1 Introduction

Lithium-ion batteries (LIBs) are regarded as a kind of green power with good cycling performance and high energy density [1,2,3,4,5], which have become one of the hottest research topics these years. At present, the LIB anodes are normally composed of graphite-like materials because of their inexpensive and high theoretical capacity (372 mAh g−1) [6]. However, due to the relatively low rate capability, it is easy to cause the formation of thick solid electrolyte interface (SEI) layers and the growth of lithium dendrites which are hindering the application of graphite-anode materials [7]. Li4Ti5O12 has been explored to overcome these disadvantages [8,9,10], which could avoid the formation of thick SEI layers and display good safety performance. However, the low theoretical capacity of Li4Ti5O12 (175 mAh g−1) limits its practical application. Therefore, it is necessary to explore novel anode material substitutes with similar advantages to Li4Ti5O12 but much higher theoretical capacities.

Recently, niobium oxide-based materials have attracted great attention due to their high theoretical capacities (374–403 mAh g−1) which are benefitted from the Nb4+/Nb5+ and Nb3+/Nb4+ multi-electron redox couples. To date, a series of niobium oxide-based anode materials have been explored by our group, such as Nb25O62 [11], MoNb12O33 [12], AlNb11O29 [13], Al0.5Nb24.5O62 [14], CrNb11O29 [15], Mg2Nb34O87 [16], GaNb11O29 [17], TiNb24O62 [18], and ZrNb24O62 [19]. FeNb11O29 [20], W3Nb14O44 [21], and K2Nb8O21 [22] were also found by other researchers to be good anode material candidates. In particularly, FeNb11O29 has a theoretical capacity of 400 mAh g−1 owing to the existence of three Fe2+/Fe3+, Nb4+/Nb5+, and Nb3+/Nb4+ redox couples (each formula unit of FeNb11O29 could transfer 23 electrons) [20]. However, the practical capacities of FeNb11O29 reported were only 168–273 mAh g−1 at 0.1C [20, 23,24,25]. Therefore, it is of great significance to enhance the specific capacity of FeNb11O29 for practical applications.

Currently, rare works were reported for improving the electrochemical performance of FeNb11O29. Zheng et al. increased the specific capacity of FeNb11O29 from 226 to 273 mAh g−1 at 0.1C by fabricating FeNb11O29 nanotubes, and the initial-cycle Coulombic efficiency was improved from 75.9 to 90.1% [20]. Lou et al. prepared Cr3+-doped FeNb11O29 (Cr0.2Fe0.8Nb11O29) with a specific capacity of 254 mAh g−1 [24]. Lou et al. further enhanced the specific capacity of FeNb11O29 to 270 mAh g−1 through generating oxygen vacancies [25]. In electrochemical experiments, there are common methods used to improve the electrochemical performance of electrode materials, such as carbon composite [26,27,28], plasmonization [29, 30], surfactant-assisted process [31], and nanosizing [32, 33]. It is believed that doping is also a common and efficient method to improve the cycling performance and rate capability of electrode materials because of its easy operation, low-consumption, and large-scale production [34,35,36,37,38,39,40,41,42,43,44,45]. Aluminum (Al) has been paid more and more attention as a substitute for various electrode materials owning to its abundant in earth, nontoxic and light characteristics [46, 47]. It has been found that Al doping could enhance the electrochemical performance of anode materials significantly because the strong Al–O bonds favor the structural stability [48,49,50]. Therefore, in this study, FeNb11O29 materials were also modified by doping with Al. The experimental results showed that Al3+ successfully and partically replaced Fe3+ and that the obtained Al0.2Fe0.8Nb11O29 materials preserved the crystal structure, grain size and morphology of FeNb11O29. Better structural stability was achieved due to the strong Al–O bonds, which greatly enhanced the electrochemical performance of FeNb11O29.

2 Experimental

2.1 Material synthesis

The AlxFe1-xNb11O29 (x = 0.1, 0.2, and 0.3) materials were synthesized via a simple solid-state reaction method by using Nb2O5 (Sinopharm, 99.9%), Al2O3 (Aladdin, 99.0%), and Fe2O3 (Aladdin, 99.0%) as the raw materials. Nb2O5, Fe2O3, and different contents of Al2O3 were mixed in ethanol, which were then ground in a high-energy ball-milling machine (SPEX 8000 M, USA) for 4 h. After drying, the ball-milled mixtures were sintered in a muffle furnace at 1300 °C in air for 4 h, forming AlxFe1-xNb11O29 materials (x = 0.1, 0.2, and 0.3). For comparison, undoped FeNb11O29 counterpart was also fabricated with a similar procedure and without the use of Al2O3.

2.2 Material characterization

The crystalline structures of the obtained samples were characterized by X-ray diffraction (XRD) on an X-ray diffractometer (German Bruker D8). The phase purity and lattice parameters were determined by Rietveld refinements, which were conducted by using the GSAS program with the EXPGUI interface [51, 52]. Morphologies, particle sizes, and microstructures of synthesized samples were recorded by scanning electron microscopy (SEM, S-4800), high-resolution transmission electron microscopy (HRTEM, Tecnai G2 F20 S-TWIN, FEI), and energy dispersive X-ray spectrometer (EDX) mapping. The specific surface areas of the samples were determined by nitrogen adsorption/desorption device (ASAP 2460).

2.3 Half-cell measurement

The electrochemical performance of AlxFe1-xNb11O29/Li (x = 0, 0.1, and 0.2) was evaluated by employing CR2016-type coin cells. Firstly, the working electrodes were prepared by mixing 70 wt% of AlxFe1-xNb11O29 (active material) powder, 20 wt% of conductive carbon (Super P, battery level), and 10 wt% of polyvinylidene fluoride (PVDF) binder, thoroughly mixed in N-methylpyrrolidone (NMP) to form a homogeneous slurry. After stirring the slurry for 8 h, the slurry was coated onto a copper foil evenly. The mass loading of AlxFe1-xNb11O29 was ~ 1.0 mg cm−2. Then, the coated foils were dried in a vacuum oven at 120 °C for 10 h and were cut into a circular plate with a diameter of 10 mm. The CR2016 coin cells were assembled in a glove box filled with argon gas (O2/H2O < 0.1 ppm) and consisted of an above as-prepared electrode, a Li foil that acted as counter and reference electrodes, a microporous polypropylene film (Celgard 2325) and electrolyte containing 1 M LiPF6 (DAN VEC) in a mixed solvent of ethylene carbonate (EC), dimethyl carbonate (DMC), and diethylene carbonate (DEC) with a volume ratio of 1:1:1. Galvanostatic discharge–charge measurements and galvanostatic intermittent titration technique (GITT) tests were performed on a Neware battery tester. Cyclic voltammetry (CV) measurements were conducted on a CHI660E electrochemical workstation. The electrochemical performance was examined within 3.0–0.8 V. To prepare the AlxFe1-xNb11O29 electrodes for ex situ XRD, the half cells at different states of discharge/charge were disassembled in the glove box, followed by washing the obtained AlxFe1-xNb11O29 electrodes by dimethyl carbonate.

2.4 Full-cell measurement

Full cells (CR2032-type coin cells) were assembled with the as-prepared Al0.2Fe0.8Nb11O29 as the anode and commercial LiFePO4 as the cathode. Similarly, LiFePO4 powder, conductive carbon (Super P), and PVDF with a weight ratio of 7:2:1 were thoroughly mixed in N-methylpyrrolidone (NMP) to form a homogeneous slurry. Then, the slurry was spread evenly on an Al foil and made into a circular plate with a diameter of 10 mm. The LiFePO4:Al0.2Fe0.8Nb11O29 weight ratio was fixed to be 2:1. Galvanostatic discharge–charge measurements were performed within 1.0–2.5 V.

3 Results and discussion

Figure 1 shows the XRD patterns of FeNb11O29 and AlxFe1-xNb11O29 (x = 0.1, 0.2, and 0.3), and the corresponding Rietveld-refined data are demonstrated in Table S1. It can be seen that the XRD patterns of AlxFe1-xNb11O29 (x = 0.1 and 0.2) completely match those of FeNb11O29 with an orthorhombic shear ReO3 crystal structure and Amma space group (JCPDS#22–0352; Fig. S1). No other phases (such as NbxOy or Fe2O3) are observed. Table S1 shows that the weighted profile residual (Rwp) values of the samples are small (about 10%), revealing that the Rietveld refinement results are credible. The fractional atomic parameters of FeNb11O29 and Al0.2Fe0.8Nb11O29 are shown in Tables S2 and S3, respectively. Since the ionic radius of Al3+ is smaller than Fe3+, it is reasonable that the lattice parameters get smaller [45, 53]. It is noteworthy that the b values of AlxFe1-xNb11O29 (x = 0.1 and 0.2) are larger than that of the previous Ti2Nb10O29 (b = 3.805 Å) reported [54], suggesting the existence of larger gap sites in AlxFe1-xNb11O29 that are more conducive to Li+ diffusion. However, when the Al3+ doping amount further increases (x = 0.3), the crystal structure of orthorhombic shear ReO3 changes, indicating that the allowable Al3+ doping amount is limited (i.e., x ≤ 0.2; Fig. 1) [53, 55].

Fig. 1
figure 1

XRD patterns and Rietveld refinement results of FeNb11O29 and AlxFe1-xNb11O29 (x = 0.1, 0.2, and 0.3) materials

Figure 2 illustrates the SEM images of FeNb11O29 and AlxFe1-xNb11O29 (x = 0.1, 0.2, and 0.3). It can be observed that all the samples consist of grains which possess a platelet morphology. The BET-specific surface area of Al0.2Fe0.8Nb11O29 is 0.27 m2 g−1 (Fig. S2b), which is slightly smaller than FeNb11O29 (0.24 m2 g−1; Fig. S2a), verifying that Al0.2Fe0.8Nb11O29 and FeNb11O29 have similar particle sizes. Thus, there are almost little influences on the grain morphology and size of FeNb11O29 through the Al3+ doping.

Fig. 2
figure 2

SEM images of a FeNb11O29, b Al0.1Fe0.9Nb11O29, c Al0.2Fe0.8Nb11O29, and d Al0.3Fe0.7Nb11O29

Figure 3 depicts the HRTEM images of FeNb11O29 (Fig. 3a) and Al0.2Fe0.8Nb11O29 (Fig. 3b), revealing their lattice fringes of 0.351 and 0.349 nm, respectively, which correspond to (311) planes. The interplanar spacing contraction in Al0.2Fe0.8Nb11O29 may be due to the smaller ionic radius of Al3+ in comparison with Fe3+. The results further verify that the crystal structure of FeNb11O29 does not change after doping with a suitable amount of Al3+, which are also consistent with the XRD characterization. The EDS elemental mapping images shown in Fig. 3c, d indicate that Fe, Nb, and O elements, and Al, Fe, Nb, and O elements, are uniformly distributed in FeNb11O29 and Al0.2Fe0.8Nb11O29 materials, respectively, further confirming the successful doping of Al3+ into FeNb11O29.

Fig. 3
figure 3

HRTEM images of a FeNb11O29 and b Al0.2Fe0.8Nb11O29. EDX elemental mapping images of c FeNb11O29 and d Al0.2Fe0.8Nb11O29

In order to understand the electrochemical mechanism of FeNb11O29 and Al0.2Fe0.8Nb11O29, CV tests with different scanning speeds were carried out on the FeNb11O29 and Al0.2Fe0.8Nb11O29-made electrodes under a potential window of 3.0–0.8 V. Figure 4a, b show the first four-cycle CV curves of these two electrodes tested at 0.2 mV s−1. The redox peak of the FeNb11O29 electrode can be attributed to Nb3+/Nb4+, Nb4+/Nb5+, and Fe2+/Fe3+ redox pairs (Fig. 4a). In particular, the Nb3+/Nb4+ peak is at ~ 1.10/1.30 V, the Nb4+/Nb5+ peak is at ~ 1.53/1.72 V, and the Fe2+/Fe3+ peak is at ~ 2.36 V. The average working potential of FeNb11O29 is calculated to be ~ 1.61 V. For the Al0.2Fe0.8Nb11O29 electrode (Fig. 4b), the Nb3+/Nb4+ peak is at ~ 1.10/1.30 V, the Nb4+/Nb5+ peak is at ~ 1.59/1.72 V, and the Fe2+/Fe3+ peak is at ~ 2.35 V. The average working potential of Al0.2Fe0.8Nb11O29 is also ~ 1.61 V. Clearly, Al0.2Fe0.8Nb11O29 and FeNb11O29 have a very similar CV characteristic, hinting the similar reaction kinetics.

Fig. 4
figure 4

The first four CV cycles of a FeNb11O29 electrode and b Al0.2Fe0.8Nb11O29 electrode at 0.2 mV s−1. The CV curves of c FeNb11O29 electrode and d Al0.2Fe0.8Nb11O29 electrode at different scan rates

Figure 4c, d illustrate the CV curves of FeNb11O29 and Al0.2Fe0.8Nb11O29 electrodes at different scanning speeds, respectively. It can be seen that with the increase of the scanning speeds, the current values of the redox front increase in both FeNb11O29 and Al0.2Fe0.8Nb11O29 electrodes. However, compared with FeNb11O29, Al0.2Fe0.8Nb11O29 has a stronger redox front, revealing that Al3+ doping enables better electrochemical-kinetics behavior.

The constant current discharge–charge curves of the FeNb11O29 and Al0.2Fe0.8Nb11O29 electrodes are demonstrated in Fig. 5a, b, respectively. Each discharge–charge curve was divided into three regions: a short slope at 3.0–1.7 V, a short platform at 1.7–1.6 V, and a long slope at 1.6–0.8 V. Solid solution reactions occur in the first and third regions, and a two-phase reaction occurs in the second region. It can be found that the FeNb11O29 electrode has a reversible capacity of 262 mAh g−1 with an initial-cycle Coulombic efficiency of 91.1%, while the Al0.2Fe0.8Nb11O29 electrode has an enhanced reversible capacity of 318 mAh g−1 with an initial-cycle Coulombic efficiency of 95.0%. Table 1 lists the electrochemical performance of the Al0.2Fe0.8Nb11O29 and M-Nb–O materials reported in literatures. It is clear that the capacity of Al0.2Fe0.8Nb11O29 is higher than those of most reported M-Nb–O materials (except for MoNb12O33 and Mg2Nb34O87), and the initial-cycle Coulombic efficiency is only lower than GaNb11O29. It is believed that the relatively high working potential of Al0.2Fe0.8Nb11O29 (~ 1.61 V) results in the less formation of SEI layer during the electrochemical reaction process, which reduces the consumption of lithium ions and leads to the high initial-cycle Coulombic efficiency [7].

Fig. 5
figure 5

Electrochemical performance of FeNb11O29 and Al0.2Fe0.8Nb11O29 electrodes: a, b discharge–charge curves at 0.1C, c, d discharge–charge curves at 0.1–10C, e rate capabilities, and f cycling stability at 10C after 1000 cycles and Coulombic efficiency of Al0.2Fe0.8Nb11O29

Table 1 Electrochemical performance of Al0.2Fe0.8Nb11O29 compared with M-Nb–O materials reported

Figure 5c, d show the discharge–charge curves of the FeNb11O29 and Al0.2Fe0.8Nb11O29 electrodes at different current density, respectively. When the current density increases from 0.1 to 10C, the specific capacity of FeNb11O29 decreases from 262 to 104 mAh g−1, while that of Al0.2Fe0.8Nb11O29 electrode decreases from 318 to 134 mAh g−1. Figure 5e demonstrates the rate capabilities of the FeNb11O29 and Al0.2Fe0.8Nb11O29 electrodes at various current densities. Both the specific capacities of the FeNb11O29 and Al0.2Fe0.8Nb11O29 electrodes reduce with increasing the current density. It can also be found that the specific capacity of the Al0.2Fe0.8Nb11O29 electrode at 10C is even higher than the FeNb11O29 electrode at 5C. Therefore, the Al0.2Fe0.8Nb11O29 electrode has achieved a better rate capability. This high rate capability is due to the fact that Al0.2Fe0.8Nb11O29 has a stronger redox front. Moreover, Fig. S3a, b illustrate the discharge–charge curves of the Al0.1Fe0.9Nb11O29 electrode. The Al0.1Fe0.9Nb11O29 electrode also exhibits an improved specific capacity (272 mAh g−1) in comparison with the undoped FeNb11O29 electrode, but lower than that of the Al0.2Fe0.8Nb11O29 electrode. This result indicates that more Al3+ doping has a more significant effect on improving the reversible capacity of FeNb11O29.

Figure 5f illustrates the cycling performance of the FeNb11O29 and Al0.2Fe0.8Nb11O29 electrodes at a current density of 10C. After 1000 cycles, the capacity retention rate of Al0.2Fe0.8Nb11O29 is 92.9%, which is obviously higher than that of FeNb11O29 (85.8%), proving that Al3+ doping is very beneficial to the cycling stability of FeNb11O29. The excellent cycling stability is due to the fact that Al0.2Fe0.8Nb11O29 owns excellent structural stability.

To study the Li+ diffusivity of the FeNb11O29 and Al0.2Fe0.8Nb11O29 electrodes, the GITT tests were employed at room temperature (Fig. 6). Figure 6a, b record the initial GITT curves of the FeNb11O29 and Al0.2Fe0.8Nb11O29 electrodes at 0.1C, respectively. The diffusion coefficient of Li+\(({D}_{L{i}^{+}})\) can be calculated according to Fick’s second law (Eq. (1)):

Fig. 6
figure 6

GITT initial curves of a FeNb11O29 electrode and b Al0.2Fe0.8Nb11O29 electrode tested at 0.1C. Calculated \({D}_{L{i}^{+}}\) from GITT curves of c FeNb11O29 electrode and d Al0.2Fe0.8Nb11O29 electrode

$$D_{Li^+}=\frac4\pi\left(\frac{m_BV_m}{M_BS}\right)^2\left(\frac{\Delta E_s}{\tau\left(dE_\tau/d\sqrt\tau\right)}\right)^2(\tau<<\frac{L^2}{D^2})$$
(1)
$$D_{Li^+}=\frac {4}{\pi\tau} \left(\frac{m_{B}V_{m}}{M_{B}S}\right)^2\left(\frac{\Delta E_{s}}{\Delta E_{\tau}}\right)^2 (\tau << \frac{L^2}{D^2})$$
(2)

where MB and mB are the molar mass and unit mass of the active material, respectively; Vm stands the molar volume of the active material; S stands the polar area; τ is titration time; L stands for the diffusion length; and ΔES and ΔEτ represent the change of equilibrium potential and the change in potential during a single-step titration (Fig. S4a, b), respectively. Since there is a linear relationship between τ0.5 and the potential during the single-step titration (Fig. S4c, d), Eq. (1) can be simplified as Eq. (2). Figure 6c, d describe how the calculated \({D}_{\text{L}{\text{i}}^{+}}\) values vary with the potential. During the Li+ insertion process, the calculated \({D}_{\text{L}{\text{i}}^{+}}\) values of the FeNb11O29 electrode are 8.03 × 10−13–3.71 × 10−11cm2 s−1 with an average value of 1.47 × 10−11 cm2 s−1, while the Al0.2Fe0.8Nb11O29 electrode has calculated \({D}_{\text{L}{\text{i}}^{+}}\) values of 1.67 × 10−12–3.41 × 10−11 cm2 s−1 with an average value of 1.76 × 10−11 cm2 s−1. For the Li+ extraction process, the calculated \({D}_{\text{L}{\text{i}}^{+}}\) values of the FeNb11O29 electrode are 3.95 × 10−12–1.83 × 10−11 cm2 s−1 with an average value of 1.28 × 10−11 cm2 s−1, while the Al0.2Fe0.8Nb11O29 electrode has calculated \({D}_{\text{L}{\text{i}}^{+}}\) values of 7.24 × 10−12–2.60 × 10−11 cm2 s−1 with an average value of 1.64 × 10−11 cm2 s−1. These similar Li+ diffusivity can be explained by the similar lattice parameters and unit-cell volumes of the two materials (Table S1).

The structural stability of Al0.2Fe0.8Nb11O29 during Li+ insertion-extraction processes was studied by XRD. Figure 7 demonstrates the XRD results of the fresh Al0.2Fe0.8Nb11O29 electrode and the Al0.2Fe0.8Nb11O29 electrodes after discharged to 0.8 V as well as charged to 3.0 V in 1st cycle and 10th cycles at 0.1C. It can be found that after the discharge–charge process, no impurity phases (such as Al2O3, Fe2O3, or NbxOy) appear in the XRD patterns, and no obvious changes are observed in the peak intensities. The XRD patterns of the FeNb11O29 electrodes after the similar discharge–charge process are shown in Fig. S5. Very similar results are obtained. Therefore, Al3+ doping does not change the crystal structure of FeNb11O29 material indeed, and the Al0.2Fe0.8Nb11O29 framework is very stable during the electrochemical reaction.

Fig. 7
figure 7

Ex situ XRD patterns of fresh Al0.2Fe0.8Nb11O29 electrode and Al0.2Fe0.8Nb11O29 electrodes after discharged to 0.8 V in 1st cycle, charged to 3.0 V in 1st and 10th cycles at 0.1C

To assess Al0.2Fe0.8Nb11O29 practical application value, we prepared a LiFePO4/Al0.2Fe0.8Nb11O29 full cell with LiFePO4 as the cathode material and Al0.2Fe0.8Nb11O29 as the anode material. Figure 8a shows the discharge–charge curves of the full cell at 0.1C. It delivered initial-cycle charge/discharge capacities of 241/206 mAh g−1, respectively. Its initial-cycle Coulombic efficiency was 85%. This result is better than the previously reported LiNi0.5Mn1.5O4/AlNb11O29 full cell with charge/discharge capacities of only 238/195 mAh g−1 and an initial-cycle Coulombic efficiency of 82% [13]. Figure 8b, c show the discharge–charge curves and rate capability of the LiPFeO4/Al0.2Fe0.8Nb11O29 full cell at different current density from 0.5 to 5C. The discharge capacities are 172, 152, 125, and 83 mAh g−1 at 0.5, 1, 2, and 5C, respectively. In particular, compared with the original current density at 0.5C, there is no significant decrease appeared in the rate-capability curve when the current density is from 5C back to 0.5C (Fig. 8c). It indicates that the full cell has good electrochemical reversibility. Figure 8d, e show that the capacity retention of the full cell reaches 84.2% after 200 cycles at 1C, and 89.8% after 1000 cycles at 5C, respectively, demonstrating good cycling stability.

Fig. 8
figure 8

Electrochemical performance of the LiFePO4/Al0.2Fe0.8Nb11O29 full cell: a discharge–charge curves at 0.1C, b discharge–charge curves at 0.1-5C, c rate capabilities, d cycling stability at 1C over 200 cycles and Coulombic efficiency, and e cycling stability at 10C over 1000 cycles and Coulombic efficiency

4 Conclusions

In summary, the electrochemical performance (especially the specific capacity) of FeNb11O29 is improved significantly by doping with Al3+. A reasonable doping amount can preserve the crystal structure, grain size, and morphology of FeNb11O29. Al0.2Fe0.8Nb11O29 shows improved electrochemical performance, with a reversible capacity of 318 mAh g−1 at 0.1 C, high initial-cycle Coulombic efficiency of 95.0%, and relatively high and safe working potential of about 1.61 V. Al0.2Fe0.8Nb11O29 further possesses excellent cycling stability with capacity retention of 92.9% over after 1000 cycles. In addition, we assembled a LiFePO4/Al0.2Fe0.8Nb11O29 full cell, which also show excellent electrochemical performance. The discharge capacity is up to 206 mAh g−1 at 0.1C. The capacity retention of the full cell reaches 84.2% after 200 cycles at 1C, and 89.8% after 1000 cycles at 5C. Therefore, Al0.2Fe0.8Nb11O29 may find practical applications in LIBs.