0% found this document useful (0 votes)
19 views21 pages

Rangaswamy Corrected

The document discusses the synthesis and application of Aurin tricarboxylic acid copper metal organic framework (ACM) as a novel electrode material for rechargeable lithium-ion batteries. The study demonstrates that ACM exhibits excellent electrochemical performance with a discharge capacity of 438.09 mA h g−1 in non-aqueous electrolytes and 169.69 mA h g−1 in aqueous electrolytes, significantly outperforming its precursor, Aurin tricarboxylic acid (ATA). Characterization techniques such as NMR, IR spectroscopy, and XRD confirm the structural integrity and enhanced conductivity of the synthesized materials.

Uploaded by

Madhu Shree
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOC, PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
19 views21 pages

Rangaswamy Corrected

The document discusses the synthesis and application of Aurin tricarboxylic acid copper metal organic framework (ACM) as a novel electrode material for rechargeable lithium-ion batteries. The study demonstrates that ACM exhibits excellent electrochemical performance with a discharge capacity of 438.09 mA h g−1 in non-aqueous electrolytes and 169.69 mA h g−1 in aqueous electrolytes, significantly outperforming its precursor, Aurin tricarboxylic acid (ATA). Characterization techniques such as NMR, IR spectroscopy, and XRD confirm the structural integrity and enhanced conductivity of the synthesized materials.

Uploaded by

Madhu Shree
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOC, PDF, TXT or read online on Scribd
You are on page 1/ 21

SYNTHESIS AND APPLICATIONS OF AURIN TRICARBOXYLIC ACID-COPPER

METAL ORGANIC FRAMEWORK FOR RECHARGEABLE LITHIUM-ION


BATTERIES

Madhushree Mallarabanavadi Ravikumar a, Vijeth Rajshekar Shetty a, Gurukar Shivappa


Suresh a *

a
Department of Chemistry and Research Centre, NMKRV College for Women, Jayanagar,
Bangalore-560 011, Karnataka, India

*Author for correspondence


Dr. G. S. Suresh
Associate Professor
Department of Chemistry and Research Centre
NMKRV College for Women
Jayanagar, III Block, Bangalore-560011, India
Tel: 91-80-22443695
Fax no:91-80-22440116
E-mail: [email protected] (G. S. Suresh)
ABSTRACT
Organic electrode materials for lithium-ion batteries are an essential tool for energy storage
systems starting from portable electronics to electric cars. Many attempts have been made to
improve the electrochemical behavior of organic electrode materials for lithium-ion batteries.
Herein we proposed a new electrode material, Aurin tricarboxylic acid synthesized via the
condensation of aldehyde and acid in the presence sodium nitrite which is further converted into
metal-organic framework using copper metal to increase its conductivity and used as an anode
material in both aqueous and non-aqueous rechargeable lithium-ion batteries. The
electrochemical performance of modified Aurin Tricarboxylic acid copper metal organic
framework (ACM) was studied using cyclic voltammetry, galvanostatic cycling with potential
limitation and Potentio electrochemical impedance spectroscopy studies and the structures were
confirmed using NMR, IR spectroscopy, and XRD techniques. It exhibits excellent cyclability
and a good discharge capacity of about 438.09 mA h g −1 in a non-aqueous electrolyte and 169.69
mA h g−1 in an aqueous electrolyte. The electrochemical activity of the ACM shows that it can be
used as electrode material in aqueous and non-aqueous rechargeable lithium-ion battery systems.

Keywords: Aurin tricarboxylic acid, Aurin Tricarboxylic acid copper metal organic framework,
Galvanostatic Cycling with Potential Limitation and Potentio Electrochemical Impedance
Spectroscopy
Introduction

Secondary batteries have attained immense success in energy storage systems due to their
significant achievement with energy densities and cyclability. Rechargeable Li-ion batteries
(RLIBs) have been distinguished from other secondary batteries by their high volumetric and
gravimetric energy, high power density, extended cycle life, least self-discharge property and
high cost per unit [1-8]. Therefore, RLIBs have been successively using day to day life as
essential power sources for daily-use portable electric devices and electric cars.
Most of the electrode materials for RLIBs are based on inorganic components such as
LiCoO2, [9] LiMn2O4, [10] and LiFePO4 [11], metal oxide [12], silicon [13] and graphite
materials. They exhibit exceptional properties like structural stability, conductivity, capacity and
high electrochemical performance. Nevertheless, their synthetic procedures are not benign to the
environment. Hence, scientists are switching towards an innovative, environmental friendly, low
cost, organic electrode material for RLIBs [14]. An organic molecule with electro active groups
like carbonyl, hydroxyl and natural product-based compounds provides prospective benefits like
structural diversity, flexibility, molecular level controllability. Quinones, [15-19] imidates, [19-
22] anhydrides [23, 24] or lithium carboxylates [25, 26], Amino-acridine [27], Naphthol bis-
indole [28] etc. have been investigated as the electrode materials for RLIBs. However an organic
electrode materials shows low specific capacity, poor redox kinetics and safety issues related to
dendrite growth of lithium metal, low cyclability, primarily owing to low conductivity, side
reactions and dissolution into electrolyte. To overcome these issue, many attempts including
polymerization [29-32], immobilization by covalent bonding [33, 34], physisorption [35, 36],
and use of radical polymers [37], metal-organic frameworks (MOFs) [38] have been proposed
and good achievements have been obtained.
In this regard, we focused our studies on MOFs which influence the redox conductivity to
increase the electrochemical performance of LIBs. MOFs have become the ideal matrix in the
conception of organic based LIBs. They are vast 1D, 2D or 3D networks resulted from the redox
processes that arise in the metal ions and in the poly functional organic molecules. This process
primarily due to the access of electrolyte ions in the lattice of MOF [38, 39]. MOF is attractive
electrode material due to their large tunable pores, surface area [40-44] etc. The first MOF
electrode material in LIB was proposed by Tarascon et al. [41, 42]. Manganese based layered
coordination polymers like [Mn-(tfbdc)(4,4-bpy)(H2O)2] [45], nickel based MOFs like
[Ni2(1,4,5,8-naphthalenetetracarboxylate)] [44], zinc-cobalt based MOFs like [Zn4O(1,3,5-
benzenetribenzoate)2] [40]. Direct pyrolysised MOFs [46], heteroatom-doped carbon materials
like nitrogen-doped graphene particles, MOF-derived metal oxides, containing iron oxide [47],
copper oxide [Cu3(btc)2] where (btc:benzene-1,3,5-tricarboxylate) MOF [48, 49], have
successfully tested as anode materials in energy storage systems. These works motivate us to
synthesis carboxylates ligands as electro active redox centers.

In the present work, a novel organic compound called Aurin tricarboxylic acid (ATA)
was first prepared via condensation method and modified to Aurin tricarboxylic acid copper
metal organic framework (ACM) through Microwave method. The structure of both ATA and
ACM are characterized and confirmed by NMR, XRD and IR spectroscopy and their
electrochemical behaviour was investigated by cyclic voltammetry, galvanostatic cycling with
potential limitation (GCPL) and Potentio electrochemical impedance spectroscopy studies in
both aqueous and non-aqueous system. The GCPL studies of ATA & ACM in full-cell
configuration with LiCoO2 as cathode [50] reveal a good reversibility and kinetic behavior. ATA
exhibits a discharge capacity of 60 mA h g−1 in aqueous system and 160.19 mA h g−1 in organic
system. The discharge capacity of ACM was found to be 169.69 and 438.09 mA h g−1 in aqueous
and non-aqueous electrolyte respectively. There is a drastic difference in the capacity of ACM to
that of ATA. ACM shows good cyclizability and a good discharge capacity in a non-aqueous
electrolyte as well as aqueous system due to the structural modification attributed by the π-π
interactions of conjugated carboxylates which stabilizes the 3D network of the Aurin
tricarboxylic acid copper metal organic framework. In this aspect, the presence of metal ions in
the metal organic framework-ACM structure would pilot to fascinating assets.

2. Experimental Section
2.1 Materials and Measurements:
All reagents are of analytic grade unless stated. ATA was synthesized by condensation of
formaldehyde with salicylic acid. The electrochemical studies was examined by biologic
potentiostat-galvanostat electrochemical workstation. The ATA and ACM working electrodes
were prepared by blending 70 wt. % of ATA or ACM with 20 wt. % of carbon black and 10 wt.
% of polyvinylidenedifluoride. The slurry was prepared by adding few drops of N-
methylpyrrolidine to the mixture, stirred overnight and coated onto thee previously weighed
collector made from stainless steel, which is of diameter 0.5 cm, dried at 80 C for 12 h and
determined the weight after drying. The electrochemical properties and performance of ATA and
ACM composite electrodes were studied using both Swagelok type and three-electrode glass
cell, and cell was assembled in a glove box (VIGOR) in an Argon atmosphere. The electrolyte
contained 1 M LiPF6 in ethylene carbonate (EC), dimethyl carbonate (DMC), and propylene
carbonate (PC) in the volume ratio of 1:1:1 respectively.

2.2 Materials Characterization


The structures of ATA and ACM were determined by 1H-NMR spectroscopy (Bruker)
and IR spectra were collected at room temperature using a FTIR Spectrometer (Perkin Elmer)
between 3000 to 450 cm-1. The ATA and ACM composites were characterized by powder X-ray
diffraction in the wide 2 theta range of 10-80 (XRD Benchtop powder diffraction system using
Cu Kα radiation as a source (λ = 0.15443 nm, Proto Manufacturing Inc.). The morphology of the
materials was observed using scanning electron microscopy (TESCAN SEM VEGA3).

2.3 Preparation of Aurin tricarboxylic acid:


The ATA was accomplished by condensation reaction [51] using commercially available
formaldehyde and salicylic acid, upon stirring with solid sodium nitrite and concentrated sulfuric
acid media. The resulted crude product was purified by recrystallizing ATA using EtOH as a
solvent to get reddish brown crystalline product of 3, 3’-[(3-Carboxy-4-oxocyclohexa-2, 5-dien-
1-ylidene) methylene] bis (6-hydroxybenzoic acid).
.

Scheme 1: Synthesis of Aurin tricarboxylic acid.

2.4 Metal organic framework of Aurin tricarboxylic acid


Cu(NO3)2. 3H2O (2.5 equivalents) and ATA (1 equivalent) were dissolved in DMF and
stirred for 10 min at room temperature then irradiated with microwave for 6 min. A red
precipitate was obtained, that was centrifuged for 15 min at 9900 rpm and then washed several
times with ethanol; the solid was soaked in CHCl 3 for 12 h and centrifuged to remove DMF.
Ultimately, the obtained product was dried using a vacuum rotary evaporator at 65 °C for 40 min
[52].

Scheme 2: Metal organic framework of Aurin tricarboxylic acid

2.5. Physical characterization

The synthesized ATA, ACM structures was elucidated by powder X-ray diffraction
technique. Figure 1a shows the XRD pattern of MOF (ACM) compared to individual ligand
(ATA) and metal salt (Cu(NO3)2). By Scherer method, the crystalline size of ATA was found to
be 1.47 nm and 13.3 nm for ACM. ATA spectra displayed the peaks at 25.59 and 18.5. This
spectral peak is disappeared in the ACM graph which shows that the synthesized MOF formed a
new crystalline compound with no noticeable quantity of reactants. Figure 1a displays the
diffraction peaks of ACM, which are indexed to 13.8, 22.32, 23.77, 26.72, 34.56, 37.46,
40.80 and 44.44 respectively [48, 49]. From the figure the peaks of ligand had been completely
increased in its intensity in MOF form, but it has retained in its structural integrity in complexes
whereas the Cu(NO3)2 produced a characteristic peak in the vicinity of 14.9, 18.1, 22.32 and
30 in its individual salt form. In the MOF form, the metal intensity has increased due to
complexation [49]. The SEM images established the MOF crystals with high dispersity and
unusual particle morphology. The ACM image (figure 1b), displays irregular shaped flakes of
crystals. The morphology of the MOF composites was reliable with SEM images published for
other Copper-MOF structures. [53]. Elemental analysis of the ACM complex was examined in
fig. 1c. The EDX spectrum for ACM exhibits multiple peaks conforming to the following
elements: C (59.88%), O (31.23%) and Cu (3.50%). The presence of these elements generates
negative charges on the surface of the as-prepared MOF complex and result in electrostatic
attraction between the MOF and electrolyte. [54]. A Fourier-transform infrared (FTIR) spectrum
of ATA and ACM was obtained using KBr pellets in the 500–3000 cm -1 region and are in Fig. 2.
The FT-IR spectrum of 3,3’-[(3-Carboxy-4-oxocyclohexa-2,5-dien-1-ylidene)methylene] bis(6-
hydroxybenzoic acid) (ATA) is as shown in fig. 2a. The ATA shows peaks with a broad and
strong band in the range of 3438 cm-1 depict the intermolecular hydrogen bond which confirms
the tautomerisation. Even, the absorption peak obtained at 1641 cm -1 is due to stretching
vibration of the C=O bonds of the carboxylic acid form of ATA molecule which facilitate the
lithiation in the tautomeric conformation. Similarly, The FT-IR spectra of ACM sample
synthesized is as shown in Fig. 2b. The spectra exhibited the peak characteristic for the main
functional groups of MOF at 1645 cm-1, 1620 cm-1, 1570 cm-1, 1550 cm-1, 1445 cm-1 and 1375
cm-1 respectively. The spectrum of ACM evidently showed an almost iso-bidentate performance
of COO- and the peaks obtained are characteristic for this kind of co-ordination mode. The latter
peaks are in line for the fact that iso- and aniso-bidentate dicopper (II) carboxylate type of
monomeric clusters in the ACM framework [55]. The prepared materials don’t show any
prominent carbonyl peak at 1710 cm-1. This statement was mainly attributed to the point that the
carboxyl anions can form 53 after the formation of the carboxylate, since both the carboxyl
oxygen tends to be equivalent, leading to a delocalized electronic cloud and presence of
characteristic peaks of carboxylate in the vicinity of 1610-1550 cm -1 and 1420-1300 cm-1 [56].
Thus, these results confirmed that the organic acids and metal salts made the target coordination
compounds. More significantly, the structure of synthesized ATA was also studied by 1H-NMR.
The NMR Spectra of ATA compound was recorded in DMSO at 400 MHz and the tetra methyl
silane [Si(CH3)4], which is used as an internal reference as shown in fig 2c. The 1H-NMR
spectrum showed three kinds of aromatic protons like Ar-H, Ar-COOH and Ar-OH. It shows
nine Ar-H exhibiting chemical shift values in the range of ~6.76 - 7.76 ppm that induces the
magnetic field via ring current effect. The two Ar-OH characteristic peaks for a de-shielded
proton was appeared about ~11.08-11.45 ppm and ~14 ppm for three Ar-COOH illustrates a
broad peak due to its acidic nature and/or exchangeable protons which will not couple with the
neighboring aromatic protons. Hence, the shrinkage in the peak was observed.

3. Electrochemical Studies
Cyclic Voltammograms of Aurin tricarboxylic acid and Aurin tricarboxylic acid copper metal
organic framework were carried out in an aqueous system at the scan rate of 0.5 mV/s and is as
shown in figure 3a. The CV of ATA illustrates a pair of weak redox peaks of which the oxidation
peak at -0.33 V and the corresponding reduction peak at -0.87 V respectively. The redox peaks
of ATA may be attributed to the structural/textural modifications by reversible electrochemical
reaction; which is influenced by the presence of carboxylate and enolate groups in ATA and
lithium in the electrolyte [57, 58]. To confirm the lithiation in the redox centres, the electrode
material was scraped out and dissolved in solvent to remove carbon and binder (PTFE). The
XRD peaks in the inset of fig 3a, shows a strong peak which attributes to the crystalline nature
and the diffraction peak for the lithiated compound gradually decreased due to the interaction of
the metal carbonyl groups. [59, 60] According to Armand et al. [26], lithium based conjugated
carboxylates commonly serves as redox centers to coordination of Li + with the COO- groups. So,
it is expected that redox contribution of the organic moiety over the COO - groups plays key role
for Li insertion/extraction in ATA.
From the figure 3a, ACM gives two cathodic peaks at the potentials of -0.04 V and 0.06
V and two anodic peaks at 0.09 V and -0.39 V in aqueous system at scan rate 0.5 mV/s. This
type of redox reaction suggests two one-electron transfer reactions [61]. The size and
morphology in ACM framework influences the lithium ion diffusion that endorses the copper
reduction [62, 63]. However, the process of reduction of cupric to cuprous mainly attributed to
the strong coordination of copper metal in the MOF structure [61] and the strong complexation
of Cu2+ with ATA. The shape of the CV curve is ascribed to the electrochemical response of
copper metal in the MOF structure, [64]. From the Figure 3a, ACM shows a better redox
behavior compared to ATA, Henceforth complete electrochemical behavior of ACM was
studied.
We have also examined ACM in different aqueous electrolytes, different pH and different
temperature to achieve practical electrode kinetics and stability of the electrode material. The CV
measurement of ACM using sat. Li3PO4, sat. LiOH, sat. LiNO3, sat. Li2SO4 electrolytes at a
voltage window between -0.6 and 0.7 V (vs. SCE) at 0.5 mV/s was carried out and depicted in
the figure 3b. The CV of ACM in sat. Li3PO4 (bold solid line), sat. LiNO3 (solid line), and sat.
Li2SO4 (square dotted line) electrolytes exhibits a pair of redox peaks respectively. In sat. LiOH
(dashed line) electrolyte, we observed a dissolution effect with no redox reaction indicates
unfavorable intercalation/deintercalation kinetics of lithium ion with LiOH electrolyte. The
redox peaks for ACM in sat.LiPO4 were obtained at -0.044 and -0.28 V, while in sat.LiNO3 were
obtained at -0.061 and -0.080 V; the peaks seems to appear at more positive potentials with poor
current response. While for sat.Li2SO4 the redox peaks were at -0.032, 0.11, -0.12 and -0.40 V
correspondingly. Using the sat.Li2SO4 solution, well defined peaks were observed, indicating the
faradaic reaction involving the intercalation/deintercalation of the lithium ions is considerably
more facile than other electrolytes [65]. On the basics of the above information, we decided that
the sat.Li2SO4 solution is the best electrolyte to perform electrochemical studies of ACM in
aqueous system.
To study the stability of ACM, pH and temperature studies were performed using
aqueous sat.Li2SO4 electrolyte as shown in Figure 3c and 3d. The influence of pH in acidic,
neutral and basic media on the redox properties of ACM is as shown in figure 3c. The Acidity
and basicity is adjusted using 0.1 M H 2SO4 and 0.1 M LiOH. In Acidic condition (pH 4), the CV
exhibits a pair of redox peaks in the range of 0.072, -0.148V and for basic media (pH 11) were at
0.27 and -0.051 V while for neutral at -0.032, 0.11, -0.12 and -0.40 V respectively. The different
pH displays variances in voltammograms, by peak current and shape of cathodic and anodic
peaks. At lower and higher pH, the obtained peaks were broadened and spread over a large
potential. The anodic current at higher pH sharply increased due to the evolution of oxygen at
higher pH. Consequently, the possibility of oxygen evolution increased, thus more interaction
between oxygen evolution and lithium intercalation/deintercalation reactions would be predicted
at higher pH. [66] However, with the increase of pH, the cathodic and anodic peak current
gradually decreases. Moreover, with increasing the pH values; the cathodic peak and the anodic
peaks shift to more positive and negative values. It advises that the quasi-reversible
electrochemical process is shifted towards the irreversible process. [68] As volttamogram with
clear anodic and cathodic peaks were obtained in sat.Li 2SO4 of pH 7, the rest of the experiments
were conducted in this condition.
Cyclic volttamogram of ACM electrodes in sat.Li 2SO4 electrolyte and pH 7 at different
temperatures of 10 ºC, RT and 50 ºC are given in Fig. 3d. At 10 ºC, ACM exhibits a redox peaks
of 0.001, -0.29V; at RT it shows two pairs of redox peaks at -0.032, 0.11, -0.12 and -0.40 V and
at 50 ºC, peaks were obtained at -0.11, 0.085, -0.197 and -0.42V respectively. The most
important effect of increasing the temperature is the appearance of a well-defined surface redox
process. [68]. The working temperature range of an LIB is mostly dictated by the electrolyte
composition, which not only affects the Lithium ion transport through the bulk electrolyte but
also determines the properties of the formed SEI. [69] The low temperature of the electrolyte
leads to the rise in viscosity successively reduces the ionic conductivity. This condition of
electrolyte will activate the polarization of ACM that would delays the lithium ion
intercalation/deintercalation kinetics [70]. While high temperature is much more complex
compared to low temperatures, since heat generation, aging and thermal runaway affects the
electrolyte. These concepts at high temperature is related with the charge transfer and chemical
reaction during intercalation/deintercalation; that leads to rise of charge transfer resistance at
electrode-electrolyte interface or SEI and causes degradation of electrodes [71, 72]. Since
distinct redox peaks of ACM were observed as we increase the temperature from 10 ºC to RT.
Hence, the rest of the electrochemical studies of ACM are carried out in room temperature at
neutral pH and in sat.Li2SO4 electrolyte.

Figure 3e shows the multiple CV cycles of ACM upto 20 cycles at scan rate of 0.1 mV/s
in the potential window of -0.7 to 0.6 V Vs. SCE respectively. The peak current sligthly
decreased from 1st cycle to 7th cycle and becomes constant at 13th cycle which is consequent to
reversible electrode kinetics. The Lithium insertion/ deinsertion of Li+ ion to/fro of MOF frame
work is in constant phenomenon confirmed by XRD studies taken before and after CV results
(figure 3a inset). As we discussed earlier the ACM was scraped out of electrode after 20 cycles
and XRD was performed. The decreased peak intensity confirms the lithiation of carboxylate
groups, Cu-O sites and attributes the crystalline nature [59, 60]. The peak current remains stable
throughout cycling subsequent to the reduction of Cu(II) to metallic Cu(0) [73] and the
development of a solid electrolyte interface (SEI) film.
To study the relationship between peak current (ip,a and ip,c) and square root of scan rate
(ν1/2) at different scan rates, CV of active material was obtained by subjecting the cell to various
scan rates from 0.1 to 1.0 mV/s in sat.Li2SO4 solution. The factors which might influence on
such studies are electron transfer reaction velocity, electrode/electrolyte reaction and the applied
potentials etc. [74]. The CVA of ACM as in aqueous electrolyte shows influence of scan rates on
the electrolyte-electrode reaction shown in fig. 3f. According to the theoretical concept, the
reversible behaviour of peaks were observed at lower scan rates, while the peaks becomes
irreversible at higher scan rates; which suggests that the intercalation/deintercalation process of
lithium ions alters from quasi-reversible to irreversible system [75]. In ACM system, reversible
peaks with two pairs of well-defined, sharp and narrow peaks were observed [76]. From the
figure, we can conclude that increase in scan rate raises the peak current. The amount of the
voltammetric current of an electrode is estimated by Randles–Sevcik equation. From the cyclic
voltammogram of ACM, reversible charge transfer process is observed; the peak current (ip)
increases with square root of scan rate (ν 1/2) and is straightly associated to the concentration of
Lithium ion in the electrolyte. By Randles–Sevcik equation, the peak current is directly
proportionate to square root of applied scan rate and the plot of ip vs. ν1/2 should be linear which
is represented in figure 3g. From the figure the linear increment of peak current with scan rate is
witnessed [77]. The diffusion layer configuration plays a dynamic role, which is fairly additional
from the commencement of the electrode reaction at lower scan rates compared to high scan
rates. Ultimately, from the figure at particular potential the peaks display the maximum current
and that is a characteristic feature of ACM with fast electrode kinetics.

The lithium-ion diffusion coefficient (DLi+) is based on Randles-Sevcik equation (1) [78]:

(1)

Where ip represents peak current, F is the Faraday constant, n is the number of electrons,
T is the temperature, R is the gas constant, D is the diffusion coefficient, A is the surface area of
the electrode, C stands for the concentration of lithium-ions in the electrolyte, and v is the
voltage scanning rate. From the figure 3g, a diffusion controlled process is taking place since the
peak current is proportional to ν1/2 [75, 79]. The DLi+ can be obtained by using the actual electrode
area. Since we can assume that, the sum of coated active material over the surface of the
electrode is equals to the definite electrode area. Ideally, the larger value of DLi+ is
comprehendible; meanwhile the smaller particle sizes of the active material and the aqueous
electrolyte render lower viscosity.
For exploring the capabilities of ATA and ACM electrode material for Li-ion battery, we
have examined the detailed mechanism of Li-ion intercalation and de-intercalation by using
cyclic voltammogram studies in non-aqueous media. To study the redox behavior of ATA in
non-aqueous system was evaluated using 1M LiPF 6 in EC/DMC/PC (1:1:1) as electrolyte; Li foil
was used as a counter and reference electrodes. In aqueous system, saturated Li 2SO4 was used as
electrolyte, Pt. was the counter electrode and SCE was the reference electrode.

The CV profile of ATA and ACM in non-aqueous system is as shown in Figure 4a. The
CV curves of ATA exhibits a broad oxidation peak at 2.41V and a blunt reduction peak in the
range of 2.01V at 0.5 mV/s. The weak redox peaks ascribes the formation of solid electrolyte
interphase [80] which restrains electrolyte degradation [79]. To increase the ionic and electronic
diffusion, high rate capability, ATA were decorated with metal-organic framework (MOF). The
MOF gives a better modification towards ATA which is designated as Aurin tricarboxylicacid
copper metal organic framework (ACM). This kind of structural modifications may increase the
conductivity and capacity behavior.
The CV curve for ACM in non-aqueous system illustrates two broad cathodic peaks at
1.32 V and at 2.07 V signifying a two-step reaction. It exhibit two reduction peaks where, the
first reduction peak at 1.32 V which was observed at the anodic scan due to irreversible
transformation and structural modification [81, 82] and the second reduction peak at 0.67 V due
to the reduction of Cu2+ to Cu0. The results obtained indicates that the active participation of
carboxylate group of ATA being weakly electron withdrawing ligands may act as electroactive
redox centers [64]. The most possible electron transfer reaction in the electrochemical reaction is
as shown below:
Cathodic reaction:

Anodic reaction:

To study the cycling stability and reversibility of ACM, the cyclic voltammetry was studied at
0.5 mV/s up to 20 cycles (figure 4b) .The peak current decreases slightly up to 6th cycle and it
becomes constant after 7th cycle which is consequent to kinetically reversible Li ion involvement
in the ACM and exhibit better electrode kinetics and good redox properties.
For further investigation of the electrochemical kinetics of ACM elctrode, CV
measurements at various scan rates (from 0.1-1.0 mV/s) was carried out. From the figure 4c,
with increase in the sacan rate, the charecteristic peaks become broader and displays similar
shape of curves; while the shift in reduction and oxidation peaks from lower or higher potentials
is noticed. Hence, in general the total current is obtained mainly from two distinct and
independent parts; diffusion-controlled process and surface induced capacitive process as
interpreted by the equation (2) [ 83].
i = avb (2)
Where v refers to scan rate (mV/s), a and b are variable parameters, and b values ranges from
0.5-1.0. When the b value is 0.5, it reveals that process is completely ascribed by diffusion
controlled process. While the b value is 1.0, denotes a total capacitive process. Correspondingly,
the b value is obtained by plotting the slope of log i vs. log v of all four oxidation and reduction
peaks respectively (figure 4d). From the fitting results of the plots of four redox peaks the b
values are 0.33, 0.85, 0.5, and 0.80. The b value is approximately equals to 1.0, for 2 peaks
showing the total capacitive process, and the b value for another 2 peaks is nearby 0.5, that
reflects the reaction is diffusion controlled. Besides, at certain scan rates, the capacity can be
distributed into capacitive (k1v) and diffuse (k2v1/2) processes at specific voltage through cycling
as given by the equation [84]:
i = k1v+ k2v1/2 (3)
i/v1/2 = k1v1/2+ k2v1/2 (4)
In equation (3), the k1v and k2v1/2 terms indicate the surface and diffusion-controlled
contributions, respectively. The k1 and k2 values at fixed potentials were determined by the
correlation between the scan rates and corresponding current signals according to equation (4)
[82]. In the electroactive species, ACM disintegration mainly depends on structural modification,
cycling and surface film modification during the electrochemical process. Hence the stability of
the electro active species can be mainly dependent on the temperature gradient, pH, and diffusion
controlled process.

3.1. Galvanostatic charge/discharge studies


To investigate the specific capacity, storage capacity and elctrochemical reversibility of
the cell systems ATA and ACM| sat.Li 2SO4| LiCoO2 in both half cell and full cell configuration;
GCPL studies were carried out at room temperature. The half cell system in aqueous media
comprises of a working electode (ATA/ACM), SCE as a reference electrode and Pt foil as a
counter electrode respectively. The galvanostatic charge/discharge studies were studied for both
two electrode and three electrode electrochemical glass cells using 10 mL of sat. Li 2SO4
electrolyte. the GCPL profiles are as displayed in figure 5.
To evaluate the electrochemical performance for ions storage in ATA charge/discharge
profiles were performed by using a three electrode system at 0.05 mA current density and is as
shown in fig. 5a. Based on CV results potential window was limited to -1.0 to 0.5 V. The
reversible capacity of ATA based on the redox activity was predicted to be around 63.45 mA h g -
1
, while ATA discharged at 60 mA h g-1 at the first discharge process and steeped to 52.9 mA h
g-1 at 100th cycle. Since the ATA exhibited the discharge capacity which is almost equal to
theoretical capacity, in certain not all the –OH and –C=O groups have reduced, which might
attribute to the repulsion amongst the charges (e -) infused in the conjugated structure. The
decline in the discharge capacity attributed mainly due to two aspects; firstly the creation of solid
electrolyte interface (SEI) film formation on surface of electrode and the other is the dissolution
of active material, through cycling process [85].
It has been broadly accepted that the improved conductivity of organic electrode
materials are credited with extended conjugated structures.[86] An extended π-conjugated ACM
shows a reversible capacity of 257.32 mA h g −1 at a current density of 0.05mA. The preferred
growing direction implies the enhanced intermolecular interactions caused by the extended π-
conjugated system. The increased discharge capacity with ACM by coupling ATA with Cu metal
prefers the enhanced intermolecular interactions by the extended π-conjugated system. ACM has
one pair of redox centres due to the presence of carboxylic acid and copper-oxygen centres
(oxidation state of Cu is Cu 2+, Cu+ and Cu0) in its framework. Fig. 5b shows the cycling
performance of ACM in three electrode system for 100 cycles in the potential window of -0.7 to
-0.8 V respectively. At the first cycle, ACM shows a discharge capacity of 285.32 mA h g -1 and a
charge capacity of 262.68 mA h g -1. This irreversible capacity at the primary cycle ascribes the
SEI formation during the initial cycle with reduced coulombic efficiency, irreversible side
reaction of the oxidative coupling of the ATA moiety [87] and from the second cycle, stable and
good reversible capacity were observed. The discharge capacity of ACM exhibits an upswing
from 257.32 mA h g−1 to 204.43 mA h g−1 at 0.05 mA before 2–100 cycles. On one hand, the
activation phenomenon may be attributed to the sluggish forming process of SEI films. While on
other hand, we suspect that Li+ has hardly any access to the interior of the helical channel at the
beginning due to the crystalline ACM materials. It needs some time for Li + to completely reach
the Cu-O and carboxylate active sites [88] or it may be due to the slow process of SEI layer
formation at interface, where the slothful migration of Li + from electrolyte to the ACM takes
place [89]. The improved discharge capacity of ACM is mainly associated to the number of
active sites in the MOF arrangement and its amorphous nature [90].The charge discharge
behaviour of ACM is far more complex due to the incorporated ATA building block [91].The
distinct multiple plateaus in the discharge profiles of ACM should be related to the different
degrees of polarization for the insertion/deinsertion of Lithium ions into the MOFs. [87]
Galvanostatic charge/discharge curve for ACM | sat.Li2SO4 | lithiated cobalt oxide cell
(full cell) is carried out at C/15 rate for 500 cycles as shown in figure 5d. The cell was
constructed using ACM as anode and commercially available LiCoO 2 as cathode material, and
GCPL was carried out at potential window of to -0.7 to 1.2 V. Good stability and stable cycle life
were attained with slight capacity fading through cycling. The charge and discharge capacity at
first cycle was about 164.09 mA h g −1 and 148.89 mAh g−1 respectively, with the capacity
retention of 90.7%. Such an irreversible capacity loss through the initial cycles can be attributed
to side reactions with the electrolyte on the accessible surface of MOF [92]. During initial redox
process, it shows irreversibility and formation of SEI layer which consequent to the irreversible
lithium ions into the ACM framework and lithium sulphate electrolyte disintegration [93-96].
The firm decline in the specific capacity is observed might be due to dissolution of electrode
material in electrolytes [97]. The ACM electrode displays a rigid discharge capacity of 136.68
mA h g−1 with stable columbic efficiency after 10 cycles and 78.07 mA h g −1 at 500 cycles. This
indicates a good rate capability and cycle stability compared to the reported Zn 4O (1, 3, 5-
tribenzoate)2 MOF network [98]. By this concern, ACM exhibited good electrochemical
performance as anode material for ARLIB which is in comparison to the current inorganic anode
materials [99, 100]. From the above results we can confirm that ACM has good specific capacity
with good cyclizability.
To study the rate capability of ACM | sat.Li 2SO4 | LiCoO2 cell at different C rates of C/15
to 3C, the GCPL studies were performed, the results are as shown in figure 5e. From the figure it
is clear that as the C rate increases, the specific discharge capacity decreases, that is, C/15, C/10,
C/5, 1C and 3C exhibits a decreasing the discharge capacity of 169.69 mA h g -1 with 94 %, 151.4
mA h g-1 with 92.05 % , 119.77 mA h g-1 with 91 %, 94.09 mA h g-1 with 90.15 % and 79.72 mA
h g-1 with 89.1 % columbic efficiency correspondingly as shown in figure 5f. It is clear that the
reversible capacity in each C rate firmly retained throughout cycling which sets a good example
for lithium ion diffusion in the organized framework [97]. The decreased capacity at higher C
rates may be ascribed to the enhancement in Li + de-insertion process from the aromatic ring
under continuous cycling which leads to the activation of the electrode. Though with steady
increase of Li+ concentration close to the working electrode with additional cycling may counter
the de-insertion process, brings the decrease in the capacity until equilibrium is reached [92]. The
electrochemical activity of MOF has been increased in comparison with the individual ligand
activity; this may be due to the crystal structure of ACM. As per the literature [ 101, 102] the
MOF moiety with exposed active sites of the ligand will enhance the electrochemical activity by
providing more surface area for its insertion/deinsertion process, such phenomenal activation
process has been reduced by ACM in our experiment.
After studying the cycling behaviour of ATA and ACM in Aqueous media; both ATA
and ACM were tested for their electrochemical performance vs. Li/Li +. With comparison to
aqueous cell, the non-aqueous cell gives a good cyclability, reversibility and stability. In Non-
aqueous system, the cycling behaviour of ATA is carried out using GCPL studies at 0.05 mA in
the potential window range of 0-3.0 V as shown in figure 5f. The discharge capacity at first cycle
is about 160.19 mA h g-1with 94% coulombic efficiency. This discharge capacity is mainly
attributed to the electron delocalization effect of COO - groups and enolate groups during redox
couple reaction [103]. After 2nd cycle, the discharge capacity decreased to 157.19 mA h g -1 and
optimized to 133.3 mA h g-1 after 100 cycles. The fade in the capacity is mainly allocated to the
dissolution of electrode material into the electrolyte. [91]
The capacity versus potential curve of ACM at 0.05 mA in non-aqueous media is
investigated (Figure 5g).The discharge capacity of 454.44 mA h g-1 and a charge capacity of
429.69 mA h g-1 at first cycle with 94.5 % coulombic efficiency. This shows the stability of ACM
material due to enolization of the carboxylate group in conjugation system as well as degree of
planarization during the redox processes [104]. The Coulombic efficiency comes up to 99% after
a few initial cycles, signifying that the obtained carbon material was stable and reversible during
the charge/discharge cycles. During 2nd cycle, the capacity is found to be 428.47 mA h g-1 with
reversible plateaus were observed and further the discharge capacity declined to 411.11 mA h g -1
during 500th cycle, the results are in good agreement compared to inorganic intercalation
compounds. The capacity fading of ACM electrode might be attributed to the structural integrity
and particle size of the ACM material and formation of solid electrolyte interphase [93]. When
compared with the inorganic intercalation material ACM exhibited good specific capacity in
non-aqueous media, with good cyclability and reversibility. From the figure, one can see the
sloping discharge curve with eventual loss of capacity from 1 st to 500 cycles; which leads to
good structural stability and quite long life cycle in non-aqueous system [105]. The possible
lithiation/delithiation sites is as shown in scheme 3, and the insertion of the Li + to the organic
moieties in the MOF includes C=O and COO- sites respectively [107].

Scheme 3: Possible Lithiation/delithiation sites for co-ordination of lithium ions with the Aurin
tricarboxylicacid.
Further, galvanostatic charge/discharge experiments were conducted at different C rates
like C/10, C/5, C/2 and 3C in non-aqueous electrolytic cell system to study the rate capability of
the electrode material and is as shown in the figure 5h. By increase in the current density at C/10,
C/5, C/2 and 3C there is a drastic decline in the discharge capacity from 431.55, 394.79, 331.87
and 271.31 mA h g-1 with a decrease in the columbic efficiency from 97.5, 95.25, 91.4 and
89.05% (figure 5i). From the above experimental result one can see that, at higher C rate
capacity decreases. The decrease in specific capacity at higher current rates is due to poor
electron conductivity of ACM molecule’s large structure [106].
From the above considerable aspect obtained for the cell system with ACM gives good
stability, cyclizability acts as a good anode material for RLIB. Additionally, ATA has less rate
capability and cycling life compared to the ACM electrode material which facilitates the lithium
ion intercalation/ deintercalation into the metal organic framework, thereby enhancing the rate
capability, and the highly conjugated bi-ligand framework of ACM helps in fast redox process.
This reveals that capacity arising from ATA is comparatively less stable and less reversible with
respect to the capacity arising from the metal organic framed ATA.
Notably the specific capacity of the MOF is superior to the ligand electrode (ATA). Tis
enhancement of the capacity are possibly due to the redox reactions of ligand associated with the
lithiation/delithiation processes [91] and the redox reactions at the paddle wheel metal clusters.
Inspite to this, the chemical processes that occurred on the surface of the electrode and interface
during the formation of SEI layer not only comprising of the redox reactions of the MOFs but
also the breakdown of the MOFs; the degradation of the electrolyte and the reactions additives.
[107]

3.2. Potentiostatic electrochemical impedance spectroscopic studies


Potentiostatic electrochemical impedance spectroscopic technique (PEIS) is used to obtain
the impedance characteristics of LIBs which gives accurate and non-damaging kinetic effects.
The specific amplitude with different frequency ranging from 100 kHz to 5 MHz was performed
to measure the Nyquist plot data. The PEIS for ACM is studied in three electrode
electrochemical cell during redox process as shown in the Figure 7 The CV is cycled for few
cycles in the potential of 0.1 mV/s and optimized before performing PEIS measurements.
Figure 6 (a-f) shows Nyquist plots of ACM in aqueous electrolyte. The cyclic
voltammogram is mainly divided into 6 domains depending on the oxidation and reduction
potentials. Figure 6 (a-c) describes the Nyquist plots of ACM a) before oxidation, b) during
oxidation and c) after oxidation potentials. Correspondingly the reductive potential is divided
into 3 potential domains such as d) before reduction, e) during reduction and f) after reduction
respectively. The impedance plots are mainly categorized into three parts: straight line which is
potential dependent in the high frequency, Warburg curve in the middle-low-frequency region
and capacitive line in the low frequency region. There is high middle frequency semi-circle arc
in the high frequency region may be due to the high rate performance or short diffusion path or
small resistance offered by less or small SEI layer [108].The SEI is less in figure 6a, 6f, this is
because of the resistive film which is not completely covered the electrode so that migration of
lithium ion through the surface film is easy and viable during electrochemical reaction. [109]
Hence electrode exhibits linear line indicating charge transfer kinetics in the vicinity of 0.07 Hz
to 45751 Hz respectively. Figure 6e shows prominent Warburg impedance line with slopes of 45
in the middle-low frequency region of 0.005 Hz to 0.07 Hz which is due to the semi-infinite
diffusion of electro active material at their peak potentials.
From the figure 6a, 6e, 6f, it is clear that the Nyquist plots after the cycling processes show
prominent Warburg impedances. The appearance of Warburg lines with slopes greater than 45
shows a near capacitive behaviour with mass transport of the active material due to finite
diffusion on the inhomogeneous distribution of the electro-active species.[110]
Similarly, figure 7 (a-f) shows Nyquist plots of ACM electrode which is divided into 6
potential domains in the organic electrolyte during the charge-discharge process. The Nyquist
plots are classified into three divisions: the low-frequency region, middle-frequency region and
high-frequency region [111]. The Nyquist plots for the ACM composite electrodes in non-
aqueous electrolyte were carried out in 2 electrode system as shown in the figure 7. The plot
describes three different frequency regions: The low-frequency region ranges from 0.005 Hz to
0.01Hz, the middle-frequency region ranges from 0.01 Hz to 3.85 Hz, and the high-frequency
region ranges from 3.85 Hz to 100024 Hz. Each frequency region describes its kinetic nature. In
the figure 7 (a-f) the high frequency region shows a potential dependent semicircle indicating
charge transfer resistance which resists the migration of lithium ion via surface film of the ACM
electrode. [112].The middle-low frequency region shows Warburg region and sloping region at
low frequency indicating the completion of the redox reaction. [113]. The Warburg line is mainly
allocated to the solid state diffusion of Li-ions into the ACM electrode and sloping line is
associated to the activity of Li-ion [114]
In 7b, 7c, 7d, 7e, the semi-circle obtained are similar to the impedance response of the SEI
layer formed on active material upon cycling, it suggests that the electrolyte could decompose
chemically on Lithium, and the decomposition products form a deposit on the electrode. [115]
To study the variation of kinetic parameters like Rs, Rct, and Rf with the electrode
polarization potential in aqueous and non-aqueous electrolyte, an equivalent circuit was fitted
using ZSimpWin software to fit the experimental Nyquist plots.
From the equivalent circuit (fig 6g), Rs indicates the solution resistance, Rct is the charge
transfer resistance, Rf is the resistance for Li+ migration via surface film, CPE represents constant
phase element, Zw for the Warburg impedance and Cdl represent double layer capacitance
correspondingly. Rs is almost constant throughout the redox process (table 1-4), it is serially
connected to the Warburg impedance and charge transfer resistance which is parallel to the
constant phase element CPE. It is used when impedance spectra shows dispersion in the low
frequency. The dispersion in the low frequency is attributed to in-homogeneity, porosity,
roughness or film substrate interface [58]. The Rct increases with increase in applied potentials
and increases at peak potentials. This shows a good electrochemical reaction. Further it is
connected to the Rf where resistance of lithium ion movement takes place via surface electrolyte
interphase and it is parallel connected to the SEI resistance.
4. Conclusion
Aurin tricarboxylic acid was synthesized via conventional method and it is found to be
good eco-friendly electrode material for aqueous rechargeable lithium ion batteries because of its
easy synthesis and low cost as compared to other materials. When ATA is modified into ACM, it
exhibits good cycling and electrochemical behavior in both aqueous and organic electrolytes.
The aqueous rechargeable lithium ion battery ACM | sat.Li 2SO4 | LiCoO2 has been fabricated and
studies the electrochemical properties and performance in both aqueous and non-aqueous
electrolyte. Good cycling behavior was achieved due to their good reversibility and rate
capability. Further studies are required in future for better understanding the role of MOFs in
LIBs.
Table 1: PEIS Parameters obtained by circuit fitting at Charging potential bias for aqueous
media
Potential in Rs in  Q in mF Rct in  W in  Q in mF Rf in 
V
0.2 13 0.7186 4135 0.00124 0.6227 543
1.27 6.069 0.6904 1.621E4 0.0005198 0.5686 1160
1.30 5.241 0.6972 1.532E4 0.0004609 0.5654 1190
1.33 6.322 0.6974 1.589E4 0.0004512 0.5692 1210
2.01 4.372 0.8454 1641 0.000448 0.5592 1770
2.04 5.977 0.799 2363 0.0004315 0.5631 1759
2.07 3.972 0.5697 1884 0.0004126 0.952 739.6
2.5 0.01 1 188.7 0.000257 0.4962 4840
3.0 1.316 0.6381 5.401E4 2.695E6 0.5117 6261
3.5 12.04 0.5106 1.049E4 0.0002321 0.7141 587
4.0 15.86 0.5824 1199 0.0004557 0.7757 6414

Table 2: PEIS Parameters obtained by circuit fitting at discharging potential bias for aqueous
media
Potential in Rs in  Q in mF Rct in  W in  Q in mF Rf in 
V
4.0 12.12 0.5351 864.5 0.0002946 0.8346 6270
3.0 14.09 0.7704 3800 0.0001419 0.5411 799
2.0 18.69 0.6123 1292 0.0004819 0.6946 2704
1.5 10.85 0.8304 445.4 0.0006211 0.4423 4972
1.35 10.27 0.8346 498.8 0.0005702 0.4322 2141
1.32 13.43 0.8016 551.5 0.0005603 0.4379 1524
1.30 13.18 0.4716 1622 0.000453 0.8313 425.8
0.65 17.48 0.9093 228 0.0008204 0.5728 497.6
0.63 19.27 0.5691 2.763E4 8.345E6 0.658 730.3
0.60 19.78 0.5752 3.046E4 147.6 0.6616 718.6
0.1 16.75 0.599 1.247E4 4.833E7 0.6146 531.9
Table 3: PEIS Parameters obtained by circuit fitting at Charging potential bias for non-aqueous
media

Potential in Rs in  Q in mF Rct in  W in  Q in mF Rf in 
V
-0.6 2.73 0.7774 3836 0.004467 0.8065 0.0106
-0.14 3.025 0,8026 558.5 0.008328 0.7939 1483
-0.13 3.038 0.7814 1477 0.003016 0.7721 5.136E7
-0.04 0.1694 0.7885 1341 0.002626 0.7282 2.96
-0.05 3.136 1 19.66 0.002705 0.8267 1952
0.2 1.865 0.7709 5550 0.001555 0.903 1.257
0.3 2.161 0.9034 35.14 0.0002708 0.8362 0.9656
0.5 3.096 0.9076 6.928E4 0.0001843 1 7.138
0.7 3.287 0.8307 4.329E4 0.001163 0.833 5.092E9
Table 4: PEIS Parameters obtained by circuit fitting at discharging potential bias for non-
aqueous media

Potential in Rs in  Q in mF Rct in  W in  Q in mF Rf in 
V
0.6 3.066 0.79838 0.08317 1.05E4 0.8256 5.48E4
0.4 3.225 1 6.509 0.002312 0.898 6.72E4
0.22 3.104 1 4.94 0.00737 0.9038 9.25E4
0.18 1.212 1 3.112 1.44E9 0.8741 9.71E4
-0.18 1.22 0.8747 272.6 1.26E-5 1 0.5513
-0.2 1.1195 0.8735 3.49E4 0.0002073 1 0.3586
-0.48 1.221 0.8721 2.03E4 0.001071 1 0.6046
-0.5 1.24 0.9208 10.96 0.000183 1 0.5762
-0.6 2.648 1 6.055 0.007007 0.9162 8602

You might also like