You are on page 1of 5

Journal of Electroanalytical Chemistry 688 (2013) 93–97

Contents lists available at SciVerse ScienceDirect

Journal of Electroanalytical Chemistry


journal homepage: www.elsevier.com/locate/jelechem

Reversible Li and Na storage behaviors of perylenetetracarboxylates as organic


anodes for Li- and Na-ion batteries
R.R. Zhao, Y.L. Cao, X.P. Ai, H.X. Yang ⇑
College of Chemistry and Molecular Science, Wuhan University, Wuhan 430072, China

a r t i c l e i n f o a b s t r a c t

Article history: Rechargeable batteries using organic materials have potential advantages of low cost and materials sus-
Available online 28 July 2012 tainability for large-scale electric storage applications. The key issue to realize such sustainable batteries
is to develop suitable organic electrode materials with sufficient redox capacity and cycling stability.
Keywords: Herein, we introduce lithium and sodium salt of 3, 4, 9,10-perylenetetracarboxylic acid (Li4C24H8O8,
Organic anodes Na4C24H8O8) as new organic anode materials for Li-ion and Na-ion batteries. The Li4C24H8O8 electrode
Perylenetetracarboxylates can deliver a reversible capacity of 200 mAh g 1 at quite low charge/discharge plateaus of 1.20/
Na-ion batteries
1.10 V, and remains 98% of its initial capacity after 100 cycles. Similarly, the Na4C24H8O8 electrode exhib-
Li-ion batteries
its a reversible Na storage capacity of 100 mAh g 1 at a low voltage region of 0.8–0.6 V (vs Na) with
almost indiscernible capacity decay during 100 cycles. These results demonstrate a potential possibility
to use the organic anodes for Li-ion and Na-ion batteries.
Ó 2012 Elsevier B.V. All rights reserved.

1. Introduction [7–9]. Particularly, there is only one report published very recently
on organic Na-storage anode [10].
Large-scale electric storage applications, such as electric vehi- In this work, we synthesized lithium and sodium perylenetetra-
cles and renewable power stations, require rechargeable batteries carboxylates (Li4C24H8O8, Na4C24H8O8) and investigated the
not only to have sufficient energy density but also to have low cost reversible bonding properties of Li and Na ions with the carboxyl-
and materials sustainability [1–3]. However, existing rechargeable ate groups of the molecules. Selection of lithium and sodium per-
batteries, either conventional Pb and Ni batteries or advanced Li- ylenetetracarboxylates as target molecules is based on the
ion batteries, are built up with transition-metal compounds, which considerations that the carboxylate groups conjugated with an aro-
lead to severe problems of cost and environmental concern. There- matic core can allow the insertion and deinsertion of Li ions
fore, it is a great challenge to develop new redox-active materials [4,11,12] and the aromatic core with larger conjugated perylene
with minimum environmental impact and abundant natural re- ring would increase the chemical stability with decreased dissolu-
sources for meeting the requirement of future large scale electric bility in organic electrolytes [13,14]. This paper describes Li- and
storage applications. Na-storage behaviors of the perylenetetracarboxylates and demon-
In the pursuit of such new materials, considerable efforts have strates their charge–discharge properties as organic anodes for Li-
been oriented to the development of organic electrode materials and Na-ion batteries.
as a substitute for the presently used transition-metal electrodes.
Aside from abundant resources, organic electrode materials also
2. Experimental
have several advantages of chemical diversity, tunable redox prop-
erty, mechanical flexibility and possible high energy density, offer-
2.1. Material synthesis and structural characterization
ing a wide selection for battery applications. In recent years, a large
number of organic molecules and polymers have been demon-
Li4C24H8O8 and Na4C24H8O8 were synthesized from 3, 4, 9, 10-
strated to give quite a high capacity and cycling stability as elec-
perylenetertracarboxylic acid dianhydride (PAD). The mixture of
trode-active materials for Li-ion batteries [4–6]. However,
5 g KOH and 5 g PAD in 100 ml redistilled water were stirred at re-
organic electrode materials for Na-ion batteries are rarely explored
flux for 12 h, and then cooled to room temperature. A dark red floc
except for a few recent works on polymeric Na-storage cathodes
was precipitated after injection of 10% sulfuric acid into the
reaction mixture. The red precipitate was collected through vac-
uum filtration and intensive washing with water and ethanol,
⇑ Corresponding author. Tel.: +86 027 68754526; fax: +86 027 87884476. and then dried in a vacuum to give perylene-3, 4, 9, 10-tetracarb-
E-mail address: hxyang@whu.edu.cn (H.X. Yang). oxylic acid as dark red powder. Both Li4C24H8O8 and Na4C24H8O8

1572-6657/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jelechem.2012.07.019
94 R.R. Zhao et al. / Journal of Electroanalytical Chemistry 688 (2013) 93–97

Scheme 1. Synthetic route of Li4C24H8O8 (or Na4C24H8O8 using NaOH instead of LiOH).

Fig. 2. (a) The charge/discharge profiles of the Li4C24H8O8 electrode at a voltage


range of 3.0–0.6 V (vs Li). (b) The charge/discharge profiles of the Na4C24H8O8
Fig. 1. (a) CV curves of the Li4C24H8O8 electrode in 1 mol L 1 LiPF6 + EC + DM- electrode at a voltage range of 2.8–0.3 V (vs Na). The data were derived from 2032
C + EMC (1: 1:1, by volume), scanned at 1 mV s 1. (b) CV curves of the Na4C24H8O8 type coin cells cycled at a constant rate of 25 mA g 1 and the capacities were
electrode in 1 mol L 1 NaPF6 + EC + EMC (1: 1, by volume), scanned at 0.05 mV s 1. calculated on the basis of the mass of Li4C24H8O8 (Na4C24H8O8).

were prepared by reaction of the PAD precursor with LiOHH2O or (wt.%), prepared by roll-pressing the mixture into ca. 0.1 mm
NaOH in 20% excess in methanol (80 mL) at 80 °C for 24 h. The thick electrode film and pressing the films onto a porous nickel foil.
products were obtained by vacuum filtration, washed with metha- The electrolytes were 1 mol L 1 LiPF6 in a mixed solvent of ethyl-
nol three times and then dried at 100 °C under vacuum. The syn- ene carbonate (EC), dimethyl carbonate (DMC) and ethyl methyl
thetic reactions can be described in Scheme 1 as follows: carbonate (EMC) (EC: DMC: EMC = 1:1:1 by volume) (Zhangjiagang
Guotai-Huarong New Chemical Materials Co., Ltd., China) and a
2.2. Electrochemical measurements NaPF6-saturated EC + DEC electrolyte (EC: DEC = 1:1, v/v). The
charge–discharge measurements were carried out using 2032 type
The Li4C24H8O8 (Na4C24H8O8) electrode for electrochemical coin cells. The cells were assembled in an argon-filled glove
characterization were consisted of 70% the anode-active materials, box, using a Li or Na metal disk as the counter electrodes and a
10% acetylene black and 10% carbon black (Super P) and 10% PTFE porous polypropylene film as separator (Cellgard 2400). The cyclic
R.R. Zhao et al. / Journal of Electroanalytical Chemistry 688 (2013) 93–97 95

0.3–2.4 V (vs Na/Na+) at a scan rate of 0.05 mV s 1. All the CV mea-


surements were carried out and recorded on an electrochemical
workstation (CHI600c, Shanghai, China). The charge/discharge
experiments were conducted on a Neware battery testing system
(BTS-53, Shenzhen, China) in the potential range of 0.6–3.0 V vs
Li/Li+ for the Li4C24H8O8 electrode at constant currents of 25, 50
and 100 mA g 1 or in a potential range of 0.3–2.8 V vs Na/Na+ for
the Na4C24H8O8 electrode at constant currents of 25 mA g 1.

2.3. Determination of the amount of cycled Na+ ions

To quantify the amount of the sodium ions cycled at different


states of charge and discharge, the coin cells at different depths
of charge and discharge were disassembled in a glove box. The
electrode samples were taken from the cells and rinsed with pure
diethyl carbonate solvent for several times. Then the electrode
samples were burnt in air at 600 °C for 6 h to give a residue of so-
dium oxides. The Na residues were dissolved in 100 mL 0.5 mol L 1
HCl for the atomic emission analysis using IRIS Intrepid III XSP
spectrometer.

3. Results and discussion

Both the as-prepared Li4C24H8O8 and Na4C24H8O8 samples ap-


pear as brown powders, almost dissoluble in most organic sol-
vents. The chemical structures of these two samples were
confirmed by 1H NMR (300 MHz, D2O) spectra of Li4C24H8O8:
d7.58 (d, J = 7.8, 4H), d8.20 (d, J = 7.5, 4H); and Na4C24H8O8:
d7.736 (d, J = 7.8, 4H), d 8.354 (d, J = 7.5, 4H). The FT-IR spectra of
these two samples show all the characteristic vibrations of salts
of perylenetetracarboxylic acid at 1590 cm 1, 1555 cm 1 (aromatic
stretching bands), and 1635 cm 1, 1425 cm 1 (stretching vibration
of ACOO ), further confirming the chemical structure of Li- and
Na-carboxylate groups conjugated with perylene ring.
Fig. 3. (a) Cycling performance of the Li4C24H8O8/Li cell at changing current rates as Fig. 1 shows the cyclic voltammograms (CVs) of the Li4C24H8O8
labeled on the capacity-retention curve. (b) Cycling curves of the Na4C24H8O8/Na and Na4C24H8O8 electrodes at a wide potential region in Li+ and Na+
cell at a constant rate of 25 mA g 1. The data were derived from 2032 type coin
electrolytes, respectively. As can be seen from Fig. 1a, the CV re-
cells.
sponse of Li4C24H8O8 at first cycle appeared quite different from
those at following cycles, exhibiting a larger cathodic branch with
two reduction peaks at +1.4 and +0.8 V and a single anodic band at
+1.3 V, which is indicative of an initial irreversible reduction reac-
tion occurring at the electrode in the first negative scan. This phe-
nomenon is usually observed from Li insertion anodes [15], due to
the electrolyte decomposition for the formation of a SEI film (Sur-
face Electrolyte Interface) on the anodes and the irreversible bond-
ing of Li+ ions onto the perylene rings. After a few cycles, the
Li4C24H8O8 electrode showed a pair of stable anodic and cathodic
peaks at +1.3 V and +0.8 V respectively, both of which have similar
areas and shapes, suggesting a couple of reversible redox reactions.
Similar CV behaviors can also be observed from the Na4C24H8O8
electrode in Na+ electrolyte. As displayed in Fig. 1b, a pair of large
redox bands emerge at +0.6 V and +0.8 V with a small cathodic
shoulder at +0.9 V, which resemble very much the CV features of
the Li4C24H8O8 electrode in Li+ electrolyte. A slight difference be-
tween the CV curves in Fig. 1a and b is the slow decrease of the
irreversible cathodic shoulder peak at +0.9 V, implying the difficult
formation of complete SEI film on the Na4C24H8O8 electrode in Na+
Fig. 4. IR spectra of the Na4C24H8O8 samples taken from (a) uncycled, (b) fully
charged and (c) completely discharged cells.
electrolyte. Although the factors affecting the chemical composi-
tions and structures of the SEI films on these two anodes are not
well understood at present, the close resemblance in the CV behav-
voltammograms (CVs) of the Li4C24H8O8 electrode were conducted iors in Fig. 1a and b suggests a similar redox mechanism, as will be
on three-electrode cells with lithium sheets as counter and refer- discussed later, operating at these two organic electrodes.
ence electrodes and cycled in a voltage range of 0.4–3.0 V (vs Li/ To evaluate their possible battery applications, the charge–dis-
Li+) at a scan rate of 1 mV s 1, while the CVs of the Na4C24H8O8 charge performances of these two organic compounds were tested
electrode was conducted on coin-type cells in a voltage range of at a constant current of 25 mA g 1 as Li- and Na-storage anodes,
96 R.R. Zhao et al. / Journal of Electroanalytical Chemistry 688 (2013) 93–97

Table 1
The average amounts of Na cycled in the Na4C24H8O8 anode at different states of charge and discharge, measured from ICP analysis a.

Concentration (ppm) Corresponding numbers of Na+ ions Chemical composition of the anode
+
Fully discharged to 0.3 V (Na insertion) 20.43 6.09 Na6.09C24H8O8
Charged to 3.0 V (Na+ extraction) 15.56 4.3 Na4.3C24H8O8
a
The electrode samples taken for ICP analysis were 11.76–11.77 mg.

Fig. 5. Molecular structure and charge–discharge mechanism of (a) Na6.09C24H8O8 and (b) Li4C24H8O8.

respectively. Fig. 2 shows the charge–discharge curves for the Li4- different depths of charge and discharge, so as to characterize
C24H8O8/Li and Na4C24H8O8/Na cells. In initial discharge, the Li4- the changes in the molecular structure and bonding state of these
C24H8O8/Li cell exhibited a voltage plateau at +1.1 V with a organic compounds during cycling. As an example, Fig. 4 gives a
capacity of 200 mAh g 1, followed with sloping discharge with comparison of the IR spectra of the Na4C24H8O8 samples in pristine,
an irreversible capacity of 250 mAh g 1. On reversed charge and fully charged and discharged states. Despite being fully charged or
successive cycles, two flat charge/discharge plateaus were ob- discharged, the Na4C24H8O8 samples taken from the cycled cells
served at 1.25/1.1 V with a stable reversible capacity of show almost identical IR bands to the pristine sample, indicating
200 mAh g 1, corresponding to 84% of the theoretical one elec- that the chemical structures of the organic cores kept intact during
tron redox capacity of Li4C24H8O8 (237 mAh g 1). Though the ini- charge and discharge cycling. Therefore, the charge storage mech-
tial coulombic efficiency was only about 45%, it rose up rapidly anism of the organic electrodes must proceed through the electron
to nearly 100% after a few cycles, which agrees very well with transfer to/from the conjugated perylenetetracarboxylic rings
the CV data (Fig. 1a) and indicates the formation of a SEI film in along with Li+ or Na+ insertion/extraction. In fact, previous studies
the first cycle. In contrast, the Na4C24H8O8/Na cell show flat of organic anodes for Li-ion batteries have revealed that Li+ ions
charge–discharge plateaus at much lower voltages of 0.6/0.4 V can be reversibly bonded to the carboxylate groups conjugated
with a smaller reversible capacity of 100 mAh g 1 as compared with a series of organic cores, thus giving a high reversible Li-stor-
with the Li4C24H8O8/Li cell. The cause for this large difference is age capacity [11,12]. Obviously, this insertion mechanism can also
not clear at the present, but may be related to the frustrated kinet- be applied for the perylenetetracarboxylate electrodes.
ics of lager Na+ ions bonding with or dissociating from the aromatic To convince this redox mechanism, we quantitatively measured
core. the amount of the sodium ions cycled in the Na4C24H8O8 electrode
Both the Li4C24H8O8 and Na4C24H8O8 cells can not only display a by ICP–AES analysis. As shown in Table 1, the electrode samples
sufficient reversible capacity, but also a strong cycleability. Fig. 3 fully discharged to 0.3 V (Na+ insertion) showed a chemical compo-
shows the cycling performances of these organic electrodes. Cy- sition of Na6.09C24H8O8, (corresponding to 1.79 Na+ ions inserted
cling at changing current rates (Fig. 3a), the Li4C24H8O8 electrode and coordinated to the carboxylate groups of the organic molecule.
delivered stable capacities of 170 mAh g 1 at 50 mA g 1 and When recharged to 3.0 V, the Na content in the electrode samples
110 mAh g 1 at 400 mA g 1, respectively. Even after 100 cycles, recovered Na4.3C24H8O8, demonstrating a complete extraction of
the reversible capacity can retain 195 mAh g 1, corresponding to the inserted Na ions from the electrode. Similarly, the Li4C24H8O8
98% capacity retention of its initial value, when the charge–dis- electrode delivered a reversible capacity almost twice as high as
charge rate was set back to 25 mA g 1. More strikingly, the Na4- the value of the Na4C24H8O8 electrode, suggesting a reversible
C24H8O8 cell exhibited very stable cycling performance with insertion of 4 Li+ ions to each perylenetetracarboxylate unit during
almost indiscernible capacity decay during successive 100 cycles charge and discharge. Thus, the molecular structures and insertion
(Fig. 3b). Compared with the smaller organic carboxylates previ- mechanisms of these organic molecules can be schematically ex-
ously reported [4,11], the Li4C24H8O8 and Na4C24H8O8 compounds pressed in Fig. 5.
displayed much better cycling stability and capacity retention,
most likely due to their larger conjugated structure of perylene
ring, which leads to an increased the chemical stability and a de- 4. Conclusions
creased solubility of the molecules in the electrolytes.
To get an insight into the redox chemistry of these electrodes, In summary, we synthesized Li4C24H8O8 and Na4C24H8O8 and
we made Infrared measurements of the organic molecules at characterized their redox properties as novel anode materials for
R.R. Zhao et al. / Journal of Electroanalytical Chemistry 688 (2013) 93–97 97

Li-ion and Na-ion batteries. The Li4C24H8O8 anode shows flat [2] V. Palomares, P. Serras, I. Villaluenga, K.B. Hueso, J. Carretero-Gonzalez, T. Rojo,
Energy Environ. Sci. 5 (2012) 5884–5901.
charge/discharge plateaus at 1.25/1.1 V with a reversible capacity
[3] X.P. Gao, H.X. Yang, Energy Environ. Sci. 3 (2010) 174.
of 200 mAh g 1, corresponding to a full one electron redox reac- [4] M. Armand, S. Grugeon, H. Vezin, S. Laruelle, P. Ribiere, P. Poizot, J.M. Tarascon,
tion. In contrast, the Na4C24H8O8 anode delivers a reversible Na- Nat. Mater. 8 (2009) 120.
storage capacity of 100 mAh g 1 at quite low charge/discharge [5] M. Zhou, J.F. Qian, X.P. Ai, H.X. Yang, Adv. Mater. 23 (2011) 4913–4917.
[6] T. Suga, S. Sugita, H. Ohshiro, K. Oyaizu, H. Nishide, Adv. Mater. 23 (2011) 751.
voltage plateaus of 0.6/0.4 V, with very stable cycling performance. [7] M. Zhou, L.M. Zhu, Y.L. Cao, R.R. Zhao, J.F. Qian, X.P. Ai, H.X. Yang, RSC Adv.
This work illustrates a great improvement in the cycling stability of (2012), http://dx.doi.org/10.1039/C2RA20666H.
the organic carboxylates for reversible Li- and Na-storage reactions [8] R.R. Zhao, L.M. Zhu, Y.L. Cao, X.P. Ai, H.X. Yang, Electrochem. Commun. (2012),
http://dx.doi.org/10.1016/j.elecom.2012.05.015.
by use of the larger conjugated aromatic cores, which leads to a [9] Y. Dai, Y.X. Zhang, L. Gao, G.F. Xu, J.Y. Xie, Electrochem. Solid-State Lett. 13
structural stability and low dissolubility of the organic molecules (2010) A22–A24.
during charge–discharge cycling. [10] L. Zhao, J. Zhao, Y.S. Hu, H. Li, Z.B. Zhou, M. Armand, L.Q. Chen, Adv. Energy
Mater. (2012), http://dx.doi.org/10.1002/aenm.201200166.
[11] W. Walker, S. Grugeon, O. Mentre, S. Laruelle, J.M. Tarascon, F. Wudl, J. Am.
Acknowledgments Chem. Soc. 132 (2010) 6517–6523.
[12] H. Chen, M. Armand, M. Courty, M. Jiang, C.P. Grey, F. Dolhem, J.M. Tarascon, P.
Poizot, J. Am. Chem. Soc. 131 (2009) 8984–8988.
The authors acknowledge the financial support by National [13] P. Novák, K. Müller, K.S.V. Santhanam, O. Haas, Chem. Rev. 97 (1997) 207.
Science Foundation of China (2097132) and the 973 Program, [14] V. Gangilenka, A. Besilva, V.R. Gangilenka, Phys. Rev. B 70 (2004) 1.
China (Grant No. 2009CB220103). [15] P. Verma, P. Maire, P. Novák, Electrochim. Acta 55 (2010) 6332.

References

[1] B. Dunn, H. Kamath, J.M. Tarascon, Science 334 (2011) 928–935.

You might also like