metal-organic framework-derived porous shuttle-like …...metal-organic frameworks (mofs), which...

15
Metal-organic framework-derived porous shuttle-like vanadium oxides for sodium-ion battery application Yangsheng Cai 1 , Guozhao Fang 1 , Jiang Zhou 1,2 ( ), Sainan Liu 1 , Zhigao Luo 1 , Anqiang Pan 1,2 ( ), Guozhong Cao 3 , and Shuquan Liang 1,2 ( ) 1 School of Materials Science and Engineering, Central South University, Changsha 410083, China 2 Key Laboratory of Nonferrous Metal Materials Science and Engineering, Ministry of Education, Central South University, Changsha 410083, China 3 Department of Materials and Engineering, University of Washington, Seattle, WA 98195-2120, USA Received: 7 November 2016 Revised: 25 April 2017 Accepted: 30 April 2017 © Tsinghua University Press and Springer-Verlag GmbH Germany 2017 KEYWORDS vanadium oxides, metal-organic frameworks, porous structure, density functional theory (DFT) calculation, sodium-ion batteries ABSTRACT Vanadium oxides with a layered structure are promising candidates for both lithium-ion batteries and sodium-ion batteries (SIBs). The self-template approach, which involves a transformation from metal-organic frameworks (MOFs) into porous metal oxides, is a novel and effective way to achieve desirable electrochemical performance. In this study, porous shuttle-like vanadium oxides (i.e., V 2 O 5 , V 2 O 3 /C) were successfully prepared by using MIL-88B (V) as precursors with a specific calcination process. As a proof-of-concept application, the as- prepared porous shuttle-like V 2 O 3 /C was used as an anode material for SIBs. The porous shuttle-like V 2 O 3 /C, which had an inherent layered structure with metallic behavior, exhibited excellent electrochemical properties. Remarkable rate capacities of 417, 247, 202, 176, 164, and 149 mAh·g 1 were achieved at current densities of 50, 100, 200, 500, 1,000, and 2,000 mA·g 1 , respectively. Under cycling at 2 A·g 1 , the specific discharge capacity reached 181 mAh·g 1 , with a low capacity fading rate of 0.032% per cycle after 1,000 cycles. Density functional theory calculation results indicated that Na ions preferred to occupy the interlamination rather than the inside of each layer in the V 2 O 3 . Interestingly, the special layered structure with a skeleton of dumbbell-like V–V bonds and metallic behavior was maintained after the insertion of Na ions, which was beneficial for the cycle performance. We consider that the MOF precursor of MIL-88B (V) can be used to synthesize other porous V-based materials for various applications. 1 Introduction With the rapid development of the energy industry, lithium-ion batteries (LIBs) have been widely applied e.g., in portable electronics and electric vehicles [1–14]. However, their use has been restricted as a result of Nano Research 2018, 11(1): 449–463 https://doi.org/10.1007/s12274-017-1653-9 Address correspondence to Jiang Zhou, [email protected]; Anqiang Pan, [email protected]; Shuquan Liang, [email protected]

Upload: others

Post on 22-Aug-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

Metal-organic framework-derived porous shuttle-like vanadium oxides for sodium-ion battery application

Yangsheng Cai1, Guozhao Fang1, Jiang Zhou1,2 (), Sainan Liu1, Zhigao Luo1, Anqiang Pan1,2 (),

Guozhong Cao3, and Shuquan Liang1,2 ()

1 School of Materials Science and Engineering, Central South University, Changsha 410083, China 2 Key Laboratory of Nonferrous Metal Materials Science and Engineering, Ministry of Education, Central South University, Changsha

410083, China 3 Department of Materials and Engineering, University of Washington, Seattle, WA 98195-2120, USA

Received: 7 November 2016

Revised: 25 April 2017

Accepted: 30 April 2017

© Tsinghua University Press

and Springer-Verlag GmbH

Germany 2017

KEYWORDS

vanadium oxides,

metal-organic frameworks,

porous structure,

density functional theory

(DFT) calculation,

sodium-ion batteries

ABSTRACT

Vanadium oxides with a layered structure are promising candidates for both

lithium-ion batteries and sodium-ion batteries (SIBs). The self-template approach,

which involves a transformation from metal-organic frameworks (MOFs)

into porous metal oxides, is a novel and effective way to achieve desirable

electrochemical performance. In this study, porous shuttle-like vanadium oxides

(i.e., V2O5, V2O3/C) were successfully prepared by using MIL-88B (V) as precursors

with a specific calcination process. As a proof-of-concept application, the as-

prepared porous shuttle-like V2O3/C was used as an anode material for SIBs. The

porous shuttle-like V2O3/C, which had an inherent layered structure with metallic

behavior, exhibited excellent electrochemical properties. Remarkable rate capacities

of 417, 247, 202, 176, 164, and 149 mAh·g−1 were achieved at current densities of

50, 100, 200, 500, 1,000, and 2,000 mA·g−1, respectively. Under cycling at 2 A·g−1, the

specific discharge capacity reached 181 mAh·g−1, with a low capacity fading

rate of 0.032% per cycle after 1,000 cycles. Density functional theory calculation

results indicated that Na ions preferred to occupy the interlamination rather

than the inside of each layer in the V2O3. Interestingly, the special layered

structure with a skeleton of dumbbell-like V–V bonds and metallic behavior

was maintained after the insertion of Na ions, which was beneficial for the cycle

performance. We consider that the MOF precursor of MIL-88B (V) can be used

to synthesize other porous V-based materials for various applications.

1 Introduction

With the rapid development of the energy industry,

lithium-ion batteries (LIBs) have been widely applied

e.g., in portable electronics and electric vehicles [1–14].

However, their use has been restricted as a result of

Nano Research 2018, 11(1): 449–463

https://doi.org/10.1007/s12274-017-1653-9

Address correspondence to Jiang Zhou, [email protected]; Anqiang Pan, [email protected]; Shuquan Liang, [email protected]

Page 2: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

| www.editorialmanager.com/nare/default.asp

450 Nano Res. 2018, 11(1): 449–463

the limited Li resources and the increasing prices of

raw materials [15]. Sodium-ion batteries (SIBs) have

become increasingly attractive because of the natural

abundance of Na and their low cost and similar

electrochemical behavior to LIBs [15–19]. However,

the radius of a Na atom (1.02 Å) is larger than that

of a Li atom (0.76 Å), which significantly limits the

electrochemical performance of SIBs, yielding a poor

cycling stability, low specific capacity, and low diffusion

coefficient [15, 20–23]. Thus, developing desirable

electrode materials for SIBs is a great challenge.

Vanadium oxides are promising candidates for SIBs

because of their safety, high specific capacity, natural

abundance, and low cost [24–34]. Among them,

vanadium pentoxide (V2O5), which has a promising

layered structure, has attracted widespread inves-

tigations as a cathode for SIBs [28–30, 35–44]. Ji et al.

fabricated novel V2O5/C nanoparticles that exhibited

an amazing rate capability (92 mA·h·g−1 at 640 mA·g−1)

as a cathode material for an SIB [28]. The investigation

of new anode materials for SIBs is urgent, as there

is currently no commericial anode. Vanadium

sesquioxides (V2O3), which has good ion intercalation

characteristics, has been widely investigated as an

anode material for LIBs [32, 33, 45–51]. Vanadium

oxides with higher valence V ion are highly toxic,

indicating that V2O3 is far less harmful than others

[33, 46, 47, 50]. Additionally, because the electrons in

the V-3d orbit can travel along the V–V chains in the

V2O3 electronic structure, the rhombohedral corundum-

type vanadium oxide exhibits metallic behavior

[47, 50, 52]. This was verified via density functional

theory (DFT) calculations in the present work. The

calculated density of states (DOS) for V2O3 is shown

in Fig. S1 in the Electronic Supplementary Materials

(ESM). There is an obvious electronic state at the Fermi

energy level, indicating that pure V2O3 is metallic

and has a good electron-transfer capability [53–56].

The metallic states are attributed to the V d-orbital

and O p-orbital, in accordance with previous reports

[50, 52]. The electrical resistance of bulk V2O3 is

approximately 10 Ω, which is far smaller than that of

other transition-metal oxides [46, 57, 58]. The intrinsic

property endows V2O3 with intriguing electrochemical

performance. For example, Mai et al. reported that

V2O3 microspheres exhibited outstanding long-life

performance as LIBs anodes [32]. However, until now,

reports concerning the application of V2O3 in SIBs have

been very limited [31].

Metal-organic frameworks (MOFs), which comprise

metal ions (or metal clusters) and organic linkers, have

a large surface area, chemical stability, and porous

properties [59–64]. They have been considered as

desirable self-templates for constructing advanced

energy materials with porous nanostructures [6, 65–70].

For example, Co3O4 hollow dodecahedra [71], Fe2O3

microboxes [72], CuO hollow octahedra [73], and

multilayer CuO@NiO hollow spheres [67] derived from

suitable MOFs exhibited enhanced electrochemical

performance in LIB applications. Our group recently

demonstrated that the three-dimensional (3D) Co3O4

Ni foam hybrid derived from Co MOF exhibited

excellent electrochemical performance, with a superior

long-term cyclic stability up to 2,000 cycles at 5 and

20 A·g−1 [6]. A Co3O4 porous hollow structure [74],

Cr2O3 nanoribbons [75], and r-GO wrapped MoO3 [69]

were prepared by employing MOFs as precursors

and exhibited excellent electrochemical performance as

electrode materials for supercapacitors. For improving

the electrochemical performance of SIBs, materials with

a high porosity and expanded interlayered structure

have been extensively investigated [22, 23, 76–79].

MOF-derived transition-metal oxides are promising

candidates as electrode materials for SIBs. However,

to our knowledge, there are no reports on the synthesis

of highly porous vanadium oxides using V-MOF as a

template.

Herein, we describe a self-template approach for

transforming MIL-88B (V) into porous shuttle-like

vanadium oxides (i.e., V2O5, V2O3) via a specific heating

treatment in different atmospheres. The MOF-derived

mechanism was investigated via a comprehensive

material characterization. The porous shuttle-like

C-coated V2O3 (V2O3/C) composites were easily produced

via the thermal reduction of the as-prepared MIL-88B

(V) in an Ar atmosphere. To test their electrochemical

performance, the porous shuttle-like V2O3/C were

applied as anode materials for SIBs, exhibiting an

outstanding rate capability and long-term cycle stability.

Importantly, first-principles calculations were employed

to investigate the insertion behaviors of Na ions

into V2O3.

Page 3: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

451 Nano Res. 2018, 11(1): 449–463

2 Experimental

2.1 Material synthesis

MIL-88B (V) was prepared according to the literature

[80]. In a typical procedure, 314 mg of VCl3 and 332 mg

of terephthalic acid (H2BDC) were mixed in a 50-mL

beaker. Then, a solution containing 2 mL of HCl

(1 mol·L–1) and 10 mL of absolute ethyl alcohol was

added under vigorous stirring at room temperature.

After 30 min, the resulting blue suspension was further

dispersed with ultrasonic treatment for 15 min. Finally,

the mixture was transferred to a 50-mL Teflon-lined

stainless-steel autoclave and heated in an electrical oven

at 120 °C for 2 day. The green powder, i.e., V-MOFs,

were obtained via centrifugation with ethanol and dried

in vacuum at 50 °C for 8 h. To obtain porous shuttle-

like V2O3/C, the V-MOFs were annealed in an Ar atmo-

sphere at 700 °C for 6 h, with a heating rate of 2 °C·min−1.

Additionally, porous V2O5 with a hollow shuttle-like

structure was prepared via heat treatment of V-MOFs

in air at 350 °C for 4 h in a heating rate of 2 °C·min−1.

2.2 Material characterization

The crystallographic phases of the specimens were

determined using a Rigaku D/max 2500 X-ray

diffractometer with Cu Kα radiation. Field-emission

scanning electron microscopy–energy dispersive X-ray

analysis (FESEM–EDS, FEI Nova Nano-SEM 230) was

performed to record the morphology of V2O3/C and

V2O5. The structures of the vanadium oxides were

analyzed via transmission electron microscopy (TEM,

JEOL-JEM-2100F transmission electron microscope). A

NOVA 4200e instrument (Quantachrome Instruments)

was used to measure the Brunauer–Emmett–Teller

(BET) surface area and pore distribution of the solids.

An ESCALAB 250Xi was employed for X-ray photo-

electron spectroscopy (XPS). A C–S Analyzer (CS-444),

inductively coupled plasma atomic emission spec-

troscopy, and a conductivity instrument (D60K, Made

in China) were utilized to measure the C content, the

content of trace elements and the conductivity of the

products, respectively.

2.3 Electrochemical measurements

CR 2016 coin cells were assembled in a professional

glove box (Mbraun, Garching, Germany) to test the

electrochemical properties of the active materials. A

viscous slurry containing 70 wt.% V2O3/C, 20 wt.%

acetylene black, and 10 wt.% polyvinylidene fluoride

binder in N-methyl-2-pyrrolidone solution was

stirred for 12 h and then coated on Cu foil. Then, the

electrodes were dried in a vacuum oven at 100 °C

for 12 h. Na half cells were constructed by using the

aforementioned electrodes as cathodes, Na metal plates

as counter electrodes, and glass fiber as a separator. A

commercial electrolyte comprising 1 M NaClO4 and

ethylene carbonate/dimethyl carbonate (1:1, v/v) was

used. Cyclic voltammetry (CV) and galvanostatic

charge/discharge measurements were performed using

a CHI604E electrochemical workstation (Chenhua,

Shanghai, China) and a Land battery tester (CT 2001A,

China). Typically, the voltage range is set as 0.01–3 V

(vs. Na/Na+), and the specific capacity and current

density are based on the mass of active materials

(V2O3/C) only.

3 Results and discussion

The porous shuttle-like vanadium oxides were

synthesized via a self-template approach with various

calcination processes (as shown in Scheme 1). An X-ray

diffraction (XRD) pattern of as-prepared MIL-88B (V)

is shown in Fig. S2 in the ESM. The diffraction peaks

are located in a low-angle range, which is consistent

with a previous report [80], indicating the successful

synthesis of MIL-88B (V). As shown in Figs. S3(a) and

S3(b) in the ESM, the MOFs exhibited a prismatic

morphology with a pyramid-like top, with lengths

and widths of approximately 10 and 1 μm, respectively.

Scheme 1 Illustration of preparation process for the shuttle-like porous V2O3/C and V2O5.

Page 4: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

| www.editorialmanager.com/nare/default.asp

452 Nano Res. 2018, 11(1): 449–463

The TEM images of the MIL-88B (V) shown in Figs. S3(c)

and S3(d) in the ESM confirmed the shuttle-like

structure.

An XRD pattern of the black compound, which was

annealed in an Ar atmosphere, is shown in Fig. 1(a).

Clearly, the patterns can be indexed to the V2O3 phase

(JCPDS card No. 34-0187), without a second impurity

phase. This indicates that vanadium trioxide was

successfully derived from MIL-88B (V). According

to the previous crystallographic data, V2O3 belongs

to the rhombohedral lattice, and its space group is     

R3_

c (167). The lattice parameters are a = 4.9525 (4) Å,

b = 4.9525 (4) Å, c = 14.0038 (2) Å, α = 90°, β = 90°, and

γ = 120°. An optimized 2 × 2 × 2 supercell of V2O3 is

depicted in Fig. 1(b). The 3D structure can be regarded

as a periodic glide of identical layers, i.e., a (012) crystal

face, along the cell diagonal. A tunneled structure com-

prising stacked V–V dumbbells and corresponding O

atoms may exhibit stability and lead to a desirable

performance for electrochemical insertion/deinsertion

[47, 50]. The length of the dumbbell-like V–V bonds

forming the skeleton of each layer is 2.5632 (3) Å. The

horizontal distance between adjacent V–V dumbbells

is 4.530 (1) Å, which is the interlamellar spacing in the

supercell.

The morphology of V2O3 was determined via FESEM

and TEM. As shown in Figs. 2(a)–2(c), the products,

which were annealed in the Ar atmosphere, maintained

the same structure as the shuttle-like MIL-88B (V).

The only difference was the uniform nanoparticles on

Figure 1 (a) XRD pattern of the shuttle-like V2O3/C. (b) Crystal structure of V2O3 (2 × 2 × 2 supercell).

the skin of the V2O3 shuttles. Accidentally, a cracked

shuttle is observed in Fig. 2(b), indicating the porous

structure of the as-derived V2O3 micro-shuttles. The

TEM images shown in Figs. 2(d) and 2(e), especially

the magnified image (Fig. 2(e)), confirms the porous

structure of the V2O3 micro-shuttles. An HRTEM

image of the as-derived V2O3 is shown in Fig. 2(f).

Because of the uniform existence of amorphous C,

the lattice fringes are not very clear. However, the

lattice fringes with spacings of 0.365, 0.219, and 0.277

nm can be accurately indexed to the (012), (006), and

(104) crystal faces, respectively.

N2 adsorption–desorption was performed to study

the surface area and pore distribution of the V2O3

shuttles. As shown in Fig. 3(a), the isotherm curves

for the V2O3 shuttles turned the corner towards the

P/P0 axis and then maintained an approximately

horizontal shape, which is typical for a Langmuir

isotherm (belong to type I N2 adsorption–desorption

isotherms). Because of the interaction of the sorbent

and V2O3, the adsorption capacity increased rapidly

at low P/P0 values, causing the micropores in the V2O3

to be filled at an extremely low relative pressure. The

distributions of the pore diameter for the V2O3 shuttles

were analyzed via the Barrett–Joyner–Halenda (BJH)

method. The BJH curves shown in Fig. 3(b) indicate a

narrow distribution of pore diameter around 3.2 nm,

which accords with the TEM results. Additionally, it

was confirmed that the products obtained by annealing

MIL-88B (V) in an Ar atmosphere were microporous

materials. As summarized in Table S1 in the ESM, the

BET surface area and pore volume of the V2O3 shuttles

reached 258.954 m2·g−1 and 0.02702 cm3·g−1, respectively.

According to previous reports [6, 78, 81, 82], this

microporous structure facilitates the penetration of

the electrolyte and the diffusion of Na ions in the

V2O3 nanocrystals, which significantly improves the

electrochemical properties for battery applications.

XPS and EDS were performed to investigate the

formation of the as-obtained V2O3. As shown in the

survey XPS spectra (Fig. 4(a)), there are three distinct

peaks at 284.19, 517.19, and 530.19 eV, which can be

assigned to C 1s, V 2p, and O 1s, respectively. The high-

resolution XPS spectrum of C 1s in Fig. 4(b) illustrates

that the peaks at 284.79 and 286.39 eV correspond to

amorphous C and C in C–O, respectively [33, 83, 84].

Page 5: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

453 Nano Res. 2018, 11(1): 449–463

The core-level spectrum of V 2p, as shown in Fig. 4(c),

revealed the feature of V in the trivalent state, with

two peaks at 516.79 and 524.27 eV, which correspond

to V 2p3/2 and V 2p1/2, respectively [33, 83]. The results

confirm the successful preparation of V2O3/C com-

posites. The elemental distribution of the V2O3 shuttles

was investigated via EDS element-mapping analysis

(Fig. 4(d)). The results confirmed the shuttle-like

structure and the uniform distribution of the elements

V, O, and C. Additionally, the C content of the porous

shuttle-like V2O3/C was measured to be 8.07 wt.%

using a C–S analyzer (CS-444). It is generally accepted

that the electronic conductivity of electrode materials

can be significantly improved via the introduction

of C-based materials [31, 48, 85–87]. The conductivity

of the shuttle-like V2O3/C products was tested and

compared with that of V2O3 without C (the morphology

is shown in Fig. S4 in the ESM) using a conductivity

instrument (D60K, Made in China). As shown in

Table S2 in the ESM, the conductivity of the shuttle-

like V2O3/C (4.93 mS/m) was higher than that of

shuttle-like V2O3 (3.51 mS/m) and V2O3 particles

(3.47 mS/m), confirming that the self-template C in

shuttle-like V2O3 can improve the conductivity of V2O3.

Furthermore, previous investigations demonstrated

that the porous C derived from MOFs improved the

electrochemical performance [88–91]. For example,

Zhang et al. [88] prepared V2O5 embedded in ZIF-67-

derived porous dodecahedron-shaped C, which

exhibited an enhanced high-rate capability. In present

study, as shown in Fig. 4(d), the V2O3 derived from

MIL-88B (V) was uniformly distributed in the self-

templated framework of porous C, which enhances

the conductivity. The formation mechanism of the

porous V2O3/C composites can be deduced as follows.

When MIL-88B (V) is heated in an Ar atmosphere

at 700 °C, the C enters the reduction state with the

carbonization of the organic frameworks, giving rise

Figure 2 (a)–(c) FESEM images; (d) and (e) TEM images; (f) HRTEM image of the shuttle-like V2O3/C.

Figure 3 (a) N2 adsorption–desorption isotherms and (b) corresponding BJH pore-size distribution curves for the shuttle-like V2O3/C.

Page 6: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

| www.editorialmanager.com/nare/default.asp

454 Nano Res. 2018, 11(1): 449–463

to numerous micropores. At this time, the V, which

maintains the +3 oxidation state, forms V–O bonds

with the O in the pristine MIL-88B (V). Finally, the

porous shuttle-like V2O3 with a uniform C coating

has been successfully produced. The in situ coated C

may improve the electronic conductivity of the V2O3

material.

Vanadium pentoxide and other C materials can

also be derived from the multipurpose MIL-88B (V).

For example, the XRD pattern of the yellow powders

is indexed to the pure V2O5 phase (space group: Pmmn

(59), a = 11.512 Å, b = 3.564 Å, c = 4.368 Å, JCPDS card

No. 77-2418) [29, 36, 38, 92], as shown in Fig. S5 in

the ESM, confirming that vanadium pentoxide can

be obtained by annealing MIL-88B (V) in an air

atmosphere. Next, the morphology of V2O5 produced

from MIL-88B (V) was recorded. FESEM images

(Figs. 5(a)–5(c)) show that the compound derived from

the shuttle-like V-MOFs well retained its prismatic

structure, whereas there were abundant micropores on

the surface of the V2O5 shuttles (Fig. 5(c)). Interestingly,

the TEM images presented in Figs. 5(d) and 5(e)

visually demonstrate the hollow structure of the V2O5

shuttles. An HRTEM image of the hollow shuttle-like

V2O5 is shown in Fig. 5(f). The distinct lattice fringe

with a spacing of 0.406 nm can be indexed to the (101)

crystal face. Moreover, the BET test was performed to

analyze the surface area and the pore distribution of

the V2O5 shuttles. As shown in Fig. S6(a) in the ESM,

the obtained N2 absorption/desorption curves for the

V2O5 exhibited a structural characteristic of a type-IV

isotherm. The BJH curves (Fig. S6(b) in the ESM)

indicate a narrow distribution of the pore diameter

around 2.39 nm, confirming the porosity of the derived

V2O5. The formation mechanism of this porous hollow

structure can be understood as follows. During the

heat treatment in the air atmosphere, the V in the

shuttle-like MIL-88B (V) is oxidized, forming vanadium

pentoxide, which constitutes the skeleton network

in the new structure. Simultaneously, the C and H

are oxidized and released from the original structure,

resulting in the final porous hollow structure.

Figure 4 (a) Survey XPS spectrum of the as-derived V2O3/C; (b) and (c) C 1s and V 2p XPS spectra of V2O3/C, respectively; (d) FESEM image and EDS maps for V, O, and C in the shuttle-like V2O3/C.

Page 7: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

455 Nano Res. 2018, 11(1): 449–463

As a proof-of-concept application, the shuttle-like

V2O3/C composites were evaluated as anode materials

for SIBs, and the electrochemical performance is

presented in Fig. 6. The CV curves for the V2O3/C

electrode are shown in Fig. 6(a). There are three obvious

reduction peaks around 1.08, 1.77, and 2.36 V in the

first cathodic scan, which are not observed in the

subsequent cycles. The phenomenon results from an

irreversible reaction associated with the formation

of a solid-electrolyte interface (SEI) film [31, 32]. The

following CV curves in Fig. 6(a) are almost identical,

indicating the excellent reversibility of the Na+

insertion/deinsertion reaction (xNa+ + V2O3 ↔ NaxV2O3)

in the V2O3/C electrode [31]. An ICP test was performed

to investigate the molar fraction of Na ions inserted

into the shuttle-like V2O3/C after the electrode was fully

discharged. As shown in Table S3 in the ESM, the

contents of V and Na were 31.32 wt.% and 22.26 wt.%,

respectively, which corresponds to an atomic ratio of

approximately 2:3. This result indicates that 3 mol of

Na ions were intercalated into V2O3/C to form Na3V2O3.

Figure 6(b) shows the discharge/charge voltage pro-

files of the V2O3/C electrode at 50 mA·g−1 for the first

discharge cycle and subsequent selected cycles. The

initial discharge capacity of the shuttle- like V2O3/C

electrode reached 630 mAh·g−1, with a Coulombic

efficiency of 64.1%. The low initial Coulombic

efficiency is related to the irreversible capacity and

is attributed to two factors. The first is the decom-

position of the electrolyte and the formation of the

SEI film [78, 93]. Secondly, the active materials and C

matrix become loose during Na ion insertion/extraction

in the initial state [78]. After the 1st cycle, the discharge/

charge curves are basically coincident, illustrating the

high capacity retention of the shuttle-like V2O3/C. An

obvious plateau around 1.1 V is observed in the first

discharge curve, and there are no visible charge/

discharge plateaus or changes in any of the curves

after the second cycle, which agrees well with the CV

observations. Additionally, discharge specific capacities

as high as 425 and 378 mAh·g−1 were achieved at the

2nd and 80th cycles, respectively, while the Coulombic

efficiency was maintained at approximately 99%

(Fig. S7 in the ESM). As shown in Fig. 6(c), the

shuttle-like V2O3/C delivered a high initial specific

discharge capacity of 257 mAh·g−1, with an initial

Coulombic efficiency of 67.8% at 200 mA·g−1, and

maintained a reversible capacity of 217 mAh·g−1 after

100 cycles. Up to 500 mA·g−1, the discharge capacity

remained around 182 mAh·g−1 from the 9th cycle to the

91st cycle and then gradually decreased to 173 mAh·g−1

at the 100th cycle. Additionally, the average Coulombic

efficiency of the V2O3/C anode was maintained at

approximately 98%, indicating good electrochemical

reversibility. To evaluate the rate performance of the

V2O3/C electrode, the coin cells were tested at mutative

current densities ranging from 50 to 2,000 mA·g−1

(Fig. 6(d)). The porous materials delivered average

discharge capacities of 417, 247, 202, 176, 164, and

149 mAh·g−1 at current densities of 50, 100, 200, 500,

Figure 5 (a)–(c) FESEM and (d) and (e) TEM images, and (f) HRTEM image of the shuttle-like V2O5.

Page 8: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

| www.editorialmanager.com/nare/default.asp

456 Nano Res. 2018, 11(1): 449–463

1,000, and 2,000 mA·g−1, respectively. The significant

capacity degradation that occurred after the initial

cycle is attributed to the irreversible formation of the

SEI film, similar to the case for most anode materials

[76, 79, 94]. However, when the current density was

decreased to 100 mA·g−1, the specific discharge capacity

stabilized at approximately 242 mAh·g−1. The capacity

retention was as high as 73.8% in the current-density

range of 200 to 2,000 mA·g−1, revealing the excellent

rate capability. Impressively, the MOF-derived V2O3/C

also exhibited excellent long-term cycling performance

at a high current density, as shown in Fig. 6(e). The

initial specific discharge capacity reached 181 mAh·g−1

at 2,000 mA·g−1. After 1,000 cycles, approximately

133 mAh·g−1 of the capacity was retained, corresponding

to a fading rate of 0.032% per cycle. The outstanding

electrochemical performance may resulted from not

only the 3D V–V framework with intrinsic metallic

behavior in the V2O3 but also the numerous micropores

and conductive C in the MOF-derived V2O3/C shuttles.

Ex-situ XRD patterns of the porous shuttle-like

V2O3/C electrodes discharged to different voltage

states and cycled 1,000 times were obtained, as shown

in Fig. S8 in the ESM. The (012), (104), (110), and (116)

diffraction peaks of V2O3 exhibited almost no change

when the electrodes were discharged to different

Figure 6 Electrochemical performance of the shuttle-like V2O3/C in NIBs. (a) The first four CV curves at a scan rate of 0.1 mV·s−1.(b) The charge/discharge profiles at 50 mA·g−1. (c) The cycling performance and corresponding Coulombic efficiency of V2O3/C electrodesat 200 and 500 mA·g−1. (d) The rate performance at different currents ranging from 50 to 2,000 mA·g−1. (e) The high-rate (2,000 mA·g−1) cycling performance and corresponding Coulombic efficiency.

Page 9: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

457 Nano Res. 2018, 11(1): 449–463

voltage states. In particular, after 1,000 cycles, the

distinct characteristic lines of V2O3 were observed,

without obvious variation (Fig. S8 in the ESM), indicating

the excellent structural stability of the V2O3. This result

indicates that V2O3 may participate in the insertion/

extraction reaction during the discharge/charge process.

As shown in Fig. S9 in the ESM, the V2O3/C electrode

maintained its rod-like morphology after long-term

cycling, although the points at both ends disappeared.

This indicates that the porous shuttle-like morphology

derived from self-templated MIL-88B (V) was highly

beneficial to the reversible electrochemical reaction

during the long-term discharging and charging. Thus,

the excellent electrochemical property the porous

shuttle-like V2O3/C was due to the combined effect of

the MOF-derived porous C with a self-templated

framework and the V2O3 with a stable structure.

To analyze the insertion behavior of the Na ions in

V2O3 in detail, DFT calculations were performed

using the Vienna Ab-initio Simulation Package [95].

A 2 × 2 × 2 supercell with 240 atoms (96 V atoms and

144 O atoms) was constructed for the rhombohedral

V2O3. During all the calculations, the projected

augmented wave approach [96, 97] and the Perdew–

Burke–Ernzerh [98] generalized gradient approximation

[99] were employed to simulate the electron–ion

interaction and the exchange-correlation energy,

respectively. The cutoff energy for the plane-wave

basis was set as 350 eV. A tolerance of 1 × 10–4 eV was

employed for the energy convergence. The Monkhorst–

Pack scheme [100] with a 3 × 3 × 3 k-point grid, which

was tested as fine for the convergence of energies and

structures, was used to sample the Brillouin-zone

integrations.

As previously discussed, the 3D structure of V2O3

can be regarded as a special layered structure com-

prising V–V dumbbells and corresponding O atoms.

According to the periodic symmetry, six possible

positions—divided into two groups—are considered

to determine the favorable insertion site of Na ions in

V2O3. In the first group, the Na ions were inserted

into the interlamination, as shown in Fig. 7. At the

sites 1, the Na ion was inserted in the middle of two

horizontal V–V dumbbells along the Z axis (Fig. 7(a)).

As shown in Fig. 7(b), at the sites 2, Na ion was just

inside the quadrangle that comprised the adjacent

two O atoms and two V atoms. Figure 7(c) plots the

sites 3, where Na ion was located between two O

atoms in different layers. In the other group (Fig. S10

in the ESM), there were another three possible insertion

sites inside each layer. The binding energy (Ebing) of

Na atoms inserted in the layered V2O3 was defined

by following Eq.[1] [101, 102]

Ebing = E(NaV2O3) – E(V2O3) – E(Na) (1)

where E(NaV2O3) and E(V2O3) are the total energies

of the V2O3 with and without Na insertion, and E(Na)

is the energy per atom of the bulk metal Na. A

negative binding energy indicates that the Na ions

are preferentially inserted into V2O3, forming a bulk

metal. The corresponding total energies and binding

energies are listed in Table S4 in the ESM. The results

indicate that the Na ions occupy the interlamination

rather than the inside of the layers in V2O3. Site 1 was

undoubtedly the most favorable insertion site for Na

atoms, as evidenced by the fact that it had the lowest

binding energy (−10.7406 eV). Notably, sites 1, with a

spacing of 4.530 (1) Å between adjacent V–V dumbbells,

provided abundant space for Na ions. Because of the

repulsive force between the Na and V ions, the Na

ions were preferentially inserted at a suitable position

(sites 1) that could easily hold the force balance.

There were slight changes in the structure of V2O3

after the Na insertion. The volume increased from

Figure 7 Considered insertion sites for Na ions in the V2O3: (a) sites 1, (b) sites 2, (c) sites 3.

Page 10: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

| www.editorialmanager.com/nare/default.asp

458 Nano Res. 2018, 11(1): 449–463

2,292.37 to 2,379.66 Å3 with a change rate of only

3.8%. This reveals that the layered structure with the

skeleton of V–V dumbbells maintained stability during

the discharging process. The charge-density differences

of V2O3 with one Na ion inserted are shown in

Fig. 8(a). It is intuitively observed that the Na ion and

neighboring O atoms were surrounded by 3D free

electron gas, indicating that the Na ion preferentially

bonded with O atoms, with significant charge transfer.

According to Bader charge analysis [103], the inserted

Na ion transferred approximately 2/3 of the valence

electrons (0.6602 e) to V2O3, confirming the formation

of a promising Na–O chemical bond. Additionally,

the DOS for the configurations after the Na insertion

was calculated to analyze the electronic structure. As

shown in Fig. 8(b), the system still exhibited metallic

behavior after the insertion of Na ions, indicating that

its good conductibility was maintained. The results

also indicate the great advantage of V2O3 as electrodes

for SIBs, i.e., the enhanced cycle stability and high-rate

capability.

Figure 8 (a) Charge-density differences of V2O3 with one Na atom inserted. (b) Calculated DOS for V2O3 after Na ion insertion.

4 Conclusions

We demonstrated a facile thermal approach for the

synthesis of porous shuttle-like vanadium oxides

(V2O5 and V2O3/C) using MIL-88B (V) as precursors.

Calcination in different atmospheres results in slight

differences in the final structure of the vanadium

oxides, but both V2O5 and V2O3/C exhibited a uniform

shuttle-like morphology. As a proof-of-concept appli-

cation, the as-obtained porous shuttle-like V2O3/C

was applied as an anode material for SIBs. Because of

its inherent layered structure with metallic behavior

and its porous shuttle-like morphology with a uniform

C coating, this material exhibited outstanding electro-

chemical performance. A low capacity fading rate

of 0.032% per cycle was observed for 1,000 cycles at

2 A·g−1, indicating capacity retention of 133 mAh·g−1.

Detailed DFT calculations revealed that Na ions may

have been inserted into the interlamination of V2O3,

and the layered structure with metallic behavior may

have been maintained after the insertion of the Na

ions. Thus, the porous shuttle-like V2O3/C is promising

as an anode material for high-performance SIBs.

Importantly, the successful preparation of highly

porous vanadium oxides via MIL-88B (V) may stimulate

the synthesis of other porous V-based materials with

great promise for various applications.

Acknowledgements

This work was supported by the National Natural

Science Foundation of China (NSFC) (Nos. 51572299

and 51374255), the National High-tech R&D Pro-

gram of China (863 Program) (No. 2013AA110106),

and the Fundamental Research Funds for the

Central Universities of Central South University

(No. 2017zzts004).

Electronic Supplementary Material: Supplementary

material (the calculated DOS of V2O3; XRD pattern

of the MIL-88B (V); FESEM and TEM images of the

MIL-88B (V); XRD pattern of the V2O5; the cycling

performance and corresponding Coulombic efficiency

of V2O3 electrodes at 50 mA·g−1; the considered insertion

sites of Na ion in the V2O3: sites 4, sites 5 and sites 6;

pore structure parameters of V2O3 shuttles; total

energy of V2O3 with and without a Na atom inserted

at representative positions, the corresponding binding

energy) is available in the online version of this article

at https://doi.org/10.1007/s12274-017-1653-9.

Page 11: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

459 Nano Res. 2018, 11(1): 449–463

References

[1] Kang, K.; Meng, Y. S.; Bréger, J.; Grey, C. P.; Ceder, G.

Electrodes with high power and high capacity for rechargeable

lithium batteries. Science 2006, 311, 977–980.

[2] Armand, M.; Tarascon, J. M. Building better batteries.

Nature 2008, 451, 652–657.

[3] Tarascon, J. M.; Armand, M. Issues and challenges facing

rechargeable lithium batteries. Nature 2001, 414, 359–367.

[4] Goodenough, J. B.; Kim, Y. Challenges for rechargeable Li

batteries. Chem. Mater. 2010, 22, 587–603.

[5] Yoshino, A. The birth of the lithium-ion battery. Angew.

Chem., Int. Ed. 2012, 51, 5798–5800.

[6] Fang, G. Z.; Zhou, J.; Liang, C. W.; Pan, A. Q.; Zhang, C.;

Tang, Y.; Tan, X. P.; Liu, J.; Liang, S. Q. MOFs nanosheets

derived porous metal oxide-coated three-dimensional

substrates for lithium-ion battery applications. Nano Energy

2016, 26, 57–65.

[7] Sun, X. L.; Si, W. P.; Xi, L. X.; Liu, B.; Liu, X. J.; Yan, C. L.;

Schmidt, O. G. In situ-formed, amorphous, oxygen-enabled

germanium anode with robust cycle life for reversible

lithium storage. ChemElectroChem 2015, 2, 737–742.

[8] Sun, X. L.; Si, W. P.; Liu, X. H.; Deng, J. W.; Xi, L. X.;

Liu, L. F.; Yan, C. L.; Schmidt, O. G. Multifunctional

Ni/NiO hybrid nanomembranes as anode materials for

high-rate Li-ion batteries. Nano Energy 2014, 9, 168–175.

[9] Wang, J. Y.; Yang, N. L.; Tang, H. J.; Dong, Z. H.; Jin, Q.;

Yang, M.; Kisailus, D.; Zhao, H. J.; Tang, Z. Y.; Wang, D.

Accurate control of multishelled Co3O4 hollow microspheres

as high-performance anode materials in lithium-ion batteries.

Angew. Chem., Int. Ed. 2013, 52, 6417–6420.

[10] Ren, H.; Sun, J. J.; Yu, R. B.; Yang, M.; Gu, L.; Liu, P. R.;

Zhao, H. J.; Kisailus, D.; Wang, D. Controllable synthesis

of mesostructures from TiO2 hollow to porous nanospheres

with superior rate performance for lithium ion batteries.

Chem. Sci. 2016, 7, 793–798.

[11] Du, J.; Qi, J.; Wang, D.; Tang, Z. Y. Facile synthesis of

Au@TiO2 core-shell hollow spheres for dye-sensitized solar

cells with remarkably improved efficiency. Energy Environ.

Sci. 2012, 5, 6914–6918.

[12] Qi, J.; Lai, X. Y.; Wang, J. Y.; Tang, H. J.; Ren, H.; Yang, Y.;

Jin, Q.; Zhang, L. J.; Yu, R. B.; Ma, G. H. et al. Multi-

shelled hollow micro-/nanostructures. Chem. Soc. Rev. 2015,

44, 6749–6773.

[13] Xu, S. M.; Hessel, C. M.; Ren, H.; Yu, R. B.; Jin, Q.;

Yang, M.; Zhao, H. J.; Wang, D. α-Fe2O3 multi-shelled

hollow microspheres for lithium ion battery anodes with

superior capacity and charge retention. Energy Environ. Sci.

2014, 7, 632–637.

[14] Xu G. Y.; Yuan, J. R.; Tao, X. Y.; Ding, B.; Dou, H.; Yan,

X. H.; Xiao, Y.; Zhang, X. G. Absorption mechanism of

carbon-nanotube paper-titanium dioxide as a multifunctional

barrier material for lithium-sulfur batteries. Nano Res. 2015,

8, 3066–3074.

[15] Dong, Y. F.; Li, S.; Zhao, K. N.; Han, C. H.; Chen,W.;

Wang, B. L.; Wang, L.; Xu, B. A.; Wei, Q. L.; Zhang, L.

et al. Hierarchical zigzag Na1.25V3O8 nanowires with

topotactically encoded superior performance for sodium-ion

battery cathodes. Energy Environ. Sci. 2015, 8, 1267–1275.

[16] Ren, W. H.; Zheng, Z. P.; Xu, C.; Niu, C. J.; Wei, Q. L.;

An, Q. Y.; Zhao, K. N.; Yan, M. Y.; Qin, M. S.; Mai, L. Q.

Self-sacrificed synthesis of three-dimensional Na3V2(PO4)3

sanofiber setwork for high-rate sodium-ion full batteries.

Nano Energy 2016, 25, 145–153.

[17] Xu, Y. N.; Wei, Q. L.; Xu, C.; Li, Q. D.; An, Q. Y.; Zhang,

P. F.; Sheng, J. Z.; Zhou, L.; Mai, L. Q. Layer-by-layer

Na3V2(PO4)3 embedded in reduced graphene oxide as superior

rate and ultralong-life sodium-ion battery cathode. Adv.

Energy Mater. 2016, 6, 1600389.

[18] Zhang, Y.; Zhao, H. Y.; Du, Y. P. Symmetric full cells

assembled by using self-supporting Na3V2(PO4)3 bipolar

electrodes for superior sodium energy storage. J. Mater.

Chem. A 2016, 4, 7155–7159.

[19] Yuan, S.; Liu, Y. B.; Xu, D.; Ma, D. L.; Wang, S.; Yang, X. H.;

Cao, Z. Y.; Zhang, X. B. Pure single-crystalline Na1.1V3O7.9

nanobelts as superior cathode materials for rechargeable

sodium-ion batteries. Adv. Sci. 2015, 2, 1400018.

[20] Palomares, V.; Serras, P.; Villaluenga, I.; Hueso, K. B.;

Carretero-González, J.; Rojo, T. Na-ion batteries, recent

advances and present challenges to become low cost energy

storage systems. Energy Environ. Sci. 2012, 5, 5884–5901.

[21] Kim, S. W.; Seo, D. H.; Ma, X. H.; Ceder, G.; Kang, K.

Electrode materials for rechargeable sodium-ion batteries:

potential alternatives to current lithium-ion batteries. Adv.

Energy Mater. 2012, 2, 710–721.

[22] Ong, S. P.; Chevrier, V. L.; Hautier, G.; Jain, A.; Moore, C.;

Kim, S.; Ma, X. H.; Ceder, G. Voltage, stability and diffusion

barrier differences between sodium-ion and lithium-ion inter-

calation materials. Energy Environ. Sci. 2011, 4, 3680–3688.

[23] Xu, X.; Yu, D. M.; Zhou, H.; Zhang, L. S.; Xiao, C. H.;

Guo, C. W.; Guo, S. W.; Ding, S. J. MoS2 nanosheets grown

on amorphous carbon nanotubes for enhanced sodium

storage. J. Mater. Chem. A 2016, 4, 4375–4379.

[24] Cheng, F. Y.; Chen, J. Transition metal vanadium oxides

and vanadate materials for lithium batteries. J. Mater. Chem.

2011, 21, 9841–9848.

[25] Wu, C. Z.; Xie, Y. Promising vanadium oxide and hydroxide

nanostructures: from energy storage to energy saving. Energy

Environ. Sci. 2010, 3, 1191–1206.

Page 12: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

| www.editorialmanager.com/nare/default.asp

460 Nano Res. 2018, 11(1): 449–463

[26] Chernova, N. A.; Roppolo, M.; Dillon, A. C.; Whittingham,

M. Stanley. Layered vanadium and molybdenum oxides:

batteries and electrochromics. J. Mater. Chemistry 2009, 19,

2526–2552.

[27] Wang, Y.; Chao, G. Z. Synthesis and enhanced intercalation

properties of nanostructured vanadium oxides. Chem. Mater.

2006, 18, 2787–2804.

[28] Raju, V.; Rains, J.; Gates, C.; Luo, W.; Wang, X. F.;

Stickle, W. F.; Stucky, G. D.; Ji, X. L. Superior cathode of

sodium-ion batteries: Orthorhombic V2O5 nanoparticles

generated in nanoporous carbon by ambient hydrolysis

deposition. Nano Lett. 2014, 14, 4119–4124.

[29] Zhang, P. F.; Zhao, L. Z.; An, Q. Y.; Wei, Q. L.; Zhou, L.;

Wei, X. J.; Sheng, J. Z.; Mai, L. Q. A high-rate V2O5 hollow

microclew cathode for an all-vanadium-based lithium-ion

full cell. Small 2016, 12, 1082–1090.

[30] Su, D. W.; Wang, G. X. Single-crystalline bilayered V2O5

nanobelts for high-capacity sodium–ion batteries. ACS Nano

2013, 7, 11218–11226.

[31] Xia, X. H.; Chao, D. L.; Zhang, Y. Q.; Zhan, J. Y.; Zhong, Y.;

Wang, X. L.; Wang, Y. D.; Shen, Z. X.; Tu, J. P.; Fan, H. J.

Generic synthesis of carbon nanotube branches on metal

oxide arrays exhibiting stable high-rate and long-cycle

sodium-ion storage. Small 2016, 12, 3048–3058.

[32] Niu, C. J.; Huang, M.; Wang, P. Y.; Meng, J. S.; Liu, X.;

Wang, X. P.; Zhao, K. N.; Yu, Y.; Wu, Y. Z.; Lin, C. et al.

Carbon-supported and nanosheet-assembled vanadium oxide

microspheres for stable lithium-ion battery anodes. Nano

Res. 2015, 9, 128–138.

[33] Jiang, L.; Qu, Y.; Ren, Z. Y.; Yu, P.; Zhao, D. D.; Zhou, W.;

Wang, L.; Fu, H. G. In situ carbon-coated yolk-shell V2O3

microspheres for lithium-ion batteries. ACS Appl. Mater.

Interfaces 2015, 7, 1595–1601.

[34] Wang J. Y.; Tang, H. J.; Zhang, L. J.; Ren, H.; Yu, R. B.;

Jin, Q.; Qi, J.; Mao, D.; Yang, M.; Wang, Y. et al. Multi-

shelled metal oxides prepared via an anion-adsorption

mechanism for lithium-ion batteries. Nat. Energy 2016, 1,

16050.

[35] Chao, D. L.; Xia, X. H.; Liu, J. L.; Fan, Z. X.; Ng, C. F.;

Lin, J. Y.; Zhang, H.; Shen, Z. X.; Fan, H. J. A V2O5/

conductive-polymer core/shell nanobelt array on three-

dimensional graphite foam: a high-rate, ultrastable, and

freestanding cathode for lithium-ion batteries. Adv. Mater.

2014, 26, 5794–5800.

[36] Zhai, T. Y.; Liu, H. M.; Li, H. Q.; Fang, X. S.; Liao, M. Y.;

Li, L.; Zhou, H. S.; Koide, Y.; Bando, Y.; Golberg, D.

Centimeter-long V2O5 nanowires: From synthesis to field-

emission, electrochemical, electrical transport, and photo-

conductive properties. Adv. Mater. 2010, 22, 2547–2552.

[37] Li, Y. W.; Yao, J. H.; Uchaker, E.; Yang, J. W.; Huang,

Y. X.; Zhang, M.; Cao, G. Z. Leaf-like V2O5 nanosheets

fabricated by a facile green approach as high energy cathode

material for lithium-ion batteries. Adv. Energy Mater. 2013,

3, 1171–1175.

[38] Wu, H. B.; Pan, A. Q.; Hng, H. H.; Lou, X. W. Template-

assisted formation of rattle-type V2O5 hollow microspheres

with enhanced lithium storage properties. Adv. Funct. Mater.

2013, 23, 5669–5674.

[39] Pan, A. Q.; Wu, H. B.; Yu, L.; Lou, X. W. Template-free

synthesis of VO2 hollow microspheres with various interiors

and their conversion into V2O5 for lithium-ion batteries.

Angew. Chem., Int. Ed. 2013, 125, 2282–2286.

[40] Parija, A.; Liang, Y. F.; Andrews, J. L.; De Jesus, L. R.;

Prendergast, D.; Banerjee, S. Topochemically de-intercalated

phases of V2O5 as cathode materials for multivalent inter-

calation batteries: A first-principles evaluation. Chem. Mater.

2016, 28, 5611–5620

[41] Xu, H. T.; Chen, J. D.; Zhang, H. J.; Zhang, Y.; Li, W. X.;

Wang, Y. Fabricating SiO2-coated V2O5 nanoflake arrays

for high-performance lithium-ion batteries with enhanced

cycling capability. J. Mater. Chem. A 2016, 4, 4098–4106.

[42] An, Q. Y.; Zhang, P. F.; Wei, Q. L.; He, L.; Xiong, F. Y.;

Sheng, J. Z.; Wang, Q. Q.; Mai, L. Q. Top-down fabrication

of three-dimensional porous V2O5 hierarchical microplates

with tunable porosity for improved lithium battery performance.

J. Mater. Chem. A 2014, 2, 3297–3302.

[43] Su, D. W.; Dou, S. X.; Wang, G. X. Hierarchical orthorhombic

V2O5 hollow nanospheres as high performance cathode

materials for sodium-ion bbatteries. J. Mater. Chem. A 2014,

2, 11185–11194.

[44] Mai, L. Q.; Dong, F.; Xu, X.; Luo, Y. Z.; An, Q. Y.; Zhao,

Y. L.; Pan, J.; Yang, J. N. Cucumber-like V2O5/poly (3,4-

ethylenedioxythiophene)&MnO2 nanowires with enhanced

electrochemical cyclability. Nano Lett. 2013, 13, 740–745.

[45] Shi, Y.; Zhang, Z. J.; Wexler, D.; Chou, S. L.; Gao, J.;

Abruña, H. D.; Li, H. J.; Liu, H. K.; Wu, Y. P.; Wang, J. Z.

Facile synthesis of porous V2O3/C composites as lithium

storage material with enhanced capacity and good rate

capability. J. Power Sources 2015, 275, 392–398.

[46] Jiang, H.; Jia, G. Q.; Hu, Y. J.; Cheng, Q. L.; Fu, Y.; Li, C.

Z. Ultrafine V2O3 nanowire embedded in carbon hybrids

with enhanced lithium storage capability. Ind. Eng. Chem.

Res. 2015, 54, 2960–2965.

[47] Li, H. Y.; Jiao, K.; Wang, L.; Wei, C.; Li, X. L.; Xie, B.

Micelle anchored in situ synthesis of V2O3 nanoflakes@C

composites for supercapacitors. J. Mater. Chem. A 2014, 2,

18806–18815.

[48] Zhang, Y. F.; Fan, M. J.; Hu, L.; Wu, W. B.; Zhang, J. C.;

Page 13: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

461 Nano Res. 2018, 11(1): 449–463

Liu, X. H.; Zhong, Y. L.; Huang, C. Fabrication of V2O3/C

core-shell structured composite and vc nanobelts by the

thermal treatment of VO2/C composite. Appl. Surf. Sci. 2012,

258, 9650–9655.

[49] Wang, Y.; Zhang, H. J.; Admar, A. S.; Luo, J. Z.; Wong,

C. C.; Borgna, A.; Lin, J. Y. Improved cyclability of

lithium-ion battery anode using encapsulated V2O3 nano-

structures in well-graphitized carbon fiber. RSC Adv. 2012,

2, 5748–5753.

[50] Sun, Y. F.; Jiang, S. S.; Bi, W. T.; Wu, C. Z.; Xie, Y. Highly

ordered lamellar V2O3-based hybrid nanorods towards

superior aqueous lithium-ion battery performance. J. Power

Sources 2011, 196, 8644–8650.

[51] Wu, C. Z.; Feng, F.; Xie, Y. Design of vanadium oxide

structures with controllable electrical properties for energy

applications. Chem. Soc. Rev. 2013, 42, 5157–5183.

[52] Taylor, J. W.; Smith, T. J.; Andersen, K. H.; Capellmann, H.;

Kremer, R. K.; Simon, A.; Schärpf, O.; Neumann, K. U.;

Ziebeck, K. R. A. Spin-spin correlations in the insulating

and metallic phases of the mott system V2O3. Eur. Phys. J. B

1999, 12, 199–207.

[53] Xu, Y. N.; Chung, S. Y.; Bloking, J. T.; Chiang, Y. M.;

Ching, W. Y. Electronic structure and electrical conductivity

of undoped LiFePO4. Electrochem. Solid State Lett. 2004, 7,

A131–A134.

[54] Shi, S. Q.; Liu, L. J.; Ouyang, C. Y.; Wang, D. S.; Wang,

Z. X.; Chen, L. Q.; Huang, X. J. Enhancement of electronic

conductivity of LiFePO4 by Cr doping and its identifica-

tion by first-principles calculations. Phys. Rev. B 2003, 68,

195108.

[55] Xu, J.; Chen, G. Effects of doping on the electronic pro-

perties of LiFePO4: A first-principles investigation. Phys. B

Condens. Matter 2010, 405, 803–807.

[56] Zhou, F.; Kang, K.; Maxisch, T.; Ceder, G.; Morgan, D.

The electronic structure and band gap of LiFePO4 and

LiMnPO4. Solid State Commun. 2004, 132, 181–186.

[57] Zheng, C. M.; Zhang, X. M.; He, S.; Fu, Q.; Lei, D. M.

Preparation and characterization of spherical V2O3 nano-

powder. J. Solid State Chem. 2003, 170, 221–226.

[58] Müller, C.; Nateprov, A. A.; Obermeier, G.; Klemm, M.;

Tidecks, R.; Wixforth, A.; Horn S. Surface acoustic wave

investigations of the metal-to-insulator transition of V2O3 thin

films on lithium niobate. J. Appl. Phys. 2005, 98, 084111.

[59] Liu, N.; Yao, Y.; Cha, J. J.; McDowell, M. T.; Han, Y.;

Cui, Y. Functionalization of silicon nanowire surfaces with

metal-organic frameworks. Nano Res. 2012, 5, 109–116.

[60] Deng, H. X.; Grunder, S.; Cordova, K. E.; Valente, C.;

Furukawa, H.; Hmadeh, M.; Gándara, F.; Whalley, A. C.;

Liu, Z.; Asahina, S. et al. Large-pore apertures in a series of

metal-organic frameworks. Science 2012, 336, 1018–1023.

[61] Hayashi, H.; Côté, A. P.; Furukawa, H.; O'Keeffe, M.;

Yaghi, O. M. Zeolite a imidazolate frameworks. Nat. Mater.

2007, 6, 501–506.

[62] Lu, W. G.; Wei, Z. W.; Gu, Z. Y.; Liu, T. F.; Park, J.; Park,

J.; Tian, J.; Zhang, M. W.; Zhang, Q.; Gentle III, T. et al.

Tuning the structure and function of metal-organic frameworks

via linker design. Chem. Soc. Rev. 2014, 43, 5561–5593.

[63] Stock, N.; Biswas, S. Synthesis of metal-organic frameworks

(MOFs): Routes to various MOF topologies, morphologies,

and composites. Chem. Rev. 2012, 112, 933–969.

[64] Xu, G. Y.; Ding, B.; Shen, L. F.; Nie, P.; Han, J. P.; Zhang,

X. G. Sulfur embedded in metal organic framework-derived

hierarchically porous carbon nanoplates for high performance

lithium-sulfur battery. J. Mater. Chem. A 2013, 1, 4490–4496.

[65] Zhang, L.; Wu, H. B.; Lou, X. W. Metal-organic-

frameworks-derived general formation of hollow structures

with high complexity. J. Am. Chem. Soc. 2013, 135,

10664–10672.

[66] Yu, X. Y.; Yu, L.; Wu, H. B.; Lou, X. W. Formation of

nickel sulfide nanoframes from metal-organic frameworks

with enhanced pseudocapacitive and electrocatalytic properties.

Angew. Chem., Int. Ed. 2015, 54, 5331–5335.

[67] Guo, W. X.; Sun, W. W.; Wang, Y. Multilayer CuO@NiO

hollow spheres: Microwave-assisted metal-organic-framework

derivation and highly reversible structure-matched stepwise

lithium storage. ACS Nano 2015, 9, 11462–11471.

[68] Huang, G.; Zhang, F. F.; Du, X. C.; Qin, Y. L.; Yin, D. M.;

Wang, L. M. Metal organic frameworks route to in situ

insertion of multiwalled carbon nanotubes in Co3O4 polyhedra

as anode materials for lithium-ion batteries. Nano 2015, 9,

1592–1599.

[69] Cao, X. H.; Zheng, B.; Shi, W. H.; Yang, J.; Fan, Z. X.;

Luo, Z. M.; Rui, X. H.; Chen, B.; Yan, Q. Y.; Zhang, H.

Reduced graphene oxide-wrapped MoO3 composites prepared

by using metal-organic frameworks as precursor for all-

solid-state flexible supercapacitors. Adv. Mater. 2015, 27,

4695–4701.

[70] Zheng, J. M.; Tian, J.; Wu, D. X.; Gu, M.; Xu, W.; Wang,

C. M.; Gao, F.; Engelhard, M. H.; Zhang, J. G.; Liu, J. et al.

Lewis acid–base interactions between polysulfides and metal

organic framework in lithium sulfur batteries. Nano Lett.

2014, 14, 2345–2352.

[71] Wu, R. B.; Qian, X. K.; Rui, X. H.; Liu, H.; Yadian, B. L.;

Zhou, K.; Wei, J.; Yan, Q. Y.; Feng, X. Q.; Long, Y. et al.

Zeolitic imidazolate framework 67-derived high symmetric

porous Co3O4 hollow dodecahedra with highly enhanced

lithium storage capability. Small 2014, 10, 1932–1938.

Page 14: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

| www.editorialmanager.com/nare/default.asp

462 Nano Res. 2018, 11(1): 449–463

[72] Zhang, L.; Wu, H. B.; Madhavi, S.; Hng, H. H.; Lou, X. W.

Formation of Fe2O3 microboxes with hierarchical shell

structures from metal-organic frameworks and their lithium

storage properties. J. Am. Chem. Soc. 2012, 134, 17388–

17391.

[73] Wu, R. B.; Qian, X. K.; Yu, F.; Liu, H.; Zhou, K.; Wei, J.;

Huang, Y. Z. MOF-templated formation of porous CuO

hollow octahedra for lithium-ion battery anode materials.

J. Mater. Chem. A 2013, 1, 11126–11129.

[74] Zhang, Y. Z.; Wang, Y.; Xie, Y. L.; Cheng, T.; Lai, W. Y.;

Pang, H.; Huang, W. Porous hollow Co3O4 with rhombic

dodecahedral structures for high-performance supercapacitors.

Nanoscale 2014, 6, 14354–14359.

[75] Ullah, S.; Khan, I. A.; Choucair, M.; Badshah, A.; Khan, I.;

Nadeem, M. A. A novel Cr2O3-carbon composite as a high

performance pseudo-capacitor electrode material. Electrochim.

Acta 2015, 171, 142–149.

[76] Qu, B. H.; Ma, C. Z.; Ji, G.; Xu, C. H.; Xu, J.; Meng, Y. S.;

Wang, T. H.; Lee, J. Y. Layered SnS2-reduced graphene

oxide composite-a high-capacity, high-rate, and long-cycle

life sodium-ion battery anode material. Adv. Mater. 2014,

26, 3854–3859.

[77] Zhang, Y. D.; Zhu, P. Y.; Huang, L. L.; Xie, J.; Zhang, S. C.;

Cao, G. S.; Zhao, X. B. Few-layered SnS2 on few-layered

reduced graphene oxide as Na-ion battery anode with

ultralong cycle life and superior rate capability. Adv. Funct.

Mater. 2015, 25, 481–489.

[78] Lu,Y. Y.; Zhao, Q., Zhang, N.; Lei, K. X.; Li, F. J.; Chen, J.

Facile spraying synthesis and high-performance sodium

storage of mesoporous MoS2/C microspheres. Adv. Funct.

Mater. 2016, 26, 911–918.

[79] Xu, D. F.; Chen, C. J.; Xie, J.; Zhang, B.; Miao, L.; Cai, J.;

Huang, Y. H.; Zhang, L. N. A hierarchical N/S-codoped

carbon anode fabricated facilely from cellulose/polyaniline

microspheres for high-performance sodium-ion batteries.

Adv. Energy Mater. 2016, 6, 1501929.

[80] Carson, F.; Su, J.; Platero-Prats, A. E.; Wan, W.; Yun, Y. F.;

Samain, L.; Zou, X. D. Framework isomerism in vanadium

metal-organic frameworks: MIL-88B(V) and MIL-101(V).

Cryst. Growth Des. 2013, 13, 5036–5044.

[81] Jiang, Y.; Yang, Z. Z.; Li, W. H.; Zeng, L. C.; Pan, F. S.;

Wang, M.; Wei, X.; Hu, G. T.; Gu, L.; Yu, Y. Nanoconfined

carbon-coated Na3V2(PO4)3 particles in mesoporous carbon

enabling ultralong cycle life for sodium-ion batteries. Adv.

Energy Mater. 2015, 5, 1402104.

[82] Li, Q. D.; Wei, Q. L.; Sheng, J. Z.; Yan, M. Y.; Zhou, L.;

Luo, W.; Sun, R. M.; Mai, L. Q. Mesoporous Li3VO4/C

submicron-ellipsoids supported on reduced graphene oxide

as practical anode for high-power lithium-ion batteries. Adv.

Sci. 2015, 2, 1500284.

[83] Moulder, J. F.; Stickle, W. F.; Sobol, P. E.; Bomben, K. D.

Handbook of X-ray Photoelectron Spectroscopy; Chastain,

J.; King, R. C. Jr, Eds.; Perkin–Elmer Corporation: Eden

Prairie, MN, USA, 1992.

[84] Tang, W.; Peng, C. X.; Nai, C. T.; Su, J.; Liu, Y. P.; Reddy,

M. V. V.; Lin, M.; Loh, K. P. Ultrahigh capacity due to

multi-electron conversion reaction in reduced graphene

oxide-wrapped MoO2 porous nanobelts. Small 2015, 11,

2446–2453.

[85] Zhang, Y. F.; Fan, M. J.; Liu, X. H.; Huang, C.; Li, H. B.

Beltlike V2O3@C core-shell-structured composite: design,

preparation, characterization, phase transition, and impro-

vement of electrochemical properties of V2O3. Eur. J. Inorg.

Chem. 2012, 2012, 1650–1659.

[86] Jiang, H.; Ren, D. Y.; Wang, H. F.; Hu, Y. J.; Guo, S. J.;

Yuan, H. Y.; Hu, P. J.; Zhang, L.; Li, C. Z. 2D monolayer

MoS2-carbon interoverlapped superstructure: Engineering

ideal atomic interface for lithium ion storage. Adv. Mater.

2015, 27, 3687–3695.

[87] Xie, X. Q.; Ao, Z. M.; Su, D. W.; Zhang, J. Q.; Wang, G. X.

MoS2/graphene composite anodes with enhanced performance

for sodium-ion batteries: The role of the two-dimensional

heterointerface. Adv. Funct. Mater. 2015, 2, 1393–1403.

[88] Zhang, Y. F.; Pan, A. Q.; Wang, Y. P.; Wei, W. F.; Su, Y. H.;

Hu, J. M.; Cao, G. Z.; Liang, S. Q. Dodecahedron-shaped

porous vanadium oxide and carbon composite for high-rate

lithium ion batteries. ACS Appl. Mater. Interfaces 2016, 8,

17303–17311.

[89] Zhang, W.; Wu, Z. Y.; Jiang, H. L.; Yu, S. H. nanowire-

directed templating synthesis of metal-organic framework

nanofibers and their derived porous doped carbon nanofibers

for enhanced electrocatalysis. J. Am. Chem. Soc. 2014, 136,

14385–14388.

[90] Tang, J.; Salunkhe, R. R.; Liu, J.; Torad, N. L.; Imura, M.;

Furukawa, S.; Yamauchi, Y. Thermal conversion of core-shell

metal-organic frameworks: A new method for selectively

functionalized nanoporous hybrid carbon. J. Am. Chem. Soc.

2015, 137, 1572–1580.

[91] Chen, Y. Z.; Wang, C. M.; Wu, Z. Y.; Xiong, Y. J.; Xu, Q.;

Yu, S. H.; Jiang, H. L. From bimetallic metal-organic

framework to porous carbon: high surface area and

multicomponent active dopants for excellent electrocatalysis.

Adv. Mater. 2015, 27, 5010–5016.

[92] Enjalbert, R.; Galy, J. A refinement of the structure of V2O5.

Acta Crystallogr. C 1986, 42, 1467–1469.

[93] Xiao, J.; Wang, X. J.; Yang, X. Q.; Xun, S. D.; Liu, G.;

Koech, P. K.; Liu, J.; Lemmon, J. P. Electrochemically

induced high capacity displacement reaction of PEO/MoS2/

Page 15: Metal-organic framework-derived porous shuttle-like …...Metal-organic frameworks (MOFs), which comprise metal ions (or metal clusters) and organic linkers, have a large surface area,

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

463 Nano Res. 2018, 11(1): 449–463

graphene nanocomposites with lithium. Adv. Funct. Mater.

2011, 21, 2840–2846.

[94] Li, D. D.; Zhang, L.; Chen, H. B.; Wang, J.; Ding, L. X.;

Wang, S. Q.; Ashman, P. J.; Wang, H. H. Graphene-based

nitrogen-doped carbon sandwich nanosheets: A new capacitive

process controlled anode material for high-performance

sodium-ion batteries. J. Mater. Chem. A 2016, 4, 8630–8635.

[95] Kresse, G.; Furthmüller, J. Efficient iterative schemes for

ab initio total-energy calculations using a plane-wave basis

set. Phys. Rev. B 1996, 54, 11169–11186.

[96] Perdew, J. P.; Chevary, J. A.; Vosko, S. H.; Jackson, K. A.;

Pederson, M. R.; Singh, D. J.; Fiolhais, C. Atoms, molecules,

solids, and surfaces: applications of the generalized gradient

approximation for exchange and correlation. Phys. Rev. B

1992, 46, 6671–6687.

[97] Blöchl, P. E.; Parrinello, M. Adiabaticity in first-principles

molecular dynamics. Phys. Rev. B 1992, 45, 9413–9416.

[98] Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient

approximation made simple. Phys. Rev. Lett. 1996, 77,

3865–3868.

[99] Methfessel, M.; Paxton, A. T. High-precision sampling for

brillouin-zone integration in metals. Phys. Rev. B 1989, 40,

3616–3621.

[100] Monkhorst, H. J.; Pack, J. D. Special points for brillouin-

zone integrations. Phys. Rev. B 1976, 13, 5188–5192.

[101] Liao, Y. H.; Park, K. S.; Xiao, P. H.; Henkelman, G.; Li,

W. S.; Goodenough, J. B. Sodium intercalation behavior

of layered NaxNbS2 (0 ≤ x ≤ 1). Chem. Mater. 2013, 25,

1699–1705.

[102] Hu, J. P.; Xu, B.; Yang, S. A.; Guan, S.; Ouyang, C. Y.;

Yao, Y. G. 2D electrides as promising anode materials for

Na-ion batteries from first-principles study. ACS Appl.

Mater. Interfaces 2015, 7, 24016–24022.

[103] Tang, W.; Sanville, E.; Henkelman, G. A grid-based bader

analysis algorithm without lattice Bias. J. Phys. Condens.

Matter 2009, 21, 084204.