Transcript
Page 1: Spinel ZnMn2O4 nanoplate assemblies fabricated via “escape-by-crafty-scheme” strategy

Dynamic Article LinksC<Journal ofMaterials Chemistry

Cite this: J. Mater. Chem., 2012, 22, 13328

www.rsc.org/materials PAPER

Publ

ishe

d on

10

May

201

2. D

ownl

oade

d on

12/

05/2

014

13:4

0:42

. View Article Online / Journal Homepage / Table of Contents for this issue

Spinel ZnMn2O4 nanoplate assemblies fabricated via‘‘escape-by-crafty-scheme’’ strategy†

Jiao Zhao,ab Fuqing Wang,bc Panpan Su,ab Mingrun Li,a Jian Chen,c Qihua Yang*a and Can Li*a

Received 11th April 2012, Accepted 10th May 2012

DOI: 10.1039/c2jm32261g

A two-step process that differs in important details from previous methods used to prepare ZnMn2O4

nanoplate assemblies has been reported. This material was prepared by thermal transformation of

metal–organic nanoparticles into metal–oxide nanoparticles based on the ‘‘escape-by-crafty-scheme’’

strategy. Firstly, the nanoscale mixed-metal–organic frameworks (MMOFs) precursor, ZnMn2–ptcda

(ptcda ¼ perylene-3,4,9,10-tetracarboxylic dianhydride), containing Zn2+ and Mn2+, was prepared by

the designed soft chemical assembly of mixed metal ions and organic ligands at a molecular scale. In

a second step, the MMOFs are thermally transformed into spinel structured ZnMn2O4 with

morphology inherited from the MMOFs precursors. The well-crystallized spinel structure can be

formed by thermal treatment of ZnMn2–ptcda at 350�C, and is formed at temperatures$450 �C using

the co-precipitation method. This ‘‘escape-by-crafty-scheme’’ strategy can be extended to the

preparation of other spinel metal–oxide nanoparticles, e.g. CoMn2O4, and NiMn2O4, with well-defined

morphology inherited from the metal–organic precursors. The ZnMn2O4 nanoplate assemblies

thermally treated at 450 �C have potential application in lithium ion batteries as anode materials, which

show high specific capacity and good cyclability.

Introduction

Rechargeable lithium ion batteries have become the dominant

power source for portable electronic devices, such as notebook

PCs, mobile phones, and camcorders. The ever-growing

demands for new electrode materials with high capacity and/or

high energy density for upcoming large-scale applications have

prompted numerous research efforts, in particular in electric

vehicles and hybrid electric vehicles.1–5 Transition metal oxides,

which can react with Li reversibly via the conversion reaction:

MxOy + 2yLi+ + 2ye� 4 xM + yLi2O, exhibit high reversible

capacities at a relatively low potential as anode materials.6,7 A

combination of two metal oxides in a spinel-like structure has

shown good capacity on cycling compared with single metal

oxides.8–12 Among various candidates, ZnMn2O4 is regarded as

a potential high-performance anode material because of its low

aState Key Laboratory of Catalysis, Dalian Institute of Chemical Physics,Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023,China. E-mail: [email protected]; [email protected]; Web: http://www.hmm.dicp.ac.cn; http://www.canli.dicp.ac.cn. Fax: +86-411-84694447;Tel: +86-411-84379552; +86-411-84379070bGraduate School of the Chinese Academy of Sciences, Beijing 100049,ChinacLaboratory of Advanced Rechargeable Batteries, Dalian Institute ofChemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road,Dalian 116023, China

† Electronic supplementary information (ESI) available: Fullinformation about EDX spectrum, XRD patterns, SEM images, andN2 adsorption–desorption isotherm. See DOI: 10.1039/c2jm32261g

13328 | J. Mater. Chem., 2012, 22, 13328–13333

cost, environmental benignancy, low operating voltages and high

energy density. Abu-Lebdeh’s group reported that ZnMn2O4

showed better performance compared to its single metal oxides,

Mn2O3 or ZnO.13 However, this material shows significant

volume change during lithiation and delithiation and severe

electron conduction restriction during the electrochemical reac-

tion. Therefore, the discharge capacity and cycling stability for

ZnMn2O4 should be improved. It was reported that metal oxides

with novel nanostructure have the advantages of high surface-

to-volume ratio and short path length for Li+ transport.14–26

ZnMn2O4 nanostructures with specific morphology have been

prepared. Zhang’s group developed a solvothermal method to

synthesize flower-like ZnMn2O4 superstructures which exhibit

higher stable capacity than nanocrystalline ZnMn2O4 synthe-

sized by a polymer-pyrolysis method.27,28 Kim and co-workers

fabricated one-dimensional ZnMn2O4 nanowires by a solid-state

method and the mechanism research reveals that there is an

additional electrochemical activity, possibly due to the enhanced

kinetics of the nanowire.29 Lou’s group reported the preparation

of ZnMn2O4 hollow microspheres from ZnCO3–MnCO3

microspheres and this material exhibits high specific discharge

capacity and cycling stability, which is attributed to the unique

hollow structure.30

Although there are several successful examples of the synthesis

of ZnMn2O4 nanomaterials as electrodes for lithium ion

batteries, a general method for the preparation of spinel struc-

tured metal oxides with well-defined composition and

morphology is still lacking. Recently, our group and others have

This journal is ª The Royal Society of Chemistry 2012

Page 2: Spinel ZnMn2O4 nanoplate assemblies fabricated via “escape-by-crafty-scheme” strategy

Publ

ishe

d on

10

May

201

2. D

ownl

oade

d on

12/

05/2

014

13:4

0:42

. View Article Online

reported the synthesis of metal oxides by a thermal trans-

formation method using nanoscale metal–organic frameworks

(MOFs) as precursors.31–39 Through a simple thermal treatment

process, the metal oxides with morphology similar to the nano-

scale MOFs could be facilely synthesized. The separated

morphology and composition control step of this method

provides more possibility for precisely controlling the structure

and morphology of the metal oxides. However, the thermal

transformation of MOFs to metal oxides has mainly focused on

MxOy, and the preparation of mixed-metal oxides using this

method is seldom reported.40,41 Herein, we present a general

method for the synthesis of spinel structured mixed-metal oxides,

such as ZnMn2O4, CoMn2O4, and NiMn2O4, with well defined

morphology and composition, by thermal treatment of nano-

scale mixed-metal–organic frameworks (MMOFs) via an

‘‘escape-by-crafty-scheme’’ strategy, as shown in Fig. 1. The key

issue for this method is to prepare MMOFs precursors with

homogeneous mixing of different metal ions in a crystal structure

with well-controlled morphology. After simple thermal treat-

ment process, the spinel structured metal oxides with

morphology similar to the MMOFs precursors could be

successfully prepared. The potential of ZnMn2O4 nanoplate

assemblies as anode materials for lithium ion batteries was also

investigated.

Experimental

Synthesis of ZnMn2–ptcda MMOFs

In a typical synthesis,39 Zn(OAc)2$2H2O (0.133 mmol) and

Mn(OAc)2$4H2O (0.267 mmol) were dissolved in 22.5 mL

deionized water, and ptcda (0.2 mmol) was dissolved in 12.5 mL

NaOH solution (0.8 mmol NaOH). The ptcda solution was

added dropwise to the mixture solution of metal acetates with

stirring. The immediate formation of a precipitate was observed.

The reaction mixture was stirred at room temperature for 30 min,

and then transferred to a Teflon-lined stainless steel vessel

(45 mL) and heated at 100 �C for 4 h. After cooling down to

room temperature, the precipitate was collected by centrifuga-

tion, washed with water and dried. ZnMn6–ptcda MMOFs were

prepared in a similar way to ZnMn2–ptcda MMOFs except

a Zn : Mn molar ratio of 1 : 6 was used. For other MMn2–ptcda

MMOFs, the synthesis process was similar to that described for

ZnMn2–ptcda except using M(II) acetate instead of zinc(II)

acetate.

Fig. 1 Illustration of the two step process for the preparation of spinel

ZnMn2O4 using ZnMn2–ptcda as a precursor by an ‘‘escape-by-crafty-

scheme’’ strategy.

This journal is ª The Royal Society of Chemistry 2012

Formation of ZnMn2O4 nanomaterials

ZnMn2–ptcda MMOFs were thermally treated in air at 450 �Cfor 1 h with a ramp of 5 �C min�1. CoMn2O4 and NiMn2O4 were

formed from CoMn2–ptcda and NiMn2–ptcda respectively,

under the same thermal treatment process.

Characterization

The thermogravimetric (TG) analysis was performed in air with

a heating rate of 5 �C min�1 by using a NETZSCH STA-449F3

thermogravimetric analyzer. The powder X-ray diffraction data

(PXRD) were collected on a Rigaku D/Max2500 PC diffrac-

tometer with Cu Ka radiation (l ¼ 1.5418 �A) in the 2q range of

3–70� with a scan speed of 5� min�1 at room temperature.

Scanning electron microscopy (SEM) was undertaken on a FEI

QUANTA 200F scanning electron microscope operating at an

acceleration voltage of 20 kV. The samples were sputtered with

gold prior to imaging. Selected-area electron-diffraction (SAED)

and high-resolution transmission electron microscopy

(HRTEM) images were recorded on a FEI Tecnai F30 micro-

scope with a point resolution of 0.20 nm operated at 300 kV. The

nitrogen sorption experiments were performed at �196 �C on

a Micromeritics ASAP 2020 system. Prior to measurement, the

samples were out-gassed at 120 �C for 6 h. The BET (Brunauer–

Emmett–Teller) specific surface areas were calculated using

adsorption data in a relative pressure range of P/P0 ¼ 0.05–0.25.

The pore size distribution curve was calculated from the

adsorption branch using the BJH (Barrett–Joyner–Halenda)

method.

Electrode fabrication and performance measurements

The ZnMn2O4 electrode slurry was prepared by mixing 75 wt%

active material, 15 wt% acetylene carbon black, 10 wt% poly-

vinylidene fluoride (PVDF) binder and an adequate amount of

N-methyl-2-pyrrolidone (NMP). The slurry was coated onto

copper foil and dried under vacuum at 85 �C overnight. The

electrode was cut into a shape with an area of 1 cm2. The loading

weight of the active materials in the electrode was about 2.0 mg

cm�2. The electrochemical performance of the ZnMn2O4 elec-

trode was evaluated in a 2016 coin-type ZnMn2O4–Li cell, in

which the lithium electrode was used as the counter electrode as

well as the reference electrode. The electrolyte was 1 M LiPF6 in

ethylene carbonate (EC) and diethyl carbonate (DEC) (1 : 1 by

volume). A Celgard 2400 (polypropylene) was used as the sepa-

rator. All the cells were assembled in an argon-filled glove box

maintaining the moisture and oxygen levels below 1.0 ppm. All

electrochemical measurements were evaluated galvanostatically

between 0.01 and 3.0 V using a Neware instrument (China). The

charge–discharge measurements were tested under a current

density of 60 mA g�1. For the rate capability measurements, the

cells were both charged and discharged at a current density of

60–3600 mA g�1.

Results and discussion

ZnMn2–ptcda, containing Zn2+ and Mn2+ ions, was prepared by

reaction of Zn2+ and Mn2+ (the molar ratio of Zn : Mn is 1 : 2)

with perylene-3,4,9,10-tetracarboxylic dianhydride (ptcda) in

J. Mater. Chem., 2012, 22, 13328–13333 | 13329

Page 3: Spinel ZnMn2O4 nanoplate assemblies fabricated via “escape-by-crafty-scheme” strategy

Fig. 3 PXRD patterns of ZnMn2–ptcda precursor and spinel ZnMn2O4

prepared by ‘‘escape-by-crafty-scheme’’ strategy using ZnMn2–ptcda as

a precursor and by co-precipitation method (450 �C, 1 h, air atmosphere).

Publ

ishe

d on

10

May

201

2. D

ownl

oade

d on

12/

05/2

014

13:4

0:42

. View Article Online

NaOH solution under hydrothermal conditions. The scanning

electron microscopy (SEM) image of ZnMn2–ptcda clearly

shows that the sample is composed of flower-like assemblies

(particle size of ca. 4 mm) constructed of nanoplates with thick-

ness of ca. 60 nm (Fig. 2a). Energy-dispersive X-ray spectroscopy

(EDX) analysis shows that ZnMn2–ptcda is composed of carbon,

oxygen, manganese and zinc elements, and the Mn : Zn molar

ratio is�2 (Fig. S1†). The EDXmappings reveal that manganese

and zinc elements are homogeneously distributed in the whole

ZnMn2–ptcda sample (Fig. 2b). Furthermore, the elemental

distribution analysis on the selected region confirms the uniform

distribution of Zn and Mn elements (Fig. 2d). This proves that

both elements are distributed homogeneously in the ZnMn2–

ptcda particle. The transmission electron microscopy (TEM) and

selected area electron diffraction (SAED) analysis of an isolated

particle display single crystal properties with periodic diffraction

spots (Fig. 2c). The powder X-ray diffraction (PXRD) pattern of

ZnMn2–ptcda is shown in Fig. 3. ZnMn2–ptcda displays strong

diffraction peaks, exhibiting a three-dimensional framework

which is isostructural with Zn–ptcda MOFs.39 The sample has

(001) crystallographic texture, a crystalline structure with an

orthorhombic crystal system and Pbam space group. All of the

diffraction peaks can be well indexed, which indicates that

ZnMn2–ptcda is a pure phase complex. The above results show

that ZnMn2–ptcda is a pure MMOFs phase with homogeneous

distribution of Zn and Mn ions. The fabrication of nanoscale

MMOFs with mixed metal ions has not been reported previ-

ously, although the post-synthesis ion-exchange method has been

used to prepare nanoscale MMOFs.34,42 Our results indicate that

nanoscale MMOFs with controllable composition, morphology

and crystal phase can be successfully synthesized by using mixed

metal ions in the initial synthesis mixture.

By thermal treatment of ZnMn2–ptcda in air at 450 �C for 1 h,

brown powders and yellow crystals were obtained. The yellow

crystal is perylene, formed by the decarboxylation of ptcda,

Fig. 2 (a) SEM image, (b) EDX mapping SEM images, (c) TEM image

of a selected particle (inset is the SAED pattern), and (d) STEM image of

a selected region and the elemental mapping of Zn and Mn of ZnMn2–

ptcda.

13330 | J. Mater. Chem., 2012, 22, 13328–13333

showing that the thermal treatment involves the ‘‘escape-by-

crafty-scheme’’ of perylene as we previously reported.41 The

brown powders can be ascertained to be tetragonal spinel struc-

tured ZnMn2O4 with I41/amd space group (JCPDS card no. 07-

0354) by XRD characterization (Fig. 3). All the diffraction peaks

can be assigned to the spinel structure phase, indicating that high

purity spinel ZnMn2O4 has been synthesized. ZnMn2O4 has

a normal spinel structurewithZn2+ occupying the tetrahedral sites

and Mn3+ occupying the octahedral sites, forming ZnO4 groups

andMnO6 groups respectively. ZnMn2O4 is isostructural with the

spinel Mn3O4. This clearly indicates that zinc is well incorporated

into the Mn3O4 lattice without phase segregation, forming

ZnMn2O4 with the replacement of Mn2+ by Zn2+.

The SEM image reveals that the resulting ZnMn2O4 material is

composed of nanoplate assemblies inherited from the ZnMn2–

ptcda precursor with flower-like morphology (Fig. 4a). The

particle size of the resulting ZnMn2O4 is about 1.6 mm, which is

smaller than that of the ZnMn2–ptcda precursor because of the

escape of the organic portion. The nanoplates, which assemble to

form the flower-like morphology, are porous under observation

of the TEM image (Fig. 4b). Furthermore, the nanoplates are

composed of nanoparticles with a particle size of about 12 nm

(Fig. 4c). Fig. 4d displays a representative HRTEM image of the

ZnMn2O4 nanoparticles. They are well-crystallized in the

tetragonal phase with the (101), (112), and (211) crystal lattices

having interplane spacings of 0.48 nm, 0.30 nm, and 0.24 nm,

respectively. ZnMn2O4 nanoplate assemblies show the uniform

distribution of Zn and Mn in the particle with a Mn : Zn molar

ratio of �2 (Fig. 4e and S1†). The porous structure of the

ZnMn2O4 nanoplate assemblies can be further confirmed by N2

sorption analysis (Fig. 4f–g). The sample exhibits a typical type

IV isotherm with a type H3 hysteresis loop, showing that this

material has a mesoporous structure. The sample has a BET

surface area of 42 m2 g�1 and uniform distribution of the meso-

pores with a pore diameter of 17.4 nm. It should be mentioned

that the spinel structured ZnMn2O4 could also be formed even at

a treatment temperature as low as 350 �C. For comparison

This journal is ª The Royal Society of Chemistry 2012

Page 4: Spinel ZnMn2O4 nanoplate assemblies fabricated via “escape-by-crafty-scheme” strategy

Fig. 4 (a) SEM, (b) TEM, and (c–d) HRTEM images of ZnMn2O4, (e)

STEM image of a selected region of a ZnMn2O4 particle and the

elemental mapping of Zn andMn, (f) N2 adsorption–desorption isotherm

and (g) pore size distribution of ZnMn2O4.

Fig. 5 (a) TG curve and (b) PXRD pattern of ZnMn2–ptcda before and

after thermal treatment at different temperatures for 1 h in air.

Publ

ishe

d on

10

May

201

2. D

ownl

oade

d on

12/

05/2

014

13:4

0:42

. View Article Online

purposes, ZnMn2O4 was also prepared using the conventional

co-precipitation method, which is a general method for the

preparation of mixed-metal oxides.13 The spinel structure could

not be formed at treatment temperatures below 450 �C. TheXRD results show that ZnMn2O4 prepared by co-precipitation

method is not as well crystallized as that prepared using ZnMn2–

ptcda as a precursor at the same thermal treatment conditions

(Fig. 3 and S2†). This is probably due to the uniform distribution

of Zn and Mn ions in a well-defined coordination state in the

crystalline ZnMn2–ptcda precursor. This suggests the advantages

of this transformation method using MMOFs as a precursor

compared with the conventional co-precipitation method.

Through varying the Zn : Mn molar ratio in the initial

mixture, ZnMn6–ptcda MMOFs (Zn : Mn molar ratio of 1 : 6)

with a nanoplate morphology could also be prepared (Fig. S3†).

However, the nanoplates with diameter about 2 mm are

uniformly distributed with only a limited amount of flower-like

assemblies. The EDX result confirms the uniform distribution of

Zn and Mn in the nanoparticles. After thermal treatment of

This journal is ª The Royal Society of Chemistry 2012

ZnMn6–ptcda in a similar way to ZnMn2–ptcda, off-stoichio-

metric Zn0.43Mn2.57O4, with the nanoplate morphology inherited

from ZnMn6–ptcda, was formed. The XRD and N2 sorption

isotherm show that Zn0.43Mn2.57O4 has a well-crystallized spinel

structure and mesoporous textural features similar to those of

ZnMn2O4 (Fig. S4†).

The thermal transformation process from MMOFs to spinel

structured mixed-metal oxide was investigated by thermogravi-

metric (TG) analysis using ZnMn2–ptcda as a model (Fig. 5a).

ZnMn2–ptcda has two consecutive weight loss steps. The

preliminary weight loss at 150 �C corresponds to the release of

coordinated water molecules. There is a well-defined weight loss

due to the escape of perylene by decarboxylation above 345 �C.Thermal treatment of ZnMn2–ptcda at temperatures below

345 �C results in the obvious {00h} diffraction peaks in its XRD

pattern, although loss of fine structure and a lattice contraction

are observed. When the temperature reaches 350 �C, ZnMn2–

ptcda is converted into spinel structured ZnMn2O4 immediately

by perylene escaping from the MMOFs through decarboxylation

(Fig. 5b). The structural and TG results suggest that ZnMn2–

ptcda, with its robust lamellar structure, acts as a precursor for

the formation of ZnMn2O4 nanoparticles with similar

morphology. Due to the complex structure of the ZnMn2–ptcda

precursor, the exact mechanism responsible for this trans-

formation from original single crystalline ZnMn2–ptcda

MMOFs to polycrystalline ZnMn2O4 is not completely under-

stood yet, though a template-engaged quasi-topotactic trans-

formation process might be plausible.39

This ‘‘escape-by-crafty-scheme’’ strategy could be extended for

the formation of various spinel structured mixed-metal oxides

J. Mater. Chem., 2012, 22, 13328–13333 | 13331

Page 5: Spinel ZnMn2O4 nanoplate assemblies fabricated via “escape-by-crafty-scheme” strategy

Publ

ishe

d on

10

May

201

2. D

ownl

oade

d on

12/

05/2

014

13:4

0:42

. View Article Online

using MMOFs as precursors. Fig. 6 displays the SEM images of

CoMn2–ptcda andNiMn2–ptcda synthesized by a similarmethod

to ZnMn2–ptcda with a Co (or Ni) : Mn molar ratio of 1 : 2.

CoMn2–ptcda has a flower-like morphology composed of nano-

plates similar to ZnMn2–ptcda (Fig. 6a) and NiMn2–ptcda is

irregular nanolamella (Fig. 6d). EDX analyses confirmed the

uniform distribution of Co (or Ni) and Mn in the MMOFs

(Fig. S5†). CoMn2–ptcda and NiMn2–ptcda were respectively

converted into CoMn2O4 and NiMn2O4 by thermal treatment at

450 �C in air. As shown in the SEM images, the resultingmaterials

maintain the morphology of the corresponding MMOFs precur-

sors (Fig. 6b and e). The Co (or Ni) signals are observed in the

EDX spectra, confirming the presence of Co (or Ni) in the spinel

structuredCoMn2O4orNiMn2O4 (Fig. S5†). ThePXRDpatterns

only show the spinel structured diffractions without cobalt oxide

or nickel oxide impurities, suggesting that the cobalt and nickel

ions are completely doped into the Mn3O4 lattice. CoMn2O4

exhibits the typical I41/amd tetragonal spinel structure, similar to

ZnMn2O4 (Fig. 6c), and NiMn2O4 shows a cubic spinel structure

in the Fd�3 m space group (Fig. 6f).

The above results suggest that the spinel structured mixed-

metal oxides could be readily synthesized using thermal trans-

formation of MMOFs via the ‘‘escape-by-crafty-scheme’’

strategy, and the morphology of the MMOFs precursors could

be precisely transferred to the resulting spinel structured mixed-

metal oxides.

The spinel structured ZnMn2O4 nanomaterials were tested as

anode materials for rechargeable lithium ion batteries. The

discharge–charge profiles of the ZnMn2O4 nanoplate assemblies

Fig. 6 (a) SEM image of CoMn2–ptcda with a Co : Mn molar ratio of

1 : 2, (b) SEM image and (c) PXRD pattern of CoMn2O4, (d) SEM image

of NiMn2–ptcda with a Ni : Mn molar ratio of 1 : 2, (e) SEM image and

(f) PXRD pattern of NiMn2O4.

13332 | J. Mater. Chem., 2012, 22, 13328–13333

for the first two cycles were recorded at a current density of 60mA

g�1 at room temperature in the potential range between 0.01 and

3.00 V. As shown in Fig. 7a, the first discharge profile comprises

a steep line above 1.50 V, a small plateau at about 1.40 V, a wide

and steady plateau at about 0.50V, and a gradual voltage decrease

in sequence. The overall capacity for the first discharge and charge

are as high as 1277 mAh g�1 and 730 mAh g�1, respectively. An

irreversible capacity of 547 mAh g�1 was measured between the

first charge and discharge, which mainly corresponds to the

formation of the solid-electrolyte interphase (SEI) layers at

the ZnMn2O4/electrolyte interface above 1.50V and the reduction

of Mn(III) to Mn(II) at about 1.40 V.29 For the second discharge,

the potential plateau shifts upward to near 0.57 V (vs.Li+/Li) with

a more sloping profile accompanied by the disappearance of the

steep line above 1.50 V and the small plateau at about 1.40 V.

Nevertheless, the overall discharge capacity of 752mAh g�1 in the

second discharge process is retained. The coulombic efficiency of

the seconddischarge–charge process is 91%and stabilizes at>97%

after eight cycles, indicating the good reversibility of the

ZnMn2O4 anode after the first discharge–charge process. Fig. 7b

shows the cycle performance of ZnMn2O4. After 30 cycles, the

capacity decays at a slower and slower rate andmaintains a nearly

constant value of about 502 mAh g�1, demonstrating the high

reversible specific capacity and long cycle life of the anode. The

excellent electrochemical performance of ZnMn2O4might benefit

from its unique architecture of assemblies of nanoplates consist-

ing of well-crystallized agglomerated nanoparticles. The porosity

of the material will enhance the electrolyte/ZnMn2O4 contact

area, shorten the Li+ ion diffusion length, and accommodate the

strain induced by the volume change during the charge and

discharge cycles.

Fig. 7 (a) Discharge–charge curves of ZnMn2O4 nanoplate assemblies,

(b) plots of discharge capacity versus cycle number for ZnMn2O4 in the

voltage range of 0.01–3.00 V at a current rate of 60 mA g�1, and (c) rate

performances of the electrode made using ZnMn2O4.

This journal is ª The Royal Society of Chemistry 2012

Page 6: Spinel ZnMn2O4 nanoplate assemblies fabricated via “escape-by-crafty-scheme” strategy

Publ

ishe

d on

10

May

201

2. D

ownl

oade

d on

12/

05/2

014

13:4

0:42

. View Article Online

Besides the high specific capacity and good cyclability, the rate

capability is also a very important property for electrode mate-

rials. To evaluate the rate capability, the ZnMn2O4 electrode was

tested at various current densities (60–3600mA g�1) in the voltage

range of 0.01–3.0 V (Fig. 7c). The discharge capacities are grad-

ually reduced with the gradual increase of the current rate. The

capacities are 488 mAh g�1 for 600 mA g�1, 426 mAh g�1 for 900

mA g�1, and 324 mAh g�1 for 1800mA g�1. Even at the rigorously

high charge–discharge rate (3600 mA g�1), the discharge capacity

is still maintained at 187 mAh g�1. After the high rate charge–

discharge cycles, the current density is reduced to 60 mA g�1 and

a discharge capacity of as high as 578 mAh g�1 could be resumed.

Conclusions

In summary, we presented a general method via ‘‘escape-by-

crafty-scheme’’ for the preparation of spinel mixed-metal oxides

MMn2O4 (M ¼ Co, Ni, Zn) using nanoscale mixed-metal–

organic frameworks (MMOFs) as precursor. The morphology

of the MMOFs precursors could be transferred precisely to the

spinel metal oxides by a simple thermal treatment process. The

MMOFs with different kinds of metal ions distributed

uniformly and with high crystallinity favor the formation of

a highly crystallized spinel structure compared with the tradi-

tional co-precipitation method. Phase pure CoMn2O4,

NiMn2O4, and ZnMn2O4 were readily synthesized using this

method, showing the wide generality of the method. ZnMn2O4,

with a unique flower-like morphology, uniform mesopores and

high crystallinity shows high reversible capacity and good

cyclability as an anode for lithium ion batteries. This work

offers exciting possibilities for the development of new func-

tional materials for lithium ion batteries.

Acknowledgements

This work was supported by the Program Strategic Scientific

Alliances between China and The Netherlands (Grant

2008DFB50130) and the National Basic Research Program of

China (Grant 2009CB623503).

References

1 J. M. Tarascon and M. Armand, Nature, 2001, 414, 359.2 Y. Idota, T. Kubota, A. Matsufuji, Y. Maekawa and T. Miyasaka,Science, 1997, 276, 1395.

3 M. Armand and J. M. Tarascon, Nature, 2008, 451, 652.4 Y. G. Guo, J. S. Hu and L. J. Wan, Adv. Mater., 2008, 20, 2878.5 P. G. Bruce, B. Scrosati and J. M. Tarascon, Angew. Chem., Int. Ed.,2008, 47, 2930.

6 P. Poizot, S. Laruelle, S. Grugeon, L. Dupont and J. M. Tarascon,Nature, 2000, 407, 496.

7 H. Li, P. Balaya and J. Maier, J. Electrochem. Soc., 2004, 151,A1878.

8 Y. Sharma, N. Sharma, G. V. S. Rao and B. V. R. Chowdari,Electrochim. Acta, 2008, 53, 2380.

This journal is ª The Royal Society of Chemistry 2012

9 P. Lavela and J. L. Tirado, J. Power Sources, 2007, 172, 379.10 Y. Sharma, N. Sharma, G. V. S. Rao and B. V. R. Chowdari, Adv.

Funct. Mater., 2007, 17, 2855.11 Y. C. Qiu, S. H. Yang, H. Deng, L. M. Jin and W. S. Li, J. Mater.

Chem., 2010, 20, 4439.12 Y. Sharma, N. Sharma, G. V. S. Rao and B. V. R. Chowdari, J. Power

Sources, 2007, 173, 495.13 F. M. Courtel, H. Duncan, Y. Abu-Lebdeh and I. J. Davidson, J.

Mater. Chem., 2011, 21, 10206.14 K. T. Nam, D. W. Kim, P. J. Yoo, C. Y. Chiang, N. Meethong,

P. T. Hammond, Y. M. Chiang and A. M. Belcher, Science, 2006,312, 885.

15 J. Hassoun, S. Panero, P. Simon, P. L. Taberna and B. Scrosati, Adv.Mater., 2007, 19, 1632.

16 Y. G. Li, B. Tan and Y. Y. Wu, Nano Lett., 2008, 8, 265.17 S. Mitra, P. Poizot, A. Finke and J. M. Tarascon, Adv. Funct. Mater.,

2006, 16, 2281.18 S. J. Han, B. C. Jang, T. Kim, S. M. Oh and T. Hyeon, Adv. Funct.

Mater., 2005, 15, 1845.19 L. Taberna, S. Mitra, P. Poizot, P. Simon and J. M. Tarascon, Nat.

Mater., 2006, 5, 567.20 Y. Yu, C. H. Chen and Y. Shi, Adv. Mater., 2007, 19, 993.21 J. Chen, L. N. Xu, W. Y. Li and X. L. Gou, Adv. Mater., 2005, 17,

582.22 X.W. Lou, D. Deng, J. Y. Lee, J. Feng and L. A. Archer,Adv.Mater.,

2008, 20, 258.23 Y. Wang and G. Z. Cao, Adv. Mater., 2008, 20, 2251.24 Y. Wang, K. Takahashi, K. Lee and G. Z. Cao, Adv. Funct. Mater.,

2006, 16, 1133.25 X. W. Lou, Y. Wang, C. L. Yuan, J. Y. Lee and L. A. Archer,

Adv.Mater., 2006, 18, 2325.26 J. S. Chen, Y. L. Tan, C.M. Li, Y. L. Cheah, D. Y. Luan, S.Madhavi,

F. Y. C. Boey, L. A. Archer and X. W. Lou, J. Am. Chem. Soc., 2010,132, 6124.

27 L. F. Xiao, Y. Y. Yang, J. Yin, Q. Li and L. Z. Zhang, J. PowerSources, 2009, 194, 1089.

28 Y. Y. Yang, Y. Q. Zhao, L. F. Xiao and L. Z. Zhang, Electrochem.Commun., 2008, 10, 1117.

29 S. W. Kim, H. W. Lee, P. Muralidharan, D. H. Seo, W. S. Yoon,D. K. Kim and K. Kang, Nano Res., 2011, 4, 505.

30 L. Zhou, H. B.Wu, T. Zhu and X.W. Lou, J. Mater. Chem., 2012, 22,827.

31 W. Cho, Y. H. Lee, H. J. Lee and M. Oh, Chem. Commun., 2009,4756.

32 W. Cho, S. Park and M. Oh, Chem. Commun., 2011, 47, 4138.33 S. Jung, W. Cho, H. J. Lee and M. Oh, Angew. Chem., Int. Ed., 2009,

48, 1459.34 W. Cho, Y. H. Lee, H. J. Lee and M. Oh, Adv. Mater., 2011, 23,

1720.35 C. C. Li, X. M. Yin, L. B. Chen, Q. H. Li and T. H. Wang, Chem.–

Eur. J., 2010, 16, 5215.36 B. Liu, X. B. Zhang, H. Shioyama, T. Mukai, T. Sakai and Q. Xu, J.

Power Sources, 2010, 195, 857.37 H. Y. Shi, B. Deng, S. L. Zhong, L. Wang and A. W. Xu, J. Mater.

Chem., 2011, 21, 12309.38 Y. X. Lu, H. Q. Cao, S. C. Zhang and X. R. Zhang, J. Mater. Chem.,

2011, 21, 8633.39 J. Zhao,M. R. Li, J. L. Sun, L. F. Liu, P. P. Su, Q. H. Yang and C. Li,

Chem.–Eur. J., 2012, 18, 3168.40 P. Mahata, D. Sarma, C. Madhu, A. Sundaresen and S. Natarajan,

Dalton Trans., 2011, 40, 1952.41 P. Mahata, T. Aarthi, G. Madras and S. Natarajan, J. Phys. Chem. C,

2007, 111, 1665.42 M. Oh and C. A. Mirkin, Angew. Chem., Int. Ed., 2006, 45,

5492.

J. Mater. Chem., 2012, 22, 13328–13333 | 13333


Top Related