You are on page 1of 8

Journal of Physics and Chemistry of Solids 135 (2019) 109082

Contents lists available at ScienceDirect

Journal of Physics and Chemistry of Solids


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

Novel synergistic in situ synthesis of lithium-ion poly(ethylene citrate)-TiO2 T


nanocomposites as promising fluorine-free solid polymer electrolytes for
lithium batteries
Fernando Pignanellia, Mariano Romeroa,∗, Jorge Castiglionib, Ricardo Faccioa,∗∗,
Alvaro W. Mombrúa
a
Centro NanoMat/CryssMat/Física – DETEMA, Facultad de Química – Universidad de la República, Montevideo, C.P. 11800, Uruguay
b
Laboratorio de Fisicoquímica de Superficies – DETEMA, Facultad de Química – Universidad de la República, Montevideo, C.P. 11800, Uruguay

A R T I C LE I N FO A B S T R A C T

Keywords: Here, we report a synergistic in situ synthesis of lithium-ion poly (ethylene citrate) with embedded titania na-
Titania noparticles in which the esterification process itself triggered the hydrolysis of the titanium alkoxide precursor to
Polyester form a promising fluorine-free lithium-ion solid polymer electrolyte. Structural characterization by means of X-
Nanocomposite ray diffraction, small angle X-ray scattering and also confocal Raman spectroscopy revealed the formation of the
Lithium-ion
polyester with embedded titania nanoparticles in agreement with our Raman spectra simulation by means of
Solid polymer electrolyte
DFT methodologies. Our theoretical modeling of the polyester also envisaged the interaction of lithium-ion via a
tridentate coordination with oxygen atoms of carboxylate (–COO-) and ester (–COO–) groups. There is a slight
increase in the degree of polymerization and thermal stability for our lithium-ion polyester with increasing
titania nanoparticles concentration. In consequence, a two-order magnitude increase in the ionic conductivity
was observed, yielding 1.74× 10−4 S cm−1 for a 20% weight fraction of titania nanoparticles. Thus, involving a
novel approach based in the synergistic in situ preparation of lithium-ion poly (ethylene citrate) with titania
nanoparticles, this procedure contributes to solve the many safety and environmental issues of the constitutive
materials of lithium-ion solid polymer electrolytes.

1. Introduction particularly; poly (ethylene succinate) [12], poly (ethylene malonate)


[13], poly (ethylene citrate) [14–16], poly (glycidyl ether carbonate)
In the last years, solid polymer electrolytes have been extensively [17] pursuing an enhancement of ionic conductivity. The strategy of
studied as promising materials to substitute the liquid electrolytes in using these polyesters is also based on the increment of the ionic-pair
lithium batteries mainly because of their flexibility, safety and light dissociation of lithium-ion due to the interaction with (–COO–) ester
weight. The solid polymer electrolytes are basically composed of an groups. However, no relevant enhancement above ∼10−5 –
electronic insulating polymer and a lithium salt with high degree of 10−4 S cm−1 was obtained for the previously mentioned cases [12–18].
ionic-pair dissociation to assure good ionic conductivity [1–11]. Among In the case of lithium-ion poly (ethylene citrate) electrolytes, the effect
all polymers tested, the most frequently used are poly (ethylene oxide) of silica nanoparticles was also reported yielding ionic conductivities up
(PEO) [1,2], poly (acrylonitrile) (PAN) [3,4] or poly (methyl metha- to ∼ 10−4 S cm−1 at room temperature [19–21]. There is also recent
crylate) (PMMA) [5,6] and the most common lithium salts are nitrate experimental and theoretical research focusing in the ability of polye-
(LiNO3), perchlorate (LiClO4), tetrafluoroborate (LiBF4) or tri- sters as promising lithium-ion solid polymer electrolytes by the inter-
fluoromethanesulfonate (LiCF3SO3) [7–11]. Typically, the ionic-pair action of different available oxygen atom groups such as (–COO–) ester,
dissociation of lithium salts is increased by the lithium-ion affinity for carbonyl (–C]O), hydroxyl (–OH) and ether (–CO–) coexisting in the
ether (–CO–), nitrile (–C^N) or carbonyl (–COO–) groups of PEO, PAN same polymer structure [19,20]. In addition to the high degree of ionic-
or PMMA, respectively [1–11]. Moreover, there have been attempts to pair dissociation, the increase in the ionic conductivity of these solid
study the use of other ecofriendly polymer hosts such as polyesters, polymer electrolytes is also related to a decrease on the crystallinity of


Corresponding authors.
∗∗
Corresponding authors.
E-mail addresses: mromero@fq.edu.uy (M. Romero), rfaccio@fq.edu.uy (R. Faccio).

https://doi.org/10.1016/j.jpcs.2019.109082
Received 30 April 2019; Received in revised form 27 June 2019; Accepted 29 June 2019
Available online 05 July 2019
0022-3697/ © 2019 Elsevier Ltd. All rights reserved.
F. Pignanelli, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109082

Fig. 1. Scheme of our synergistic in situ preparation of lithium-ion ethylene citrate polyester with TiO2 nanoparticles.

the polymer and to the increase in the dissociation of lithium-ion salt by 253081) were dissolved separately in ethylene glycol and slowly
adding ceramic nanoparticles as fillers. In this sense, several studies poured to the previous solution; the volumes of titanium propoxide
performed by adding ceramic nanoparticles such as alumina (Al2O3) were calculated to obtain TiO2/polyester nano-composites of theore-
[21,22], silica (SiO2) [23] and titania (TiO2) [24,25] yielded to an tical compositions equal to: 0%, 15%, 20% and 25% by mass (named
improve in the room temperature conductivity of lithium-ion up Li–P-X with X = 0, 15, 20, 25 respectively).
to ∼ 10−4–10−3 S cm−1. In addition, there are significant contribu- The mixtures were maintained under magnetic stirring at 80 °C for
tions for the successful in situ sol-gel preparation of hybrid titania na- 12 h and a pH close to neutrality (pH∼7) was observed. Then, the
nostructures embedded in polymer hosts such as polyaniline [26], temperature was set to 150 °C in order to slowly evaporate the excess of
polypyrrole [27] and poly (ethylene oxide) [28]. Inspired by this pre- solvent and favor the polyesterification process until a transparent
vious works and also contributing to solve the many safety and en- light-yellowish gel was obtained in each case. Finally, the samples were
vironmental issues of the constitutive materials of solid polymer elec- dried for 6 h in a vacuum oven at 100 °C to remove the presence of
trolytes, our strategy involves a novel approach based in the synergistic residual solvent from our samples.
in situ preparation of a fluorine-free lithium-ion poly (ethylene citrate)
polyester with titania nanoparticles. In order to achieve this task, we
2.2. Characterization of the Li–P-X samples
use lithium carbonate salt as a lithium-ion source and titanium alkoxide
as precursor, which in situ hydrolysis to form titania nanoparticles is
Confocal Raman microscopy imaging and spectroscopy was per-
synergistically triggered by the polymer esterification process, as re-
formed using a WITec Alpha 300-RA confocal Raman spectrometer; the
presented in the scheme shown in Fig. 1. In addition, we provide a full
laser excitation wavelength used in this experiment was 785 nm in
characterization and discussion of the synthesis mechanism, structure,
order to avoid sample fluorescence and power was set to < 10 mW to
microstructure and ionic conductivity analysis for our samples.
diminish the risk of thermal decomposition by local heating. Raman
spectra were obtained by averaging a set of 150 measurements with
2. Materials and methods 0.5 s as integration time. Raman spectra simulation was performed
using density functional theory (DFT) [29,30] using the hybrid ex-
2.1. Preparation of Li–P-X samples change correlation potential B3LYP [31–34] for a 6–31 G (d,p)+ basis
set, as implemented in Gaussian 09 [35]. X-ray powder diffraction
Citric acid monohydrate (CA) with Sigma-Aldrich ID: C1909 (XRD) measurements were performed using a Rigaku Ultima IV dif-
(1.7794 g) and lithium carbonate (LC) with Sigma-Aldrich ID: 255823 fractometer, utilizing CuKα radiation, operating in the
(0.0632 g) in a CA:LC ∼ 10:1 M ratio were dissolved in 40 mL of ethy- 2θ = 5.00–70.00° range using steps of 0.02° with a 12 s integration time
lene glycol (anhydrous, 99.8%) under magnetic stirring at 80 °C. per step. Grazing incidence small angle X-ray scattering (GI-SAXS)
Appropriate quantities of titanium (IV) propoxide (Sigma-Aldrich ID: measurements were taken in parallel beam configuration in the range

2
F. Pignanelli, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109082

Fig. 2. Raman microscopy images of polyester-rich (blue) and TiO2-rich


(yellow) zones for Li–P-X with X = 0, 15, 20 and 25 samples.

q = 0.01–1.80 Ε−1 and the fixed incident angle at +0.2° above the
critical angle. Thermal decomposition of samples was analyzed by
thermogravimetric analysis (TGA) using Shimadzu TGA-50, with a
50 mL min−1 air flow and a heating rate of 5°Cmin-1 in the 25–900 °C
temperature range. Differential Scanning Calorimetry (DSC) analysis
was obtained using Shimadzu DSC-60 with a 10 °C/min ramp rate in the
25–300 °C temperature range using sealed aluminum crimps with static
air atmosphere. Electrochemical impedance spectroscopy (EIS) mea-
surements were performed for the pellets with 1.0 cm surface/thickness
ratio sputtered gold blocking electrodes. EIS data was obtained using a
10 mV AC voltage amplitude in the 0.1 Hz − 1.0 MHz frequency range
at 20 °C. DC polarization using a Gamry Reference 3000 impedance
analyzer and data was processed using Echem Analyst software.

Fig. 3. Confocal Raman spectra of selected polyester-rich (blue) and TiO2-rich


3. Results and discussion
(yellow) zones for Li–P-X with X = 0, 15, 20 and 25 samples. Raman spectra for
Li2CO3 (grey), CA (red) and EG (green) precursors are shown as references
3.1. - Confocal Raman microscopy (REFs) and theoretical Raman spectra calculation for CA (red), EG (green) and
EG-(AC/Li)-EG (blue) models using Density Functional Theory (DFT) are shown
Confocal Raman microscopy images for Li–P-X with X = 0, 15, 20 with their corresponding assignation of vibrational modes. The dotted lines
and 25 samples are shown in Fig. 2. The confocal Raman images were indicate selected Raman peaks position of precursor references as a guide to the
obtained for a 10 × 10 μm2 area by collecting 40 × 40 Raman spectra eye to show the Raman shifting of peaks corresponding Li–P-X with X = 0, 15,
with ∼270 nm/pixel resolution totalizing 1600 Raman spectra. The 20 and 25 samples.
contrast between the two complementary zones shown in the images
were filtered by the relative intensity of a sharp peak at ∼150 cm−1 anions when reacting with the CA residual protons in aqueous media,
typically ascribed to the Ti–O stretching mode of TiO2 anatase poly- possibly as carbon dioxide (CO2) gas, as previously proposed in litera-
morph [36] and on the other hand, a peak at ∼ 1460 cm−1 typically ture [14–16]. In order to assign the vibrational modes for our polyester
associated to the C–H bending mode of aliphatic chains of polymers, samples with more accuracy, we performed simulations of Raman
already observed for ethylene citrate polyester [37], defining a TiO2- spectra by means of DFT calculations. In Fig. 3, we showed the calcu-
rich (yellow) and polymer-rich (blue) zones in the images, respectively. lated Raman spectra for theoretical optimized structures of CA, EG and
The presence of TiO2 anatase nanoparticles is evidenced by the a small oligomer based in the esterification of two molecules of EG with
presence of yellow spots in the confocal Raman images, showing rela- one molecule of CA and a lithium-ion bonded to the free carboxylate
tively good homogeneity in their distribution embedded in the polymer group of CA; named as EG (AC-Li)EG, as displayed in Fig. 4a.
matrix. It is also observed that the increasing concentration of TiO2 The DFT optimized structure of EG (AC-Li)EG showed the interac-
nanoparticles (X) also lead to an increment in the agglomeration of tion of lithium-ion via a tridentate coordination with carbonyl (–C]O)
TiO2 nanoparticles evidenced by the formation of larger sized clusters. groups from (–COO-) carboxylate and (–COO–) ester groups. The most
Raman spectra associated to each zone in the confocal Raman images relevant features of the theoretically simulated Raman spectra of EG
for Li–P-X with X = 0, 15, 20 and 25 samples together with the Raman (AC-Li)EG are basically, the splitting of the C]O stretching mode of CA
spectra of lithium carbonate salt (Li2CO3), citric acid (CA) and ethyle- from 1730 cm−1 to 1790, 1710 and 1630 cm−1 for EG (AC-Li)EG, the
neglycol (EG) precursors are shown in Fig. 3. At first inspection, we slight shift of the CH2 scissoring modes of EG at ∼ 1500 cm−1 and CA
observe the broadening and splitting of some Raman peaks in our Li–P- at 1480 cm−1 towards ∼ 1466 cm−1 for EG (AC-Li)EG, the splitting of
X samples, respect to the isolated precursors and no clear sign of the the CH2 wagging mode of EG at ∼ 1380 cm−1 and CA at ∼ 1240 cm−1
sharp peak of carbonate anion stretching at ∼1100 cm−1 in the Li–P-X towards ∼ 1330, 1300, 1280 cm−1 for EG (AC-Li)EG, the splitting of
samples. This could be suggesting the partial elimination of carbonate

3
F. Pignanelli, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109082

estimation of the mean TiO2 nanoparticles size using the Beaucage


equation and assuming a spherical morphology was D ∼50 nm and no
drastic variations were observed for any composition. At high-q region,
the presence of low amounts of TiO2 nanoparticles (X = 15) showed the
suppression of the peaks associated to the correlation distances of the
polyester, suggesting that the nanoparticles are probably filling the
polyester interdomain available free space. However, the increasing
amounts of TiO2 nanoparticles showed the emergence of broad peaks at
q ∼0.6 and 0.4 nm−1, with associated d ∼10 and 13 nm correlation
distances, for X = 20 and 25, respectively. These correlation distances
are probably associated to a modification of the average interdomain
spacing of the polyester but can also be related to the interspacing of
TiO2 nanoparticles when agglomerated to form large-sized clusters
embedded in the polyester.
X-ray powder diffraction patterns for Li–P-X with X = 0, 15, 20 and
25 samples are shown in Fig. 5b. The unloaded polyester (X = 0) pat-
tern showed the presence of a broad halo at 2θ ∼ 20.5° which can be
associated to a mean correlation distance (dC) between the polyester
backbone chains which can be estimated using the Bragg law yielding
Fig. 4. (a) Lithium-ion ethylene glycol citrate model, named as EG (AC-Li)EG, dC ∼ 0.43 nm. The presence of increasing amounts of TiO2 nano-
after optimization using DFT calculations. (b) Pictorial scheme for the Li–P-X particles lead to the emergence of peaks associated to the TiO2 anatase
samples depicting the interchain (dC) and interdomain (d) distances for the polymorph, which are evident for X = 25. In addition, the broad halo
polyester. ascribed to the polyester backbone interchain distance showed a slight
decrease in intensity and a shift towards higher 2θ values with the in-
the C–C backbone stretching mode at 1080 cm−1 of CA and the C–O crease of TiO2 nanoparticles. The slight decrease in the relative in-
stretching mode at 1000 cm−1 towards several peaks with high relative tensity of the broad halo could be suggesting a declining of the degree
intensity at ∼ 1000–1100 cm−1 for EG (AC-Li)EG, the shifting of the of order of the polyester structure. In addition, the broad peak slight
C–C–C backbone bending at ∼ 920 cm−1 to 970 cm−1 for EG (AC-Li) shift towards 2θ ∼22.2°, with an associated dC ∼0.40 nm mean in-
EG, the non-modification of HOC–C stretching mode of EG at terchain distance, is probably suggesting that the increasing amounts of
830 cm−1, the central C–COO stretching mode of CA from ∼750 to TiO2 nanoparticles could be yielding to a slight compression of the
840 cm−1 for EG (AC-Li)EG and finally the shift of the OCO–H mode of polyester chains, as schematized in Fig. 4b.
CA from ∼550 towards 610 cm−1 for EG (AC-Li)EG. In general, the
predictions of the simplified EG (AC-Li)EG model are in quite good 3.3. – Thermogravimetric analysis (TGA) and differential scanning
agreement with the trend experimentally observed for Li–P-X samples. calorimetry (DSC)
It is interesting to remark that the increasing amounts of TiO2 nano-
particles lead to a more notorious shifting and splitting of the Raman Thermogravimetric analysis (TGA) profiles for Li–P-X with X = 0,
peaks ascribed to the polyester, in agreement with the trends envisaged 15, 20 and 25 samples are shown in Fig. 6a. The unloaded lithium-ion
by our theoretical calculations for the EG (AC-Li)EG model. This could ethylene glycol citrate polymer (X = 0) showed a two-step thermal
be suggesting that the presence of increasing amounts of TiO2 nano- decomposition profile with a first ∼ 75% weight loss at T = 200–250 °C
particles lead to a higher degree of polymerization of the polyester and and a second 25% weight loss at T = 500–550 °C. The first weight loss
can be basically explained by the higher consumption of water mole- at T = 200–250 °C is possibly ascribed to the decarboxylation of the
cules for the titanium alkoxide hydrolysis, displacing the equilibrium to polyester while the second weight loss at T = 500–550 °C is typically
the esterification process. associated to the decomposition of the polyester backbone [37,40].
However, in addition to the two-stage thermal decomposition observed
for the unloaded polyester, Li–P-X with X = 20 and 25 samples also
3.2. - Small angle X-ray scattering (SAXS) and X-ray diffraction (XRD) showed a ∼10% weight loss at T = 120–200 °C, typically ascribed to
the presence of residual ethylene glycol. The first thermal decomposi-
Small angle X-ray scattering (SAXS) patterns for Li–P-X with X = 0, tion stage in the T = 220–450 °C temperature range showed an ap-
15, 20 and 25 samples are shown in Fig. 5a. The unloaded lithium-ion parent slight shift to higher temperatures with increasing TiO2 nano-
ethylene glycol citrate polymer (X = 0) showed the presence of two particles concentration, while the second thermal decomposition in the
well-defined peaks at q1 ∼ 1.50, q2 ∼ 0.83 nm−1 and a broader peak at T = 450–550 °C temperature range showed a slight shift to lower
q3 ∼ 0.54 nm−1. temperatures. Differential scanning calorimetry (DSC) studies were also
These well-defined peaks are probably associated to correlation performed to get better insight, especially on the first thermal decom-
distances (d) between polymer domains which can be roughly esti- position stage of Li–P-X samples, as depicted in Fig. 6b. DSC profiles for
mated using the Bragg's law as d = 2π/qi, with qi being the q value of a all samples consist of an endothermic peak at T = 200–250 °C, at al-
maximum peak in the SAXS pattern, as schematized in Fig. 4b. The most the same temperature range where the weight loss associated to
presence of a broad peak at q3 ∼ 0.54 nm−1 in the SAXS pattern, with the first stage thermal decomposition was observed in the TGA analysis.
an associated d3 ∼ 11 nm average inter domain spacing was also ob- This is corroborating that the weight loss observed in this temperature
served for other polyesters [38]. Moreover, the well-defined two peaks range can be associated to an endothermic process such as the dec-
at q1 and q2 are suggesting the presence of other shorter-range corre- arboxylation of the polyester, which involves the loss of mass, probably
lation lengths d1 ∼ 4 and d2 ∼ 7 nm, respectively, that can be probably in the form of carbon dioxide gas [37,40]. Although the endothermic
associated to interdomain distances with some degree of coherence for heat associated to the decarboxylation stage showed no drastic mod-
our titania unloaded lithium-ion polyester sample, as also observed for ifications for all cases, the decarboxylation temperature showed a shift
other polymers [39]. The increasing concentration of TiO2 nano- towards higher temperatures from 200 to 230 °C with the increase of
particles showed the progressive emergence of a bump at low q-region TiO2 nanoparticles concentration.
which can be associated to the formation of TiO2 nanoparticles. The This could be suggesting that there is an apparent increase in the

4
F. Pignanelli, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109082

Fig. 5. (a) Small angle X-ray scattering (SAXS) patterns indicating with arrows a rough estimation of the mean correlation lengths and (b) X-ray diffraction (XRD)
patterns indicating with arrows the mean coherence lengths for polyester and Bragg reflections of TiO2 anatase polymorph for Li–P-X samples.

decarboxylation temperatures [38]. The presence of heavier TiO2 na-


noparticles loadings (X = 25%) lead to no further increase but a slight
decrease of the decarboxylation temperature that could be related to a
slight decrease of the polyester degree of order for 25% in comparison
with X = 20%. This could be inducing the presence of more accessible
carboxylate groups yielding a slight decrease of the decarboxylation
temperature for larger amounts of TiO2 nanoparticles [38].

3.4. – Electrochemical impedance spectroscopy (EIS)

The total impedance (Z) can be defined in terms of the real part (Z’)
and the imaginary part (Z”) of the impedance response by the following
equation:
Z = Z′ − jZ"

Impedance spectroscopy analysis displayed as Nyquist plots (-Z” vs.


Z’) collected for Li–P-X with X = 0, 15, 20 and 25 samples are shown in
Fig. 7a. In all cases, the Nyquist plots showed a single semicircle-like arc
at the intermediate frequency regime (f > 103 Hz) associated to the
ionic transport followed by a straight line at the low frequency regime
(f < 103 Hz) which is typically associated to a sample-electrode dif-
fusion process [41]. The circuit model used to fit our impedance data
consists on the parallel combination of a resistor (R) and a constant
phase element (CPE), connected in serie with an additional resistor (Rs)
to describe the wire connections contribution and an additional CPE
element to describe the low frequency capacitive processes (CPEd) at
the gold blocking electrode interphase, as depicted in Fig. 7a. The im-
pedance of a CPE element is defined by the following equation:
1
ZCPE =
A0 (ωi)n

where A0 and n are frequency independent parameters, with n = 1 re-


sembling an ideal capacitor [41]. The most relevant fitting parameters
Fig. 6. (a) Thermogravimetric analysis (TGA) expressed as weight loss (straight
using the previously described circuit model are shown in Table 1.
lines) and corresponding first derivative (dotted lines) versus temperature for
Li–P-X samples. (b) Differential scanning calorimetry analysis (DSC) for Li–P-X As observed, the increasing amounts of TiO2 nanoparticles lead to a
samples. drastic decrease in the fitted ionic transport resistance (R) from
R ∼ 169 kohm down to a minimum value of R ∼ 5.8 kohm observed for
X = 20. The decrease in the resistance is directly related to the notor-
degree of polymerization of the polyester with increasing TiO2 nano-
ious decrease of the semicircle-like arc observed in the Nyquist plots, as
particles concentration, as also envisaged by the Raman spectroscopy
depicted in Fig. 7a and 7a-inset.
studies. The higher degree of polymerization in polyesters con-
The ionic conductivity (σ) was calculated using the following
comitantly leads to the decrease of carboxylate species at terminal lo-
equation:
cations, thus only allowing the decarboxylation of the more stable
carboxylate groups from the backbone structure with associated higher t
σ=
R. A

5
F. Pignanelli, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109082

Fig. 7. (a) Impedance spectroscopy data collected at room temperature (300 K) for Li–P-X samples displayed as Nyquist plots (-Z″ vs. Z′) and corresponding fittings
using the depicted circuit model. The inset shows a zoom in the low impedance region. (b) Lithium-ion conductivity at room temperature versus TiO2 nanoparticles
weight fractions for Li–P-X samples with the lines as a guide to the eye.

where t is the thickness and A is the electrode area of the solid polymer
electrolytes. As it was expected from the fitted resistance values, the
increasing amounts of TiO2 nanoparticles yielded an increment of al-
most two-order of magnitude from σ ∼10−6 up to a maximum value of
σ = 1,74 × 10−4 S cm−1 for X = 20. The drastic enhancement on the
ionic conductivity (or decrease of resistance) with increasing TiO2 na-
noparticles up to 20% could be due to the increment of the polyester
structure disorder promoted by the presence of TiO2, as already ob-
served for similar systems [21–25]. Since only a slight decrease in the
degree of order was evidenced by XRD analysis, another possible ex-
planation could be based in an enhancement of the ionic-pair dis-
sociation of lithium ion due interactions with TiO2 nanoparticles sur-
face. However, it also important to note that heavier TiO2 nanoparticles
loadings above X = 20 yielded a slight decrease in the conductivity,
suggesting an ionic percolation threshold at X = 20, as typically ob-
served for lithium-ion solid polymer electrolytes with other nano-
particles [42,43]. Arrhenius plots of ionic conductivity, displayed as ln Fig. 8. Arrhenius plots of lithium-ion conductivity for Li–P-X samples.
(σ) vs. 1000/T, for Li–P-X with X = 0, 15, 20 and 25 samples are shown
in Fig. 8. In all cases, a good linear behavior was observed for all cases,
suggesting no drastic deviations from a typical Arrhenius behavior, as monitored as a function of time on application of fixed DC voltage
expressed by the following equation: across the sample with blocking electrodes. The results of DC polar-
ization measurements on our Li–P-X samples by applying a 2.0 V DC
Ea bias voltage at 300 K are shown in Fig. 9, suggesting that the electronic
ln (σ ) = ln(σ0) −
kB T residual conductivity is negligible compared to the ionic transport for
all the Li–P-X electrolytes, as similarly observed in the literature
where σ0 is the pre-exponential factor, Ea is the activation energy and [45,46]. Finally, in order to estimate the ionic residual conductivity
kB is the Boltzmann constant. The activation energies obtained from the associated to other ionic species in our polyester samples; possibly re-
slope of Arrhenius plots were Ea = 0.36, 0.31, 0.24 and 0.27 eV for sidual protons, a lithium-ion unloaded sample was also prepared and
X = 0, 15, 20 and 25. This is suggesting that there is a decrease in the measured as a reference. For this purpose, we used another carbonate
activation energy with increasing TiO2 nanoparticles loadings yielding salt with the aim to preserve the pH close to neutrality and avoid extra
a minimum for X = 20; the same concentration for which the highest acidity of the medium if no carbonate is added.
lithium-ion conductivity was observed. Thus, to have a better comparison with our samples, the reference
We estimate the residual electronic conductivity by means of was prepared using barium carbonate instead of lithium carbonate but
Wagner's polarization technique [44]. Here, the DC current is in the same concentration to reproduce, as close as possible, the

Table 1
Most relevant impedance spectroscopy fitting parameters for Li–P-X with X = 0, 15, 20 and 25.
X=0 X = 15 X = 20 X = 25

R (103 Ω) 169 ± 1 32.0 ± 0.2 5.76 ± 0.08 6.66 ± 0.07


A0 (10−11S.sn) 2.00 ± 0.02 34.2 ± 0.2 45.0 ± 0.7 52.0 ± 0.7
n 0.909 ± 0.008 0.904 ± 0.005 0.89 ± 0.01 0.88 ± 0.01
σ (10−4S.cm−1) 0.0592 ± 0.0004 0.313 ± 0.002 1.74 ± 0.02 1.50 ± 0.02

The thickness to electrode area ratio (t/A) for our electrolytes was normalized to t/A ∼ 1 cm−1. The circuit model used for fitting analysis is shown in Fig. 6a-inset
and for all cases, Rs showed values < 7 ohms

6
F. Pignanelli, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109082

3864–3873.
[2] S. Choudhary, R.J. Sengwa, Effect of different anions of lithium salt and MMT
nanofiller on ion conduction in melt-compounded PEO–LiX–MMT electrolytes,
Ionics 18 (2012) 379–384.
[3] R. Sengwa, S. Choudhary, Dielectric properties and fluctuating relaxation processes
of poly (methyl methacrylate) based polymeric nanocomposite electrolytes, J. Phys.
Chem. Solids 75 (2014) 765–774.
[4] W. Liu, N. Liu, J. Sun, P.-C. Hsu, Y. Li, H.-W. Lee, Y. Cui, Ionic conductivity en-
hancement of polymer electrolytes with ceramic nanowire fillers, Nano Lett. 15
(2015) 2740–2745.
[5] D. Ostrovskii, P. Jacobsson, Concentrational changes in PAN-based polymer gel
electrolyte under current flow: in situ micro-Raman investigation, J. Power Sources
97 (2001) 667–670.
[6] (a) M. Ulaganathan, C.M. Mathew, S. Rajendran, Highly porous lithium-ion con-
ducting solvent-free poly (vinylidene fluoride-co-hexafluoropropylene)/poly (ethyl
methacrylate) based polymer blend electrolytes for Li battery applications,
Electrochim. Acta 93 (2013) 230–235;
(b) N. Shukla, A.K. Thakur, Ion transport model in exfoliated and intercalated
polymer–clay nanocomposites, Solid State Ionics 181 (2010) 921–932.
[7] M. Deka, A. Kumar, Enhanced electrical and electrochemical properties of
PMMA–clay nanocomposite gel polymer electrolytes, Electrochim. Acta 55 (2010)
1836–1842.
[8] S. Ramesh, K.C. Wong, Conductivity, dielectric behaviour and thermal stability
studies of lithium ion dissociation in poly(methyl methacrylate)-based gel polymer
electrolytes, Ionics 15 (2009) 249–254.
Fig. 9. DC polarization current versus time for Li–P-X samples. [9] S. Ahmad, T.K. Saxena, S. Ahmad, S.A. Agnihotry, The effect of nanosized TiO2
addition on poly(methylmethacrylate) based polymer electrolytes, J. Power Sources
159 (2006) 205–209.
synthesis conditions of our samples. The use of barium carbonate is [10] A.M.M. Ali, M.Z.A. Yahya, H. Bahron, R.H.Y. Subban, M.K. Harun, I. Atan,
mainly based in the expected low conductivities down to 10−11 S cm−1 Impedance studies on plasticized PMMA-LiX [X: CF3SO3-, N(CF3SO2)2-] polymer
electrolytes, Mater. Lett. 61 (2007) 2026–2029.
observed for the heavy Ba2+ ion in a similar poly (ethylene oxide) host
[11] S. Ramesh, L. Wen, Investigation on the effects of addition of SiO2 nanoparticles on
[47]. In addition, it is important to remark that we did not observe ionic conductivity, FTIR, and thermal properties of nanocomposite
detectable differences on the crystallinity between this reference PMMA–LiCF3SO3–SiO2, Ionics 16 (2010) 255–262.
sample and our Li–P-X samples that could induce indirect variations in [12] M. Watanabe, M. Rikukawa, K. Sanui, N. Ogata, H. Kato, T. Kobayashi, Z. Ohtaki,
Ionic conductivity of polymer complexes formed by poly(ethylene succinate) and
the conductivity [48,49]. Thus, with the use of barium carbonate in the lithium perchlorate, Macromolecules 17 (1984) 2902–2908.
reference sample, apart from reproducing as close as possible the [13] Y.C. Lee, M.A. Ratner, D.F. Shriver, Ionic conductivity in the poly(ethylene mal-
synthesis conditions of our Li–P-X samples, it is possible to consider that onate)/lithium triflate system, Solid State Ionics 138 (2001) 273–276.
[14] F.L. Souza, P.R. Bueno, E. Longo, E.R. Leite, Sol–gel nonhydrolytic synthesis of a
the major contribution to the ionic transport will be that ascribed to hybrid organic–inorganic electrolyte for application in lithium-ion devices, Solid
ions with higher mobility such as residuals protons [50,51]. The ionic State Ionics 166 (2004) 83–88.
conductivity of the lithium-ion unloaded polyester reference showed a [15] E.R. Leite, F.L. Souza, P.R. Bueno, S. de Lazaro, E. Longo, Hybrid
Organic−Inorganic Polymer: a new approach for the development of decoupled
σ ∼10−8 S cm−1 conductivity, which is two orders of magnitude lower polymer electrolytes, Chem. Mater. 171 (2005) 4561–4563.
than the lowest observed for lithium-ion conductivities in our samples [16] C.E. Tambelli, J.P. Donoso, C.J. Magon, E.C. Pereira, A.V. Rosario, NMR, con-
(σ ∼ 10−6 – 10−4 S cm−1). Finally, by making a rough extrapolation ductivity and DSC study of Li+ transport in ethylene glycol/citric acid polymer gel,
Electrochim. Acta 53 (2007) 1535–1540.
with all the suppositions described above, this is suggesting that the [17] Y. Tominaga, T. Shimomura, M. Nakamura, Alternating copolymers of carbon di-
contribution of residual protons in our Li–P-X samples is considerably oxide with glycidyl ethers for novel ion-conductive polymer electrolytes, Polymer
lower than that associated to lithium-ions. 51 (2010) 4295–4298.
[18] E.M. Masoud, Citrated porous gel copolymer electrolyte composite for lithium ion
batteries application: an investigation of ionic conduction in an optimized crystal-
4. Conclusions line and porous structure, J. Alloy. Comp. 651 (2015) 157–163.
[19] M.A. Webb, Y. Jung, D.M. Pesko, B.M. Savoie, U. Yamamoto, G.W. Coates,
In the present manuscript, a synergistic in situ preparation of li- N.P. Balsara, Z.G. Wang, T.F. Miller, Systematic computational and experimental
investigation of lithium-ion transport mechanisms in polyester-based polymer
thium-ion poly (ethylene citrate) with titania nanoparticles as pro- electrolytes, ACS Cent. Sci. 1 (2015) 198–205.
mising fluorine-free lithium-ion solid polymer electrolyte is reported. [20] E. Yildirim, D. Dakshinamoorthy, M.J. Peretic, M.A. Pasquinelli, R.T. Mathers,
Confocal Raman spectroscopy analysis revealed the formation of the Synthetic design of polyester electrolytes guided by hydrophobicity calculations,
Macromolecules 49 (2016) 7868–7876.
polyester with embedded titania nanoparticles in agreement with our [21] P. Dhatarwal, S. Choudhary, R.J. Sengwa, Electrochemical performance of Li+-ion
Raman spectra simulation. Our DFT theoretical modeling of the conducting solid polymer electrolytes based on PEO–PMMA blend matrix in-
polyester envisaged the interaction of lithium-ion via a tridentate co- corporated with various inorganic nanoparticles for the lithium ion batteries,
Composites Communications 10 (2018) 11–17.
ordination with carbonyl (–C]O) groups from (–COO-) carboxylate and [22] E.M. Masoud, A.-A. El-Bellihi, W.A. Bayoumy, M.A. Mousa, Organic–inorganic
(–COO–) ester groups. The degree of polymerization slightly increases composite polymer electrolyte based on PEO–LiClO4 and nano-Al2O3 filler for li-
and the thermal stability also improves for our lithium-ion polyester thium polymer batteries: dielectric and transport properties, J. Alloy. Comp. 575
(2013) 223–228.
with increasing titania nanoparticles concentration. In consequence, a [23] J. Shin, S. Passerini, PEO LiN(SO2CF2CF3)2 polymer electrolytes V. Effect of fillers
two-order magnitude increase in the ionic conductivity was observed on ionic transport properties, J. Electrochem. Soc. 151 (2004) A238–A245.
yielding 1.74× 10−4 S cm−1 for a 20% weight fraction of titania na- [24] B. Scrosati, F. Croce, L. Persi, Impedance spectroscopy study of PEO‐based nano-
composite polymer electrolytes, J. Electrochem. Soc. 147 (2000) 1718–1721.
noparticles.
[25] F. Croce, R. Curini, A. Martinelli, L. Persi, F. Ronci, B. Scrosati, R. Caminiti, Physical
and chemical properties of nanocomposite polymer electrolytes, J. Phys. Chem. B
Acknowledgements 103 (1999) 10632–10638.
[26] D. Mombrú, M. Romero, R. Faccio, A.W. Mombrú, Microstructure evolution,
thermal stability and fractal behavior of water vapor flow assisted in situ growth
The authors wish to thank the Uruguayan ANII, CSIC and PEDECIBA poly(vinylcarbazole)-titania quantum dots nanocomposites, J. Phys. Chem. Solids
funding institutions. 111 (2017) 199–206.
[27] G.J.A.A. Soler-Illia, E. Scolan, A. Louis, P.A. Albouy, C. Sanchez, Design of meso-
structured titanium oxo based hybrid organic–inorganic networks, New J. Chem. 25
References (2001) 156–165.
[28] S. Roux, G.J.A.A. Soler-Illia, S. Demoustier-Champagne, P. Audebert, C. Sanchez,
[1] Y. Kumar, S. Hashmi, G. Pandey, Ionic liquid mediated magnesium ion conduction Titania/polypyrrole hybrid nanocomposites built from in‐situ generated organically
in poly(ethylene oxide) based polymer electrolyte, Electrochim. Acta 56 (2011) functionalized nanoanatase building blocks, Adv. Mater. 15 (2003) 2117–2221.

7
F. Pignanelli, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109082

[29] W. Kohn, L.J. Sham, Self-consistent equations including exchange and correlation Technology 6 (2015) 221–230.
effects, Phys. Rev. 140 (1965) A1133. [41] J.T.S. Irvine, D.C. Sinclair, A.R. West, Electroceramics: characterization by im-
[30] P. Hohenberg, W. Kohn, Density functional theory, Phys. Rev. B 136 (1964) pedance spectroscopy, Adv. Mater. 2 (1990) 132–138.
864–876. [42] M. Romero, R. Faccio, S. Vázquez, A.W. Mombrú, Enhancement of lithium con-
[31] R.G. Parr, W. Yang, Density-Functional Theory of Atoms and Molecules, Oxford ductivity and evidence of lithium dissociation for LLTO-PMMA nanocomposite
University Press, New York, 1989. electrolyte, Mater. Lett. 172 (2016) 1–5.
[32] A.D. Becke, Thermochemistry Density‐Functional Iii, The role of exact exchange, J. [43] F. Pignanelli, M. Romero, R. Faccio, L. Fernández-Werner, A.W. Mombrú,
Chem. Phys. 98 (1993) 5648–5652. Enhancement of lithium-ion transport in poly (acrylonitrile) with hydrogen titanate
[33] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson, D.J. Singh, nanotube fillers as solid polymer electrolytes for lithium-ion battery applications, J.
C. Fiolhais, Atoms, molecules, solids, and surfaces: applications of the generalized Phys. Chem. C 122 (2018) 1492–1499.
gradient approximation for exchange and correlation, Phys. Rev. B 46 (1992) 6671. [44] J.B. Wagner, C. Wagner, Electrical conductivity measurements on cuprous halides,
[34] P.J. Hay, W.R. Wadt, Ab initio effective core potentials for molecular calculations. J. Chem. Phys. 26 (1957) 1597.
Potentials for the transition metal atoms Sc to Hg, J. Chem. Phys. 82 (1985) [45] D. Kumar, S.A. Hashmi, Ionic liquid based sodium ion conducting gel polymer
270–283. electrolytes, Solid State Ionics 181 (2010) 416–423.
[35] M.J. Frisch, et al., Gaussian, Inc., Wallingford CT, (2009). [46] Z. Osman, M.I. Mohd Ghazali, L. Othman, K.B. Md Isa, AC ionic conductivity and
[36] F. Tian, Y. Zhang, J. Zhang, C. Pan, Raman spectroscopy: a new approach to DC polarization method of lithium ion transport in PMMA–LiBF4 gel polymer
measure the percentage of anatase TiO2 exposed (001) facets, J. Phys. Chem. C 116 electrolytes, Results in Physics 2 (2012) 1–4.
(2012) 7515–7519. [47] J.J. Fontanella, M.C. Wintersgill, J.P. Calame, C.G. Andeen, Electrical and differ-
[37] M. Romero, R. Faccio, J. Martínez, H. Pardo, B. Montenegro, C.C. Plá Cid, A.A. Pasa, ential scanning calorimetry studies of poly(ethylene oxide) complexed with alka-
A.W. Mombrú, Effect of lanthanide on the microstructure and structure of line‐earth thiocyanates, Journal of Polymer Science B, Polymer Physics 23 (1985)
LnMn0.5Fe0.5O3 nanoparticles with Ln=La, Pr, Nd, Sm and Gd prepared by the 113–120.
polymer precursor method, J. Solid State Chem. 221 (2015) 325–333. [48] B. Bhattacharya, J.Y. Lee, J. Geng, H.T. Jung, J.K. Park, Effect of cation size on solid
[38] S.L. Chang, T.L. Yu, C.C. Huang, W.C. Chen, K. Linliu, T.L. Lin, Effect of polyester polymer electrolyte based dye-sensitized solar cells, Langmuir 25 (2009)
side-chains on the phase segregation of polyurethanes using small-angle X-ray 3276–3281.
scattering, Polymer 39 (1998) 3479–3489. [49] T.M.W.J. Bandara, W.J.M.J.S.R. Jayasundara, M.A.K.L. Dissanayake, M. Furlani,
[39] X. Wang, M. Goswami, R. Kumar, B.G. Sumpter, J. Mays, Morphologies of block I. Albinsson, B.E. Mellander, Effect of cation size on the performance of dye sen-
copolymers composed of charged and neutral blocks, Soft Matter 8 (2012) sitized nanocrystalline TiO2 solar cells based on quasi-solid state PAN electrolytes
3036–3052. containing quaternary ammonium iodides, Electrochim. Acta 109 (2013) 609–616.
[40] M. Romero, H. Pardo, R. Faccio, L. Suescun, S. Vázquez, I. Laborda, L. Fernández- [50] R.D. Shannon, Revised effective ionic radii and systematic studies of interatomic
Werner, A. Acosta, J. Castiglioni, A.W. Mombrú, A study on the polymer precursor distances in halides and chalcogenides, Acta Crystallogr. A32 (1976) 751.
formation and microstructure evolution of square-shaped (La0.5Ba0.5) [51] A.W. Adamson, Textbook of Physical Chemistry, Academic Press inc., London,
(Mn0.5Fe0.5)O3 ceramic nanoparticles, Journal of Ceramic Science and 1973.

You might also like