Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

NJC

View Article Online


PAPER View Journal | View Issue

Nanostructured Co(II)-based MOFs as promising


Published on 06 September 2016. Downloaded by University of Windsor on 19/07/2017 15:49:35.

anodes for advanced lithium storage†


Cite this: New J. Chem., 2016,
40, 9238
Danhua Ge,a Jie Peng,a Genlong Qu,a Hongbo Geng,a Yaoyao Deng,a Junjie Wu,a
Xueqin Cao,a Junwei Zhengab and Hongwei Gu*a

A novel kind of Co-based metal–organic frameworks (MOFs) are firstly prepared by coordination assembly
of Co(II) and 1,3,5-benzenetricarboxylic acid (H3BTC) through a hydrothermal route (referred to as Co-BTC
MOF). Such Co-BTC MOF have a zigzag vermicular-like shape with a length of 126 nm and a width of 55 nm.
Co-BTC MOFs without calcination treatment are applied as anode materials for lithium-ion batteries (LIBs)
1
and show superior electrochemical performance, including high lithium storage capacity (1739 mA h g
1
in the first discharge process at a current density of 100 mA g ), excellent coulombic efficiency (above
Received (in Montpellier, France)
1
18th August 2016, 99% in the 199th cycle) and cycling performance (4750 mA h g over 200 cycles) in the potential range
Accepted 5th September 2016 +
of 0.01–3.0 V vs. Li/Li . Such excellent electrochemical performance may be derived from the conjugated
DOI: 10.1039/c6nj02568d carboxylate moiety and the linker structure. More importantly, the facile approach may also be extended
to other metal-based MOFs and the intriguing results show that nano-sized MOFs will be regarded as
www.rsc.org/njc futuristic anode materials for advanced lithium storage.

1. Introduction As a new type of porous crystalline materials, MOFs have


received tremendous research interest owing to several advan-
In the past few years, rechargeable lithium-ion batteries (LIBs) have tages including high surface area, tunable pores, good stability
attracted intensive interest as a promising power source for and chemical functionality.1a,4 As we all know, MOFs, which are
hybrid electric vehicles (HEVs), medical devices and electrical also regarded as ‘coordination polymers’, have been constructed
energy storage (EES) due to their excellent cycling performance, from the assembly of metal ions or clusters and organic bridging
high energy density, lightweight and environmentally friendly ligands by the metal–ligand coordination mechanism along with
features.1 Graphite as a traditional anode material cannot meet intramolecular interactions to form 2D–3D networks.5 MOFs
the emerging requirements due to its low capacity of 372 mA h g 1. show a variety of applications in fields including catalysis,
As a consequence, various efforts have been devoted to designing sensing devices, energy engineering, gas adsorption, drug delivery,
novel electrode materials which could deliver higher capacities etc.6 Also, MOFs can work well as precursors and templates for
and exhibit better cycling stability and high current performance, the preparation of porous metal oxides and transition metal
such as transition metal oxides, nitrides, polymers, sulfide, oxides, which demonstrate high electrochemical performance
graphitic/nongraphitic carbon and so on.1d,f,2 In addition, organic and excellent cycling stability.7 Unfortunately, pyrolysis of MOFs
ionic plastic crystals as a new type of electrolytes could also provide resulting in metal oxides usually makes them lose their original
durability and cycling stability in fuel cells and batteries.2k–n 3D ordered networks so that the uniqueness and potential of the
Yet, preparation of most of these materials could suffer from porous materials would be reduced.8 It is demonstrated that the
laborious procedures and formation of some by-products. In 3D porous networks and high uniformity of MOFs with abun-
order to overcome these problems, we selected metal–organic dant chemical functionalities in virtue of vast structural versa-
frameworks (MOFs) as anode materials in LIBs owing to their tility can offer unique symbiotic effects and generate orthogonal
flexible synthetic routes and comparatively inexpensive precur- interactions central to realizing supramolecular self-assembly.9
sors in contrast with other porous materials.3 In addition, metal cations inside MOF structures can provide a
pathway for electrons and function as active sites for redox
a
Key Laboratory of Organic Synthesis of Jiangsu Province, College of Chemistry, behavior while the linker structure in the frameworks could
Chemical Engineering and Materials Science & Collaborative Innovation Center of promote efficient and reversible charge transfer, resulting in
Suzhou Nano Science and Technology, Soochow University, Suzhou 215123,
advanced electrochemical properties involving higher diffusion
China. E-mail: hongwei@suda.edu.cn; Fax: (+86)512-65880905
b
Institute of Chemical Power Sources, College of Physics, Optoelectronics and
rates and better electrolyte accessibility.8,10 Therefore, MOFs
Energy, Soochow University, Suzhou 215006, P. R. China have directly emerged as novel alternative lithium storage
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6nj02568d materials with special properties.

9238 | New J. Chem., 2016, 40, 9238--9244 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2016
View Article Online

Paper NJC

In recent years, MOFs have successively emerged as promising


electrodes for Li-ion batteries.10d,11 Chen’s group synthesized
polycrystalline MOF-177 provided with different morphologies
via the solvothermal method, which was the first report that the
MOF was used as an electrode for lithium storage.11a However,
MOF-177 cannot exhibit excellent electrochemical reversibility, as
a result of which, it cannot be well applied in lithium storage. Soon
after, Tarascon et al. reported MIL-53(Fe) as a battery material for Scheme 1 The preparation of the cobalt-containing metal–organic
LIBs based on Fe2+–Fe3+ redox properties with gravimetric respon- frameworks (referred to as Co-BTC MOF).
Published on 06 September 2016. Downloaded by University of Windsor on 19/07/2017 15:49:35.

sivity (75 mA h g 1) superior to that of MIL-68(Fe).11b–d The use of


HF was necessary in this case, but it could greatly hinder the further
over 200 cycles) and excellent coulombic efficiency (above 99%
development.3b Following the pioneering works, many efforts have
in the 199th cycle).
been devoted to exploring novel MOFs either as an anode or a
cathode material to investigate the electrochemical performance of
LIBs.10d Among them, formate-based MOFs such as Co3(HCOO)6 2. Results and discussion
and Zn3(HCOO)6 were efficient anode materials through conversion
Co-BTC MOF were prepared by coordination polymerization of
reaction with the capacities of 410 and 560 mA h g 1 after
Co(II) and H3BTC in DMF and ethanol. The morphology and size of
60 runs, respectively.11e More recently, a 3D Co2(OH)2BDC frame-
the resulting product are investigated by scanning electron micro-
work has been synthesized by assembling Co(II) ions with
scopy (SEM) and transmission electron microscopy (TEM) in Fig. 1.
1,4-benzenedicarboxylic acid (H2BDC) through a solvothermal
Fig. 1a and b clearly present the as-prepared Co-BTC MOF with a
route. When used as an anode in LIBs, the MOF showed large
rough surface and vermicular-like morphology. TEM images
reversible capacity (650 mA h g 1) after 100 cycles at a current
(Fig. 1c and d) further prove the existence of zigzag vermicular
density of 50 mA g 1.11f Based on the published work, MOFs can
structure. Fig. 1d (inset) further confirms the results of SEM and
be employed as the potential LIB electrode. However, it is necessary
TEM images, which show the MOF with a length of 126 nm and a
to develop novel MOFs as battery materials with better rate
width of 55 nm. In addition, we explored the influence of the
performance and higher electrochemical stability in the charge–
reaction time on the morphology and structure. When the reaction
discharge processes, especially some novel cobalt(II)-containing
time was decreased to 4 h, the morphology of the resulting product
metal–organic frameworks.
would be changed into unpurified and inhomogeneous spheres
According to the previous works, the conversion mechanism
(Fig. S1, ESI†). The chemical composition of the Co-BTC MOF is
in MOFs is usually owing to the redox behavior of transition
determined by energy dispersive X-ray spectroscopy (EDS) and
metal ions. However, 3D frameworks of MOFs would be pulver-
infrared radiation analyses (IR). The peaks of Co, O and C are
ized due to structure variations during the lithiation/delithia-
found in the EDS profile (other peaks originated from the ITO
tion process. Armand et al. have demonstrated that the weakly
substrate) (Fig. S2, ESI†).
electron-withdrawing carboxylates can be also provided with
As shown in Fig. 2a, the absence of the characteristic bands
low-voltage redox activity.12 Then, it is possible that the organic
at ca. 3084 cm 1, 1719 cm 1, and 537 cm 1 indicates the
linkers of ligands play an important role in the redox process,
in which the MOF network is sustained during cycling due to
the flexibility of organic linkers and rigidity of the inorganic
clusters. Co(II) ions as centers would be linked to 1,3,5-
benzenetricarboxylic acid (H3BTC) as organic ligands to form
a framework by condensation polymerization.13 Hence, Co(II)
acetate hydrate and H3BTC have been employed as precursors
to achieve violet Co-containing MOF, labelled as Co-BTC MOF
in this study. H3BTC is viewed as a classic ligand having
constructed a suite of MOF materials. To the best of our knowl-
edge, this is a novel Co-based MOF material with vermicular-like
structure derived from the assembly of Co(II) ions and H3BTC.
In brief, synthesis of Co-BTC MOF involves the dissolution of
polyvinylpyrrolidone (PVP) in the mixed solvent including N,N-
dimethylformamide (DMF) and ethanol, followed by the sub-
sequent addition of equimolar precursors Co(OAc)2 and H3BTC
via solvothermal treatment as shown in Scheme 1.14 Co-BTC
MOF are then characterized by different analytical techniques,
and the electrochemical performance of Co-BTC MOF as an
anode material in LIBs is investigated. The as-prepared Co-BTC Fig. 1 (a and b) SEM images, (c and d) TEM images of the as-prepared
MOF can exhibit superior cycling performance (4750 mA h g 1 Co-BTC MOF at 200 1C for 8 h (inset: the TEM image of a single MOF).

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2016 New J. Chem., 2016, 40, 9238--9244 | 9239
View Article Online

NJC Paper

to not subjecting the material to any activation process,10d,16


a suitable surface area always leads to less side reactions with
the electrolytes.16b
Thermogravimetric analysis (TGA) was performed under an air
atmosphere at a ramp rate of 10 1C min 1 and the curves reveal
three prominent weight loss steps. As shown in Fig. 2d, the TG
curve shows that the thermally stable temperature is B450 1C due
to the decomposition of tricarboxylate linkers. The breakdown of
the framework takes place at B590 1C with the formation of mixed
Published on 06 September 2016. Downloaded by University of Windsor on 19/07/2017 15:49:35.

valence Co3O4. From the TGA curve, the wt% of Co is calculated to


be 19.3. The mole ratio of Co and H3BTC was 10 : 7 based on the
results of EDS (Fig. S2, ESI†). Additionally, the first weight loss
(7.5%) is attributed to water molecules. In combination with TGA
and EDS, we surmise that the formula for the material is
Co3(BTC)23H2O. Unfortunately, we failed to achieve the accurate
structure after various efforts. As a control, we prepared Co3O4
Fig. 2 (a) FTIR spectra of Co-BTC MOF and H3BTC; (b) Co 2p XPS spectra nanoparticles through the pyrolysis of Co-BTC MOF at 500 1C in air
of the as-prepared Co-BTC MOF before and after full discharge; (c) N2
for 4 h. The SEM and TEM images of nanoparticles have shown
adsorption–desorption isotherms (inset: the BJH pore size distribution)
and (d) TG curve of Co-BTC MOF. uniformity in size (Fig. 3a). The phase purity and crystalline
structure of nanoparticles were also tested by XRD, and the
diffraction peaks were in accordance with the standard XRD data
disappearance of carbonyl groups of BTC. The signals at 1569 and card of Co3O4 (JCPDS No. 43-1003) (Fig. 3b). In addition, the EDX
1437 cm 1 can be attributed to the asymmetric and symmetric provided further information on the chemical composition and
stretching vibrations of –COO .11f,15 And more importantly, the the main peaks of Co and O were observed (Fig. S5, ESI†). XPS
appearance of the band at 766 cm 1 has resulted from ring-out-of- illustrated the chemical composition of Co3O4 nanoparticles.
plane vibration of the 1,3,5-substituted benzene.10d The FTIR The Co 2p XPS spectra in Fig. 3c exhibit two peaks at 794.9 and
spectrum of Co-BTC MOF could prove the successful coordination 779.6 eV with a spin-orbit splitting of 15.3 eV, while Fig. 3d shows
of Co(II) and H3BTC. The phase purity and crystalline structure of the O 1s peak at 529.7 eV corresponding to the oxygen species in
Co-BTC MOF are analyzed by X-ray diffraction (XRD) (Fig. S3, Co3O4 nanoparticles.
ESI†). It is obvious that well defined diffraction peaks do not The electrochemical performance of Co-BTC MOF as an anode
appear in the XRD pattern of the resulting Co-BTC MOF. The material for LIBs was evaluated by cyclic voltammograms (CVs)
solvents used in the synthesis are important, which will affect the within the potential window of 0.01–3.0 V and the first five
crystallinity.16 If the solvent is heat removed, the changes of pore successive CV scans are exhibited in Fig. 4a. In the first cycle
structures will result in significant amorphization.10d Moreover, two cathodic peaks appear at 0.64 V and 1.25 V, corresponding to
the chemical composition of Co-BTC MOF is further investigated the lithiation of the Co-BTC MOF electrode with Li. Only an
by X-ray photoelectron spectroscopy (XPS). As shown in Fig. S4 anodic peak ascribed to the delithiation reaction is recorded at
(ESI†), the XPS survey spectrum of the Co-BTC mainly shows four
dominant peaks at 285.2, 400.2, 532.2, and 781.2 eV, corres-
ponding to C 1s, N 1s, O 1s and Co 2p, respectively.17 Meanwhile,
the Co 2p peak is analysed by high-resolution XPS (Fig. 2b).
The Co 2p1/2 and Co 2p3/2 peaks of Co-BTC are at 797.0 eV and
780.9 eV with a 16.1 eV peak-to-peak separation, which illustrates
that the oxidation state of Co is maintained as +2.18 It is speculated
that the binding energy of Co 2p is slightly higher than that in the
previous research studies due to the coordination of Co2+ and
–COO . The porosity of the Co-BTC MOF is further characterized
by N2 adsorption–desorption isotherms shown in Fig. 2c. The
curve of the material describes a small hysteresis loop as a typical
type-IV isotherm. Moreover, the pore size distribution of Co-BTC
MOF was estimated by using the Barrett–Joyner–Halenda (BJH)
method utilizing the desorption part in Fig. 2c inset. It could have
resulted from the formation of the slit pores owing to the
accumulation of vermicular-like particles. Although the BET
surface area of Co-BTC MOF (18.5 m2 g 1) with a pore volume of
0.13 cm3 g 1 is not high, which could result from rearrange- Fig. 3 (a) SEM image (inset: TEM image); (b) XRD pattern; (c) Co 2p XPS
ment of pores under thermal removal of solvent as well as due spectra and (d) O 1s XPS spectra of the as-prepared Co3O4 nanoparticles.

9240 | New J. Chem., 2016, 40, 9238--9244 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2016
View Article Online

Paper NJC

cycle of Co-BTC MOF at a current density of 100 mA g 1 in the


voltage range between 0.01 and 3.0 V. The discharge and charge
capacities are 1739 and 622 mA h g 1 respectively, with a low
coulombic efficiency of 36% (Fig. 4c). Similar to most electrodes, a
large initial irreversible reaction takes place in the first cycle owing
to the decomposition of the electrolyte during the discharge
process to form a solid electrolyte interface (SEI).11e,f,19a,20 The
second-discharge curve differs from the first curve in shape with
disappearance of the plateau at about 0.7 V. The second discharge
Published on 06 September 2016. Downloaded by University of Windsor on 19/07/2017 15:49:35.

and charge capacities are 562 and 445 mA h g 1 respectively, and


the coulombic efficiency is remarkably increased to 79% (Fig. 4c).
As seen in Fig. 4c, it is obvious that the discharge capacity begins
to slowly increase and remains stable at above 800 mA h g 1 after
the initial 11 cycles. Even after 200 runs, the reversible capacity
could remain above 750 mA h g 1. The increasing specific capacity
Fig. 4 (a) Cyclic voltammograms (CVs); (b) voltage profiles at various
cycles; (c) cycle life (black) and coulombic efficiencies (red) of electrodes presented may be on account of the sluggish amorphization and
of the as-prepared Co-BTC MOF at a current density of 100 mA g 1; equilibration of the Co-BTC MOF electrode. The coulombic
(d) cycling performance of the Co-BTC MOF and Co3O4 nanoparticles at efficiency achieved is as high as 499% after 12 cycles, which
a current density of 100 mA g 1 after 60 cycles. illustrates that the material delivers an excellent reversible
capacity for long term cycling. The above results illustrate the
superior cycle life and high capacity of Co-BTC MOF, which are
0.98 V. A large irreversible lithiation exists in the first cycle, which of importance to explain the electrochemical behavior of the
could be due to the formation of a solid electrolyte interface electrode material. In addition, we studied the electrochemical
(SEI).11f The other peaks could be attributed to some side performance of Co-BTC MOF and Co3O4 nanoparticles as anodes
reactions, and a similar kind of situation has also been reported in LIB at a current density of 100 mA g 1 after 60 cycles (Fig. 4d).
in some related literature studies.11g In the second cycle, the It is found that the Co-BTC MOF delivers a higher reversible
cathodic peaks are shifted to 0.72 and 1.28 V, respectively, and discharge capacity than Co3O4 nanoparticles, which further
the main anodic peak remains unchanged. Yet, the peak intensity proves its superiority in LIB. Additionally, the TEM image of
is somewhat lower than the initial cycle. The integral areas and the Co-BTC MOF electrode shows that the MOF is still ‘‘alive’’
peak intensity of the subsequent cycles are the same as that of the after cycling (Fig. S6, ESI†).
second cycle. The above results illustrate the electrochemical The electrochemical behaviors of Co-BTC MOF are different
reversibility of the Co-BTC MOF after the first cycle. However, it from those of most reported electrodes during the Li insertion/
is difficult to ascertain the Li storage mechanism through the CV desertion process, which involves physical change and electro-
method. A series of cobalt oxides with different structures have chemical reduction/oxidation.11a When comparing the capacity
been studied as electrodes for lithium insertion/deinsertion to of the first and second cycles (Fig. 4b), it is noted that the capacity
demonstrate high performance.19 So, we were interested in inves- deterioration reaches as high as two-thirds. To the best of our
tigating the possibility of employing the as-prepared Co-BTC MOF knowledge, the large number of cavities in Co-BTC MOF and
as lithium storage material.11a The electrochemical performance guest molecules such as DMF and H2O could contribute to high
of the sample was evaluated as an anode for LIBs using the capacity in the first discharge. During the discharge/charge
standard MOF/Li half-cell configuration, in which lithium foil and process two kinds of lithium (intercalated Li+ and resultant Li)
MOF work as negative and positive electrodes, respectively. Fig. 4b entering and staying in the cavities can react with DMF and H2O
shows the discharge–charge curves of the material as electrodes in containing active protons. The reaction process could receive or
the 1st, 2nd, 10th, 50th, 100th and 150th cycles over the potential release electrons from the anode via the outer circuit. An irrever-
range of 0.01–3.0 V vs. Li/Li+ at a current density of 100 mA g 1. sible electrochemical process only occurs in the first discharge
The discharge capacities of the composite in the first and second with high capacity. In the presence of a higher content of guest
cycles are 1739 and 562 mA h g 1 respectively. The theoretical molecules, a larger capacity will be achieved.11a Co-BTC MOF
capacity of one Li+ insertion is 85 mA h g 1, so the Co-BTC MOF could also provide reactive sites for the electrode process due to
molecule is able to accept about 20 Li ions at a current density of the consumption of electrons. The capacity after the 10th cycle
100 mA g 1 due to the discharge capacity of 1739 mA h g 1 in the slowly increases, which may be derived from the delayed wetting
first cycle, while about 7 Li ions could be inserted in each Co-BTC of the electrolyte.21 Sequentially, we explored the Li storage
MOF molecule in the second cycle.11f Before the 10th cycle, the mechanism by XPS measurements of the electrode under the
discharge capacity is decreased to 308 mA h g 1. However, the fully discharged condition. Fig. 2b (red) confirms the presence
discharge capacity picks up in the following cycles until the 100th of Co2+ state due to the binding energy positions and binding
cycle with a discharge capacity of 865 mA h g 1, while in the 150th energy satellites. Armand et al. have reported that the weakly
cycle the discharge capacity shows a slight reduction. Fig. 4c electron withdrawing ligands could provide redox centers for
demonstrates the corresponding discharge/charge capacity versus the coordination of COO and Li. As shown in Scheme 2, it is

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2016 New J. Chem., 2016, 40, 9238--9244 | 9241
View Article Online

NJC Paper

PAN-alytical) with a Cu Ka X-ray source (l = 1.540598 Å).


Thermogravimetric analysis (TGA) was carried out on a Perkin
Elmer TGA 4000 thermogravimetric analyser. IR spectra were
measured on a Nicolet 5700 FTIR, using KBr pellets. XPS data
were obtained using a KRATOS Axis ultra-DLD X-ray 90 photo-
electron spectrometer with a monochromatic Mg Ka X-ray
Scheme 2 The probable lithiation/delithiation sites for coordination with
source (1283.3 eV). The specific surface area was calculated
Li in the as-prepared cobalt-containing metal–organic frameworks (Co-BTC
MOF).
by the Brunauer–Emmett–Teller (BET) method. The pore size
distribution was calculated by the Barret–Joyner–Halenda
Published on 06 September 2016. Downloaded by University of Windsor on 19/07/2017 15:49:35.

(BJH) method.
possible that the organic moiety is involved in the Li storage
mechanism through Li deinsertion from the COO groups and 4.2 Synthesis of Co-BTC MOF
benzene rings.12 Therefore, it is very necessary and challenging The zigzag MOF was prepared by a hydrothermal method
to design novel kinds of MOFs as electrode materials for lithium according to a previous study.12 Polyvinylpyrrolidone (PVP K15,
storage due to the irreversible capacity in the initial discharge 3 g) was dissolved in the mixed solvent of 96 mL DMF and 48 mL
process and further research is needed to comprehend the exact ethanol. Then equimolar (0.25 mmol) Co(OAc)24H2O and H3BTC
mechanism. were added into the above transparent solution under ultrasonic
conditions for 20 min. Afterwards, the purple solution was
transferred into a 200 mL Teflon-lined stainless steel autoclave
3. Conclusions and heated at 200 1C for 8 h. Subsequently, the product was
In summary, we have successfully prepared a novel kind of obtained by centrifugation at 7000 rpm for 5 min by washing
Co-containing metal–organic frameworks (Co-BTC MOF) via with ethanol for three times. The violet solids were soaked in
solvothermal treatment, which is firstly applied as an anode CH2Cl2 for 12 h to remove the guest molecules. The Co3O4
material in lithium-ion batteries. Co(II) ions as centers are linked to nanoparticles were prepared by pyrolysis of the Co-BTC MOF in
1,3,5-benzenetricarboxylic acid (H3BTC) as organic ligands to form air at 500 1C.
a framework by condensation polymerization. The composite
4.3 Electrochemical measurements
without calcination treatment as an anode material delivers
high lithium storage capacity in the first discharge process The electrochemical measurements of the hollow and porous
(1739 mA h g 1 at a current density of 100 mA g 1) due to the products were carried out using a coin-type half cell (CR 2016).
existence of guest molecules and many cavities. And nano-sized The working electrode was prepared by mixing 80 wt% active
MOF could provide reactive sites for the electrode process. material, 10 wt% conductive material (acetylene black, AB) and
Superior cycling performance (4750 mA h g 1 over 200 cycles) 10 wt% polymer binder (polyvinylidenedifluoride, PVDF) with the
and excellent coulombic efficiency (above 99% in the 199th N-methyl-2-pyrrolidone (NMP) solvent to form a homogeneous
cycle) could be achieved. In addition, it is found that the linker slurry, followed by spreading onto copper foil. Finally, the copper
structure promotes efficient and reversible charge transfer and foil was dried in a vacuum at 120 1C overnight. The electrolyte
the conjugated carboxylates possess a stronger p–p interaction, was 1 M LiPF6 in a mixed solution of ethylene carbonate (EC) and
which both appear to play a key role in the excellent Li storage diethyl carbonate (DEC) in the ratio of EC–DEC = 1 : 1 (w/w,
mechanism. Further studies on structure transformation and provided by Guotaihuarong, Zhangjiagang, China). Cell assembly
the inner electrochemical mechanism are under way. These results was performed in an argon-filled glovebox in the presence of
make Co-BTC MOF a promising anode in LIBs, which will provide moisture and oxygen below 0.1 ppm. The cells were charged and
opportunities for design and development of next generation high- discharged between 3.0 V and 0.01 V at room temperature.
performance electrodes in LIBs.
Acknowledgements
4. Experimental section H. W. G. acknowledges financial support from the National
4.1 Material characterization Natural Science Foundation of China (No. 21373006). The project
IR spectra were recorded on a Nicolet 5700 FTIR using KBr was funded by the Priority Academic Program Development of
pellets. The morphologies and sizes were examined by scanning Jiangsu Higher Education Institutions (PAPD).
electron microscopy (SEM) and transmission electron microscopy
(TEM). SEM was performed on a Hitachi S-4700 cold field- References
emission scanning electron microscope operated at 30 KV, and
TEM was performed on a TecnaiG220 microscope (FEI, USA) 1 (a) S. J. Yang, S. Nam, T. Kim, J. H. Im, H. Jung, J. H. Kang,
fitted with a Gatan CCD794 camera operated at 200 KV. The S. G. Wi, B. Park and C. R. Park, J. Am. Chem. Soc., 2013, 135,
as-prepared Co-BTC MOF were characterized by X-ray diffraction 7394–7397; (b) P. G. Bruce, B. Scrosati and J.-M. Tarascon,
(XRD) using an X’Pert-Pro MPD diffractometer (Netherlands Angew. Chem., Int. Ed., 2008, 47, 2930–2946; (c) B. Scrosati,

9242 | New J. Chem., 2016, 40, 9238--9244 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2016
View Article Online

Paper NJC

J. Hassoun and Y.-K. Sun, Energy Environ. Sci., 2011, 4, X. Wang, Angew. Chem., Int. Ed., 2014, 53, 1034–1038;
3287–3295; (d) C. Li, T. Q. Chen, W. J. Xu, X. B. Lou, (e) T. Zhang and W. Lin, Chem. Soc. Rev., 2014, 43,
L. K. Pan, Q. Chen and B. W. Hu, J. Mater. Chem. A, 2015, 5982–5993; ( f ) C. C. Sun, J. Yang, X. H. Rui, W. N. Zhang,
3, 5585–5591; (e) B. Scrosati, Nature, 1995, 373, 557–558; Q. Y. Yan, P. Chen, F. W. Huo, W. Huang and X. C. Dong,
( f ) P. Poizot, S. Laruelle, S. Grugeon, L. Dupont and J.-M. J. Mater. Chem. A, 2015, 3, 8483–8488; ( g) D. Farrusseng,
Tarason, Nature, 2000, 407, 496–499; (g) R. F. Nelson, J. Power S. Aguado and C. Pinel, Angew. Chem., Int. Ed., 2009, 48,
Sources, 2000, 91, 2–26; (h) J.-M. Tarascon and M. Armand, 7502–7513; (h) O. M. Yaghi, H. L. Li and T. L. Groy, J. Am.
Nature, 2001, 414, 359–367; (i) R. A. Huggins, Solid State Ionics, Chem. Soc., 1996, 118, 9096–9101; (i) F. F. Cao, M. T. Zhao,
2002, 152, 61–68; ( j) B. Guo, C. S. Li and Z.-Y. Yuan, J. Phys. Y. F. Yu, B. Chen, Y. Huang, J. Yang, X. H. Cao, Q. P. Lu,
Published on 06 September 2016. Downloaded by University of Windsor on 19/07/2017 15:49:35.

Chem. C, 2010, 114, 12805–12817; (k) N. Yuca, H. Zhao, X. Zhang, Z. C. Zhang, C. L. Tan and H. Zhang, J. Am. Chem.
X. Y. Song, M. Dogdu, W. Yuan, Y. B. Fu, V. S. Battaglia, X. C. Soc., 2016, DOI: 10.1021/jacs.6b02540.
Xiao and G. Liu, ACS Appl. Mater. Interfaces, 2014, 6, 5 (a) S. B. Wang and X. C. Wang, Small, 2015, 11, 3097–3112;
17111–17118; (l ) Y. M. Chen, L. Yu and X. W. (David) Lou, (b) A. Banerjeea, U. Singha, V. Aravindanc, M. Srinivasanc
Angew. Chem., Int. Ed., 2016, 55, 5990–5993; (m) N. Yan, and S. Ogale, Nano Energy, 2013, 2, 1158–1163.
F. Wang, H. Zhong, Y. Li, Y. Wang, L. Hu and Q. W. Chen, 6 (a) A. R. Millward and O. M. Yaghi, J. Am. Chem. Soc., 2005,
Sci. Rep., 2013, 3, 1568. 127, 17998–17999; (b) E. D. Bloch, W. L. Queen, R. Krishna,
2 (a) J. Cabana, L. Monconduit, D. Larcher and M. R. Palacin, J. M. Zadrozny, C. M. Brown and J. R. Long, Science, 2012, 335,
Adv. Mater., 2010, 22, E170–E192; (b) J. S. Wu, X. H. Rui, 1606–1610; (c) B. Chen, L. Wang, Y. Xiao, F. R. Fronczek,
G. K. Long, W. Q. Chen, Q. Y. Yan and Q. C. Zhang, Angew. M. Xue, Y. Cui and G. Qian, Angew. Chem., Int. Ed., 2009, 48,
Chem., Int. Ed., 2015, 54, 7354–7358; (c) J.-M. Tarascon, Nat. 500–503; (d) C. Wang, T. Zhang and W. Lin, Chem. Rev., 2012,
Chem., 2010, 2, 510; (d) P. Poizot and F. Dolhem, Energy 112, 1084–1104; (e) M. D. Allendorf, C. A. Bauer, R. K. Bhakta
Environ. Sci., 2011, 4, 2003–2019; (e) L.-X. Yuan, Z.-H. Wang, and R. J. T. Houk, Chem. Soc. Rev., 2009, 38, 1330–1352;
W.-X. Zhang, X.-L. Hu, J.-T. Chen, Y.-H. Huang and ( f ) J. D. Rocca, D. Liu and W. Lin, Acc. Chem. Res., 2011, 44,
J. B. Goodenough, Energy Environ. Sci., 2011, 4, 269–284; 957–968; (g) C. D. Wu, A. Hu, L. Zhang and W. Lin, J. Am.
( f ) W. Yuan, M. Y. Wu, H. Zhao, X. Y. Song and G. Liu, Chem. Soc., 2005, 127, 8940–8941.
J. Power Sources, 2015, 282, 223–227; (g) H. Su, Y. W. Li, 7 (a) S. J. Yang, S. Nam, T. Kim, J. H. Im, H. Jung, J. H. Kang,
K. Yue, Z. Wang, P. T. Lu, X. Y. Feng, X.-H. Dong, S. Zhang, S. Wi, B. Park and C. R. Park, J. Am. Chem. Soc., 2013, 135,
S. Z. D. Cheng and W.-B. Zhang, Polym. Chem., 2014, 5, 7394–7397; (b) R. Li, X. Ren, X. Feng, X. Li, C. Hu and
3697–3706; (h) S. H. Xuan, F. Wang, Y.-X. J. Wang, J. C. Yu B. Wang, Chem. Commun., 2014, 50, 6894–6897; (c) S. Qiu,
and K. C.-F. Leung, J. Mater. Chem., 2010, 20, 5086–5094; M. Xue and G. Zhu, Chem. Soc. Rev., 2014, 43, 6116–6140;
(i) L. Hu, H. Zhong, X. R. Zheng, Y. M. Huang, P. Zhang and (d) B. Kong, J. Tang, Z. Wu, J. Wei, H. Wu, Y. Wang, G. Zheng
Q. W. Chen, Sci. Rep., 2012, 2, 986; ( j) Y. Pan, J. H. Gao, and D. Zhao, Angew. Chem., Int. Ed., 2014, 53, 2888–2892;
B. Zhang, X. X. Zhang and B. Xu, Langmuir, 2010, 26, 4184; (e) T. Y. Ma, S. Dai, M. Jaroniec and S. Z. Qiao, J. Am. Chem.
(k) D. R. MacFarlane and M. Forsyth, Adv. Mater., 2001, Soc., 2014, 136, 13925–13931.
13, 957; (l ) J. S. Luo, O. Conrad and I. F. J. Vankelecom, 8 T. An, Y. H. Wang, J. Tang, Y. Wang, L. J. Zhang and G. F.
J. Mater. Chem. A, 2013, 1, 2238–2247; (m) J. S. Luo, Zheng, J. Colloid Interface Sci., 2015, 445, 320–325.
A. H. Jensen, N. R. Brooks, J. Sniekers, M. Knipper, D. Aili, 9 (a) A. K. Chaudhari, I. Han and J.-C. Tan, Adv. Mater., 2015,
Q. F. Li, B. Vanroy, M. Wübbenhorst, F. Yan, L. V. Meervelt, 27, 4438–4446; (b) M. R. Ryder and J. C. Tan, Mater. Sci.
Z. G. Shao, J. H. Fang, Z.-H. Luo, D. E. D. Vos, K. Binnemans Technol., 2014, 30, 1598–1612; (c) S. Furukawa, J. Reboul,
and J. Fransaer, Energy Environ. Sci., 2015, 8, 1276–1291; S. Diring, K. Sumida and S. Kitagawa, Chem. Soc. Rev., 2014,
(n) X. L. Chen, H. L. Tang, T. Putzeys, J. Sniekers, 43, 5700–5734; (d) A. Heeres, C. van der Pol, M. C. A. Stuart,
M. Wübbenhorst, K. Binnemans, J. Fransaer, D. E. D. Vos, A. Friggeri, B. L. Feringa and J. van Esch, J. Am. Chem. Soc.,
Q. F. Li and J. S. Luo, J. Mater. Chem. A, 2016, 4, 12241–12252; 2003, 125, 14252–14253; (e) H. Hofmeier and U. S. Schubert,
(o) D. Ge, J. Wu, G. Qu, Y. Deng, H. Geng, J. Zheng, Y. Pan and Chem. Commun., 2005, 2423–2432.
H. Gu, Dalton Trans., 2016, 45, 13509–13513. 10 (a) Y. Kobayashi, B. Jacobs, M. D. Allendorf and J. R. Long,
3 (a) S. E. Burkhardt, J. Bois, J. M. Tarascon, R. G. Hennig and Chem. Mater., 2010, 22, 4120–4122; (b) L. Kang, S.-X. Sun, L.-B.
H. D. Abruna, Chem. Mater., 2013, 25, 132–141; (b) J. W. Shin, Kong, J.-W. Lang and Y.-C. Luo, Chin. Chem. Lett., 2014, 25,
M. Kim, J. Cirera, S. Chen, G. J. Halder, T. A. Yersak, 957–961; (c) J. N. Behera, D. M. D’Alessandro, N. Soheilnia and
F. Paesani, S. M. Cohen and Y. S. Meng, J. Mater. Chem. A, J. R. Long, Chem. Mater., 2009, 21, 1922–1926; (d) S. Maiti,
2015, 3, 4738–4744; (c) Q. H. Yang, Q. Xu, S.-H. Yu and A. Pramanik, U. Manju and S. Mahanty, ACS Appl. Mater.
H.-L. Jiang, Angew. Chem., Int. Ed., 2016, 55, 3685–3689. Interfaces, 2015, 7, 16357–16363.
4 (a) Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu and Z. Li, 11 (a) X. X. Li, F. Y. Cheng, S. N. Zhang and J. Chen, J. Power
Angew. Chem., Int. Ed., 2012, 51, 3364–3367; (b) J. L. Wang, Sources, 2006, 160, 542–547; (b) G. Ferey, F. Millange,
C. Wang and W. Lin, ACS Catal., 2012, 2, 2630–2640; M. Morcrette, C. Serre, M.-L. Doublet, J.-M. Greneche and
(c) S. L. Li and Q. Xu, Energy Environ. Sci., 2013, 6, J.-M. Tarascon, Angew. Chem., Int. Ed., 2007, 46, 3259–3263;
1656–1683; (d) S. Wang, W. Yao, J. Lin, Z. Ding and (c) F. Alexandra, H. Patricia, D. Thomas, S. Christian,

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2016 New J. Chem., 2016, 40, 9238--9244 | 9243
View Article Online

NJC Paper

M. Jerome, G. Jean-Marc, M. Mathieu, T. Jean-Marie, Pan, Q. Chen, M. Shen and B. W. Hu, RSC Adv., 2016, 6,
M. Guillaume and F. Gerard, Eur. J. Inorg. Chem., 2010, 61319–61324.
3789–3794; (d) J. Overgaard, F. K. Larsen, B. Schiott and 17 H. B. Geng, Q. Zhou, Y. Pan, H. W. Gu and J. W. Zheng,
B. B. Iversen, J. Am. Chem. Soc., 2003, 125, 11088–11099; Nanoscale, 2014, 6, 3889–3894.
(e) K. Saravanan, M. Nagarathinam, P. Balaya and J. J. Vittal, 18 (a) S. C. Petitto, E. M. Marsh, G. A. Carson and M. A. Langell,
J. Mater. Chem., 2010, 20, 8329–8335; ( f ) L. Gou, L.-M. Hao, J. Mol. Catal. A: Chem., 2008, 281, 49–58; (b) X. Wang, X.-L.
Y.-X. Shi, S.-L. Ma, X.-Y. Fan, L. Xu, D.-L. Li and K. Wang, Wu, Y.-G. Guo, Y. T. Zhong, X. Q. Cao, Y. Ma and J. N. Yao,
J. Solid State Chem., 2014, 210, 121–124; ( g) H. Ming, J. Ming, Adv. Funct. Mater., 2010, 20, 1680–1686; (c) Y. M. Sun,
S.-M. Oh, S. Tian, Q. Zhou, H. Huang, Y.-K. Sun and X. L. Hu, W. Luo and Y. H. Huang, J. Phys. Chem. C, 2012,
Published on 06 September 2016. Downloaded by University of Windsor on 19/07/2017 15:49:35.

J. W. Zheng, ACS Appl. Mater. Interfaces, 2014, 6, 15499–15509. 116, 20794–20799; (d) J. F. Moulder, W. F. Sticlke, P. E. Sobol
12 M. Armand, S. Grugeon, H. Vezin, S. Laruelle, P. Ribiere, and K. D. Bomben, in Handbook of X-ray Photoelectron
P. Poizot and J.-M. Tarascon, Nat. Mater., 2009, 8, 120–125. Spectroscopy, ed. J. Chastain, PerkinElmer Corporation,
13 N. L. Rosi, J. Kim, M. Eddaoudi, B. L. Chen, M. O’Keeffe and Eden Prairie, MN, 1992.
O. M. Yaghi, J. Am. Chem. Soc., 2005, 127, 1504–1518. 19 (a) D. H. Ge, H. B. Geng, J. Q. Wang, J. W. Zheng, Y. Pan, X. Q.
14 M. T. Zhao, K. Deng, L. C. He, Y. Liu, G. D. Li, H. J. Zhao and Cao and H. W. Gu, Nanoscale, 2014, 6, 9689–9694; (b) Z. S. Wu,
Z. Y. Tang, J. Am. Chem. Soc., 2014, 136, 1738–1741. W. Ren, L. Wen, L. Gao, J. Zhao, Z. Chen, G. Zhou, F. Li and
15 (a) Y.-Y. Liu, J. Zhang, L.-X. Sun, F. Xu, W.-S. You and H. M. Cheng, ACS Nano, 2010, 4, 3187–3194; (c) X. X. Zou, J. Su,
Y. Zhao, Inorg. Chem. Commun., 2008, 11, 396–399; R. Silva, A. Goswami, B. R. Satheab and T. Asefa, Chem.
(b) M. Kurmoo, H. Kumagai, M. A. Green, B. W. Lovett, Commun., 2013, 49, 7522–7524.
S. J. Blundell, A. Ardavan and J. Singleton, J. Solid State 20 (a) A. K. Rai, J. Gim, L. T. Anh and J. Kim, Electrochim. Acta,
Chem., 2001, 159, 343–351; (c) K. Nakamoto, Infrared and 2013, 100, 63–71; (b) J. S. Chen, T. Zhu, Q. H. Hu, J. J. Gao,
Raman Spectra of Inorganic and Coordination Compounds, F. B. Su, S. Z. Qiao and X. W. (David) Lou, ACS Appl. Mater.
Wiley, New York, 1986. Interfaces, 2010, 2, 3628–3635.
16 (a) H. Reinsch and N. Stock, CrystEngComm, 2013, 15, 21 H. Wu, G. H. Yu, L. J. Pan, N. A. Liu, M. T. McDowell,
544–555; (b) T. Li, C. Li, X. S. Hu, X. B. Lou, H. P. Hu, L. K. Z. N. Bao and Y. Cui, Nat. Commun., 2013, 4, 1943.

9244 | New J. Chem., 2016, 40, 9238--9244 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2016

You might also like