You are on page 1of 10

Optik - International Journal for Light and Electron Optics 208 (2020) 164558

Contents lists available at ScienceDirect

Optik
journal homepage: www.elsevier.com/locate/ijleo

Original research article

Influence of 1-methyl-3-propylimidazolium iodide ionic liquid on


T
the performance of dye-sensitized solar cell using hexanoyl
chitosan/poly(vinyl chloride) based polymer electrolyte
F.H. Muhammada, Tan Winieb,c,*
a
Centre of Foundation Studies, Universiti Teknologi MARA, 43800, Dengkil, Selangor, Malaysia
b
Faculty of Applied Sciences, Universiti Teknologi MARA, 40450, Shah Alam, Selangor, Malaysia
c
Institute of Science, Universiti Teknologi MARA, 40450, Shah Alam, Selangor, Malaysia

A R T IC LE I N F O ABS TRA CT

Keywords: Polymer electrolytes based on hexanoyl chitosan/poly(vinyl chloride) (PVC) were prepared by
Hexanoyl chitosan employing sodium iodide (NaI) as the doping salt and 1-methyl-3-propylimidazolium iodide
PVC (MPImI) as the ionic liquid. The concentration of MPImI was varied from 2 to 10 wt.% while the
Polymer electrolyte amount of polymer and salt were kept constant. The effect of MPImI on the crystallinity, struc-
Conductivity
tural and conductivity properties of hexanoyl chitosan/PVC-NaI have been investigated. XRD
Dye-sensitized solar cells
results show that MPImI disrupts the crystallinity of polymer matrix while FTIR results show the
existence of interactions among components in hexanoyl chitosan/PVC-NaI-MPImI. Polymer
electrolyte without MPImI shows the highest conductivity of 1.5 × 10−5 S cm−1 at 303 K. Upon
incorporation of MPImI, increment in conductivity can be observed. The hexanoyl chitosan/PVC-
NaI-MPImI electrolytes were assembled into dye-sensitized solar cells (DSSCs) and the depen-
dence of DSSCs performance on MPImI concentration has been investigated. DSSC with 8 wt.% of
MPImI shows the energy conversion efficiency, η of 4.55 % with short circuit current density, Jsc
of 10.34 mA cm-2 and open circuit voltage, Voc of 0.74 V.

1. Introduction

Dye-sensitized solar cells (DSSCs) are emerging as promising candidates for the next generation solar cells, owing to their ease of
fabrication, design flexibility and possibility for indoor application [1–3]. A typical DSSC consists of a photoanode, an electrolyte
with I−/I3− redox couple and a counter electrode. The electrolyte, separating the photoanode and counter electrode, acts as an
important medium for transportation of redox ions. Efficiency of DSSCs can be improved by optimizing the properties of electrolyte
such as ionic conductivity [4–6].
Liquid electrolytes are usually used in DSSCs as they provide high conversion efficiency. However, they are likely to evaporate,
leak, corrode the counter electrode and cause photodegradation of dye, which make them unstable for long-run. To overcome these
shortcomings, polymer electrolytes have emerged as potential alternatives to liquid electrolytes [7–11].
Polymer electrolytes comprise of polymer matrixes, in which numerous salts are dissolved. Most of the binary systems of polymer
and salt do not exhibit sufficient conductivity for applications [12–14]. Hence, several strategies to improve the conductivity of
polymer electrolytes have been adopted. This includes addition of plasticizer [15,16], inorganic filler [17–19] and ionic liquid


Corresponding author at: Faculty of Applied Sciences, Universiti Teknologi MARA, 40450, Shah Alam, Selangor, Malaysia.
E-mail address: tanwinie@uitm.edu.my (T. Winie).

https://doi.org/10.1016/j.ijleo.2020.164558
Received 9 January 2020; Received in revised form 26 February 2020; Accepted 11 March 2020
0030-4026/ © 2020 Elsevier GmbH. All rights reserved.
F.H. Muhammad and T. Winie Optik - International Journal for Light and Electron Optics 208 (2020) 164558

[20–22].
Our strategy is to use ionic liquid as conductivity enhancer. Ionic liquids are composed of organic cations and inorganic anions
[23,24]. Ionic liquids have several advantages over organic plasticizing solvents such as non-flammability, non-volatility, high
chemical and thermal stabilities [25–27]. These properties make ionic liquids good candidates for application in DSSCs.
In our previous work, hexanoyl chitosan/poly(vinyl chloride) (PVC)-sodium iodide (NaI) solid polymer electrolytes have been
prepared and tested in DSSCs [28]. The efficiency achieved was 2.93 %. Continuous effort to improve the efficiency of DSSCs has
prompted us to incorporate 1-methyl-3-propylimidazolium iodide (MPImI) ionic liquid into the hexanoyl chitosan/PVC-NaI. As
anticipated, the short circuit current density, Jsc and efficiency have been enhanced. This paper sheds some lights on the efficiency
enhancement brought about by the MPImI. This paper also reports the effect of MPImI on the crystallinity and conductivity as well as
reveals the interactions among components in hexanoyl chitosan/PVC-NaI-MPImI. To the best of our knowledge, this is the first
report on using the MPImI ionic liquid in hexanoyl chitosan/PVC blends polymer electrolytes for DSSCs.

2. Experimental

2.1. Polymer electrolyte preparation and characterization

The preparation of hexanoyl chitosan/PVC-NaI electrolytes is similar to our previous work [28]. MPImI (Aldrich) was added to
the highest conducting sample [28]. The concentration of MPImI was varied from 2 to 10 wt.% while the amounts of hexanoyl
chitosan/PVC and NaI were kept constant. The composition of hexanoyl chitosan and PVC was kept at 90:10.
ATR-FTIR spectroscopic studies were carried out using a Thermo Fisher Scientific Nicolet iS10 spectrophotometer (USA). The
FTIR spectra were recorded in the frequency range between 600 and 4000 cm−1 with a resolution of 2 cm−1. X-ray diffraction (XRD)
studies were carried out using an X-Pert PRO XRD. The electrolyte samples were scanned with a beam of monochromatic CuKα of
wavelength λ =1.5418 Å between a 2θ angles of 5° and 80°. Conductivities were determined from the impedance measurements
using a HIOKI 3532-50 LCR Hi-tester impedance spectrometer in the frequency range of 50 Hz - 1 MHz.

2.2. DSSC assembly and characterization

TiO2 photoanodes and Pt counter electrodes were prepared according to our earlier works [28]. The TiO2 photoanodes were
immersed in a cis-bis(isothiocyanato)bis(2,2′-bipyridyl-4,4′-dicarboxylato) ruthenium (II) dye solution (N3) for 24 h. The dye-coated
TiO2 photoanodes were then taken out from the dye solution and dried at room temperature.
Iodine chips (I2) were added to the electrolyte solution of hexanoyl chitosan/PVC-NaI-MPImI and stirred until homogeneous. The
concentration of iodine was one tenth of the concentration of iodide, I−. For DSSC assembly, the film of hexanoyl chitosan/PVC-NaI-
MPImI-I2 was sandwiched between the TiO2 photoanode and the Pt counter electrode. The effective area of the cell was 0.20 cm2. The
current-voltage (J-V) characteristics and impedance measurements of the DSSCs were measured under illumination of 1000 W m−2
Xenon light source, from an Oriel Newport LCS-100 solar simulator, with a Metrohm Autolab potentiostat (PGSTAT128N).

Fig. 1. Overlaid FTIR spectra of MPImI, hexanoyl chitosan/PVC-NaI and hexanoyl chitosan/PVC-NaI-MPImI.

2
F.H. Muhammad and T. Winie Optik - International Journal for Light and Electron Optics 208 (2020) 164558

Fig. 2. Overlaid FTIR spectra of hexanoyl chitosan/PVC-NaI and hexanoyl chitosan/PVC-NaI-MPImI in the region from 1700–3200 cm−1.

3. Results and discussion

3.1. FTIR studies

Fig. 1 shows the overlaid FTIR spectra of MPImI, hexanoyl chitosan/PVC-NaI and hexanoyl chitosan/PVC-NaI-MPImI with 8 wt.%
MPImI. Their corresponding peaks are assigned based on the literature [29–34]. In the spectrum of hexanoyl chitosan/PVC-NaI, the
CH2 asymmetric stretching, vas(CH2) and CH2 symmetric stretching, vs(CH2) of hexanoyl chitosan are observed at 2933 and 2866
cm−1, respectively. The OCOR and O = C(NCOR)2 of hexanoyl chitosan are located at 1745 and 1715 cm−1, respectively. Other
characteristic peaks of hexanoyl chitosan are observed at 1162 cm−1 (CH2 twisting, τ(CH2)) and 721 cm−1 (CH2 rocking, ρ(CH2)).
Peaks observed at 1433 and 1239 cm−1 are corresponding to the CH2 wagging, ω(CH2) and CeH rocking of CHCl, ρ(CeH) of PVC.
Characteristic peaks of MPImI due to CeH stretching, v(CH) in CHee3 group, v(CH) in CHee2 group, CN] stretching in imidazole
ring, vring(C]N), CH bending in side CHee2 group, δ(CH), CHee bending in imidazole ring, δring(CeH) and CHe stretching of
imidazole ring, vring(CeH) are observed at 3075, 2963, 1570, 1456, 1168 and 750 cm−1, respectively.
Fig. 2 shows the FTIR spectra of hexanoyl chitosan/PVC-NaI and hexanoyl chitosan/PVC-NaI-MPImI in the region between 1700
and 3200 cm−1. Upon addition of MPImI, peaks due to hexanoyl chitosan at 2933 (vas(CH2)), 2866 (vs(CH2)), 1745 (OCOR) and 1715
cm−1 (OC](NCOR)2) are found to shift to 2938, 2869, 1748 and 1717 cm−1, respectively, with increase in their relative intensities.
Shown in Fig. 3 is the same spectra in the region between 1150 and 1550 cm−1. Presence of MPImI shifts the ρ(CeH) and ω(CH2) of
PVC from 1239 to 1241 cm−1 and 1433 to 1435 cm−1, respectively. In the region between 1150 and 1200 cm−1, MPImI exhibits a
peak at 1168 cm−1 (δring(CHe) in imidazole ring) (cf. Table 1) and hexanoyl chitosan/PVC-NaI exhibits a peak at 1162 cm−1, which

Fig. 3. Overlaid FTIR spectra of hexanoyl chitosan/PVC-NaI and hexanoyl chitosan/PVC-NaI-MPImI in the region from 1150–1550 cm−1.

3
F.H. Muhammad and T. Winie Optik - International Journal for Light and Electron Optics 208 (2020) 164558

Table 1
FTIR peak assignment for hexanoyl chitosan/PVC-NaI, MPImI, hexanoyl chitosan/PVC-NaI-MPImI and MPImI-NaI.
Component Wavenumber (cm−1) Assignments

Hexanoyl chitosan/PVC -NaI 2933 vas(CH2) of hexanoyl chitosan


2866 vs(CH2) of hexanoyl chitosan
1745 OCOR of hexanoyl chitosan
1715 O]C(NCOR)2 of hexanoyl chitosan
1625 O]CNeHR of hexanoyl chitosan
1433 ω(CH2) of PVC
1239 ρ(CeH) of CHCl of PVC
1162 τ(CH2) of hexanoyl chitosan
721 ρ(CH2) of hexanoyl chitosan
MPImI 3075 v(CeH) in CeH3
2963 v(CeH) in CeH2
1570 vring(C]N) in imidazole ring
1456 δ(CeH)
1168 δring(CeH) in imidazole ring
750 vring(CeH) in imidazole ring
Hexanoyl chitosan/PVC-NaI -MPImI 2938 vas(CH2) of hexanoyl chitosan (shifted from 2933 cm−1)
2869 vs(CH2) of hexanoyl chitosan (shifted from 2866 cm−1)
1748 OCOR of hexanoyl chitosan (shifted from 1745 cm−1)
1717 ]OC(NCOR)2 of hexanoyl chitosan (shifted from 1715 cm−1)
1435 ω(CH2) of PVC (shifted from 1433 cm−1)
1241 ρ(CHe) of CHCl of PVC (shifted from 1239 cm−1)
1165 Superimposition of τ(CH2) of hexanoyl chitosan with δring(CeH) of MPImI
NaI-MPImI (2:1) 3071 v(CeH) in CeH3 of MPImI (shifted from 3075 cm−1)
2956 v(CH) in CeeH2 of MPImI (shifted from 2963 cm−1)
1562 vring(CN]) in imidazole ring of MPImI (shifted from 1570 cm−1)
1451 δ(CeH) of MPImI (shifted from 1456 cm−1)
1162 δring(CHe) in imidazole ring of MPImI (shifted from 1168 cm−1)
744 vring(CeH) in imidazole ring of MPImI (shifted from 750 cm−1)

is attributed to the τ(CH2) of hexanoyl chitosan. Incorporation of MPImI into hexanoyl chitosan/PVC-NaI has resulted in the su-
perimposition of 1168 and 1162 cm−1 peaks to form an intense peak at 1165 cm−1.
The perturbation of peaks of MPImI, subsequent to the addition of NaI is shown in Fig. 4. The characteristic peaks of MPImI at
3075, 2963, 1570, 1456, 1168 and 750 cm−1 are found to shift to 3071, 2956, 1562, 1451, 1162 and 744 cm−1, respectively. Table 1
summarizes the interactions among components in the hexanoyl chitosan/PVC-NaI-MPImI.

3.2. XRD studies

In our previous works, we have reported that blending semicrystalline hexanoyl chitosan with amorphous PVC hindered the
crystallinity of hexanoyl chitosan [28] and incorporation of NaI further decreased the crystallinity of the polymer matrix [33]. Fig. 5
shows the X-ray diffractograms of hexanoyl chitosan/PVC-NaI with various MPImI concentrations. All diffractograms can be char-
acterized by a sharp crystalline peak at 2θ = 6.9° and a broad amorphous halo centered at 2θ = 25°. No peaks corresponding to the

Fig. 4. Overlaid FTIR spectra of MPImI and MPImI-NaI.

4
F.H. Muhammad and T. Winie Optik - International Journal for Light and Electron Optics 208 (2020) 164558

Fig. 5. Deconvoluted X-ray diffractograms of hexanoyl chitosan/PVC-NaI with (a) 0 wt.%; (b) 4 wt.%; (c) 6 wt.%; (d) 8 wt.% and (e) 10 wt.% of
MPImI.

NaI salt and MPImI ionic liquid indicates complete dissolution of NaI and MPImI in the polymer matrix. The diffractograms were
deconvoluted in order to obtain the values of IC and IT. IC is the area under the peak at 2θ = 6.9° and IT is the total area of the
diffractogram from 2θ = 5 to 50°. The obtained values of IC and IT were used to calculate the degree of crystallinity, XC according to

IC
Xc = × 100%
IT (1)

Fig. 6. Degree of crystallinity for hexanoyl chitosan/PVC-NaI with respect to MPImI concentration.

5
F.H. Muhammad and T. Winie Optik - International Journal for Light and Electron Optics 208 (2020) 164558

Fig. 7. Influence of MPImI concentration on the conductivity of the hexanoyl chitosan/PVC-NaI at 303 K.

Presented in Fig. 6 is the variation of the degree of crystallinity with respect to MPImI concentration. The degree of crystallinity is
observed to decrease from 24 % to 22 % when 4 wt.% of MPImI is added. The degree of crystallinity continues to decrease to 17 % as
the amount of MPImI increases to 8 wt.%. The decrease of crystallinity upon addition of MPImI may be attributed to the destruction
of the ordered phases of polymer matrix by MPImI. The polymer matrix consists of immiscible blend of hexanoyl chitosan and PVC
[28]. The ordered phases of polymer matrix are maintained by the hexanoyl chitosan because PVC is amorphous in nature. Hexanoyl
chitosan has a layered structure [35]. The polysaccharide main chains form layers and layers are separated by hydrocarbon side
chains. The polysaccharide chains are not coplanar and have different rotation angles relative to each other. The rotational freedom
of polysaccharide chains is constrained under hydrogen bonds, maintaining the ordered phases of hexanoyl chitosan. In the presence
of MPImI, some hydrogen bonds may be broken to surround the MPImI, causing destruction of the ordered phases of hexanoyl
chitosan.
XRD results show that MPImI ionic liquid is able to decrease the crystalline phase of polymer and therefore provides a larger
amorphous region for ionic conduction. However, addition of MPImI beyond the optimum concentration may result in agglomerate
formation which increases the crystalline phase of polymer. In amorphous region, rapid segmental motions of polymer chain increase
the mobility of ions. Therefore, changes in the crystallinity may be used to interpret the conductivity behaviour of polymer elec-
trolytes.

3.3. Conductivity studies

The conductivity values obtained at 303 K for hexanoyl chitosan/PVC-NaI with various concentrations of MPImI are presented in
Fig. 7. The conductivity is found to increase from 1.5 × 10−5 to 1.3 × 10-4 S cm-1 upon addition of 8 wt.% MPImI. The increment in
conductivity is because MPImI provides additional ions namely MPIm+ and I- ions for conduction besides Na+ and I- ions from NaI
salt. This increases the number of ions, n for conduction. The MPIm+ is bulky cation and hence its movement in the polymer matrix is
easily hindered by the entanglement of polymer chains. Hence, contribution to conductivity from MPIm+ is minimal. The con-
tribution to conductivity is mostly from I- anions. The conductivity increment can also be attributed to the plasticization effect of
MPImI. As shown in section 3.2, MPImI increases the amorphous phase of polymer, which increases the segmental motions of
polymer chain and hence increases the mobility of ions, μ.

3.4. DSSC performance

Fig. 8 shows the J-V characteristics of DSSCs having hexanoyl chitosan/ PVC-NaI electrolytes with various concentrations of
MPImI. Their performance parameters are listed in Table 2. The conversion efficiency, η and fill factor, ff of the DSSC were calculated
according to equations shown in [1].
As can be seen from Table 2, the photovoltaic performance particularly Jsc and η is strongly influenced by the MPImI con-
centration while the ff seems not dependent on the MPImI concentration. In the absence of MPImI, the Jsc exhibited by the DSSC is
8.62 mA cm−2. With addition of 6 wt.% MPImI, the Jsc increases to 8.80 mA cm−2. The Jsc continues to increase to 10.34 mA cm−2 as
MPImI concentration is increased to 8 wt.%, but thereafter, the Jsc decreases to 9.02 mA cm−2 upon addition of MPImI beyond 8 wt.
%. The increase of Jsc with MPImI concentration perhaps are associated with the increase in the iodide conductivity of the electrolyte
(cf. Section 3.3) and the absorption of the cations on the TiO2 electrode. Na+ cations with radius of 1.02 Å [36] can absorb to the
TiO2 surface and shift the Fermi level of TiO2 towards the redox potential. Such Fermi level shift increases the electron injection rate
from dye to the conduction band of TiO2 and leads to an increase in Jsc. On the other hand, for big MPIm+ cations, shift of the Fermi
level of TiO2 towards the redox potential is less as the big cations do not absorb easily to the TiO2 surface, but it can generate more I-

6
F.H. Muhammad and T. Winie Optik - International Journal for Light and Electron Optics 208 (2020) 164558

Fig. 8. J-V characteristics of (a) cell A, (b) cell B, (c) cell C and (d) cell D.

Table 2
Performance parameters for DSSCs assembled with hexanoyl chitosan/PVC-NaI-MPImI electrolytes.
Cell MPImI (wt.%) Jsc (mA cm−2) Voc (V) ff η (%)

A 0 8.62 0.58 0.59 2.93


B 6 8.80 0.70 0.61 3.82
C 8 10.34 0.74 0.60 4.55
D 10 9.02 0.71 0.61 4.17

ions because the ionic dissociation of big cation is greater. Higher Jsc value results in higher efficiency of DSSC. The highest efficiency
of 4.55 % exhibited by cell C is due to its highest Jsc value.
Bulky MPIm+ can accumulate near the TiO2 electrode and hence blocks the redox ions to the TiO2 electrode. This reduces
recombination of injected electrons with redox ions. The effect of MPImI on the kinetics of recombination was further investigated by
EIS. Fig. 9 shows the Nyquist plots of cell A without MPImI and cell C with 8 wt.% MPImI.
There are two semi-circles in the plots, one at low frequencies representing the recombination resistance, Rct and CPE capacitance,
Qct at the TiO2 electrode; the other one at high frequencies representing the charge transfer resistance, Rce and CPE capacitance, Qce at
the Pt counter electrode. The equivalent circuit representation is shown in inset Fig. 9. The plots can be fitted using the following
equation [37]:
1 1
Ztotal = Rs + +
⎡ R1 + Qce ω βce cos
⎣ ce ( )
βce π
2
⎤ + j ⎡Qce ω βce sin
⎦ ⎣ ( )
βce π
2


⎡ R1 + Qct ω βct cos
⎣ ct ( ) ⎤⎦ + j ⎡⎣Q ω
βct π
2 ct
βct sin ( ) ⎤⎦
βct π
2 (2)
where Z is the impedance of the circuit, Rs is the ohmic series resistance, β is the deviation of the semicircle diameter from the Zi axis
and ω is the angular frequency. Subscripts ce and ct denotes Pt electrode-electrolyte interface and TiO2 electrode-electrolyte interface,
respectively. The parameters of the equivalent circuit are tabulated in Table 3.
The values of Rct for cells A and C are 36.1 and 53.9 Ω, respectively. Cell C exhibits higher value of Rct indicates higher re-
combination resistance at the TiO2 electrode. Higher Rct value is beneficial to the performance of DSSC because it reduces re-
combination between injected electrons and redox ions.
Fig. 10 shows the Bode plots of cells A and C. The maximum frequency of the Bode plot, fmax is inversely proportional to the
electron recombination lifetime, τe.
1
τe =
2πfmax (3)

Fig. 9. Nyquist plots of (a) cell A and (b) cell C.

7
F.H. Muhammad and T. Winie Optik - International Journal for Light and Electron Optics 208 (2020) 164558

Table 3
The parameters of equivalent circuit for cells A and C.
Cell MPImI (wt.%) Rs (Ω) Rce (Ω) Rct (Ω) τe (ms)

A 0 31.2 4.2 36.1 18


C 8 24.8 13.5 53.9 39

Fig. 10. Bode plots of (a) cell A and (b) cell C.

The values of τe for cells A and C are 18 and 39 ms, respectively. The τe is longer in cell C. This means that in cell C, the excited
electrons are not easy to recombine with the redox ions due to the blockage by MPImI. Thus, lower recombination loss in cell C as
compared to cell A. Combination of higher Jsc and lower recombination loss leads to better performance of cell C with 8 wt.% of
MPImI with efficiency of 4.55 %.
Table 4 compares the performance of the DSSC obtained in this work with other polymer electrolyte-based DSSCs. It can be seen
that, generally, the efficiency obtained using polymer blend is higher than single polymer and efficiency obtained with ionic liquid is
higher than ionic liquid-free.

4. Conclusions

Polymer electrolytes based on hexanoyl chitosan/PVC-NaI-MPImI have been prepared and tested in DSSCs. XRD results show the
influence of MPImI on the crystallinity of the polymer matrix. Interactions among components in the hexanoyl chitosan/PVC-NaI-

Table 4
Performance of DSSCs with various polymer electrolytes.
Gel polymer electrolyte Sensitizer η (%) References

(PVDF-HFP)-KI-TBAI-I2 N3 1.88 [38]


PVDF-HFP/PEO-KI-TBAI-I2 N3 2.46
PVDF-KI-I2 N3 1.80 [39]
PVDF/PEO-KI-I2 N3 2.70
PVDF-KI-I2-BNIN N3 3.60
PVDF/PEO-KI-I2-BNIN N3 7.30
P(MMA-co-MAA)-NaI-I2-EC-PC N719 2.34 [40]
P(MMA-co-MAA)-NaI-I2-EC-PC-HMImI N719 5.30
PEO-LiI-I2 N3 2.06 [41]
PEO/PMMA-LiI-I2 N3 4.90
(P[VP-co-VAc])-KI-I2-EC-PC N3 2.94 [42]
(P[VP-co-VAc])-KI-I2-EC-PC-MPImI N3 4.67
PVC-g-POEM-LiI-I2 N3 0.80 [43]
PVC-g-POEM-LiI- I2-MPII N3 3.60
Phthaloyl chitosan/PEO-TPAI-I2-EC N3 5.82 [44]
Phthaloyl chitosan/PEO-TPAI-I2-EC-BMII N3 8.74
Hexanoyl chitosan/PVC-NaI-I2 N3 2.93 [28]
Hexanoyl chitosan/PVC-NaI-I2-MPImI N3 4.55 Present work

Notes: N3(Cis-bis(isothiocyanato)bis(2,2′-bipyridyl-4,4′-dicarboxylato)ruthenium(II) [Ru(4,4′-dicarboxy-2,2′-bipyridine)2(NCS)2]), N719 (Di-


tetrabutylammonium cis-bis(isothiocyanato)bis(2,2′-bipyridyl-4,4′-dicarboxylato)ruthenium(II)).

8
F.H. Muhammad and T. Winie Optik - International Journal for Light and Electron Optics 208 (2020) 164558

MPImI have been manifested experimentally in FTIR. Incorporation of MPImI increases the conductivity of hexanoyl chitosan/PVC-
NaI by 1) providing additional ions for conduction and 2) increasing the amorphous phase of polymer, which increases the segmental
motions of polymer chain and hence increases the movement of ions. MPImI has enhanced the efficiency of the DSSC. The highest
efficiency of 4.55 % is obtained with 8 wt.% MPImI. The efficiency enhancement is due to higher Jsc and lower recombination loss
brought about by the MPImI.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgements

The authors wish to thank the Ministry of Higher Education Malaysia for supporting this work through FRGS/1/2018/STG07/
UITM/02/12 and F.H. Muhammad thanks the University for the scholarship awarded.

References

[1] A. Hagfeldt, G. Boschloo, L. Sun, L. Kloo, H. Pettersson, Dye-sensitized solar cells, Chem. Rev. 110 (2010) 6595–6663.
[2] S. Venkatesan, I.-P. Liu, W.-N. Hung, H. Teng, Y.-L. Lee, Highly efficient quasi-solid-state dye-sensitized solar cells prepared by printable electrolytes for room
light applications, Chem. Eng. J. 367 (2019) 17–24.
[3] G.R. Mutta, S.R. Popuri, J.I.B. Wilson, N.S. Bennett, Sol-gel spin coated well adhered MoO3 thin films as an alternative counter electrode for dye-sensitized solar
cells, Solid State Sci. 61 (2016) 84–88.
[4] L.P. Teo, T.S. Tiong, M.H. Buraidah, A.K. Arof, Effect of lithium iodide on the performance of dye sensitized solar cells (DSSC) using poly(ethylene oxide) (PEO)/
poly(vinyl alcohol) (PVA) based gel polymer electrolytes, Opt. Mater. 85 (2018) 531–537.
[5] Y. Cui, X. Zhang, J. Feng, J. Zhang, Y. Zhu, Enhanced photovoltaic performance of quasi-solid-state dye-sensitized solar cells by incorporating a quaternized
ammonium salt into poly(ethylene oxide)/poly(vinylidene fluoride-hexafluoropropylene) composite polymer electrolyte, Electrochim. Acta 108 (2013)
757–762.
[6] N. Pavithra, D. Velayutham, A. Sorrentino, S. Anandan, Poly(ethylene oxide) polymer matrix coupled with urea as gel electrolyte for dye sensitized solar cell
applications, Synth. Met. 226 (2017) 62–70.
[7] J.H. Kim, M.-S. Kang, Y.J. Kim, J. Won, Y.S. Kang, Poly(butyl acrylate)/NaI/I2 electrolytes for dye-sensitized nanocrystalline TiO2 solar cells, Solid State Ion. 176
(2005) 579–584.
[8] A.F. Nogueira, J.R. Durrant, M.A.D. Paoli, Dye-sensitized nanocrystalline solar cells employing a polymer electrolyte, Adv. Mater. 13 (2001) 826–830.
[9] M. Olisha, S. Lobregas, D.H. Camacho, Gel polymer electrolyte system based on starch grafted with ionic liquid: synthesis, characterization and its application in
dye-sensitized solar cell, Electrochim. Acta 298 (2019) 219–228.
[10] Rahul, P.K. Singh, B. Bhattacharya, Z.H. Khan, Environment approachable dye sensitized solar cell using abundant natural pigment based dyes with solid
polymer electrolyte, Optik 165 (2018) 186–194.
[11] S.V. Kuppu, A.R. Jeyaraman, P.K. Guruviah, S. Thambusamy, Preparation and characterizations of PMMA-PVDF based polymer composite electrolyte materials
for dye sensitized solar cell, Curr. Appl. Phys. 18 (2018) 619–625.
[12] M. Ravi, S. Song, K. Gu, J. Tang, Z. Zhang, Electrical properties of biodegradable poly(ε-caprolactone): lithium thiocyanate complexed polymer electrolyte films,
Mater. Sci. Eng. B-Adv. 195 (2015) 74–83.
[13] T. Winie, A. Jamal, F.I. Saaid, T.-Y. Tseng, Hexanoyl chitosan/ENR25 blend polymer electrolyte system for electrical double layer capacitor, Polym. Adv.
Technol. 30 (2019) 726–735.
[14] N.K. Karan, D.K. Pradhan, R. Thomas, B. Natesan, R.S. Katiyar, Solid polymer electrolytes based on polyethylene oxide and lithium trifluoro-methane sulfonate
(PEO-LiCF3SO3): ionic conductivity and dielectric relaxation, Solid State Ion. 179 (2008) 689–696.
[15] T. Winie, A.K. Arof, Hexanoyl chitosan-based gel electrolyte for use in lithium-ion cell, Polym. Adv. Technol. 17 (2006) 552–555.
[16] P. Pal, A. Ghosh, Investigation of ionic conductivity and relaxation in plasticized PMMA-LiClO4 solid polymer electrolytes, Solid State Ion. 319 (2018) 117–124.
[17] S.U. Rehman, M. Noman, A.D. Khan, A. Saboor, M.S. Ahmad, H.U. Khan, Synthesis of polyvinyl acetate/ graphene nanocomposite and its application as an
electrolyte in dye sensitized solar cells, Optik 202 (2020) 163591.
[18] P.K. Singh, B. Bhattacharya, R.K. Nagarale, Effect of nano TiO2 dispersion on PEO polymer electrolyte property, J. Appl. Polym. Sci. 118 (2010) 2976–2980.
[19] C.R.M. Pila, E.P. Cappe, N.D.S. Mohallem, O.L. Alves, M.A.A. Frutis, N.S. Ramírez, R.M. Torresi, H.L. Ramírez, Y.M. Laffita, Effect of the LLTO nanoparticles on
the conducting properties of PEO-based solid electrolyte, Solid State Sci. 88 (2019) 41–47.
[20] C.M.S. Prasanna, S.A. Suthanthiraraj, Effective influences of 1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide (EMIMTFSI) ionic liquid on the ion
transport properties of micro-porous zinc-ion conducting poly (vinyl chloride)/poly(ethyl methacrylate) blend-based polymer electrolytes, J. Polym. Res. 23 (7)
(2016) 140, , https://doi.org/10.1007/s10965-016-1043-0.
[21] J. Tang, R. Muchakayala, S. Song, M. Wang, K.N. Kumar, Effect of EMIMBF4 ionic liquid addition on the structure and ionic conductivity of LiBF4-complexed
PVdF-HFP polymer electrolyte films, Polym. Test. 50 (2016) 247–254.
[22] T. Winie, A. Amisha, M.D. Rozana, Ionic liquid effect for efficiency improvement in poly(methyl acrylate)/poly(vinyl acetate)-based dye-sensitized solar cells,
High Perform. Polym. 30 (2018) 937–948.
[23] M. Galiński, A. Lewandowski, I. Stępniak, Ionic liquid as electrolytes, Electrochim. Acta 51 (2006) 5567–5580.
[24] L.N. Sim, S.R. Majid, A.K. Arof, Effects of 1–butyl–3–methyl imidazolium trifluoromethanesulfonate ionic liquid in poly(ethyl methacrylate)/ poly(vinylide-
nefluoride–co–hexafluoropropylene) blend based polymer electrolyte system, Electrochim. Acta 123 (2014) 190–197.
[25] J.Y. Yuan, D. Mecerreyes, M. Antonietti, Poly(ionic liquid)s: an update, Prog. Polym. Sci. 38 (2013) 1009–1036.
[26] M.E. Armand, F. Endres, D.R. Macfarlene, H. Ohno, B. Scrosati, Ionic liquid materials for the electrochemical challenges of the future, Nat. Mater. 8 (2009)
621–629.
[27] C.P. Lee, K.C. Ho, Poly(ionic liquid)s for dye-sensitized solar cells: a mini-review, Eur. Polym. J. 108 (2018) 420–428.
[28] F.H. Muhammad, R.H.Y. Subban, T. Winie, Solid solutions of hexanoyl chitosan/poly(vinyl chloride) blends and NaI for all-solid-state dye-sensitized solar cells,
Ionics 25 (2019) 3373–3386.
[29] Q. Li, Q. Tang, N. Du, Y. Qin, J. Xio, B. He, H. Chen, L. Chu, Employment of ionic liquid-imbibed polymer gel electrolyte for efficient quasi-solid-state dye-
sensitized solar cells, J. Power Sources 248 (2014) 816–821.
[30] I. Jerman, V. Jovanovski, A.S. Vuk, S.B. Hocevar, M. Gaberscek, A. Jesih, B. Orel, Ionic conductivity, infrared and Raman spectroscopic studies of 1-methyl-3-
propylimidazolium iodide ionic liquid with added iodine, Electrochim. Acta 53 (2008) 2281–2288.
[31] T. Winie, A.K. Arof, FT-IR studies on interactions among components in hexanoyl chitosan-based polymer electrolytes, Spectrochim. Acta A. 63 (2006) 677–684.
[32] M. Beltran, M. Marcilla, Fourier transform infrared spectroscopy applied to the study of PVC decomposition, Eur. Polym. J. 33 (1997) 1135–1142.

9
F.H. Muhammad and T. Winie Optik - International Journal for Light and Electron Optics 208 (2020) 164558

[33] F.H. Muhammad, A. Jamal, T. Winie, Study on factors governing the conductivity performance of acylated chitosan-NaI electrolyte system, Ionics 23 (2017)
3045–3056.
[34] F.H. Muhammad, R.H.Y. Subban, T. Winie, Charge carrier density and mobility of poly(vinyl chloride)-based polymer electrolyte using impedance spectroscopy,
Mater. Today Proc. 4 (2017) 5130–5137.
[35] Z. Zong, Y. Kimura, M. Takahashi, H. Yamane, Characterization of chemical and solid state structures of acylated chitosans, Polymer 41 (2000) 899–906.
[36] B. Bhattacharya, J.Y. Lee, J. Geng, H.-T. Jung, J.-K. Park, Effect of cation size on solid polymer electrolyte based dye-sensitized solar cells, Langmuir 25 (2009)
3276–3281.
[37] T.S. Tiong, M.H. Buraidah, L.P. Teo, A.K. Arof, Conductivity studies of poly(ethylene oxide)(PEO)/poly(vinyl alcohol)(PVA) blend gel polymer electrolytes for
dye-sensitized solar cells, Ionics 22 (2016) 2133–2142.
[38] R.A. Senthil, J. Theerthagiri, J. Madhavan, A.K. Arof, Performance characteristics of guanine incorporated PVDF-HFP/PEO polymer blend electrolytes with
binary iodide salts for dye-sensitized solar cells, Opt. Mater. 58 (2016) 357–364.
[39] S. Ganesan, B. Muthuraaman, Vinod Mathew, M. Kumara Vadivel, P. Maruthamuthu, M. Ashokkumar, S. Austin Suthanthiraraj, Influence of 2,6 (N-pyrazolyl)
isonicotinic acid on the photovoltaic properties of a dye-sensitized solar cell fabricated using poly(vinylidene fluoride) blended with poly(ethylene oxide)
polymer electrolyte, Electrochim. Acta 56 (2011) 8811–8817.
[40] V. Sundararajan, N.K. Farhana, H.M. Ng, S. Ramesh, K. Ramesh, Efficiency enhancement study on addition of 1-hexyl-3-methylimidazolium iodide ionic liquid to
the poly(methyl methacrylate-co-methacrylic acid) electrolyte system as applied in dye-sensitized solar cells, J. Phys. Chem. Solids 129 (2019) 252–260.
[41] E. Aram, M. Ehsani, H.A. Khonakdar, Improvement of ionic conductivity and performance of quasi-solid-state dye sensitized solar cell using PEO/PMMA gel
electrolyte, Thermochim. Acta 615 (2015) 61–67.
[42] H.M. Ng, S. Ramesh, K. Ramesh, Efficiency improvement by incorporating 1-methyl-3-propylimidazolium iodide ionic liquid in gel polymer electrolytes for dye-
sensitized solar cells, Electrochim. Acta 175 (2015) 169–175.
[43] D.K. Roh, J.T. Park, F. Sung Hoon Ahn, Hyungju Ahn, Du Yeol Ryu, Jong Hak Kim, Amphiphilic poly(vinyl chloride)-g-poly(oxyethylene methacrylate) graft
polymer electrolytes: interactions, nanostructures and applications to dye-sensitized solar cells, Electrochim. Acta 55 (2010) 4976–4981.
[44] M.H. Buraidah, Shahan Shah, L.P. Teo, Faisal I. Chowdhury, M.A. Careem, I. Albinsson, B.-E. Mellander, A.K. Arof, High efficient dye sensitized solar cells using
phthaloylchitosan based gel polymer electrolytes, Electrochim. Acta 245 (2017) 846–853.

10

You might also like