Download as pdf or txt
Download as pdf or txt
You are on page 1of 39

Journal Pre-proofs

Tunable dielectric and memory features of ferroelectric layered perovskite


Bi4Ti3O12 nanoparticles doped nematic liquid crystal composite

Anu, Depanshu Varshney, Kamlesh Yadav, Jai Prakash, Harikesh Meena,


Gautam Singh

PII: S0167-7322(22)02359-5
DOI: https://doi.org/10.1016/j.molliq.2022.120820
Reference: MOLLIQ 120820

To appear in: Journal of Molecular Liquids

Received Date: 19 April 2022


Revised Date: 8 November 2022
Accepted Date: 15 November 2022

Please cite this article as: Anu, D. Varshney, K. Yadav, J. Prakash, H. Meena, G. Singh, Tunable dielectric and
memory features of ferroelectric layered perovskite Bi4Ti3O12 nanoparticles doped nematic liquid crystal
composite, Journal of Molecular Liquids (2022), doi: https://doi.org/10.1016/j.molliq.2022.120820

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2022 Elsevier B.V. All rights reserved.


Author Statement

The authors have no competing interests for the article entitled “Tunable dielectric and
memory features of ferroelectric layered perovskite Bi4Ti3O12 nanoparticles doped nematic
liquid crystal composite”.

Highlights

 The impact of Bi4Ti3O12 (BT4) nanoparticles (NPs) on dielectric and electro-optical


properties of 5CB is investigated.
 The BT4 NPs were synthesised via a microwave-assisted chemical method and
characterized using various instruments.
 The agglomeration-free and uniform dispersion of BT4 NPs in the 5CB matrix was
confirmed by optical textures.
 The dielectric parameters and activation energy of the composite (5CB-BT4) are
estimated using dielectric spectroscopy.
 The composites would certainly be useful in the fabrication of NLC based tunable
devices such as optical memory etc.

Tunable dielectric and memory features of ferroelectric layered perovskite Bi4Ti3O12

nanoparticles doped nematic liquid crystal composite

Anu1, Depanshu Varshney2, Kamlesh Yadav1, 3, *, Jai Prakash2, *, Harikesh Meena4,5, and
Gautam Singh5
1Department of Physics, School of Basic Sciences, Central University of Punjab, Bathinda-
151401, Punjab, India
2Department of Physics, Aligarh Muslim University, Aligarh (U.P.), India
3Department of Physics, University of Allahabad, Prayagraj-211002, Uttar Pradesh,
India
4Department of Physics and Electronics, Rajdhani College, University of Delhi, Raja Garden,

New Delhi-110015, India


5Department of Applied Physics, Amity Institute of Applied Sciences, Amity University, Uttar
Pradesh, Noida-201313, India
*Email: jpsphysics@gmail.com (J. Prakash), kamlesh.yadav@cup.edu.in (K. Yadav)

Abstract

Herein, we report the synthesis of ferroelectric layered perovskite Bi4Ti3O12 (BT4) nanoparticles

(NPs) and the temperature-dependent dielectric and electro-optical (especially memory effect)

properties of 4-pentyl-4̍-cyanobiphenyl (5CB) nematic liquid crystal (NLC) doped with 1 wt %

BT4 NPs (i.e. 5CB-BT4 composite) using polarising optical microscopy and frequency-

dependent dielectric spectroscopy techniques. BT4 NPs were synthesised via a microwave-

assisted chemical method and characterised using various instrumental techniques, which

confirmed the formation of a non-stoichiometric and oxygen-deficient orthorhombic crystal

phase. The agglomeration-free and uniform dispersion of BT4 NPs in the 5CB matrix was

confirmed by optical textures. The optical memory studied by bias voltage-dependent (ON-OFF)

optical textures is decreased by 2.6 times in the 5CB-BT4 composite compared with 5CB.

Moreover, dielectric parameters such as dielectric permittivity, dielectric loss, loss tangent,

conductivity, and activation energy of 5CB and composite (5CB-BT4) are estimated using

dielectric spectroscopy. The dielectric anisotropy is decreased, whereas no shift in the clearing

temperature is observed in the 5CB-BT4 composite compared to the 5CB sample. Also, the DC

conductivity of 5CB-BT4 composite is found to be increased by approximately four times

compared to the 5CB. Our studies clearly demonstrate the tunability of the dielectric and optical

memory features of NLC (5CB) matrix by dopant BT4 NPs, without significantly affecting the

molecular alignment of the NLC molecules. Such composites would certainly be useful in the

fabrication of NLC based tunable devices such as optical memory and conductivity switches.
Keywords: Nematic liquid crystal, Ferroelectric, Bismuth titanate, Dielectric anisotropy,

Memory effect, Dielectric and Electro-Optical Properties

1. Introduction

Liquid crystals (LCs) are intermediate states of solids and liquids and are generally considered to

be the fourth state of matter. It shares some properties with solids, such as crystallinity and

anisotropy, while exhibiting fluid-like properties of isotropic liquids. Depending on their

molecular arrangement, LCs can be further classified as smectic, nematic, or cholesteric. LCs

have diverse applications in electro-optical devices such as display devices (laptops,

wristwatches, calculators, and smartphones), biosensors, thermal imaging, optical computing,

LC-based antennas, and waveguide [1-3]. Usually, the primary requirements for the use of LCs

in industrial applications are a low threshold voltage, quick response time, thermal stability, and

high optical performance, which are directly dependent on the electro-optical and dielectric

characteristics of LCs. However, all LCs contains free mobile ions generated during the

production process. Excess mobile ions are considered impurities that cause long-term sticking

effects, delayed response times, grey-level shifts, and short-term glimmer effects [4-6]. Nematic

liquid crystals (NLCs), long-range orientationally ordered fluids, are gaining recognition because

of their unique properties, such as deliberate light transmission as a function of the external

electric field, frustrated topological memory effect, and low cost, which facilitate their

application in numerous electronic devices [4, 7]. However, unlike other LCs, unwanted ionic

impurities present in NLCs degrade their performance in NLC display devices. Nevertheless,

these devices still require fast response time, low threshold voltage, and better optical contrast

and electrical conductivity, which are directly reliant on the dielectric and electro-optical

properties of LCs and require enhancement to a certain extent [4, 8].


To improve the properties of NLCs, the assimilation of different kinds of nanomaterials

(metal/metal oxides, carbon, graphene, semiconductor, ferroelectric, etc.) in NLCs has resulted

in enhanced electro-optical, optical, and dielectric properties [9-15]. For instance, it is found that

the doping of oxide nanoparticles TiO2, W18O49, CoO, Cu: ZnO, and MgO can alter the

concentration of impurity ions and the distribution of intermolecular interaction energy in NLCs

[16-20]. The doping of CdSe/ZnS quantum dots (size ~ 5.6 nm) in 5CB NLC increases the

intermolecular interaction between LC molecules and quantum dots which increases the mobile

ion density, dielectric permittivity, dielectric loss, conductivity, and isotropic to nematic phase

transition temperature [21].

Recently, ABO3-type ferroelectric oxide nanoparticles (FONPs) such as BiFeO3, BaTiO3,

LiNbO3, SrTiO3, and MnTiO3 with high polarisability and large permanent dipole moment doped

NLCs demonstrated strong electro-optical and dielectric responses that are desirable for

industrial applications [4, 22-26]. For instance, a significant decrease in the conductivity of NLC

(5CB) is observed upon doping with FONPs (BaTiO3, size  50 nm) and is attributed to the

strong trapping of mobile ions due to the high intrinsic electric field of the utilised NPs [4]. A

significant enhancement in the clearing temperature and dielectric anisotropy (Δε') of the NLCs

is observed in FONP-NLC composites [27-29]. Conversely, a decrease in dielectric anisotropy

and order parameter is also observed in the FONPs doped NLCs (i.e. 5CB and cybotactic

nematic phase) owing to the suppression of polarisation caused by the adsorption of ionic

impurities on the surface of FONPs [23, 30]. In 1wt% MnTiO3 NPs (size: <50 nm) doped NLC

(6CHBT), a decrease in threshold voltage and splay elastic constant and an increase in dielectric

anisotropy and clearing temperature has been reported [25]. The observed results for the FONP-

NLC composites could be explained by the plausible interaction between NLC molecules and
dopants based on the segregation effect and anchoring between the NLC molecules and the

embedded dopant [31, 32]. In addition to these simple ABO3-type ferroelectric perovskite

oxides, another class of layered ferroelectrics with the general formula (An-1BnO3n+1)2–(Bi2O2)2+

are known as Aurivillius compounds, which have yet not been explored much as dopants in

liquid crystal applications. Their crystal structure comprises perovskite blocks (An–1BnO3n+1)2–

and fluorite structure layers (Bi2O2)2+ alternatively stacked along the c-axis, where A, B, and n

represent (mono-, di-, trivalent or a mixture of these) cations, (tri-, tetra-, or pentavalent) cations

and the number of perovskite units, respectively. Bi4Ti3O12 (BT4) is a typical 3-layered

Aurivillius compound with a high dielectric constant and high remnant polarisation (~ 2 P =

40 μC/cm2) and a small direct band gap (~ 3 eV). Because of its layered crystallographic

structure, BT4 facilitates easy charge transportation and acts as a preferential adsorptive site to

capture more LC ionic impurities compared to other ABO3-type perovskites [33-39]. The

distinctive optical and electronic properties of the BT4 dopant can improve the electro-optical

properties of the host LC. Moreover, ferroelectricity is size dependent and disappears below a

certain critical size [40, 41]. The orthorhombic phase of BT4 is ferroelectric, whereas its

tetrahedral phase does not exhibit any ferroelectricity [42, 43]. The size-driven phase transition

from orthorhombic to tetrahedral (ferroelectric to paraelectric) occurs in BT4 NPs at a critical

crystallite size of < 48 nm [42, 43]. However, the orthorhombic BT4 phase synthesised by

various methods shows ferroelectricity below a particle size of < 30 nm [44-49], which is

beneficial for LC device applications. Recently, dielectric anisotropy and memory enhancement

have been reported in Bi2Ti2O7/Bi4Ti3O12 (BT2/BT4, ~ 42 nm) nanocomposite doped 5CB NLC

[50]. Interestingly, an increase in the dielectric constant and ac conductivity in ferroelectric

layered perovskite BT4 NPs (~16-18 nm) doped lyotropic liquid crystals has been reported and
attributed to the orientation of BT4 on the surface of the LC [51]. However, the effects of BT4

NPs doping on the dielectric and memory characteristics of 5CB NLC have not been reported.

In the present study, we report the synthesis of ferroelectric layered perovskite BT4 NPs and

study the effect of 1wt% BT4 NPs doping on the dielectric and electro-optical (memory) features

of 5CB NLC, using dielectric spectroscopy and electro-optical techniques. The observed

modulation in the dielectric and memory in the 1 wt% 5CB-BT4 composite compared to 5CB

NLC is greatly indicative of their plausible application in conductivity switches and tunable

memory devices.

2. Experimental details

2.1. Synthesis of Ferroelectric Layered Perovskite BT4 NPs

The BT4 NPs were synthesised by modifying previously reported simple hydrothermal

and microwave (MW) synthesis methods [52, 53]. First, a stoichiometric solution of

Bismuth (III) nitrate pentahydrate [Bi(NO3)3.5H2O, Aldrich, ≥ 98%] and Titanium (IV)

isopropoxide [Ti(OCH(CH3)2)4], Aldrich, >97%], was prepared in 40 ml double-distilled

water containing few drops of concentrated HNO3 (69%) to adjust pH of the solution to

1. This solution was stirred for an hour using a magnetic stirrer and sonicated for 15 min

to uniformly disperse the precursors. This uniform dispersion was transferred into an 80

ml glass vial for MW heating at 200 °C for 1 h using a microwave synthesiser (Discover,

CEM, USA). MW heating causes rapid and uniform heating of the reactants, which

reduces the reaction time. This method provides a narrow particle size distribution,

uniform morphology, high yield, and pure phase of nanoparticles [54]. White

precipitates appeared after MW treatment. The white precipitates were collected after
centrifugation and then dried in a vacuum oven at 80°C for 12 h. The dried powder was

sintered at 800 °C for 5 h to obtain pure phase of BT4 NPs.

2.2. Used nematic liquid crystal

The NLC material used in this study was 5CB (Merck, Kenilworth, New Jersey, USA).

Its chemical formula is 𝐶18𝐻19𝑁. It exhibits a phase transition from crystalline to

nematic phase at ~22.5 °C and transforms from nematic to isotropic phase at ~35 ˚C. It

exhibits a positive dielectric anisotropy of ~11.5 at 24°C when measured at an applied

field frequency of 1kHz [30, 50]. The significant electric dipole moment in 5CB along

the longer axis arises from the presence of the cyano group [55].

2.3. Preparation of 5CB-BT4 Composite

First, we prepared a dispersion solution of BT4 NPs in the organic solvent

isopropanol. A sufficient amount of BT4 NPs was dissolved in isopropanol and

ultrasonicated for an hour to obtain a uniform dispersion. Subsequently, the

calculated amount of this solution was mixed with 5 mg of 5CB NLC to achieve a 1

wt% concentration of BT4 NPs in 5CB. It was sonicated for 2 h at 50 °C (well above

the isotropic phase of the 5CB) to obtain a uniform and stable dispersion of BT4

NPs in 5CB. Here, we have considered 1 wt% concentration of BT4 NPs as an

additive in 5CB NLC because of the optimum ion absorption capability at this

concentration, as reported in our recent study on an almost similar system [50]. It

has been reported that BT2/BT4 doped 5CB LC shows the highest dielectric

permittivity and dielectric anisotropy at 0.1 wt%. However, the highest ion

adsorption and enhanced memory effect are observed at a 1wt% concentration.

Therefore, 1wt% BT4 is used as the optimum concentration in the present study so
that NPs can adsorb more ionic impurities. In addition, 1 wt% FNPs doped NLC has

been reported as the optimum concentration by many researchers [23, 56, 57].

2.4 Liquid crystal sample cells details

For the dielectric and memory measurements, we used commercially available

S100A050uG180 LC cells with a thickness of 5 µm and an ITO active area of 100

mm2 manufactured by INSTEC (INSTEC Inc., Colorado, Boulder, USA). These

cells consisted of a glass substrate coated with indium tin oxide (ITO) films with a

sheet resistance of 10 Ω/sq to acquire capacitor-type geometry. The size of the lower

glass sheet (at which contact of both electrodes is situated) was 25 × 22 mm2, and

the size of the upper glass sheet was 20 × 20 mm2. The polyimide layers were

coated on these ITO films and rubbed in parallel (or planar, pre-tilt angle 1 to 3o)

to achieve homogeneous alignment of the NLC molecules. The LC cells were filled

with the 5CB and 5CB-BT4 via capillary action at an isotropic temperature for

dielectric characterisation. The electro-optical, i.e. memory measurements were

performed on the homogeneously aligned sample cell placed under a cross-polarised

optical microscope at 26°C by applying a bias voltage of 40 V and then after

removal of this bias voltage at different time periods.

2.4. Characterisation of BT4 NPs

2.4.1. Structural and morphological measurement of BT4 NPs

The as-synthesised BT4 sample was characterised by powder X-ray diffraction

(Panalyticals XPert Pro) using Cu Kα (λ) = 0.1546 nm, with a scan rate of 1° per min

in the 2θ range (5°-90°). The shape and size of the BT4 NPs were observed by field

emission scanning electron microscopy (FESEM) (Merlin compact 6073, Carl


Zeiss) coupled with an energy dispersive X-ray analyser (EDX) to analyse the

elemental composition. The Fourier-transform infrared spectroscopy (FTIR)

transmission pattern was recorded using a Bruker Tensor 27 FTIR spectrometer in

the wavenumber range of 600-2500 cm-1 using KBr pellets as a reference. The UV-

visible absorption spectrum was recorded using a UV-Visible Spectrometer (UV-

2450, Shimadzu Japan) with water as a reference in the wavelength range of 200-

800 nm.

2.4.2. Dielectric and electro-optical measurements

Dielectric and electro-optical memory measurements of the 5CB and 5CB-BT4

composite samples filled in the homogeneously aligned LC sample cells were

carried out using an LCR meter (E4980A, Keysight, USA) coupled with a

temperature controller in the frequency range of 20 Hz to 2 MHz. Various dielectric

parameters, i.e. the real and imaginary parts of the dielectric permittivity (ε', ε'') and

loss factor (tanδ) were observed under the effect of 500 mV AC voltage along with

different values of DC biasing as per the experimental requirements.

To analyse the experimental dielectric data, the real (ε') and imaginary (εʺ) parts of

the complex relative permittivity (𝜀 ∗ ) have been fitted according to the generalised

Cole-Cole equation [58, 59], which is commonly used to describe dielectric

relaxation and is given as:

𝜀′(0) ― 𝜀′(∞) 𝐴 𝜎𝑖𝑜𝑛


𝜀 ∗ (𝜔) = 𝜀′ ―𝑗𝜀ʺ = 𝜀′(∞) + ∑1 + (𝑗𝜔𝜏)(1 ― 𝛼) + 𝜔𝑛 ―𝑗𝜀 𝜔𝑘 ―𝑗𝐵𝜔𝑚 (1)
0

Where 𝜀′(0) and 𝜀′(∞) are the low and high frequency limiting values of the

dielectric permittivity, respectively. (𝜀′(0) ― 𝜀′(∞)), ω, 𝜀0, τ, and α represent the

dielectric strength, angular frequency, permittivity of free space, relaxation time,


and distribution parameter (0 ≤ 𝛼 ≤ 1), respectively. When α=0, the Cole-Cole

model reduces to the Debye model, whereas α > 0 signifies a stretched relaxation.

The third term of equation (1) signifies the electrode polarisation capacitance at low

frequencies having A and n as the fitting parameters [30]. The fourth imaginary term

of equation (1) represents the contribution of ionic conductivity (𝜎𝑖𝑜𝑛) at low

frequencies with k as a fitting parameter having value usually 1 for dc conductivity.

The fifth imaginary term B𝜔𝑚 partially accounts for the effect of finite surface

resistance of the electrodes and inductance of the connecting wires with B and m as

the fitting constants in the measured dielectric data at >100kHz [60]. The ε' and εʺ

parts of 𝜀 ∗ (𝜔) can be expressed separately as:

(𝜀′(0) ― 𝜀′(∞))(1 + (𝜔𝜏)(1 ― 𝛼)𝑠𝑖𝑛 (𝛼𝜋 2)) 𝐴


𝜀′ = 𝜀′(∞) + ∑ + 𝜔𝑛 (2)
1 + 2(𝜔𝜏)(1 ― 𝛼)𝑠𝑖𝑛 ( 𝛼𝜋
)
2 + (𝜔𝜏)
2(1 ― 𝛼)

𝛼𝜋
(𝜀′(0) ― 𝜀′(∞))(𝜔𝜏)(1 ― 𝛼)𝑐𝑜𝑠 ( 2) 𝜎𝑖𝑜𝑛
ʺ
𝜀 =∑ + 𝜀 𝜔𝑘 + 𝐵𝜔𝑚 (3)
1 + 2(𝜔𝜏)(1 ― 𝛼)𝑠𝑖𝑛 ( 𝛼𝜋
)
2 + (𝜔𝜏)
2(1 ― 𝛼) 0

The measured ε' and εʺ parts of relative dielectric permittivity are fitted with

equations (2) and (3), respectively, using a program developed with Origin software.

The best fit of measured data is estimated by the value of chi-square (𝜒2) and

correlation coefficient (𝑅2), which in turn should tend to be 0 and 1, respectively, for

the best-fitted data. Further, the frequency dependece of real part of conductivity is

estimated using the equation 𝜎(𝜔) = 𝜔𝜀0𝜀ˮ; where we have used the parasitic effect

free 𝜀ˮ obtained by subtracting the low and high-frequency correction terms


𝜎𝑖𝑜𝑛
calculated using the fitted parameters (i.e., 𝜀0𝜔𝑘
𝑎𝑛𝑑 𝐵𝜔𝑚), respectively from the
measured data. The dielectric loss tangent (tan𝛿) is calculated using the relation,
εʺ
tanδ = and fitted according to the following equation [61]:
ε

𝛼′
tanδ = (
((𝜀′(0) 𝜀′(∞)) ― 1)ωτ

(𝜀′(0) 𝜀′(∞)) + (ωτ)2 ) (4)

Here 𝛼′ is a shape parameter with a value of 0 ≤ 𝛼′ ≤ 1. If 𝛼'=1, the eq. (4) reduces

to the ideal Debye equation for a single particle and non-interacting system, while 𝛼′

<1 represents a system having multi-type dipole polarisation.

3. Results and Discussion

3.1. Structural and optical properties of BT4 NPs

Figure 1 (a) shows a typical XRD pattern of the synthesised BT4 NPs. The diffraction

peaks were indexed to the orthorhombic crystal structure of BT4 (JCPDS file No. 00-

036-1486) [62]. The crystallite size (Γ) was estimated using the Scherrer equation

considering the most intense diffraction peak (117) as:


0.89λ
Γ = βcosθ (5)

Where λ, β and θ represent the wavelength of the incident Cu Kα X-ray, full width at half

maximum (FWHM) of the diffraction peak and angle of diffraction, respectively. The

crystallite size was estimated as ~ 71 nm. Rietveld refinement using FullProf Suite

software was performed on the experimental XRD data of the BT4 sample, as shown in

Figure 1 (b). The refinement pattern is in good agreement with the experimental

diffraction pattern. The refined XRD pattern further suggests the formation of a pure

orthorhombic crystal structure of BT4 NPs with the space group Fmmm. The Refined
lattice parameters, unit cell volumes, and conventional Rietveld factors are listed in

Table 1.

Figure 1: (a) XRD pattern of Bi4Ti3O12 NPs and (b) Rietveld refinement of the XRD pattern of

Bi4Ti3O12 NPs.

Table 1: Refined structural parameters of Bi4Ti3O12 NPs.

Refined lattice parameters Unit cell Refinement parameters

(Å) volume

a b C (Å3) Rp Rwp Rexp χ2

5.44683 5.40907 32.82320 967.05 21.5 17.7 15.51 1.3

Figure 2 (a) and its inset show FESEM images of the BT4 NPs at low and high

magnifications, respectively. It consists of irregular-shaped nanoparticles of an average

particle size of ~ 85 nm. Figure 2(b) shows the EDX pattern of the BT4 NPs, which

confirms the presence of Bi, Ti, and O and the formation of a pure BT4 sample. The

inset of Figure 2 (b) shows the experimental weights and atomic % of BT4 NPs. The
theoretical and experimental weights and atomic % indicate the formation of non-

stoichiometric BT4 NPs (see Table 2). The highly volatile nature of Bi causes the

formation of Bi-ion vacancies in the crystal lattice, which produce oxygen vacancies in

the crystal lattice to maintain the overall charge neutrality of BT4. Hence, this primarily

accounts for the formation of a non-stoichiometric BT4 sample [63].

Figure 2: (a) FESEM image of Bi4Ti3O12 NPs, (b) EDX spectrum of Bi4Ti3O12 NPs, and

the corresponding inset shows the experimental weight and atomic %.

Table 2: Theoretical and experimental weight and atomic % of Bi4Ti3O12 NPs.

Elements Theoretical Experimental Theoretical Experimental


weight % weight % atomic % atomic %
Bi 71.35 64.71 21.05 14.98
Ti 12.26 10.78 15.79 10.89
O 16.38 24.51 63.16 74.12

Figure 3 shows the FTIR transmission spectrum of the BT4 NPs in the wavenumber

range of 600-2500 cm-1. The fundamental bands observed at 662 cm-1 and 817 cm-1

correspond to the stretching vibrations of the Ti-O bonds, which confirms the

formation of a titanate structure [62-64]. The peaks appearing at 950 cm-1 and 1077
cm-1 can be related to the different types of C-O bonds, whereas the peaks at ~1292

cm-1 and 1656 cm-1 can be attributed to the presence of -NO2 stretching vibrations.

In addition, the peaks at 1388 cm-1 and 1580 cm-1 can be assigned to the symmetric

and asymmetric stretching vibrations of [COO]-, respectively [65, 66].

1077cm-1

1388cm-1

1656cm-1
Transmittance (%)
660cm-1

1292cm-1

1580cm-1
952cm-1

Bi4Ti3O12
817cm-1

600 800 1000 1200 1400 1600 1800 2000


Wavenumber (cm-1)

Figure 3: FTIR spectra of Bi4Ti3O12 NPs.

Figure 4 shows the UV-visible absorption spectra of BT4 NPs. Previous reports have

confirmed a direct band gap for BT4 NPs [67, 68]. The direct band gap (Eg) was

estimated using the Tauc formula [36], ( ∈ hν)2 ∝ (hν ― Eg); ( ∈ = (2.303 × À)/𝑑),

where, ∈ , h, ν, À, and d, represent the absorption coefficient, Plank constant, frequency

of the incident photon, absorbance, and thickness of the quartz cuvette, respectively. The

inset of Figure 4 shows a Tauc plot. Eg was estimated by extrapolating the linear portion

of the Tauc plot to zero absorbance. It is found to be 2.7 eV, which is comparatively

smaller than the reported band gap in stoichiometric bulk BT4 (3.1 eV) [67, 69, 70]. This

is attributed to the presence of oxygen vacancies in the crystal structure of the BT4 NPs,
as previously reported [68]. The refractive index (μ) of BT4 NPs is calculated using the
1

Moss relation, 𝜇 = ( ) is 2.51, which is comparable to that reported in the literature


108
𝐸𝑔
4

[63].

0.8

(h) (eVcm-1) 2
100
Bi4Ti3O12
0.7 80
Absorbance (a.u.)

60

0.6 40

20
0.5 0
1 2 3 4 5 6 7
h eV
0.4

0.3

200 300 400 500 600 700 800


Wavelength (nm)

Figure 4: UV-visible absorption spectra of Bi4Ti3O12 NPs and the corresponding inset show the

Tauc plot.

3.2. Liquid crystal data analysis

3.2.1 Polarising optical micrographs analysis

Figure 5 (a-d) shows polarising optical micrographs of 5CB and 5CB-BT4 (1 wt.%) in a

bright (0V bias) and dark (40V bias) state at 26°C. In the bright optical micrographs

(Figure 5 (a,c)), uniformly distributed black dots can be seen in both samples owing to

the presence of spacer balls in the purchased LC sample cells. The number of spots was

greater in both the bright and dark images of the 5CB-BT4 composite compared to 5CB,

which is due to the scattering of light around the BT4 NPs, and hence confirms the
presence of BT4 NPs in the 5CB matrix. In addition, the uniform colour of the optical

micrographs of the 5CB-BT4 composite also suggests an undisturbed, homogeneous,

and aggregation-free molecular alignment of nematogens in the 5CB-BT4 composite

(see Fig 5(c-d)).

Figure 5: Polarising optical micrographs of 5 μm thick homogeneously aligned LC


sample cells filled with (a-b) 5CB and (c-d) 1 wt% 5CB-BT4 composite, in the bright
and dark states, respectively, at 26°C.

3.2.2 Dielectric Data Analysis

Figure 6 (a-b) shows a plot of the real part of dielectric permittivity (ε') versus frequency. This

plot is fitted according to equation (2) at different temperatures, covering the nematic to isotropic

phases. The values of the fitted parameters are listed in Table 1S (supplementary data). The
experimental data did not fit well at high frequencies (> 106 Hz). This happens due to the

contribution of general parasitic effects, such as the high surface resistance of the electrodes and

the resistance and inductance of the connecting wires [59]. Moreover, 𝜀′(∞) is the lowest value

of the permittivity and is equal to the square of the refractive index of the materials and likely to

arise at optical frequencies [59]. Its value for pure 5CB is reported to be ~1.5-1.8 [71]. However,

in the present case, the value of fitting parameter 𝜀′(∞) is found to be <1 for 5CB and 5CB-BT4

composite, which further indicates the presence of parasitic effect at high frequencies. Therefore,

INSTEC LC cells are not suitable for dielectric measurements at high frequencies (> 106 Hz) [25,

59]. ε' decreases rapidly and then attains a small constant value with an increase in frequency in

both the 5CB and 5CB-BT4 composite samples, which is a well-known dielectric behaviour. The

large dielectric permittivity at low frequencies generally arises owing to interfacial space charge

polarisation. This is attributed to the fast drift of the LC ions toward the electrodes compared to

the rate of change in polarity of the applied electric field. At high frequencies, the dipoles cannot

follow the high frequency of the applied field and lag the field and hence the orientational

polarization ceases, therefore ε′ decreases, and attains a constant value at high frequencies [72].

In addition, the ε' value increased with temperature in both samples because of the easy transport

of ions at high temperatures. The value of ε' of 5CB-BT4 composite is found to be relatively

smaller than 5CB in the entire frequency and temperature range except at the frequencies ≤ 100

Hz for T ≤ 28°C. The change in the dielectric permittivity of a system depends mainly on the

number of mobile ions reaching the electrode surface and the alignment of the permanent dipoles

of the system along the direction of applied field. The alignment of most doped NPs dipole

moments along the nematic director causes an enhancement in polarisation, which leads to an

increase in the dielectric permittivity of the system [73-75]. However, systems with anti-parallel
alignment of most ferroelectric NPs dipole moments to the LC molecules direction show a

suppression of polarisation, which causes a decrease in dielectric permittivity [23, 30, 76]. In

addition, the adsorption of ionic impurities present in the LC on the surface of doped NPs causes

a decrease in the number of mobile ions resulting in a decrease in ε' [50]. In the present case,

owing to the formation of oxygen vacancies in BT4 NPs, the surface energy and intrinsic

chemical reactivity increased, which increased the ionic impurity adsorption ability of BT4 NPs.

Moreover, the strong intrinsic electric field of ferroelectric BT4 NPs and their layered crystal

structure further facilitate the preferential adsorption of impurities [4, 77-79]. Therefore, there is

the possibility of adsorption of most of the impurity ions present in 5CB on the surface of BT4

NPs, which causes a small decrease in ε' of the 5CB-BT4 composite compared to 5CB as also

observed previously in NP doped NLC [50, 77]. Furthermore, the dipoles of BT4 NPs align anti-

parallel to the LC molecule dipole in the present studied samples, further causing the suppression

of polarisation. Therefore, the significant decrease in ε' of the 5CB-BT4 composite compared to

that of 5CB is attributed to the combined effect of the reduction in ionic impurities and the anti-

parallel alignment of the doped BT4 NPs dipole moments to the nematic director. Additionally,

the larger value of ε' of 5CB-BT4 composite compared to 5CB at very low frequencies ( ≤ 100

Hz for T ≤ 28°C) indicates more contribution of dominant space charge effect at very low

frequencies in composite sample than 5 CB. Moreover, Figure 6 (c) shows the temperature

variation of the fitted parameters 𝜀′(0), calculated for the 5CB and 5CB-BT4 samples obtained

after fitting ε' versus ω data according to equation 2. The smaller value of 𝜀′(0), in the case of

5CB-BT4 composite compared to 5CB, further confirms the suppression of polarisation in the

5CB-BT4 composite sample.


Figure 6: (a-b) Temperature-dependent fitted ε' versus frequency plot of 5CB LC and 1 wt%
5CB-BT4composite under 0V DC biasing, (c) Plot of fitted parameters 𝜀′(0) with temperature of
5CB LC and 1wt% 5CB-BT4 composite.
Figure 7 (a-b) shows the fitted frequency-dependent dielectric loss (εʺ) of 5CB and 5CB-BT4

composite at various temperatures. The fitting parameters are listed in Table 2S (Supplementary

data). It can be comprehended clearly that the magnitude of εʺ in 5CB-BT4 composite is larger

than 5CB for T ≤ 28°𝐶, however, it is smaller than 5CB for T ≥ 30°C. This larger εʺ in 5CB-BT4

composite than 5CB for T ≤ 28°𝐶 arises because of the improved ion density and ion mobility in

the composite. However, the smaller εʺ of the composite than 5CB for T ≥ 30°C may be caused

due to increased resistivity. Also, the trapping of ions on the surface of BT4 NPs was lower at
higher temperatures. In addition, a broad high-frequency relaxation peak with increasing

magnitude and a slight shift towards lower frequency with the increase in temperature is

observed in both samples (see Inset of Fig 7 (a-b)), as also reported [21, 50, 80]. This high

molecular relaxation peak may belong to the ITO relaxation process because we used a

commercial INSTEC LC cell that uses ITO as an electrode [21, 80]. Figure 7 (c-d) shows the

variation of the fitted tanδ with frequency for the 5CB and 5CB-BT4 samples at different

temperatures. The fitted parameters are listed in Table 3S (Supplementary data). A low-

frequency relaxation peak appears in both samples, which is attributed to ion diffusion and

interfacial space charge polarisation [21, 50, 75]. The magnitude of such a low-frequency

relaxation peak is smaller in the 5CB-BT4 composite than in 5CB, suggesting lower electrical

energy dissipation in the composite than in 5CB. This is due to the reduced ionic impurities and

availability of free space, which leads to easy charge transportation, and hence small electrical

energy dissipation in 5CB-BT4 compared to 5CB, as also recently reported for

Bi2Ti2O7/Bi4Ti3O12 doped 5CB composite [50]. Moreover, this low relaxation peak in both

samples shifted toward the higher frequency side with an increase in temperature. This suggests

a reduction in resistivity (enhancement in conductivity) with increasing temperature owing to

easy ion transportation and an increase in kinetic energy. Furthermore, the relaxation peak of

5CB-BT4 composite appeared at a slightly higher frequency compared to the relaxation peak of

5CB at T < 34°C, while at T ≥ 34°C, the relaxation peak of 5CB-BT4 composite appeared at a

slightly lower frequency than the relaxation peak of 5CB. Figure 7 (e) shows the variation in the

relaxation time (τ) with temperature corresponding to the relaxation peak for both samples. τ is

estimated by fitting tanδ using equation 4. The smaller value of τ for the 5CB-BT4 composite as

compared to 5CB in the nematic phase suggests the high conductivity of the composite as
compared to 5CB because of the increased density and mobility of ions in the 5CB-BT4

composite and the strong interaction between the BT4 NPs and 5CB LC molecules, which is

consistent with the previous discussion. In the isotropic phase, the value of τ of the 5CB-BT4

composite is relatively larger than that of 5CB because of the decrease in the conductivity of the

5CB-BT4 composite with increasing temperature, as discussed previously. Figure 7 (f) shows the

variation in the shape parameter (α') with temperature. It was estimated by fitting tanδ according

to Equation (4). It is observed that the value of the fitted parameter α' for both 5CB and 5CB-

BT4 is slightly less than 1, which signifies multi-type dipole polarisation in both samples [61,

81]. In addition, the value of α' decreases with increasing temperature for both samples, which

signifies enhanced multi-type polarisation with an increase in temperature. Further, the 5CB-BT4

sample shows a smaller value of 𝛼' compared to 5CB in the overall temperature range, and hence

signifies the plausible interaction of 5CB and BT4 NPs dipoles.


Figure 7: (a-b) Fitted εʺ versus frequency with temperature (c-d) fitted tanδ versus frequency, (e-
f) Variation of fitted parameters τ and α', of 5CB LC and 1 wt% 5CB-BT4 composite with
temperature, respectively, under 0V DC biasing.
To understand the motion of the ions more clearly, frequency dependence of real part of

conductivity was estimated using the relationship 𝜎(𝜔) = 𝜔𝜀0𝜀ˮ for the 5CB and 5CB-BT4

composite samples. Figure 8 (a-b) shows the fitting of 𝜎(𝜔) versus ω data using the Jonscher's

universal power law (JPL) equation [82, 83],

σ(ω) = σd.c. + Bωp (6)

where σ(ω) is the total conductivity, σd.c. is d.c. conductivity independent of the frequency, B is

the temperature-dependent constant, and p is the fraction component that determines the order of

interaction between mobile ions and lattice. The fitted experimental data confirmed that the

conduction process followed the JPL. The fitted parameter p > 1 for both samples confirms the

presence of impurities (See Table. 4S as supplementary data) [84]. In addition, the value of the

fitted parameter ‘p’ is smaller in the 5CB-BT4 composite compared to 5CB (see Table 4S). This

further suggests a reduction in the impurities in the 5CB-BT4 composite owing to the adsorption

of ionic impurities by the BT4 NPs, as discussed previously. Figure 8 (c) shows the variation in

σd.c. with temperature for the 5CB and 5CB-BT4 composite samples. The value of σd.c. for 5CB

was found to be 10-9-10-10 mho cm-1 which is comparable to the reported literature [21].

Enhancement of σd.c. was approximately four times higher in the 5CB-BT4 composite than in

5CB. This is contrary to our observation of the adsorption of ions on the surface of BT4 NPs,

which decreases conductivity [21, 50, 56]. The conductivity of a sample mainly depends upon

the density and mobility of ions present in the system. The enhanced mobility of ions is expected

since the ions get a better pathway due to the plausible interaction between dipole moments of

BT4 NPs and LC molecules, as also observed previously [21, 30, 60, 85]. However, in order to
explain the many-fold increased conductivity of 5CB-BT4 compared to 5CB, we should have to

consider the ion capturing and ion releasing process (ions adsorption and desorption) happening

over the surface of NPs [86-88]. Previously, Yadav et al. [17] examined the TiO2 NPs doped

NLC and found increased conductivity due to the presence of impurities in TiO2 NPs before

dispersing them into the NLC. In the present case of BT4 NPs, there is also a quite possibility

that the BT4 NPs are contaminated (i.e., carry some ions on their surface) prior dispersing into

the LC matrix and release the impurity ions from the surface of BT4 NPs into the LC matrix,

which are adsorbed on the surface of BT4 during the synthesis and contributes to the

enhancement of conductivity, as also reported previously [86-88]. The synthesis of BT4 NPs

using the MW-assisted method results in the formation of a larger number of surface defects

compared to NPs prepared using other conventional heating methods [89]. In addition, sintering

of BT4 NPs at high temperatures causes the loss of Bi ions because of their volatile nature,

which leads to the formation of oxygen vacancies and subsequent adsorption of unreacted metal

ions on the surface of the NPs [63]. Therefore, the approximately four-fold increased

conductivity in composite compared to 5CB is attributed to the combined effect of BT4 NPs

ionic additive behavior and the availability of free space in 5CB-BT4 composite eventually

facilitated the easy transport of ions in the samples, as also observed previously [17, 85, 90]. The

Arrhenius equation is used to determine the activation energy (𝐸𝑎) of free charge carriers from

the variation of 𝜎𝑑.𝑐 with temperature as [91]:

Ea
𝜎𝑑.𝑐 = σ0exp ( ― kBT) (7)

Ea
ln (𝜎𝑑.𝑐.) = ― kBT +ln (σ0) (8)
Where 𝜎0, T, and kB represent the constant pre-exponential factor, absolute temperature and

Boltzmann constant, respectively. Figure 8 (d) shows the Arrhenius plot of linear fitted ln (𝜎𝑑.𝑐.)

versus 1000/T in the nematic phase. The activation energy is calculated using the slope of this

Arrhenius plot and is found to be 44.43 kJ/mol and 21.97 kJ/mol for 5CB and 5CB-BT4

composite samples, respectively. The calculated activation energy of 5CB is comparable to that

reported in the literature [21]. The smaller value of the activation energy of the 5CB-BT4

composite compared to that of 5CB further suggests increased conductivity in the composite

compared to the 5 CB sample.


Figure 8: (a-b) Frequency-dependent plot of conductivity at different temperatures, (c)

temperature variation of 𝜎𝑑.𝑐., and (d) Arrhenius plot showing the variation of ln (𝜎𝑑.𝑐.) versus

inverse of temperature, for 5CB LC and 1 wt% 5CB-BT4 composite.

Dielectric anisotropy (∆𝜀′) is a parameter that determines how effectively LC molecules

are reoriented in the presence of an electric field. It is given by 𝜀′ǁ ― 𝜀′ ⊥ , where 𝜀′ǁ and

𝜀′ ⊥ are the dielectric components parallel and perpendicular to the NLC director [92].

Figure 9 (a) shows the temperature-dependent perpendicular and parallel components of

the dielectric permittivity measured by changing the alignment of the sample cell from

homogeneous (0 V) to homeotropic (40 V). The estimated ∆𝜀′ values for the 5CB and

5CB-BT4 samples are shown in Figure 9(b). The magnitude of 𝛥𝜀′ of 5CB is comparable

to that reported in the literature [21, 50, 93]. It is observed that the 𝛥𝜀′ decreases with

increasing temperature for both samples. This is attributed to the reduction in polarisation

owing to the enhancement in disorder at high temperatures [92]. In addition, 𝛥𝜀′ values of

the 5CB-BT4 composite are smaller than those of the 5CB NLC at various temperatures.

This decrease in 𝛥𝜀′ with the incorporation of ferroelectric NPs into NLC has also been

reported [23, 30, 94]. This decrease in 𝛥𝜀′ can be better understood by considering the

theory proposed by Maier-Meier [95],

∆ε′ =
NHF
ε0 (∆α ― F 2
2𝑘𝐵Tμ (1 )
― 3cos2β) S (9)

Here N, ε0, Δα, μ, β, kB, T, F, S and H represent the density of the LC molecules,

permittivity of free space, anisotropy of the polarisability, the resultant dipole moment of

the molecules, angle between the dipole moment and the long axis of the molecule,

Boltzmann constant, temperature, feedback factor, nematic order parameter, and H= (3ε0
/(2ε0+1)), respectively. According to eq. 9, the 𝛥𝜀′ significantly depends on N and S. The

decrease in N is expected due to a decrease in the number density of 5CB molecules in

5CB-BT4, which may cause a slight decrease in 𝜀′ ⊥ , 𝜀′ǁ and hence, Δε'. However, it does

not seem as the mere reason for the observed large decrease in the value of 𝜀′ǁ and Δε'. In

addition, no significant change in 𝜀′ ⊥ and no change in the phase transition temperature

of 5CB-BT4 compared to 5CB indicate no appreciable change in the order parameter (S).

Furthermore, the strong electric dipole moment of single-phase ferroelectric BT4 NPs

causes anti-parallel alignment of its dipoles to dipole moments of NLC molecules, which

leads to a reduction in the value of 𝜀′ǁ, and hence decreases Δε' significantly in 5CB-BT4

composite as compared to 5CB NLC. Similar results have been reported previously [23,

30]. Contrary to this result, in our previous work [50], doping with similar kinds of NPs

Bi2Ti2O7/Bi4Ti3O12 (BT2/BT4) in 5CB NLC causes an increase in Δε'. BT2/BT4 contains

only 10% BT4 ferroelectric and 90% non-ferroelectric BT2 phases. Therefore, an

inadequate fraction of the BT4 phase does not cause a noticeable amount of anti-parallel

alignment of its dipole to the dipoles of the 5CB molecule. Therefore, its reducing effect

is not reflected in Δε'. The small increase in Δε' is due to the net polarisation effect of the

BT2/BT4 nanocomposite.

Another important LC parameter is the clearing temperature (TN-I) at which a LC shows a

transition from the nematic to the isotropic phase. The dielectric anisotropy of both

samples becomes approximately zero at 36°C (see Figure 9 (b)), which shows TN-I at 36°C

in both samples. This result indicates that there is no change in TN-I after incorporating

BT4 NPs into the 5CB matrix. Mixed reviews have been reported in the literature on the

effect of ferroelectric NPs on the TN-I of NLCs. A few reports show an increase [74, 96,
97], while other reports show a decrease [23, 30, 94, 98] in TN-I. In addition, in some

reports, no change is reported in TN-I with doping [50]. The shift in TN-I results from

competition between three major effects, (i) permanent polarisation of ferroelectric NPs,

(ii) ionic impurity adsorption effect, and (iii) orientation interaction between LC

molecules and incorporated ferroelectric NPs [32, 98]. The enhancement in TN-I is

attributed to the permanent polarisation of ferroelectric NPs [32]. However, the

adsorption of ionic impurities on the surface of ferroelectric NPs screens the polarisation

effect. If the concentration of ionic impurities is >1023 m-3 in the NLC, then it suffices to

screen the polarisation effect of the ferroelectric NPs completely. However, in 5CB NLC,

the concentration of ionic impurities has been reported to be <1021 [21]. Therefore, this

effect cannot screen the polarisation effect of ferroelectric NPs completely. Therefore,

these two factors (i and ii) favour the increase in TN-I. Nevertheless, because of the

orientational effect, a decrease in TN-I is observed due to the anti-parallel alignment of

ferroelectric NPs to NLC molecules, as reported [30]. Therefore, the increase in TN-I

because of the net polarisation effect is compensated by the orientational effect in the

present studied samples resulting in no shift in TN-I of 5CB-BT4 composite with respect

to 5CB. Moreover, owing to the combined effects of these three processes, incorporating

BT4 NPs into 5CB NLC has no overall effect on the nematic director, indicating no

change in the liquid-crystal order parameter (S). However, the inset of Figure 9 (b) shows

the dielectric anisotropy versus temperature plot of the 5CB and 5CB-BT4 samples in the

isotropic phase. It is observed that the dielectric anisotropy of the 5CB-BT4 composite is

approximately two times higher than that of 5CB at 36°C. This indicates that although

there is no change in TN-I after the incorporating BT4 NPs in the 5CB matrix, the intrinsic
electric field of BT4 NPs stabilises the formation of local pseudo-nematic domains (i.e.,

orientational ordering) in the isotropic phase, which results in a nonzero higher dielectric

anisotropy of the 5CB-BT4 composite than that of 5CB at 36°C [32, 99]. Similar kind of

nonzero dielectric anisotropy in the isotropic phase has been observed previously in

BaTiO3 FNPs doped 5CB LC [99]. Moreover, as the temperature increases the thermal

energy eliminates these orientational ordering and cause approximately zero dielectric

anisotropy [32], as seen at 38-40°C in the inset of Figure 9 (b).

Figure 9: Temperature-dependent (a) parallel and perpendicular component of dielectric


permittivity (b) dielectric anisotropy of 5CB and 1 wt% 5CB-BT4 composite at 10 kHz. The inset
of (b) shows the dielectric anisotropy versus temperature of 5CB and 5CB-BT4 composite in the
isotropic phase.
4. Memory effects

4.1 Dielectric Memory

The signature of the memory effect can be observed by applying 0V and 40 V and after

removing the bias, as observed previously for doped NLCs [50]. Figure 10 (a-b) shows

the variation of ε' in 5CB and 5CB-BT4 samples with frequency by applying 0 V and 40

V biased voltage and then removing the bias of 40 V at different time intervals such as 0
s, 25 s, 45 s, 65 s and 95 s at 28 °C. The dielectric responses of the 5CB and 5CB-BT4

composite samples were recorded immediately after the removal of the external bias

superimposed with the dielectric response studied before applying bias (0 V) in both

samples. Hence, it indicates the absence of a dielectric memory in the 5CB and 1 wt%

5CB-BT4 composite samples. It has been reported that 5CB LC shows negligible

dielectric hysteresis because of the weak H-bonds formation among their molecules [100,

101]. Incorporating BT4 NPs into 5 CB leads to anti-parallel alignment of BT4 and 5CB

dipoles; however, no hydrogen bonds are formed between BT4 NPs and LC molecules.

Therefore, network formation between BT4 and LC molecules is absent [100]. Moreover,

the anti-parallel alignment of BT4 NPs dipoles to 5CB dipoles further accelerated the

switching of nematic dipoles to their initial state after the removal of the external field.

Thus, the 5CB-BT4 sample does not exhibit a dielectric memory.

Figure 10: Frequency-dependent dielectric permittivity of (a) 5CB and (b) 1 wt% 5CB-BT4
composite at 0V, 40V, and after removal of the bias. The temperature of the sample was kept at
28 oC.
4.2 Optical memory
The phenomenon of the reorientation of the nematic director with the application of an

electric field from one optical state (i.e., bright) to the other states (i.e., dark) can be used

to investigate the optical memory in the 5CB and 5CB-BT4 composite samples. Figure

(11 and 12) shows the polarising optical micrographs of the 5CB and 5CB-BT4 samples.

With a LC containing a single domain free from disordering, the monodomain switches at

a single time, and its micrograph shows a uniform change in the sample while returning

to its original scattering state. However, in the present case, the sample micrographs show

a particular pattern of returning to the original scattering state. This indicates

multiple/some disordered nematic domains within the system, which switch at different

paces and times and hence follow a particular pattern of returning to their original

scattering state [102]. It is inferred from the optical textures recorded after removing the

DC bias at different time intervals that 5CB takes  80 s to return to its original bright

state after removing the 40V DC bias, which suggests an optical memory of  80 s in

5CB, NLC. However, from the polarising optical micrographs of the 1 wt% 5CB-BT4

composite, it can be concluded that it takes only  30 s to return to its original bright state

after the removal of the bias voltage, which indicates an optical memory of 30s in the

composite. This is ~ 2.6 times lesser than that of 5CB. This is due to the strong restoring

force in the 1 wt% 5CB-BT4 composite, which is attributed to the anti-parallel coupling

of the strong ferroelectric polarisation of BT4 NPs to NLC molecules over the dipole-

dipole coupling in the NLC system, which eventually facilitates the faster flip-flop

motion of 5CB molecules in the 5CB-BT4 composite about their short axes [22].
Figure 11: Polarising optical micrographs of 5CB NLC at 0 V and 40 V DC biasing
voltage, and for different times at 0 V, (a) 0 V, (b) 40 V, (c) again 0 V, (d) 0 V after 10 s,
(e) 0 V after 20 s and (f) 0 V after 30 s (g) 0V after 40s, (h) 0V after 50s, (i) 0V after 60s
(j) 0V after 70s (k) 0V after 80 s, at 26°C.

Figure 12: Polarising optical micrographs of 1 wt% 5CB-BT4 composite at 0 V and 40 V


DC biasing voltage, and for different times at 0 V, (a) 0 V, (b) 40 V, (c) again 0 V, (d) 0 V
after 10 s, (e) 0 V after 20 s and (f) 0 V after 30 s (g) 0 V after 40 s , at 26°C.

5. Conclusion

In summary, we have synthesised the single-phase non-stoichiometric Bi4Ti3O12 (i.e. BT4)

layered perovskite oxide NPs with an optical direct band gap of 2.7eV using a microwave-

assisted chemical route method. The BT4 NPs are dispersed uniformly in the 5CB NLC
matrix, and the effect of doping on the dielectric and electro-optical (i.e. memory) properties

of NLC has been reported. The frequency-dependent dielectric permittivity reveals the

adsorption of LC impurities on the surface of BT4 NPs and the anti-parallel alignment of

BT4 NPs dipoles to LC dipoles responsible for the suppression of polarisation. Low-

frequency dielectric relaxation is observed in both samples owing to the combined effects of

ion diffusion and interfacial space charge polarisation. The magnitude of this relaxation peak

is smaller in the 1 wt% 5CB-BT4 composite than in 5CB suggesting low electrical energy

dissipation in the composite owing to the reduced ionic impurities and enhanced NLC

ordering, which is favourable for easy charge transportation in the NLC matrix. The DC

conductivity was estimated from the Jonscher power law equation and was found to be

approximately four times higher in the 1 wt% 5CB-BT4 composite than in 5CB, which is due

to the contribution of the ionic additive BT4 NPs and their interaction with LC molecules

which enhances the nematic ordering and facilitates the movement of ions. The activation

energy corresponding to the low-frequency tangent relaxation associated with ions diffusion

and space charge polarisation is found to be 44.43 kJ/mol and 21.97 kJ/mol for the 5CB and

1 wt% 5CB-BT4 composite samples, respectively. The dielectric anisotropy of the composite

decreases, which is attributed to anti-parallel dipole packing. The intrinsic electric field of the

ferroelectric BT4 NPs stabilises the local pseudo-nematic domains in the isotropic phase. No

shift in the clearing temperature in the 1 wt% 5CB-BT4 compared to the 5CB sample is

explained in detail by considering the polarisation, ionic and orientational alignment effects.

The optical memory is reduced in the 1 wt% 5CB-BT4 composite compared to that in the

5CB sample because of the faster flip-flop motion of 5CB molecules in the composite about
their short axes. We anticipate that such composite systems will certainly be useful for device

applications such as conductivity switching and optical memory.

Acknowledgements

The authors are grateful to the Central Instrumentation Laboratory and Department of Physics,
Central University of Punjab, India, for providing research facilities. One of the authors (Anu) is
grateful to the University Grant Commission (UGC), India, for financial support in the form of a
Senior Research Fellowship. Depanshu Varshney would also like to thank the Department of
Science and Technology (DST), India, for providing the INSPIRE fellowship for financial
support. Dr. Jai Prakash, one of the authors, is grateful to the DST for supporting this work under
the Science and Engineering Research Board (SERB) funded EMR Project (EMR/2016/006142).
The author (GS) thanks Amity University Uttar Pradesh (AUUP), Noida, India, for continuous
encouragement in his research work.

References:

1. Stephen, M.J. and J.P. Straley, Physics of liquid crystals. Reviews of Modern Physics, 1974. 46(4):
p. 617-704.
2. Binnemans, K., Ionic Liquid Crystals. Chemical Reviews, 2005. 105(11): p. 4148-4204.
3. Pauluth, D. and K. Tarumi, Advanced liquid crystals for television. Journal of Materials Chemistry,
2004. 14(8): p. 1219-1227.
4. Basu, R. and A. Garvey, Effects of ferroelectric nanoparticles on ion transport in a liquid crystal.
Applied Physics Letters, 2014. 105(15): p. 151905.
5. Prakash, J., et al., Metal oxide-nanoparticles and liquid crystal composites: A review of recent
progress. Journal of Molecular Liquids, 2020. 297: p. 112052.
6. Shen, Y. and I. Dierking, Perspectives in Liquid-Crystal-Aided Nanotechnology and Nanoscience.
Applied Sciences, 2019. 9(12).
7. Khoo, I.-C., Liquid crystals. Vol. 64. 2007: John Wiley & Sons.
8. Stark, H., Physics of colloidal dispersions in nematic liquid crystals. Physics Reports, 2001. 351(6):
p. 387-474.
9. Kumar, A., D. Pratap Singh, and G. Singh, Recent progress and future perspectives on carbon-
nanomaterial-dispersed liquid crystal composites. Journal of Physics D: Applied Physics, 2021.
55(8): p. 083002.
10. Supreet and G. Singh, Recent advances on cadmium free quantum dots-liquid crystal
nanocomposites. Applied Materials Today, 2020. 21: p. 100840.
11. Singh, G., M. Fisch, and S. Kumar, Emissivity and electrooptical properties of semiconducting
quantum dots/rods and liquid crystal composites: a review. Reports on Progress in Physics, 2016.
79(5): p. 056502.
12. Singh, G., M.R. Fisch, and S. Kumar, Tunable polarised fluorescence of quantum dot doped
nematic liquid crystals. Liquid Crystals, 2017. 44(3): p. 444-452.
13. Garbovskiy, Y. and A. Glushchenko, Ferroelectric Nanoparticles in Liquid Crystals: Recent
Progress and Current Challenges. Nanomaterials, 2017. 7(11).
14. Yadav, S.P. and S. Singh, Carbon nanotube dispersion in nematic liquid crystals: An overview.
Progress in Materials Science, 2016. 80: p. 38-76.
15. Pandey, F.P., et al., Dielectric and electro-optical properties of zinc ferrite nanoparticles dispersed
nematic liquid crystal 4’-Heptyl-4-biphenylcarbonnitrile. Liquid Crystals, 2020. 47(7): p. 1025-
1040.
16. Ranjkesh, A., et al., Temperature-dependent dielectric property of a nematic liquid crystal doped
with two differently–shaped tungsten oxide (W18O49) nanostructures. Journal of Molecular
Liquids, 2022. 348: p. 118024.
17. Yadav, S.P., R. Manohar, and S. Singh, Effect of TiO2 nanoparticles dispersion on ionic behaviour
in nematic liquid crystal. Liquid Crystals, 2015. 42(8): p. 1095-1101.
18. Vafaie, R., et al., Dielectric and electro optical properties of 6CHBT nematic liquid crystals doped
with MgO nanoparticles. Liquid Crystals, 2021. 48(10): p. 1417-1428.
19. Varshney, D., A. Parveen, and J. Prakash, Effect of cobalt oxide nanoparticles on dielectric
properties of a nematic liquid crystal material. Journal of Dispersion Science and Technology,
2022. 43(1): p. 42-49.
20. Tripathi, P.K., et al., Improved dielectric and electro-optical parameters of ZnO nano-particle (8%
Cu2+) doped nematic liquid crystal. Journal of Molecular Structure, 2013. 1035: p. 371-377.
21. Rani, A., S. Chakraborty, and A. Sinha, Effect of CdSe/ZnS quantum dots doping on the ion
transport behavior in nematic liquid crystal. Journal of Molecular Liquids, 2021. 342: p. 117327.
22. Nayek, P. and G. Li, Superior electro-optic response in multiferroic bismuth ferrite nanoparticle
doped nematic liquid crystal device. Scientific Reports, 2015. 5(1): p. 10845.
23. Derbali, M., et al., Dielectric, electrooptic and viscoelastic properties in cybotactic nematic phase
doped with ferroelectric nanoparticles. Journal of Molecular Liquids, 2020. 319: p. 113768.
24. Park, E.-G., C.-W. Oh, and H.-G. Park, Improvement of the electro-optical properties of nematic
liquid crystals doped with strontium titanate nanoparticles at various doping concentrations.
Liquid Crystals, 2020. 47(1): p. 136-142.
25. Elkhalgi, H.H.M., et al., Effects of manganese (II) titanium oxide nano particles on the physical
properties of a room temperature nematic liquid crystal 4-(trans-4′-n-hexylcyclohexyl)
isothiocyanatobenzene. Journal of Molecular Liquids, 2018. 268: p. 223-228.
26. Parveen, A., J. Prakash, and G. Singh, Impact of strontium titanate nanoparticles on the
dielectric, electro-optical and electrical response of a nematic liquid crystal. Journal of Molecular
Liquids, 2022. 354: p. 118907.
27. Imamaliyev, A.R., M.A. Ramazanov, and S.A. Humbatov, Effect of ferroelectric BaTiO3 particles on
the threshold voltage of a smectic A liquid crystal. 2018. 9: p. 824-828.
28. Al-Zangana, S., M. Turner, and I. Dierking, A comparison between size dependent paraelectric
and ferroelectric BaTiO3 nanoparticle doped nematic and ferroelectric liquid crystals. Journal of
Applied Physics, 2017. 121(8): p. 085105.
29. Ouskova, E., et al., Dielectric relaxation spectroscopy of a nematic liquid crystal doped with
ferroelectric Sn2P2S6 nanoparticles. Liquid Crystals, 2003. 30(10): p. 1235-1239.
30. Mishra, M., R.S. Dabrowski, and R. Dhar, Thermodynamical, optical, electrical and electro-optical
studies of a room temperature nematic liquid crystal 4-pentyl-4′-cyanobiphenyl dispersed with
barium titanate nanoparticles. Journal of Molecular Liquids, 2016. 213: p. 247-254.
31. Emdadi, M., et al., Investigation of Nematic Liquid Crystals Doped with Spherical Multiferroic
Nanoparticles in the Presence of a Magnetic Field. Brazilian Journal of Physics, 2018. 48(5): p.
433-441.
32. Lopatina, L.M. and J.V. Selinger, Theory of Ferroelectric Nanoparticles in Nematic Liquid Crystals.
Physical Review Letters, 2009. 102(19): p. 197802.
33. Newnham, R.E., R.W. Wolfe, and J.F. Dorrian, Structural basis of ferroelectricity in the bismuth
titanate family. Materials Research Bulletin, 1971. 6(10): p. 1029-1039.
34. Withers, R.L., J.G. Thompson, and A.D. Rae, The crystal chemistry underlying ferroelectricity in
Bi4Ti3O12, Bi3TiNbO9, and Bi2WO6. Journal of Solid State Chemistry, 1991. 94(2): p. 404-417.
35. Long, C., et al., High oxide ion conductivity in layer-structured Bi4Ti3O12-based ferroelectric
ceramics. Journal of Materials Chemistry C, 2019. 7(29): p. 8825-8835.
36. Liu, L., et al., Cooperation of oxygen vacancies and 2D ultrathin structure promoting CO2
photoreduction performance of Bi4Ti3O12. Science Bulletin, 2020. 65(11): p. 934-943.
37. He, H., et al., A controllable photoresponse and photovoltaic performance in Bi4Ti3O12
ferroelectric thin films. Journal of Alloys and Compounds, 2017. 694: p. 998-1003.
38. Noguchi, Y. and M. Miyayama, Large remanent polarization of vanadium-doped Bi4Ti3O12.
Applied Physics Letters, 2001. 78(13): p. 1903-1905.
39. Anu and K. Yadav, Optical and dielectric properties of Bi2Ti2O7/Bi4Ti3O12 nanocomposite.
Materials Today: Proceedings, 2020. 28: p. 153-157.
40. Hong, J. and D. Fang, Size-dependent ferroelectric behaviors of BaTiO3 nanowires. Applied
Physics Letters, 2008. 92(1): p. 012906.
41. Tsunekawa, S., et al., Critical size and anomalous lattice expansion in nanocrystalline BaTiO3
particles. Physical Review B, 2000. 62(5): p. 3065-3070.
42. Du, Y.L., et al., Size effect and evidence of a size-driven phase transition in Bi4Ti3O12 nanocrystals.
Solid State Communications, 2002. 124(3): p. 113-118.
43. Du, Y.L., G. Chen, and M.S. Zhang, Grain size effects in Bi4Ti3O12 nanocrystals investigated by
Raman spectroscopy. Solid State Communications, 2004. 132(3): p. 175-179.
44. Lorenz, A., et al., Doping the nematic liquid crystal 5CB with milled BaTiO3 nanoparticles. Physical
Review E, 2012. 86(5): p. 051704.
45. Marikani, A., et al., Ferroelectric, dielectric, and optical properties of Nd-substituted Bi4Ti3O12
nanoparticles synthesized by sol-gel method. Progress in Natural Science: Materials
International, 2016. 26(6): p. 528-532.
46. Selvamurugan, V., et al., Ferroelectric and Dielectric Behavior of Samarium-Substituted Bi4Ti3O12
Nanomaterials Synthesized by Gel Combustion Method. Transactions of the Indian Institute of
Metals, 2017. 70(4): p. 903-908.
47. Meng, X.B., et al., Enhanced ferroelectric and UV photocatalytic properties in a Bi4Ti3O12@ZnO
core–shelled nanostructure. Journal of Materials Science: Materials in Electronics, 2014. 25(3): p.
1423-1428.
48. Makovec, D., et al., Ferroelectric bismuth-titanate nanoplatelets and nanowires with a new
crystal structure. Nanoscale, 2022. 14(9): p. 3537-3544.
49. Bhardwaj, S., et al., Electroactive Phase Induced Bi4Ti3O12–Poly(Vinylidene Difluoride) Composites
with Improved Dielectric Properties. Journal of Electronic Materials, 2015. 44(10): p. 3710-3723.
50. Varshney, D., et al., Probing the impact of bismuth-titanate based nanocomposite on the
dielectric and electro-optical features of a nematic liquid crystal material. Journal of Molecular
Liquids, 2022. 347: p. 118389.
51. Shukla, R.K. and K.K. Raina, Structural and dielectric behaviors of Bi4Ti3O12 – lyotropic liquid
crystalline nanocolloids. Phase Transitions, 2018. 91(3): p. 301-307.
52. Zhang, Z., et al., Low-temperature synthesis of Bi 4 Ti 3 O 12 nanocrystals by hydrothermal
method. Journal of Materials Science: Materials in Electronics, 2018. 29(9): p. 7453-7457.
53. Hao, H., et al., Lead-Free SrBi4Ti4O15 and Bi4Ti3O12 Material Fabrication Using the Microwave-
Assisted Molten Salt Synthesis Method. Journal of the American Ceramic Society, 2007. 90(5): p.
1659-1662.
54. Kumar, A., et al., Microwave chemistry, recent advancements, and eco-friendly microwave-
assisted synthesis of nanoarchitectures and their applications: a review. Materials Today Nano,
2020. 11: p. 100076.
55. Ratna, B.R. and R. Shashidhar, Dielectric properties of 4′-n-alkyl-4-cyanobiphenyls in their
nematic phases. Pramana, 1976. 6(5): p. 278-283.
56. Joshi, T., et al., Alumina nanoparticles find an application to reduce the ionic effects of
ferroelectric liquid crystal. Journal of Physics D: Applied Physics, 2011. 44(31): p. 315404.
57. Podoliak, N., et al., Elastic constants, viscosity and response time in nematic liquid crystals doped
with ferroelectric nanoparticles. RSC Advances, 2014. 4(86): p. 46068-46074.
58. Cole, K.S. and R.H. Cole, Dispersion and Absorption in Dielectrics I. Alternating Current
Characteristics. The Journal of Chemical Physics, 1941. 9(4): p. 341-351.
59. Dhar, R., Comments on the fitting of Cole-Cole/Havriliak-Negami equation with the dielectric
data under the influence of parasitic effects in order to extract correct parameters of the
materials. Journal of Molecular Liquids, 2021. 343: p. 117682.
60. Singh, U.B., et al., Influence of low concentration silver nanoparticles on the electrical and
electro-optical parameters of nematic liquid crystals. Liquid Crystals, 2013. 40(6): p. 774-782.
61. Arya, A. and A.L. Sharma, Effect of salt concentration on dielectric properties of Li-ion conducting
blend polymer electrolytes. Journal of Materials Science: Materials in Electronics, 2018. 29(20):
p. 17903-17920.
62. Golda, R.A., A. Marikani, and D.P. Padiyan, Mechanical synthesis and characterization of
Bi4Ti3O12 nanopowders. Ceramics International, 2011. 37(8): p. 3731-3735.
63. Anu, et al., Effect of oxygen vacancies, lattice distortions and secondary phase on the structural,
optical, dielectric and ferroelectric properties in Cd-doped Bi2Ti2O7 nanoparticles. Materials
Research Bulletin, 2021. 141: p. 111373.
64. Araghi, M.A., N. Shaban, and M. Bahar, Synthesis and characterization of nanocrystalline barium
strontium titanate powder by a modified sol-gel processing. Materials Science-Poland, 2016.
34(1): p. 63-68.
65. Zhang, H., S. Jiang, and K. Kajiyoshi, Preparation and characterization of sol–gel derived sodium–
potassium bismuth titanate powders and thick films deposited by screen printing. Journal of
Alloys and Compounds, 2010. 495(1): p. 173-180.
66. Basith, D.M., et al., Enhanced photocatalytic dye degradation and hydrogen production ability of
Bi25FeO40-rGO nanocomposite and mechanism insight. Scientific Reports, 2018. 8: p. 8:11090.
67. Singh, D.J., S.S.A. Seo, and H.N. Lee, Optical properties of ferroelectric Bi4Ti3O12. Physical Review
B, 2010. 82(18): p. 180103.
68. Wang, J., et al., Realizing semiconductivity by a large bandgap tuning in Bi4Ti3O12 via inserting
La1-xSrxMnO3 perovskite layers. Applied Physics Letters, 2017. 110(21): p. 212102.
69. Zhao, X., et al., Synthesis and theoretical study of large-sized Bi4Ti3O12 square nanosheets with
high photocatalytic activity. Materials Research Bulletin, 2018. 107: p. 180-188.
70. Xu, G., et al., Hydrothermal synthesis and formation mechanism of the single-crystalline Bi4Ti3O12
nanosheets with dominant (010) facets. CrystEngComm, 2016. 18(13): p. 2268-2274.
71. Pan, R.-P., et al., Temperature-dependent optical constants and birefringence of nematic liquid
crystal 5CB in the terahertz frequency range. Journal of Applied Physics, 2008. 103(9): p. 093523.
72. Ozerov, R.P. and A.A. Vorobyev, 4 - Dielectric Properties of Substances, in Physics for Chemists,
R.P. Ozerov and A.A. Vorobyev, Editors. 2007, Elsevier: Amsterdam. p. 251-304.
73. Shukla, R.K. and K. Raina, Structural and dielectric behaviors of Bi4Ti3O12–lyotropic liquid
crystalline nanocolloids. Phase Transitions, 2018. 91(3): p. 301-307.
74. Lin, Y., et al., A comparative study of nematic liquid crystals doped with harvested and non-
harvested ferroelectric nanoparticles: phase transitions and dielectric properties. RSC advances,
2017. 7(56): p. 35438-35444.
75. Dubey, R., et al., Electric behaviour of a Schiff's base liquid crystal compound doped with a low
concentration of BaTiO3 nanoparticles. Journal of Molecular Liquids, 2017. 225: p. 496-501.
76. Vardanyan, K.K. and B. Bates, Dielectric and Material Properties of 5CB Nematic Dispersed with
Graphene Nanoparticles. Liquid Crystals, 2021: p. 1-13.
77. Mukherjee, P.K., Impact of ferroelectric nanoparticles on the dielectric constant of nematic liquid
crystals. Soft Materials, 2021. 19(1): p. 113-116.
78. Ruiz Puigdollers, A., et al., Increasing Oxide Reducibility: The Role of Metal/Oxide Interfaces in
the Formation of Oxygen Vacancies. ACS Catalysis, 2017. 7(10): p. 6493-6513.
79. Chen, H., et al., Oxygen vacancy regulation in Nb-doped Bi2WO6 for enhanced visible light
photocatalytic activity. RSC Advances, 2019. 9(39): p. 22559-22566.
80. Peterson, M.S.E., et al., Dielectric analysis of the interaction of nematic liquid crystals with
carbon nanotubes. Liquid Crystals, 2018. 45(3): p. 450-458.
81. Bokov, A.A., et al., Dielectric spectra and Vogel-Fulcher scaling in Pb(In0.5Nb0.5)O3relaxor
ferroelectric. Journal of Physics: Condensed Matter, 1999. 11(25): p. 4899-4911.
82. Jonscher, A.K., The ‘universal’ dielectric response. Nature, 1977. 267(5613): p. 673-679.
83. Uttam, R., S. Kumar, and R. Dhar, Magnified charge carrier conduction, permittivity, and
mesomorphic properties of columnar structure of a room temperature discotic liquid crystalline
material due to the dispersion of low concentration ferroelectric nanoparticles. Physical Review
E, 2020. 102(5): p. 052702.
84. Dyre, J.C. and T.B. Schrøder, Universality of ac conduction in disordered solids. Reviews of
Modern Physics, 2000. 72(3): p. 873-892.
85. Pandey, A.S., et al., Enhancement of the display parameters of 4′-pentyl-4-cyanobiphenyl due to
the dispersion of functionalised gold nano particles. Liquid Crystals, 2011. 38(1): p. 115-120.
86. Garbovskiy, Y., Nanomaterials in Liquid Crystals as Ion-Generating and Ion-Capturing Objects.
Crystals, 2018. 8(7): p. 264.
87. Garbovskiy, Y., Adsorption/desorption of ions in liquid crystal nanocolloids: the applicability of
the Langmuir isotherm, impact of high electric fields and effects of the nanoparticle’s size. Liquid
Crystals, 2016. 43(6): p. 853-860.
88. Garbovskiy, Y. and I. Glushchenko, Nano-Objects and Ions in Liquid Crystals: Ion Trapping Effect
and Related Phenomena. Crystals, 2015. 5(4): p. 501-533.
89. Reddy, B.M., et al., Microwave-assisted Synthesis and Structural Characterization of Nanosized
Ce0.5Zr0.5O2 for CO Oxidation. Catalysis Letters, 2009. 130(1): p. 227-234.
90. Krishna Prasad, S., et al., Electrical conductivity and dielectric constant measurements of liquid
crystal–gold nanoparticle composites. Liquid Crystals, 2006. 33(10): p. 1121-1125.
91. Srivastava, S.L. and R. Dhar, Effect of γ-irradiation on liquid crystalline properties of cholesteryl
pelargonate (nonanoate). Radiation Physics and Chemistry, 1996. 47(2): p. 287-293.
92. Kocakülah, G., G. Algül, and O. Köysal, Effect of CdSeS/ZnS quantum dot concentration on the
electro-optical and dielectric properties of polymer stabilized liquid crystal. Journal of Molecular
Liquids, 2020. 299: p. 112182.
93. Jahanbakhsh, F., et al., Dispersion of multiferroic BiFeO3 nanoparticles in nematic liquid crystals.
Applied Physics A, 2019. 125(12): p. 877.
94. Paul, S.N., et al., Change in Dielectric and Electro-Optical Properties of a Nematic Material
(6CHBT) Due to the Dispersion of BaTiO3 Nanoparticles. Molecular Crystals and Liquid Crystals,
2011. 545(1): p. 105/[1329]-111/[1335].
95. Maier, W. and G. Meier, A simple theory of the dielectric are some homogeneous criteria
oriented liquid crystal phases of nematic type. Zeitschrift für Naturforschung A, 1961. 16(3): p.
262-267.
96. Lin, Y., et al., Electric field effects on phase transitions in the 8CB liquid crystal doped with
ferroelectric nanoparticles. Physical Review E, 2016. 93(6): p. 062702.
97. Lin, Y., et al., Correlation between dielectric properties and phase transitions of 8CB/Sn2P2S6
liquid crystal nanocolloids. Journal of Molecular Liquids, 2017. 232: p. 123-129.
98. Lin, Y., et al., On the phase transitions of 8CB/Sn2P2S6 liquid crystal nanocolloids. The European
Physical Journal E, 2015. 38(9): p. 103.
99. Basu, R., Soft memory in a ferroelectric nanoparticle-doped liquid crystal. Physical Review E,
2014. 89(2): p. 022508.
100. Gdovinová, V., et al., Memory effect in nematic phase of liquid crystal doped with magnetic and
non-magnetic nanoparticles. Journal of Molecular Liquids, 2019. 282: p. 286-291.
101. Kempaiah, R., et al., Giant soft-memory in liquid crystal nanocomposites. Applied Physics Letters,
2016. 108(8): p. 083105.
102. Prakash, J., A. Chandran, and A.M. Biradar, Scientific developments of liquid crystal-based optical
memory: a review. Reports on Progress in Physics, 2016. 80(1): p. 016601.

You might also like