycling stability [37]. However, improving the limited Li-ion diffusion coefficients and electrical conductivities of V2O5 remains challenging. Hence, there is a need for a new method to address the issues associated with the existing V2O5 material and enhance its electrochemical performance as an anode for LIBs.

In this study, a novel method involving the introduction of Zn doping into V2O5 was proposed for the first time. The Zn doping effectively addressed issues in the original V2O5, thereby improving its electrochemical performance. To elucidate the reasons behind the improvement in electrochemical performance due to Zn doping and to confirm the presence of zinc doping, several structural and electrochemical analyses were conducted.

2. Experimental

2.1. Chemicals

Vanadium pentoxide (V2O5, 99.6%) powder and ammonia solution (28-30%) were obtained from Sigma-Aldrich and utilized without additional refinement. Zinc nitrate hexahydrate (Zn(NO3)2∙6H2O, 98%) was supplied by Samchun and utilized without additional purification.

2.2. Material Synthesis

The Zn-doped V2O5 material (namely, the Zn-V2O5 sample) was synthesized using a two-step sol-gel procedure. Initially, 20 mmol of vanadium pentoxide and 1 mmol of zinc nitrate hexahydrate were dissolved in deionized (DI) water and agitated for 24 h to obtain a homogeneous solution. Subsequently, an ammonia solution was added into the prepared solution to adjust the pH to 8.5, and the resulting solution was stirred for an additional 24 h to create a uniform solution. Finally, the resulting solution was subjected to centrifugation and dried at 80°C for 12 h. Next, commercially available V2O5 powder without any modification (referred to as the V2O5 sample) was subjected to various analyses to serve as a reference.

2.3. Material Characterization

To examine the crystal structures of the samples, the samples were analyzed by X-ray diffraction (XRD) using a Rigaku Ultima IV instrument. The morphology of the prepared specimens was assessed by field emission scanning electron microscopy (FE-SEM) using a JEOL 7500 microscope as well as by high-resolution transmission electron microscopy (HR-TEM) using a JEOL ARM-200F instrument. Additionally, X-ray photoelectron spectroscopy (XPS) was employed to analyze the chemical compositions of the samples.

2.4. Electrochemical Measurements

The electrode fabrication process for all the samples was conducted as follows. First, the prepared active material (either V2O5 or Zn-V2O5 sample), a conducting agent (Super P), and a binder (polyvinylidene fluoride, PVDF, Kynar 2801), in a ratio of 60 : 30 : 10 by weight, were dissolved in N-methyl-2-pyrrolidone (NMP) solution. Next, the resulting slurry was applied onto a copper foil and subjected to vacuum drying at 120°C for a duration of 12 h (hereafter, the V2O5 and Zn-V2O5 electrodes are called as electrodes prepared from the V2O5 and Zn–V2O5 samples, respectively). Additionally, Li3V2O5 and Zn–Li3V2O5 electrodes were fabricated through lithiation of the V2O5 and Zn–V2O5 electrodes in a half cell operated at a current density of 0.1 A g–1, respectively.

In the half-cell testing, a Li foil was used as the counter/reference electrode, and the prepared electrodes served as the working electrodes. The cell configuration included the prepared electrode, an electrolyte composed of 1 M LiPF6 in a mixture of ethylene carbonate and diethyl carbonate (in a 3 : 7 volume ratio), an Li foil, and a polyethylene separator. Galvanostatic testing of the cells was conducted using a battery cycler (Basytec CTS-Lab) within a cutoff voltage range from 0.01 to 2.0 V (vs. Li+/Li). All the operations were carried out under an inert atmosphere, specifically inside an Ar-filled glovebox.

To prepare an LiNi0.8Mn0.1Co0.1O2 cathode for full-cell tests, slurries were prepared by dissolving the active material (LiNi0.8Mn0.1Co0.1O2, 80% by weight), a conducting agent (Super P, 10% by weight), and a binder (PVDF, 10% by weight) in NMP. These slurries were then applied onto an Al-foil current collector. Following the coating process, the electrodes underwent compression and were subsequently dried in an oven at a temperature of 120°C for 12 h.

Cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) analyses were carried out using a potentiostat (BioLogic VSP). The CV measurements were conducted across a potential range of 0.01 to 2.0 V (vs. Li+/Li) with varying scan rates of 0.2, 0.4, 0.6, 0.8, and 1.0 mV s–1 at room temperature (25°C). Simultaneously, the EIS measurements were performed over a frequency range from 1 MHz to 0.1 Hz with an amplitude of 15 mV. To determine the Li-ion diffusion coefficients in the different materials, the galvanostatic intermittent titration technique (GITT) was employed. During the GITT test, charge and discharge cycles were executed within a potential range of 0.01 V to 2.0 V (vs. Li+/Li) using a pulse current of 0.1C ( mAh g–1) followed by a rest period of 3 h.

3. Results and Discussion

Figure 1 displays the XRD patterns of the V2O5 and Zn-V2O5 samples, incorporating reference values based on the standards established by the Joint Committee on Powder Diffraction Standards (JCPDS). In the pattern of the V2O5 sample (used as the precursor before the Zn doping), distinct peaks corresponding to the orthorhombic structure of V2O5 are observed (JCPDS No. 041-1426) [39]. The same peaks, which correspond to the orthorhombic structure of V2O5, are visible in the XRD profile of the Zn-V2O5 sample, suggesting that the orthorhombic structure of V2O5 is preserved even after the introduction of Zn. Furthermore, the obtained XRD patterns were magnified to investigate whether Zn doping could impact the cell parameters. The corresponding results are illustrated in Figure S1. In contrast to the V2O5 sample, the (001) and (110) reflections in the Zn-V2O5 sample exhibited a slight shift towards a lower angle. This indicates an expanded V2O5 lattice, attributed to the larger radius of Zn2+ (0.74 Å) compared to V5+ (0.68 Å). These findings suggest the incorporation of Zn2+ ions into the crystal structure of V2O5. The particle size of both samples from the XRD results based on the Debye-Scherrer equation can be calculated, and the results are shown in Table S1.

The Zn-V2O5 sample was analyzed using multiple characterization methods to gain insights into its morphology. Figure S2 displays the FE-SEM image of the V2O5 sample, corresponding to the typical FE-SEM image of a commercial V2O5 sample [40]. Figure 2(a) exhibits the FE-SEM image of the Zn-V2O5, showing aggregation of the primary particles into a secondary particle. Figure 2(b) shows the TEM image with the energy dispersive spectroscopy (EDS) results (Figure 2(c)) of the Zn–V2O5 sample. The EDS graph is illustrated in Figure S3. Based on the EDS analysis results, it is confirmed that the V and O elements are uniformly distributed throughout the entire particle. Additionally, trace amounts of Zn (0.63 at%) are also detected, indicating the successful incorporation of Zn in this sample. Figure 2(d) presents the HR-TEM images, and Figures 2(e) and 2(f) show the fast Fourier transform (FFT) patterns of the samples. The measured lattice spacing of the crystallites is 3.40 and 4.39 Å, corresponding to the (110) and (001) reflections of the orthorhombic structure of V2O5, respectively [41]. The particle size distribution from the TEM images was obtained, and the results are given in Figure S4.

Subsequently, XPS analyses were conducted to discern the distinct chemical states of the V2O5 and Zn-V2O5 samples. Figure 3(a) displays the survey spectra of the V2O5 and Zn-V2O5 samples, and Figure 3(b) shows the Zn 2p core-level spectra of both the samples. No discernible peaks are visible in the spectrum of the undoped sample, whereas distinct peaks corresponding to ZnO (at 1022.1 and 1045.6 eV) appear in the spectrum of the Zn-doped sample [42, 43]. Figures 3(c) and 3(d) depict the V 2p3/2 core-level spectra of the V2O5 and Zn-V2O5 samples, respectively, with deconvoluted profiles, and the corresponding areal ratios are presented in Table 1. The V 2p3/2 spectra can be separated into two peaks corresponding to the V5+ (at 517.0 eV) and V4+ (at 515.7 eV) states [44, 45]. For the undoped sample, the determined areal ratios of V5+ and V4+ are 97.3% and 2.7%, respectively. By contrast, in the Zn-doped sample, the areal ratio of V5+ decreases from 97.3% to 94.2%, while that of V4+ increases from 2.7% to 5.8%. Figure S5a and b shows the O 1s core-level spectra of the V2O5 and Zn-V2O5 samples, respectively, with deconvoluted profiles. The corresponding areal ratios for the V2O5 and Zn-V2O5 samples are shown in Table S2. The O 1s spectra can be distinguished into two peaks originating from the V5+ (529.9 eV) and V4+ (531.2 eV) states [45]. The deconvoluted O 1s spectra are similar to those of V 2p3/2 spectra shown in Figure 3. Specifically, before the Zn doping, the areal ratios for V5+ and V4+ are 92.5% and 7.5%, respectively. By contrast, following the Zn doping, the areal ratio of V5+ decreases from 92.5% to 88.6%, while that of V4+ increases from 7.5% to 11.4%. The increased V4+/V5+ ratio after zinc doping can be ascribed to the substitution of V5+ by Zn2+ ions, indicating the successful doping of Zn into V2O5. Moreover, the occupation of V5+ sites by Zn2+ leads to the formation of oxygen vacancies, creating rapid channels for Li+ intercalation/extraction [30].

To assess the ability of the Zn-V2O5 sample to undergo electrochemical prelithiation to yield ω-Li3V2O5, we conducted ex situ XRD measurements at a cutoff voltage of 1.9 V. Figure S6 presents the ex situ XRD patterns of the V2O5 and Zn-V2O5 electrodes discharged to 1.9 V at 0.1 A g–1. Evidently, both the V2O5 and Zn-V2O5 electrodes exhibit identical peaks at approximately 38.16°, 44.42°, and 64.44°, corresponding to the (111), (200), and (220) crystal planes of ω-Li3V2O5, respectively [35]. This result indicates that similar to V2O5, Zn–V2O5 can indeed undergo electrochemical prelithiation to form ω-Li3V2O5.

Figure 4 presents the electrochemical properties of the Li3V2O5 and Zn-Li3V2O5 electrodes. The voltage profiles of these electrodes shown in Figure 4(a) indicate that for the Li3V2O5 electrode, the initial discharge and charge capacities are 309.3 and 277.4 mAh g–1, respectively, with an initial coulombic efficiency of 89.7%. Conversely, for the Zn-Li3V2O5 electrode, the initial discharge and charge capacities at a 2C rate ( mA g–1) are 321.4 and 298.1 mAh g–1, respectively, with an initial coulombic efficiency of 92.8%. Notably, after the Zn doping, the capacities and initial coulombic efficiencies improved. Figure 4(b) illustrates the Nyquist plots of the Li3V2O5 and Zn-Li3V2O5 electrodes, indicating that the resistance of the Zn-Li3V2O5 electrode is lower than that of the Li3V2O5 electrode. This reduction in resistance can be attributed to the Zn doping-induced oxygen vacancies, which reduce the charge-transfer resistance [31]. Figure 4(c) compares the rate performances of the Li3V2O5 and Zn-Li3V2O5 electrodes. Notably, the capacity of the Zn-Li3V2O5 electrode surpasses that of the Li3V2O5 electrode at various rates, and this improved rate capability can be assigned to the Zn doping effect. Lastly, Figure 4(d) shows the cycling performances of both the Li3V2O5 and Zn-Li3V2O5 electrodes at 2C. The Zn-Li3V2O5 electrode consistently outperforms the Li3V2O5 electrode in terms of capacity retention over extended cycling periods. Overall, the enhancement in the electrochemical characteristics of the Zn-V2O5 sample can be attributed to the oxygen vacancies created by the introduction of Zn. These vacancies increase the number of active sites available for the intercalation and deintercalation of the Li+ ions [30, 31].

To assess the practical applicability of the proposed Zn-doped anode, a full cell was assembled using the Zn-Li3V2O5 anode, which was prepared using the electrochemical lithiation method described earlier, and an LiNi0.8Co0.1Mn0.1O2 (NCM 811) cathode (Figure S7). Figure S7a shows the voltage profiles of the full cell subjected to cycling at a consistent current of 0.1 A g–1 within a voltage range of 1.5 V to 3.9 V. Figure S7b shows the cycling performance of the full cell at a rate of 0.1 A g–1, demonstrating its robust stability over 200 cycles without noticeable degradation.

To investigate the Li+ storage mechanism of the Li3V2O5 and Zn-Li3V2O5 electrodes, we conducted CV measurements at different scan rates ranging from 0.2 to 1.0 mV s–1. Figures 5(a) and 5(b) show the CV results of the Li3V2O5 and Zn-Li3V2O5 electrodes, respectively. The CV curves of both the electrodes exhibit distinct peaks at approximately 0.3–0.8 V (vs. Li+/Li), which can be attributed to the processes of Li insertion and extraction [37]. Notably, the shapes and positions of these CV curves remain consistent, irrespective of the presence of the Zn dopants or variations in the scan rate. According to the power law, the relationship between peak current and scan rate can be expressed using the following equation [37, 4648]: where is the peak current, and are constant parameters, and is the scan rate. The Li+ storage mechanisms can be elucidated by assessing the values of the parameter “.” Specifically, approaching 0.5 implies that the Li+ storage mechanism is predominantly governed by a diffusion-limited process. Conversely, a value of 1.0 signifies that the Li+ storage mechanism is entirely controlled by a capacitive process [46]. The computed values, based on the CV curves displayed in Figures 5(a) and 5(b), are presented in Figure S8a and b for the Li3V2O5 and Zn-Li3V2O5 samples, respectively. Evidently, the values exhibited by the Zn–Li3V2O5 electrode for both anodic and cathodic peaks are higher than that shown by the Li3V2O5 electrode. This result suggests that unlike the reactions that occur in the Li3V2O5 electrode, those in the Zn-Li3V2O5 electrode are more likely controlled by a capacitive process.

Moreover, the peak current () originates from two different mechanisms, i.e., capacitive and diffusion-limited processes, as described in the following equation [47]: where is the peak current at a certain potential and and represent the capacitive and diffusion-limited processes, respectively. The capacitive and diffusion contributions calculated based on the above equation at a scan rate of 0.2–1.0 mV s–1 for both the electrodes are shown in Figures 5(c) and 5(d). The capacitive contributions determined from the areal portions in the CV curves at 1.0 mV s–1 are displayed in Figures 5(e) and 5(f). In the Li3V2O5 electrode, the capacitive contribution at a scan rate of 0.2 mV s–1 is 56.6%, which gradually increases to 74.4% at a scan rate of 1.0 mV s–1. By contrast, the capacitive contributions in the Zn-Li3V2O5 electrode at scan rates of 0.2 and 1.0 mV s–1 are 65.1% and 80.6%, respectively, which are higher than those in the Li3V2O5 electrode. This result implies that Zn-V2O5 is superior to V2O5 in terms of rapid Li+ storage kinetics. The accelerated Li+ storage kinetics observed in the Zn-V2O5 sample can be attributed to the creation of oxygen vacancies through the introduction of Zn. Presence of oxygen vacancies is advantageous for forming a more open structure, which facilitates easy penetration and movement of Li-ions [30, 31].

Additionally, we employed the GITT to analyze the Li-ion diffusion coefficients, and the corresponding results obtained for the Li3V2O5 and Zn-Li3V2O5 electrodes are shown in Figures 6(a) and 6(b), respectively. The Li-ion diffusion coefficient can be calculated using the following equation [4952]: where , , , and denote the electrode thickness, pulse duration, steady-state voltage change, and total voltage change (eliminating iR drop), respectively. The Li-ion diffusion coefficients, determined from the GITT curves during the Li insertion and extraction processes, are illustrated in Figures 6(c) and 6(d). Specifically, during the Li insertion process, the average Li-ion diffusion coefficients for the Li3V2O5 and Zn-Li3V2O5 electrodes are computed as 2.03 and  cm2 s–1, respectively. Conversely, during the Li extraction process, the average Li-ion diffusion coefficients for the Li3V2O5 and Zn-Li3V2O5 electrodes decrease to 1.97 and  cm2 s–1, respectively. The observed high average Li-ion diffusion coefficients of the Zn–Li3V2O5 electrode confirm that the Zn doping into V2O5 improved the Li+ diffusion kinetics in the doped electrode. The electrochemical kinetic investigation results confirm the excellent capacitive behavior and high Li-ion diffusion coefficients of the Zn-V2O5 composite; these factors collectively enhance its rate performance as well as capacity.

4. Conclusions

In this study, we introduced, for the first time, the concept of doping Zn into V2O5 using a two-step sol-gel method. Zn doping into V2O5 was employed to enhance the conductivity and Li-ion diffusion rates of V2O5. Consequently, the electrochemical performance of V2O5 as an anode for LIBs was improved through zinc doping. The incorporation of Zn into V2O5 was verified using multiple analytical techniques, such as EDS and XPS. In addition, multiple structural and electrochemical analyses confirmed that Zn incorporation improved the Li+ ion storage kinetics of V2O5, and the resultant Zn-V2O5 material consequently exhibited strong capacitive behavior. Lastly, these materials exhibited excellent cycling durability, sustaining their performance for over 1000 cycles at a high discharge rate of 2C, while delivering superior rate performances along with increased capacity compared to the V2O5 sample. The zinc doping method can open up new possibilities for developing electrodes in high-power energy storage.

Data Availability

The data that supports the findings of this study are available from the corresponding author upon a reasonable request.

Conflicts of Interest

There are no conflicts to declare.

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korean Government (2021M3H4A3A02086910, 2022R1A2C1011181, and 2022R1A5A7000765).

Supplementary Materials

Figure S1: enlarged XRD patterns of the Zn-V2O5 and V2O5 samples. Figure S2: FE-SEM image of the V2O5 sample. Figure S3: EDS results of the Zn-V2O5 sample. Figure S4: (a, c) TEM images of the V2O5 and Zn-V2O5 samples. (b, d) Particle size distribution plots obtained from the TEM images of the V2O5 and Zn-V2O5 samples. Figure S5: O 1s XPS core-level spectra for the (a) V2O5 and (b) Zn-V2O5 samples. Figure S6: ex situ XRD profiles of the V2O5 and Zn-V2O5 electrodes at a cutoff voltage of 1.9 V. Figure S7: (a) voltage profiles and (b) cycling performances of Zn-Li3V2O5||NCM 811 full cell at 100 mA g–1. Figure S8: values calculated from the log (scan rate) vs. log (peak current) curves. Table S1: crystallite size calculation using the Debye-Scherrer equation. Table S2: abundance ratio for the O valence states observed in the O 1s XPS spectra of the V2O5 and Zn-V2O5 samples. (Supplementary Materials)