You are on page 1of 14

Journal of Molecular Liquids 355 (2022) 118925

Contents lists available at ScienceDirect

Journal of Molecular Liquids


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

Comprehensive characterization of glycerol/ZnO green nanofluids for


advances in multifunctional soft material technologies
R.J. Sengwa ⇑, Mukul Saraswat, Priyanka Dhatarwal
Dielectric Research Laboratory, Department of Physics, Jai Narain Vyas University, Jodhpur 342 005, India

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

Article history: Nanofluids (NFs) are increasingly used in most of the areas of soft condensed matter technologies, espe-
Received 5 January 2022 cially for heat transfer, energy harvesting, electrical insulation, and in the development of a variety of
Revised 1 March 2022 optoelectronic devices and controllable systems. Nowadays, green nanofluids are of considerable interest
Accepted 11 March 2022
to researchers from an environmental safety point of view. Therefore, in this work, green nanofluids hav-
Available online 15 March 2022
ing biodegradable glycerol (GLY) as base fluid with homogeneous dispersion of eco-friendly zinc oxide
(ZnO) nanoparticles (concentration range 0.01–1.00 wt%) were prepared and comprehensively character-
Keywords:
ized by employing various advanced techniques for exploring their promising multifunctional properties.
Glycerol
Zinc oxide
The X-ray diffraction study of these GLY/ZnO green nanofluids, which is the first of its kind on nanofluids,
Green nanofluids confirmed the existence of monodispersed ZnO nanoparticles in the arranged 3D supramolecular struc-
Structural properties tural network of the glycerol fluid. The ZnO nanoparticles having higher suspension stability in the GLY
Dielectric dispersion fluid were able to exhibit all the characteristic crystalline diffraction peaks for the concentration  0.3
Thermophysical parameters 0 wt%, at which the visible range photons absorption became almost saturated. The detailed analysis
of the UV–Vis absorbance spectra of these nanofluids with increasing ZnO concentration explained sub-
stantial alterations in the electronic transition energy for the lone electron pairs of the hydroxyl oxygen
atoms of glycerol molecules and also the valence band electrons of the ZnO nanomaterial. Dielectric and
electrical spectra of these nanofluid materials are carried out in the harmonic field of 20 Hz to 1 MHz
range, at 25 °C. A small change in the static dielectric permittivity of these GLY/ZnO nanofluids with
the increase of ZnO concentration reveals some modifications in the parallel arranged dipolar ordering
of the glycerol molecules. The conductivity relaxation process ruled the electrical conductivity of the
GLY/ZnO nanofluids, but the low conductivity of these materials confirmed them as potential candidates
for electric insulating nanodielectrics. The decreased viscosity and the increased thermal conductivity
with some constructive changes in the acoustic parameters of these nanofluids were noted with the
increase of ZnO concentration. The experimental results on the structural, optical, dielectric, electrical,
and thermophysical properties of the GLY/ZnO nanofluids are technologically appealing and recognize
these as innovative multifunctional soft materials.
Ó 2022 Elsevier B.V. All rights reserved.

1. Introduction liquid and solid phases functional materials including numerous


novel nanofluids for the heat transfer systems [6–12], optical
In the last two decades, the ZnO material with a variety of crys- materials for energy harvesting devices [3,10,13,14], photocataly-
talline nanostructures has attracted much attention as an sis [5,15,16], sensors [17], antibacterial and drug release controller
advanced smart material due to its extraordinary electronic and in cosmetics and pharmaceutical products [9,18], and polymer
optical properties [1–5]. This metal oxide is a well-established nanocomposites for flexible optoelectronic and microelectronic
non-toxic, bio-safe, and biocompatible compound semiconductor devices [19–23].
of wide energy band gap (Eg ffi 3.3 eV) [1,4,5]. The ZnO crystals In regards to the nanofluids dispersed with ZnO nanoparticles,
are prominently used in the development of extensive advanced these are largely prepared using water, ethylene glycol, and oils
as base fluids, and are mostly characterized for their thermal con-
ductivity and viscosity behaviour in regards to evaluating their
⇑ Corresponding author. suitability in heat transfer systems [6–8,10–12,24–29]. To the best
E-mail addresses: rjsengwa@rediffmail.com, rjs.ph@jnvu.edu.in (R.J. Sengwa).

https://doi.org/10.1016/j.molliq.2022.118925
0167-7322/Ó 2022 Elsevier B.V. All rights reserved.
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

of our knowledge, the glycerol (GLY) and ZnO-based GLY/ZnO 2. Experimental


nanofluids are less attempted [30] especially to establish a correla-
tion between their structural, optical, and dielectric properties 2.1. Materials
with the variation of ZnO concentration for the technological
advances. Glycerol (GLY) of purity greater than 99.0% for spectroscopy
Glycerol (propane-1,2,3-triol) is categorized as a green polar grade was purchased from HiMedia Laboratories Pvt. Ltd., India.
solvent owing to its biodegradable, biocompatible, water soluble, The nanopowder sample of zinc oxide (ZnO) having particle sizes
and non-toxic characteristics, and therefore it has been realized less than 100 nm, density 5.6 g/cc, and surface area 10–25 m2/g
for versatile technological and industrial uses which are thor- was procured from Sigma-Aldrich Co., USA.
oughly demonstrated in the recent reviews [31–34]. Three alco-
holic hydroxyl groups (–OH) in the GLY molecular formula OH–
2.2. Preparation of nanofluids
CH2–CH(OH)–CH2–OH make it highly interactive with all kinds
of inorganic/organic additives. Furthermore, the high viscosity,
The GLY/x (wt%) ZnO nanofluids comprises GLY as base fluid
elevated boiling point, low vapour pressure, and appreciable
and different weight percent (x (wt%)) amounts of ZnO nanopow-
dielectric polarization strength altogether realize the GLY as an
der to the weight of GLY (x = 0.00, 0.01, 0.02, 0.05, 0.30, and
attractive green base fluid for the preparation of nanofluids hav-
1.00) were prepared following a standard two-step method. In
ing better suspension stability of the nanoparticles. Investigations
the first step, the GLY was taken into a stoppered glass bottle by
on glycerol-based nanofluids have established that these are
the required amount of volume and simultaneously its weight
extremely promising materials from technological points of view
was measured. In the second step, the correct weight of the ZnO
because of their high thermal stability and also good suspension
nanopowder for a particular weight concentration x (wt%) was
durability of the dispersed nanoparticles in the high viscous
taken and then added to the GLY base fluid previously filled in
GLY fluid [35–38], which are critical characteristics required to
the bottle. An analytical electronic balance (Wensar Weighing
maintain longer workability performance of the nanofluid
Scales Limited, Chennai, India, model: MAB 250) with a standard
[26,29,39].
uncertainty of 0.01 mg was used for the weight measurements.
A careful review of literature pointed out that there is a research
To get the homogeneous dispersion of ZnO nanoparticles into the
gap on the XRD study of nanofluids for their structural characteri-
GLY fluid medium, firstly the GLY/ZnO material was continuously
zation and the confirmation of nanoparticles suspension stability,
stirred for about 30 min at about 25 °C, with gradual increasing
especially in the case of crystalline nanoparticles. Mostly the XRD
the speed of the bar magnet up to 1000 rpm by employing a mag-
traces of the powder nanoparticles are reported for their structural
netic bar stirrer (Sudarshan Chemical, Kota, India, model: AlphaTM).
confirmation [11,25,27,40,41] which were then used for the prepa-
Thereafter, this bar magnet stirred solution was sonicated for
ration of nanofluids, but such nanofluids have not been character-
10 min using a high-energy probe sonicator operated with con-
ized in detail with the XRD technique, so far. Additionally, in most
trolled power for 15 s On/Off through cyclic timer (Trans-O-
of the researches, the UV–Vis measurements are performed for
sonic, Mumbai, India; model: TOS Ultrasonic Disintegrator - power
checking the homogeneity and stability of nanoparticles in the
250 W and frequency 20 kHz). In the probe sonication process,
nanofluids with repetitive measurements after appropriate inter-
high pressure cavitations were generated by the directly inserted
vals [11,25,28,42–44]. However, the detailed study of electronic
probe into the sample, and their transmission crushed the aggre-
transitions and energy band gaps of the nanofluids having a sus-
gated nanoparticles and turned the nanofluid sample of very high
pension of wide band gap nanoparticles with the concentration
homogeneity. During the sonication, the heat was generated in
variations are less attempted to confirm their suitability in liquid
the medium by absorbing the propagating ultrasonic wave energy
phase optical devices and also the photovoltaic/thermal systems
(acoustic energy) and hence, the temperature of the sample solu-
[10,44,45]. Furthermore, a correlation has not been made yet,
tion was increased by about 60 °C. All the GLY/ZnO nanofluids with
between the particles suspension stability and homogeneity in
different concentrations of ZnO were prepared by following the
nanofluids with the simultaneous analysis of their XRD and UV–
same steps.
Vis spectroscopic results. In addition to this research gap, the
Digital picture of these GLY/x (wt%) ZnO nanofluids (x = 0.00,
dielectric and electrical characterization with frequency variation
0.01, 0.02, 0.05, 0.30, and 1.00 wt%) is shown in Fig. 1. This picture
of the harmonic electric field using dielectric relaxation spec-
shows that the glycerol sample is a clear transparent fluid, and the
troscopy (DRS) is less performed on the nanofluids to explore the
GLY/ZnO nanofluid degree of transparency reduced largely as the
character of molecule-nanoparticle interaction and their effect on
ZnO concentration was increased from 0.01 to 0.05 wt%. At higher
fluid dipolar ordering in the case of H-bonded base fluid, and also
ZnO concentrations (i.e., 0.30 wt% and 1.00 wt%), these nanofluids
the changes in dielectric relaxation processes [46,47]. However,
appeared dense milky and turned opaque. Furthermore, sedimen-
relative dielectric permittivity and electrical conductivity studies
of some important nanofluids were performed and thoroughly
analyzed to explore the effect of particles sizes and the type of base
fluids on the dielectric polarization and charge conduction beha-
viour [48,49].
Therefore, after the deliberation of all the above-mentioned
facts, the present paper deals with XRD, UV–Vis, and DRS charac-
terization techniques employed simultaneously on the GLY/ZnO
nanofluids, having different concentrations of ZnO, to understand
in detail their structural properties and promising multifunctional-
ity for advances in condensed matter technologies. Additionally,
the thermophysical properties like viscosity, density, ultrasonic
velocity, thermal conductivity, and several important acoustic
parameters of these nanofluids were also characterized for explor-
ing their appropriateness as conventional energy transporting/har- Fig. 1. Digital photograph of GLY/x (wt%) ZnO nanofluids (x = 0.00, 0.01, 0.02, 0.05,
vesting materials. 0.30, and 1.00).

2
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

tation of the ZnO nanoparticles in these nanofluids was found measurements, calibration of the microprocessor controlled this
insignificant when kept for a longer duration and seen with naked viscometer was done following the manufacturer instructions.
eyes. The assessment of sedimentation was experimentally per- During the sample measurements, the required torque was main-
formed using the UV–Vis absorbance spectra and the XRD traces tained in the range provided by the manufacturer (i.e., 15% to
of these materials which were thoroughly discussed in the follow- 100%) for the standard uncertainty of 1%.
ing ‘Results and discussion’ section. These experimental results The density measurements of the nanofluid samples were per-
confirmed that the nanoparticles suspension stability level is high formed by filling them in a precision relative density bottle of
due to which the homogeneity was retained and therefore, there is 10 ml capacity capped with a fine capillary Teflon cap (Borosil
no challenge in alteration of their performance with time elapse. Glass Works Ltd. Bharuch, India; Catalogue Code 1624006), and
weighing up to an accuracy of 0.01 mg using analytical electronic
2.3. Measurements balance (Wensar, India, model: MAB 250). Prior to sample mea-
surements, the density bottle was tested using deionized water
The crystallographic characteristics of the GLY/ZnO nanofluids at 25 °C and noted the standard uncertainty of 0.005 kg/m3. The
were explored from their XRD traces recorded by employing an refractive index of these nanofluid samples was measured with a
X-ray diffractometer (Malvern PANalytical B.V., Almelo, Nether- resolution of 0.0001 using an Abbe refractometer (Mittal Enter-
lands, model: X0 pert Pro Multi-Purpose Diffractometer). The Cu- prises Co., New Delhi, India, model: MA-02).
Ka monochromatic X-ray beam of wavelength k = 0.15406 nm The temperature of the sample, during the measurements with
was generated by the X-ray tube at 1800 W with a bias supply of the LCR meter, ultrasonic nanofluid interferometer, viscometer,
45 kV and 40 mA. For the XRD measurements, the nanofluid sam- density bottle, and Abbe refractometer was thermostated at
ple was filled in the powder sample holder and then recorded its 25 °C by employing a temperature controller of accuracy 0.01 °C
XRD trace in reflection mode operation of the diffractometer over (Thermo-Haake, Germany, model: Haake DC 10) fitted along with
a wide angular range 2h from 10° to 70° with a scan step of heating and cooling devices in the water circulating bath (Escy
0.05 deg/s. The XRD traces of the nanofluids were recorded filling Enterprises, Pune, India, model: Z).
them one by one in the sample holder at the ambient temperature.
The optical characterizations of the nanofluids are performed
from the ultraviolet–visible (UV–Vis) absorbance spectra recorded 3. Results and discussion
in the wavelength range from 200 nm to 800 nm with an accuracy
of 1 nm. A dual-beam UV–Vis spectrophotometer (Agilent Tech- 3.1. Structural properties
nologies Pvt. Ltd., USA, model: Cary 60 controlled with Cary WinUV
software) was used for UV–Vis absorbance measurements at ambi- The XRD pattern of the ZnO nanopowder has intense and sharp
ent temperature. For these measurements, each sample was filled diffraction peaks (see the upper layer of Fig. 2) confirming the crys-
one after another in the quartz cuvette of 3 ml volume and 1 cm talline structure of its nanoparticles [20] with a high degree of
width, and the experiment was executed at a scan rate of 10 nm/ crystallinity that is about 93% as reported previously [51]. The
s with subtraction of baseline corrections. hkl indices marked on the peaks of various planes of ZnO identified
Dielectric spectra of the GLY/ZnO samples were determined its hexagonal wurtzite crystalline structures explained in the ear-
from capacitance and resistance measurements using a precision lier study [20]. The XRD pattern of highly viscous polyol glycerol
inductance-capacitance-resistance (LCR) meter of the harmonic fluid (labeled with x = 0.00 in the lower layer of Fig. 2) exhibited
field frequency range from 20 Hz to 1 MHz. For the electrical mea- a substantial intense peak at 2h of 20.87° suggesting some amount
surements with ASTM D150 standard, the 4284A LCR meter having of crystalline phase present in its molecular structure. It seems that
compatible 16452A parallel plates four-terminals liquid test fix- the ordered chain structure of the H-bonded GLY molecules real-
ture of Agilent Technologies Pvt. Ltd., USA, was used. The complex ized the diffraction plane of some crystalline phases. This beha-
dielectric permittivity e*(x) = e0 – je00 values as a function of fre- viour is found similar to that of the H-bonded chain structure in
quency, at a fixed temperature of 25 °C, were determined with the flexible semicrystalline poly(vinyl alcohol) matrix [20]. The
the relative combined uncertainty of less than 1%. The values of study on dipole moment, static dielectric permittivity, Kirkwood
various transformed complex electrical functions viz. alternating correlation factor (a measure of the degree of parallel aligned
current (ac) electrical conductivity r*(x) = r0 + jr00 , electric mod- dipoles in the H-bonded chain-like molecular structure), and a
ulus M*(x) = M0 + jM00 , and impedance Z*(x) = Z0 – jZ00 were com- molecular reorientation dynamics explained that the glycerol
puted by the mutual relations 1/e*(x) = M*(x) = jxC0Z* molecules in its high viscous liquid phase form a dense H-boned
(x) = jxe0/r*(x) as described in detail elsewhere [50]. three-dimensional (3D) ordered supramolecular structural net-
The ultrasound velocity u for the acoustic wave of frequency work [52,53]. These findings support the appearance of diffraction
f = 2 MHz propagating in the nanofluid sample was measured by peak in the XRD pattern of the liquid glycerol system noted in this
employing a Nanofluid Interferometer and its compatible TC–2 test study. Additionally, a low-temperature crystallographic XRD data
cell (Mittal Enterprises Co., New Delhi, India, model: NF-10X). The collection and their detailed analysis with the help of crystal struc-
standing wave pattern was produced in the sample filled ther- ture software package established that the glycerol molecules
mostated test cell from which the wavelength k in the liquid sam- assembled through H-bonding have formed 2D wave-like sheet
ple was noted. The velocity of the propagating ultrasound was structures, and these sheet structures also connected through H-
computed using wave equation u = f k with the standard uncer- bonding resulted into the 3D structure of crystalline glycerol [54]
tainty of 0.2 m/s. Various acoustic parameters and thermal conduc- also favouring the XRD result of this study.
tivity of the samples were determined using the ultrasonic velocity Fig. 2 explains that when the ZnO nanoparticles were dispersed
based relations which are demonstrated in the following section in the glycerol 3D network structure, then several structural mod-
3.4.3. ifications occurred which are reflected from the comparative
The dynamic viscosity of these GLY/ZnO nanofluids was mea- changes in the shape of the XRD traces of GLY/x (wt%) ZnO nanoflu-
sured at a fixed shear rate of 4 s1 by employing a rotational vis- ids. Firstly, the glycerol peak was significantly modified with
cometer (rheometer) equipped with a compatible APM sample reduced intensity and also an alteration in the width confirming
adapter cell of 8 ml capacity and a TL-5 cylindrical spindle (Fungi- the formation of electrostatic interaction between the glycerol
lab S.A., Barcelona, Spain, model: Alpha series L). Before nanofluids molecules and the ZnO nanoparticles, and as a result of this fact,
3
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

Secondly, the ZnO peaks are little noticeable in the glycerol fluid
at 0.05 wt% revealing that there were no aggregates and sedimen-
tation of the dispersed nanoparticles, and the nanofluid bears high
homogeneity. Thirdly, at 0.30 wt% and 1.00 wt% concentrations,
the ZnO peaks appeared in the XRD traces of these nanofluids con-
firming that the suspended nanoparticles have settled in a highly
arranged planar structure with appreciable suspension stability.
Due to this fact, all the conditions of Bragg’s law satisfied and the
characteristic crystal diffraction peaks of the ZnO have appeared
reasonably intense. Fourthly, there is more than three times
increase in the intensity of the ZnO peaks with the increase of its
concentration from 0.30 wt% to 1.00 wt% in the glycerol base fluid
revealing an increase in density of orderly arranged suspended ZnO
nanoparticles which are as expected due to higher concentration of
loaded nanoparticles as concluded from the XRD results of numer-
ous flexible-type polymer nanocomposites [20,21,23]. Further-
more, all the crystalline peaks of the ZnO are clearly noticeable
in the nanofluid and their relative intensities are also found in
the order that is of pure ZnO powder when 1.00 wt% ZnO nanopar-
ticles are suspended in GLY fluid. This finding evidenced an appro-
priate formation of three-dimensional crystallographic
arrangement of the suspended nanoparticles in the base fluid.
And lastly, the positions of the ZnO diffraction peaks slightly
shifted to higher 2h values in the nanofluids as compared to that
of the pristine ZnO powder, which infers that the interaction of
glycerol molecules with ZnO particles also modified the surface
area of the nanoparticles.
With the consideration of these XRD results on the GLY/ZnO
nanofluids, herein, we propose a structural network of the GLY
fluid-based nanofluid having ZnO concentration  0.30 wt% as
shown in Fig. 3. In this figure, the 2D wave-like H-bonded chains
of glycerol molecules assembled in sheet form turned into the 3D
supramolecular network by mutual H-bonding of the 2D sheets.
The ZnO nanoparticles are set orderly arranged between the glyc-
erol molecular sheets and satisfy all the Bragg’s law conditions of
the crystalline ZnO material. Due to the high viscosity of the GLY
fluid and the molecule-particle electrostatic interaction, the GLY
supramolecular structural network is capable to hold the dispersed
ZnO nanoparticles with greater suspension stability and producing
the characteristic diffraction peaks of the ZnO in the base fluid. The
fluid-nanoparticle structural arrangement, as sketched in Fig. 3 for
the GLY/ZnO nanofluids is exclusive in regards to the high viscosity
base fluid and the dispersed crystalline nanoparticles. This struc-
tural arrangement may be interesting to understand the structural
assembly of other nanofluids with similar physicochemical
properties.

Fig. 2. XRD patterns of the ZnO nanopowder and the GLY/x (wt%) ZnO nanofluids
(x = 0.00, 0.05, 0.30, and 1.00).
3.2. Optical properties

The UV–Vis characterization of the nanofluids is frequently con-


ducted to have an idea about the stability of the suspended
some alterations in the glycerol ordered 3D supramolecular struc- nanoparticles and aggregation/sedimentation with the elapsed
ture occurred which further changed with the variation of ZnO time [11,25,28,42–44] and also to explore the usable properties
concentration in these nanofluids. For the confirmation of alter- related to their photovoltaic/thermal applications [10,43,44,55].
ation in glycerol crystallite structure, the values of peak position But in addition to these applications, the nanofluids having the
2h, d-spacing between the glycerol H-bonded 2D sheet structure optoelectronic nanomaterial can be characterized in detail their
d (d = k/2sinh), crystallite size L (L = 0.94k/bcosh), and interchain optical properties by UV–Vis spectroscopy for the identification
separation distance R (R = 5k/8sinh) were determined for the of suspended optical nanoparticles, their relative concentration,
GLY/ZnO nanofluids and these structural parameters are listed in and also the behaviour of electronic transitions in such liquid
Table 1. It can be read from this table that there are significant vari- phase materials. Keeping these facts in mind, herein, we performed
ations in these crystal structure parameters of GLY with the the UV–Vis measurements of the GLY/ZnO nanofluids for the
increase of ZnO concentration which evidence alteration in the detailed optical characterization, and also to explore their possible
supramolecular structural network of the glycerol molecules due optoelectronic applications, and also to identify a correlation with
to their interaction with ZnO nanoparticles. some of the structural properties confirmed from the XRD mea-
surements in the previous section.
4
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

Table 1
Values of the glycerol crystallographic parameters (diffraction peak position 2h, d-spacing between H-bonded molecular chains sheets d, full width at half maximum FWHM,
crystallite size L, and interchain separation distance R) in the GLY/x (wt%) ZnO nanofluids (NFs; x = 0.00, 0.05, 0.30, and 1.00).

NFs 2h d FWHM L R
x (wt%) (deg) (nm) (10–3 rad) (nm) (nm)
0.00 20.89 0.425 80.74 1.82 0.531
0.05 21.53 0.412 64.44 2.29 0.516
0.30 21.20 0.419 61.76 2.38 0.523
1.00 20.75 0.428 76.86 1.92 0.535

Fig. 3. The sketch of hydrogen–bonded 3D supramolecular structure of glycerol molecules along with orderly arranged ZnO nanoparticles. The electrostatic interaction
between the glycerol molecules and ZnO nanoparticles provided high suspension stability to the nanoparticles in the highly viscous media. This structure satisfied the Bragg’s
law conditions and produced the characteristic diffraction peaks of ZnO for the GLY/x (wt%) ZnO nanofluid samples for x  0.30. In the sketched GLY/ZnO nanofluid structure,
the sizes of the constituent are not in exact molecular and particle scales.

The UV–Vis absorption spectra of the GLY/x (wt%) ZnO nanoflu- bance attained steady-state up to the UV range of about 240 nm
ids are shown in Fig. 4a. These spectra infer several appealing fea- which is the onset wavelength value of the GLY absorbance band.
tures related to the characteristic optical behaviour of the glycerol These changes are related to the characteristic optical properties
base fluid and the optoelectronic properties of the zinc oxide nano- of the ZnO nanomaterial dispersed in the glycerol fluid. When the
material. The UV–Vis spectrum of pure glycerol confirms that the concentration of ZnO was increased from 0.01 wt% to 0.02 wt%,
GLY is a transparent fluid covering the entire visible wavelength the absorbance further increased in the visible range and appeared
range of photons and, also showing insignificant absorbance in almost twice in the UV region with displaying the identical charac-
the higher wavelength UV radiations (UV-A and UV-B bands). This teristic of ZnO peak and a sharp absorbance band related to the
optical behaviour of the glycerol signifies that the energy of the electronic transition from the oxygen atoms of glycerol molecules.
electromagnetic radiations of this wavelength range is not suffi- This finding evidenced the double concentration of the suspended
cient to excite the lone pair electrons of the oxygen atoms present ZnO nanoparticles, which is 0.02 wt% in comparison to 0.01 wt%
in the hydroxyl groups of glycerol molecules. However, at a lower concentration and supports the similar higher suspension stability
wavelength UV radiation, an absorbance band has appeared in two of the dispersed nanoparticles. At 0.05 wt% concentration of ZnO,
steps (initial onset around 240 nm and the later step onset around there is extremely high absorption for the visible range photons
215 nm) which signifies the n ? r* electronic transitions associ- and the increasing trend of absorption value has appeared in two
ated with the lone pair electrons of the oxygen atoms which has steps with decreased wavelength. Initially, there is a large increase
been also reported and discussed for some other alcohols in the in absorption with a reducing wavelength from 800 nm to 600 nm
previous literature [56]. range, and then a gradual increase in the 600 nm to 400 nm range.
In comparison to the pure glycerol UV–Vis absorbance spec- Lastly, the sample exhibited the ZnO characteristic plasmon peak at
trum, the GLY/x (wt%) ZnO nanofluids exhibited extremely different about 375 nm and then turned saturated in the remaining UV wave-
characteristic absorption spectra (see Fig. 4a). Initially, with a small length range. The ZnO peak position in the UV–Vis spectra of these
addition of ZnO nanoparticles (0.01 wt%) in the glycerol fluid, a sig- nanofluids remained unaltered besides the increase in its concen-
nificant absorption in the visible region was exhibited. The absor- tration confirming that there was no aggregation of the nanoparti-
bance value was increased exponentially when the wavelength of cles because in such cases the peak shifts relatively to a higher
the photon was decreased, and at the start of the UV region, the wavelength side [25].
absorption turned into a sharp peak at around 375 nm representing Fig. 4a illustrates that when the ZnO concentration increased to
the ZnO plasmonic characteristics [20,23,25]. Thereafter, absor- 0.30 wt% the absorbance turned almost saturated (possible maxi-

5
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

Fig. 4. (a) UV–Vis absorbance spectra of GLY/x (wt%) ZnO nanofluids (x = 0.00, 0.01, 0.02, 0.05, 0.30, and 1.00), (b) plots of absorbance (Abs.) versus x (wt%) of ZnO for the
nanofluids at fixed wavelengths (i.e., k = 600 nm and 800 nm), (c) Tauc plots ((ahm)2 versus hm) of the nanofluids showing the extrapolated dashed lines for determination of
various direct energy band gap (Egd) values, and (d) plots of energy band gaps (Egd) versus x (wt%) amounts of ZnO for the nanofluids.

mum level) from the beginning of the higher wavelength visible fluid are arranged in a more compact three-dimensional arrange-
region i.e., 800 nm, which evidenced that the arrangement of ment and as a result of this fact the intensities of ZnO diffraction
nanoparticles turned sufficiently dense and are capable to shield peaks were noted significantly increased in the XRD trace. The vari-
the transmission of visible photons. This UV–Vis absorbance result ation of absorbance (Abs.) as a function of ZnO concentration x (wt
on the nanofluid also favours its finding from the XRD trace in %) at fixed wavelengths 800 nm and 600 nm for the GLY/x (wt%)
which all the characteristic ZnO diffraction peaks are noted for ZnO nanofluids are shown in Fig. 4b. This figure explains that the
the 0.30 wt% concentration (see Fig. 2). In other words, we can sug- increase of absorbance is almost linear with the increase of ZnO
gest that a nanofluid having crystalline nanomaterial can display concentration up to 0.05 wt%, and thereafter, it turned saturated.
the characteristic diffraction peaks of the suspended nanoparticles From these interpretations, we conclude that besides the use of
only when the visible range absorption became almost saturated. UV–Vis spectroscopy for the confirmation of nanoparticles suspen-
For the 1.00 wt% concentration nanofluid there is complete satura- sion stability in the nanofluids, the XRD measurements also pro-
tion of the absorption over the entire visible range and also the UV vided analogous and accurate information about the suspended
range which explains that the ZnO nanoparticles in the glycerol particles when they are of crystalline structures.

6
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

The optical material of energy band gap Eg in principle follow


the power law relation (ahm)m = B(hm–Eg) based on the incident
photons energy hm and absorption coefficient a. Here B is a con-
stant related to the extent of band tailing and the variable m value
is taken 2 for the direct energy band gap. This relation is in general
analyzed in the form of Tauc’s plots for the determination of Eg val-
ues of such optical material demonstrated in the literature [20,22].
The Tauc plots ((ahm)2 versus hm) of the GLY/x (wt%) ZnO nanoflu-
ids, as shown in Fig. 4c, were used for the determination of their
direct energy band gap Egd values. The appropriate tangents
(shown by dashed lines) were drawn on these plots and the esti-
mated Egd values (viz. Egd1, Egd2, and Egd3) by the extrapolation
method for these nanofluids are listed in Table 2 and also plotted
against x (wt%) of ZnO in Fig. 4d for the interest of readers. Two dif-
ferent Egd values (Egd1 = 5.92 eV and Egd2 = 5.35 eV) were obtained
for the pure glycerol. The Egd1 represents the characteristic band
gap of glycerol that can be assigned to the n ? r* electronic tran-
sition from the oxygen atoms lone pair electrons. In addition to this
characteristic, a fewer probability of the higher energy electrons of
the oxygen atoms may get excited to the conduction band at the
EM-radiations of energy equal to Egd2 which is slightly lower than
that of the Egd1.
The GLY/ZnO nanofluid having 0.01 wt% concentration of ZnO
exhibited three energy band gaps (Egd1 = 5.43 eV, Egd2 = 5.05 eV,
and Egd3 = 3.15 eV). Out of these three, Egd1 and Egd2 are associated
with the electronic transition related to glycerol molecules
whereas Egd3 represents the energy band gap of ZnO material in
the nanofluid. An appreciable decrease in Egd1 and Egd2 values as
compared to that of the corresponding pure glycerol infers that
the interaction of ZnO nanoparticles with the glycerol molecules
promoted the transition of lone pair electrons of the oxygen atoms
of the glycerol hydroxyl groups. The Egd3 value is found a little
lower than that of the pure ZnO (Eg ffi 3.3 eV) which also evidences
that the glycerol interaction slightly favoured the transitions of
valence band electrons of the ZnO material. When the ZnO concen-
tration increased from 0.01 wt% to 0.02 wt% in the glycerol base
fluid, then only two band gaps viz. Egd2 = 4.34 eV and Egd3 = 3.13 eV
were observed. A huge decrease in Egd2 value as compared to the
pristine glycerol confirms that the molecule-nanoparticle interac-
tions are highly favourable for the glycerol electronic transition
and also to some extent for the ZnO transitions. At 0.05 wt% and
also the higher concentrations (0.30 wt% and 1.00 wt%) of the
ZnO, only Egd3 band gap is obtained which suggests the dominant
behaviour of ZnO electronic transition and these nanofluids of
ZnO concentration  0.05 wt% behave as pristine ZnO material as
far as their optical characteristics are concerned.
Fig. 5. Complex dielectric permittivity (e0 and e00 ) and dielectric loss tangent (tand)
spectra of GLY/x (wt%) ZnO nanofluids (x = 0.00, 0.01, 0.02, 0.05, 0.30, and 1.00), at
25 °C. Inset of (a) shows the enlarged view of e0 spectra in the frequency range from
3.3. Dielectric and electrical spectra of nanofluids 5 kHz to 1 MHz, and inset of (b) shows the variation in e00 values with x (wt%)
amounts of ZnO for the nanofluids at 1 kHz.
3.3.1. Dielectric dispersion
Fig. 5 presents the 20 Hz to 1 MHz frequency range spectra of
dielectric functions (e0 , e00 , and tand = e00 /e0 ) for the GLY/x (wt%) ues showed firstly a gradual decrease and then turned steady state
ZnO nanofluids, at 25 °C. In the 100 Hz to 1 MHz range, the e0 val- in the high frequency range which represents the static dielectric
permittivity es of these nanofluid materials (see inset of Fig. 5a).
The es values noted from the steady-state range of the spectra,
Table 2 attributed to the molecular dipolar polarization which is a measure
Values of direct energy band gaps (Egd1, Egd2, and Egd3) for the GLY/x (wt%) ZnO
of energy storing ability of these dielectric nanofluids, and these
nanofluids (NFs; x = 0.00, 0.01, 0.02, 0.05, 0.30, and 1.00).
values are listed in Table 3. The es value of pure GLY is found in
NFs Egd1 Egd2 Egd3 good agreement with the literature [57,58]. Fig. 5a confirms high
x (wt%) (eV) (eV) (eV)
e0 values of these materials at the initial frequencies of the spectra
0.00 5.92 5.35 – and the rapid decrease with sweeping the frequency from 20 Hz to
0.01 5.43 5.05 3.15 about 100 Hz, which represents a large contribution of the elec-
0.02 – 4.34 3.13
0.05 – – 2.71
trode polarization (EP) process. The effect of the EP process
0.30 – – 2.72 immensely suppresses when the frequency of the harmonic elec-
1.00 – – 2.68 tric field was increased. This type of e0 dispersion in the lower fre-

7
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

Table 3
Values of static dielectric permittivity es, dc electrical conductivity rdc, conductivity relaxation time sr, dc resistance Rdc, and refractive index nD for the GLY/x (wt%) ZnO
nanofluids (NFs; x = 0.00, 0.01, 0.02, 0.05, 0.30, and 1.00), at 25 °C.

NFs es rdc sr Rdc nD


x (wt%) (10–8 S/cm) (ls) (kX)
0.00 41.52 4.22 88.4 215 1.4671
0.01 42.21 3.46 101.6 254 1.4661
0.02 42.05 3.63 99.1 241 1.4660
0.05 41.37 4.10 89.4 214 1.4660
0.30 41.33 6.98 54.6 125 1.4640
1.00 41.20 9.24 34.5 94 1.4631

quency range is a characteristic behaviour of the polar fluids, their


mixtures, and also of NF materials [58–64].
Linear behaviour of e00 versus f plots on the log–log scale
(Fig. 5b) evidenced that the dielectric losses in these nanofluids
are due to ohmic type charge conduction behaviour which is usu-
ally exhibited in the polar liquids over the static permittivity fre-
quency regime [58]. The inset given in this figure confirms that
at a fixed frequency of 1 kHz, the variation in energy losses are rel-
atively low for the nanofluids having a lower concentration of ZnO
(x  0.05 wt%) as compared to that of the pure glycerol, but it
increased significantly for the ZnO amount of 0.30 wt% and then
further increased for 1.00 wt%. In contrast to the linear behaviour
of e00 (f) plots, the tand plots (Fig. 5c) have non-linear rise with
the decrease of frequency below 1 kHz, and the nanofluids having
0.30 wt% and 1.00 wt% amounts of ZnO exhibited a relaxation peak
which can be assigned to the EP relaxation process in these mate-
rials. From the relative variation in the shape of these tand spectra,
it seems that the nanofluids having ZnO concentration less than
0.30 wt% may have a relaxation peak of the EP process below
20 Hz. Furthermore, in the 1 kHz to 1 MHz range, the tand values
are insignificant which also confirms that the dipolar losses are
negligible in these nanofluid materials which credited their appro-
priateness for energy storage over this experimental broader fre-
quency range where the e0 values remained almost frequency
independent.

3.3.2. AC electrical conductivity


The charge carriers concentration and their mobility in princi-
ple govern the electrical conductivity of the nanofluid materials
[49]. The study of electrical conductivity of nanofluids under the
harmonic field has importance for judging their suitability as nan-
Fig. 6. Complex alternating current (ac) electrical conductivity (r0 and r00 ) spectra
odielectric fluid for electrical insulations [7,46,47,65,66]. To under-
for the GLY/x (wt%) ZnO nanofluids (x = 0.00, 0.01, 0.02, 0.05, 0.30, and 1.00), at
stand the charge conduction mechanism in the GLY/x (wt%) ZnO 25 °C. Horizontal solid lines in (a) represent the dc conductivity rdc plateau and the
nanofluids, their r0 and r00 values were determined as a function inset of (b) shows the variation of r00 with x (wt%) amounts of ZnO for the
of frequency, at 25 °C, and these are plotted in Fig. 6. Over the nanofluids at 1 kHz.
20 Hz to several kHz frequency range, the r0 values are found fre-
the charge transport mechanism from dc to ac in these nanofluids
quency independent and therefore, the steady-state r0 values are
which are also observed in several polar fluids [50,60,62].
assigned to the direct current (dc) electrical conductivity rdc of
The r00 spectra of GLY/x (wt%) ZnO nanofluids have appeared
these nanofluids. The rdc values, noted from the steady-state of
overlapped on the logarithmic scale and show linear behaviour
r0 spectra of the GLY/x (wt%) ZnO nanofluids are given in Table 3
with the frequency variation of the applied harmonic electric field.
which are the magnitude of 10–8 S/cm confirming the high insula-
A little deviation in r00 values from the linearity at the lower fre-
tion of these materials. Fig. 6 reveals that there is less change in r0
quencies is due to the contribution of the EP effect [61–63]. Inset
values for the nanofluids containing the ZnO concentration up to
of Fig. 6b evidences that, at a fixed frequency (1 kHz), there is
0.05 wt%, but a significant increase in r0 was noted at 0.30 wt%
anomalous variation in the r00 values with the increase of ZnO
concentration which further increased when the ZnO amount
amount in the glycerol fluid but this change is relatively small
was increased to 1.00 wt% in the glycerol base fluid. These results
for these nanofluids, and therefore all the r00 spectra seem over-
on electrical conductivity suggest that there is either an increase in
lapped on the log–log scale.
charge carriers concentration or a higher probability of the forma-
tion of favourable charge conductive paths which promote the
charge mobility at higher ZnO amounts in these nanofluids. A 3.3.3. Electric modulus spectra
noticeable enhancement in the r0 values with an increase of fre- The study of complex electric modulus spectra (M0 and M00 ) of
quency in the end range of the spectra is an indication of turning the dielectric materials has significance in the confirmation of
8
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

the conductivity relaxation process [60–62,64]. With this fact, the peak value of M00 is half of the maximum M0 value which is an iden-
M0 and M00 spectra as a function of f for the GLY/x (wt%) ZnO tification of the Debye dispersion of these functions [60,61]. More
nanofluids, at 25 °C, are computed and presented in Fig. 7a. These understanding about this behaviour was confirmed by the complex
electric modulus spectra are mainly considered for the determina- plane plot (M00 versus M0 ) drawn for this nanofluid sample in
tion of conductivity relaxation time related to the charge conduc- Fig. 7c. The appeared semicircle shape confirms the Debye type
tion process in these nanofluid materials because these spectra dispersion behaviour of the electric modulus values and it applies
have numerous advantages over the dielectric spectra, as explained to all these GLY/x (wt%) ZnO nanofluids. With this fact, taking the
in the earlier literature [60,61,64]. All these nanofluid materials values of frequency fp corresponding to the M00 peaks, the conduc-
showed the identical shape of their modulus spectra. The M0 spec- tivity relaxation time sr of these nanofluids were determined from
tra of GLY/ZnO nanofluids have initially almost zero values at the relation sr = 1/2pfp [60]. The sr values of these nanofluids, at
lower frequencies which enhanced sharply in the middle fre- 25 °C, are listed in Table 3 which lies in the range of  35 ls to
quency range and lastly attained a steady-state at higher frequen- 102 ls. These sr values are considered in the next section 3.3.5.
cies, whereas the M00 spectra have a sharp peak of conductivity to demonstrate their correlation with the rdc values of the GLY/x
relaxation process in the middle range of experimental frequencies (wt%) ZnO nanofluids.
(i.e., 1 kHz to 10 kHz). Furthermore, one can see that the M0 spectra
of the nanofluids have appeared almost overlapped up to 0.05 wt% 3.3.4. Impedance spectra
of ZnO but thereafter a separation is noted with their shift towards In addition to the electrical conductivity spectra of the nanoflu-
the high frequency side for the 0.30 wt% and 1.00 wt% concentra- ids, their impedance spectra are also characterized from the elec-
tions. This change reflects some difference in conduction mecha- trical insulation point of view, and for the determination of both
nism in these nanofluids having ZnO concentration up to 0.05 wt the conductivity and EP relaxation times of the polar fluids
% and the other systems of higher concentrations  0.30 wt%. [50,63,64]. The dispersion of complex impedance (Z0 and Z00 ) with
The shifting in M00 spectra with ZnO concentration variation is the frequency of harmonic electric field for the GLY/x (wt%) ZnO
found identical to their M0 spectra for these nanofluid materials, nanofluids containing x values 0.01 wt% and 1.0 wt% are provided
however, the height of the relaxation peak is found almost equal in Fig. 8a as representative samples, which reveal that the impe-
for all the nanofluids. dance functions of these materials also have Debye dispersion.
In Fig. 7b, the M0 and M00 values are plotted together on the same These Z00 (f) plots of both the representative nanofluid samples
frequency scale for the GLY based nanofluid of 1.00 wt% ZnO con- exhibited relaxation peaks, and it is noted that the corresponding
centration as a representative sample. This figure realizes that the peak frequencies perfectly match with the values obtained from

Fig. 7. (a) Complex electric modulus (M0 and M00 ) versus frequency (f) plots for the GLY/x (wt%) ZnO nanofluids (x = 0.00, 0.01, 0.02, 0.05, 0.30, and 1.00) at 25 °C, (b)
simultaneous representation of M0 and M00 versus f, and (c) Debye type complex plane plot (M00 versus M0 ) for the GLY/1.00 (wt%) ZnO nanofluid, at 25 °C.

9
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

the peaks of M00 (f) plots. This finding confirms that the conductivity correctly analyzed from the Z00 (f) spectrum of a nanofluid provided
relaxation mechanism can be correctly explained from the impe- that the used experimental frequency range appropriately covers
dance spectra of these nanofluids as reported earlier for a variety both the relaxation processes.
of polar solvents and their mixtures [50,60,64]. In the Z00 (f) spec- The Z00 versus Z0 complex plane plots for all the samples of GLY/x
trum of 1.00 wt% ZnO containing nanofluid sample, a dip has also (wt%) ZnO nanofluids are drawn in Fig. 8b, which are appeared as
appeared near the lower frequency end, which is highlighted in the semicircles of different sizes revealing the Debye dispersion beha-
inset of Fig. 8a. The frequency value (marked as fEP) corresponding viour of the impedance spectra of these materials. The impedance
to the minimum (i.e., dip) is found the same as that noted from its spectroscopic study on numerous high viscosity fluids also con-
tand (f) spectrum. This correlation demonstrates that the relax- firmed their Debye dispersion in this experimental frequency
ation time of the charge conductivity and the EP process can be regime [50,60,64]. The values of dc resistance Rdc for these nano-
fluid materials were taken from the intercept point on the real axis
(i.e., Z0 axis) as marked for 0.30 wt% ZnO containing nanofluid, as a
representative sample in Fig. 8b, and these Rdc values are listed in
Table 3. An anomalous variation in Rdc values of the nanofluids is
noted for lower ZnO concentrations (x  0.05 wt%), and a large
decrease for the higher ZnO concentrations (x  0.30 wt%). Further,
the high Rdc values suggest the appropriateness of these nanofluids
for electrical insulation purposes in addition to their other dielec-
tric applications in soft condensed matter technologies.

3.3.5. Comparison of es, rdc, sr, and nD values of the nanofluids


The dependency of es, rdc, sr, and nD values for the GLY/x (wt%)
ZnO nanofluids on the ZnO concentration are shown in Fig. 9. One
can read from Fig. 9a that the es values increased a little at 0.01 wt%
and 0.02 wt% concentrations of ZnO in these nanofluids as com-
pared to that of pure glycerol, but for the nanofluids of ZnO con-
centrations x  0.30 wt% exhibited a small decrease which infers
that the loading of ZnO nanoparticles influenced the parallel
aligned dipolar ordering of the glycerol molecules in their 3D
supramolecular structural network. At ambient temperature, the
dielectric constant of ZnO material under the influence of 1 MHz
harmonic electric field, is 10.3 [51], which is much lower than that
of the pure glycerol (es = 41.52). Besides this fact, a small increase
in es values of the nanofluids having 0.01 wt% and 0.02 wt% con-
centrations of ZnO was noted which evidence an increase in the
number of parallel aligned dipoles. Whereas, at higher ZnO concen-
trations, it seems that there is a little structural disordering pro-
duced in the dipolar arrangement of the glycerol supramolecular
structure. This finding also favours the XRD results on these
nanofluids demonstrated in the previous section.
It can be seen from Fig. 9 (b and c) that the rdc and sr values
have relatively an inverse variation with the increase of ZnO con-
centration in these GLY/x (wt%) ZnO nanofluids. This result
asserted that the electrical conductivity of these nanofluids is pri-
marily ruled by the conductivity relaxation time, which means the
relative increase in sr value reduced the rdc value of the nanofluid
and vice-versa. Furthermore, the changes in rdc values of these
nanofluids are relatively less for the ZnO concentration up to
0.05 wt%, but thereafter the increase in rdc values is significant
and progressive with the further increase of the ZnO concentration.
Comparatively different behaviour of the rdc and sr values for the
nanofluid samples having 0.30 wt% and 1.00 wt% of ZnO amounts
in comparison to that of the nanofluid up to 0.05 wt% also supports
the saturation in UV–Vis absorbance over the entire visible region.
Fig. 9d explains that initially at the lesser ZnO concentration, there
is a small decrease in refractive index nD of the nanofluids which
appreciably reduced at higher concentration of ZnO. All these
dielectric parameters (es, rdc, sr, and nD) of the GLY/x (wt%) ZnO
Fig. 8. (a) Complex impedance (Z0 and Z00 ) spectra of GLY/x (wt%) ZnO nanofluids nanofluids have distinct behaviour for the ZnO concentration
for x = 0.01 and 1.00, and the inset shows the enlarged view of Z00 spectrum for
1.00 wt% ZnO in the frequency range from 10 Hz to 500 Hz, and (b) Debye type
ranges from 0.01 wt% to 0.05 wt% in comparison to that from
complex plane plots (Z00 versus Z0 ) for the GLY/x (wt%) ZnO nanofluids (x = 0.00, 0.01, 0.30 wt% to 1.00 wt%, which found consistent with the results
0.02, 0.05, 0.30, and 1.00), at 25 °C. obtained from the XRD and UV–Vis characterizations.
10
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

in Table 4 and also plotted in Fig. 10a. This figure infers that the g
values of these nanofluids decreased largely with the increase of
ZnO concentration up to 0.05 wt%, and thereafter a relatively less
decrease is noted at higher concentrations. The lowering in g val-
ues suggest that the ZnO nanoparticles electrostatic interaction
with the glycerol molecules reduced the strength of H-bonded
molecular layers. The decrease in g values was also reported for
the nanofluids of H-bonded ethylene glycol molecules with the
increase of dispersed ZnO nanoparticles concentration [28,67]. This
outcome on the viscosities of the GLY/ZnO nanofluids specifies the
suitability of such decreased viscosity nanofluids in the fluid
pumping systems [26,28,39].

3.4.2. Density
Density q values of the GLY/x (wt%) ZnO nanofluids were deter-
mined at 25 °C and reported in Table 4. The study of q values of the
nanofluids has significance in the computation of some important
parameters like pressure loss, friction factor, Reynold number,
Nusselt number, Prandtl number, etc. which are related to the heat
transfer processes involved in the thermal systems [11,12,26,68].
The q values of these GLY/x (wt%) ZnO nanofluids are plotted as
a function of ZnO concentration in Fig. 10b. It has been noted from

Fig. 9. Plots of es, rdc, sr, and nD with x (wt%) amounts of ZnO for the GLY/x (wt%)
ZnO nanofluids, at 25 °C.

3.4. Thermophysical properties of nanofluids

3.4.1. Viscosity
Viscosity g of a nanofluid has an important role in energy trans-
portation, especially in flow systems defined previously [26,39,40].
Therefore, with the consideration of this fact, the viscosity values
of GLY/x (wt%) ZnO nanofluids as a function of ZnO concentration Fig. 10. Plots of g and q with x (wt%) amounts of ZnO for the GLY/x (wt%) ZnO
were determined at a fixed temperature of 25 °C, and are reported nanofluids, at 25 °C.

Table 4
Values of viscosity g, density q, ultrasonic velocity u, adiabatic compressibility ba, acoustic impedance Za, thermal conductivity k, and viscoacoustic relaxation time sga for the
GLY/x (wt%) ZnO nanofluids (NFs; x = 0.00, 0.01, 0.02, 0.05, 0.30, and 1.00), at 25 °C.

NFs g q u ba  1010 Za  10–6 k (W/mK) sga  1012


x (wt%) (mPa.s) (kg/m3) (m/s) (m2 N1) (kg m2 s1) (s)
0.00 733 1259.7 1920.4 2.15 2.419 0.2847 210
0.01 726 1262.4 1905.0 2.18 2.405 0.2853 211
0.02 713 1261.8 1921.5 2.15 2.425 0.2851 204
0.05 702 1261.9 1925.7 2.14 2.430 0.2858 200
0.30 699 1261.6 1928.1 2.13 2.432 0.2868 199
1.00 692 1267.1 1924.7 2.13 2.439 0.2867 197

11
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

this figure that the q values of these nanofluids are slightly higher ciation. When the ZnO nanoparticles are added to the glycerol fluid
than that of the pure glycerol density for the ZnO amount up to and their amount gradually increased the interaction of these par-
0.30 wt%, which also varies a little and anomalously with the ticles with the glycerol molecules (as revealed from the XRD, UV–
increase of ZnO concentration, but at 1.00 wt% a significant Vis, and dielectric results) alter the H-bonded molecular structural
increase in q value is noted. The increase in the density of nanoflu- packing which probably varies unevenly with the variation of ZnO
ids is expected because the density of ZnO nanopowder is much concentration, and as a result of this fact, the observed density of
higher than that of pure glycerol. But the density of pure glycerol these GLY/x (wt%) ZnO nanofluids varies anomalously.
is strongly related to its strength of H-bonded intermolecular asso-

3.4.3. Ultrasonic velocity and acoustic parameters


Using the measured values of ultrasound wave velocity u, the
acoustic parameters i.e., adiabatic compressibility ba = 1/qu2,
acoustic impedance Za = qu, and viscoacoustic relaxation time
sga = (4/3)gba and also the thermal conductivity k = 2.8ud(NA/
V)2/3KB of the GLY/x (wt%) ZnO nanofluids were determined at
25 °C using the respective relations described in the previous liter-
ature [30,69–75]. The values of these thermophysical parameters
at different ZnO concentrations for the GLY/ZnO nanofluids are
provided in Table 4. In addition to the density and viscosity, these
acoustic parameters are also important to get insight into the phys-
ical behaviour of the nanofluids and to identify their suitability in
advances of soft condensed matter technologies, including the heat
transfer systems.
Fig. 11 shows the variation of all these different acoustic param-
eters as a function of ZnO amount for the GLY/x (wt%) ZnO nanoflu-
ids. It can be seen from this figure that u, ba, and Za values vary
irregularly with the increase of ZnO concentration in these
nanofluids. Further, it can be noted from Fig. 11 that the u and Za
values of these nanofluids are higher than that of the pure glycerol
whereas the ba and sga values are low, except for the nanofluid
sample of 0.01 wt% ZnO concentration. The k values of all the
GLY/x (wt%) ZnO nanofluids are found higher than pure glycerol
and also increased unevenly with the increase of ZnO concentra-
tion up to 0.30 wt%, and thereafter the variation is small at
1.00 wt%. The increased thermal conductivity of different nanoflu-
ids containing ZnO nanoparticles is reported in several studies and
reviews [25,28,29,67]. The appreciable k values of GLY/x (wt%) ZnO
nanofluids confirm them as potential thermal materials to be used
in a variety of heat transfer systems. These results of thermophys-
ical parameters infer that the nanoparticle-molecular interactions
play an important role in the settlement of various acoustic param-
eters of the GLY/x (wt%) ZnO nanofluids.

4. Conclusions

This paper reports in detail the structural, optical, dielectric,


electrical, and thermophysical properties of the GLY/x (wt%) ZnO
nanofluids having ZnO concentrations ranging from 0.01 to
1.00 wt%. The XRD results established that the H-bonded ordered
3D supramolecular structure network of glycerol molecules is
highly efficient to hold the suspended ZnO nanoparticles with high
suspension stability. Additionally, the XRD results evidenced that
due to homogeneity and monodispersity of the suspended ZnO
nanoparticles in the glycerol network emerged as a highly ordered
crystallographic arrangement with excellent continuity when the
ZnO concentration exceeded 0.05 wt%. The characteristics of UV–
Vis absorbance also supported the results of nanoparticles homo-
geneity in these GLY/ZnO nanofluids. For the ZnO
concentration  0.05 wt%, at which the visible range absorption
became saturated, the ZnO nanoparticles showed their characteris-
tic diffraction peaks in these nanofluids. Furthermore, the tunable
absorbance at different wavelengths of the UV–Vis radiations, and
also the regulating behaviour of the energy band gaps with the
Fig. 11. Variation of u, ba, Za, k, and sga with x (wt%) amounts of ZnO for the GLY/x variation of ZnO amount confirmed the appropriateness of these
(wt%) ZnO nanofluids, at 25 °C. GLY/ZnO nanofluids as a liquid optical filter, photosensor, photo-
12
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

voltaic material for solar energy converter, and the UV–Vis blocker [8] M. Moosavi, E.K. Goharshadi, A. Youssefi, Fabrication, characterization, and
measurement of some physicochemical properties of ZnO nanofluids, Int. J.
and shielder.
Heat Fluid Flow 31 (4) (2010) 599–605.
The dependency of dielectric and electrical properties on the [9] R. Jalal, E.K. Goharshadi, M. Abareshi, M. Moosavi, A. Yousefi, P. Nancarrow,
ZnO concentration of the GLY/x (wt%) ZnO nanofluids was thor- ZnO nanofluids: green synthesis, characterization, and antibacterial activity,
oughly characterized. It was explored that a significant increase Mater. Chem. Phys. 121 (2010) 198–201.
[10] L. Huaxu, W. Fuqiang, L. Dong, Z. Jie, T. Jianyu, Optical properties and
in electrical conductivity occurred when the amount of suspended transmittances of ZnO-containing nanofluids in spectral splitting photovoltaic/
ZnO nanoparticles is  0.30 wt% in the glycerol fluid, and it is pre- thermal systems, Int. J. Heat Mass Transf. 128 (2019) 668–678.
dominantly governed by the conductivity relaxation time over the [11] W. Ahemed, Z.Z. Chowdhury, S.N. Kazi, M.R. Johan, N. Akram, C.S. Oon, Effect of
ZnO-water based nanofluids from sonochemical synthesis method on heat
entire ZnO concentration range. The electrical behaviour of these transfer in a circular flow passage, Int. Commun. Heat Mass Transf. 114 (2020)
nanofluid materials was also examined from their complex impe- 104591.
dance spectra which obeyed the Debye dispersion. With the help [12] T. Wen, L. Lu, S. Zhang, H. Zhong, Experimental study and CFD modelling on
the thermal and flow behavior of EG/water ZnO nanofluid in multiport mini
of dielectric and electrical parameters, the significance of these channels, Appl. Therm. Eng. 182 (2021) 116089.
materials to be used as liquid nanodielectric insulators and electri- [13] D.K. Pandey, H.L. Kagdada, A. Materny, D.K. Singh, Hybrid structure of ionic
cal energy storage was confirmed. The viscosity of these materials liquid and ZnO nano clusters for potential application in dye-sensitized solar
cells, J. Mol. Liq. 322 (2021) 114538.
reduced largely with the increase of ZnO amount up to 0.05 wt% [14] D. Damberga, R. Viter, V. Fedorenko, I. Iatsunskyi, E. Coy, O. Graniel, S.
but exhibited a relatively less decrease at higher concentrations. Balme, P. Miele, M. Bechelany, Photoluminescence study of defects in ZnO-
The values of density, ultrasound wave velocity, adiabatic com- coated polyacrylonitrile nanofibers, J. Phys. Chem. C 124 (17) (2020) 9434–
9441.
pressibility, acoustic impedance, and viscoacoustic relaxation time
[15] D. Cheng, Y. Zhang, C. Yan, Z. Deng, X. Tang, G. Cai, X. Wang, Polydopamine-
of the GLY/x (wt%) ZnO nanofluids are reported. The changes in assisted in situ growth of three-dimensional ZnO/Ag nanocomposites on PET
these thermophysical parameters with the variation in ZnO con- films for SERS and catalytic properties, J. Mol. Liq. 338 (2021) 116639.
centration were found to be influenced by the molecule- [16] C. Chen, Z. Li, H. Lin, G. Wang, J. Liao, M. Li, S. Lv, W. Li, Enhanced visible light
photocatalytic performance of ZnO nanowires integrated with CdS and Ag2S,
nanoparticle electrostatic interaction. An increase in thermal con- Dalton Trans. 45 (9) (2016) 3750–3758.
ductivity and appreciably decreased viscosity with the increase [17] M.S. Choi, M.Y. Kim, A. Mirzaei, H.-S. Kim, S.-il Kim, S.-H. Baek, D.W. Chun, C.
of ZnO concentration also realized the suitability of these GLY/ Jin, K.H. Lee. Selective, sensitive, and stable NO2 gas sensor based on porous
ZnO nanosheets, Appl. Surf. Sci. 568 (2021) 150910.
ZnO nanofluids in a variety of heat transfer systems. [18] S. Jiang, K. Lin, M. Cai, ZnO Nanomaterials: Current advancements in
antibacterial mechanisms and applications, Front. Chem. 8 (2020) 580.
[19] D. Ponnamma, J.J. Cabibihan, M. Rajan, S.S. Pethaiah, K. Deshmukh, J.P. Gogoi, S.
CRediT authorship contribution statement K.K. Pasha, M.B. Ahamed, J. Krishnegowda, B.N. Chandrashekar, A.R. Polu, C.
Cheng, Synthesis, optimization and applications of ZnO/polymer
nanocomposites, Mater. Sci. Eng. C 98 (2019) 1210–1240.
R.J. Sengwa: Conceptualization, Supervision, Methodology,
[20] S. Choudhary, R.J. Sengwa, ZnO nanoparticles dispersed PVA–PVP blend matrix
Data curation, Project administration, Resources, Writing – original based high performance flexible nanodielectrics for multifunctional
draft, Writing – review & editing. Mukul Saraswat: Conceptualiza- microelectronic devices, Curr. Appl. Phys. 18 (9) (2018) 1041–1058.
[21] R.J. Sengwa, P. Dhatarwal, Polymer nanocomposites comprising PMMA matrix
tion, Methodology, Data curation, Writing – original draft.
and ZnO, SnO2, and TiO2 nanofillers: a comparative study of structural, optical,
Priyanka Dhatarwal: Conceptualization, Methodology, Data cura- and dielectric properties for multifunctional technological applications, Opt.
tion, Writing – original draft. Mater. 113 (2021) 110837.
[22] P. Dhatarwal, R.J. Sengwa, Poly(vinyl pyrrolidone) matrix and SiO2, Al2O3,
SnO2, ZnO, and TiO2 nanofillers comprise biodegrable nanocomposites of
Declaration of Competing Interest controllable optical properties for optoelectronic applications, Optik 241
(2021) 167215.
[23] P. Dhatarwal, R.J. Sengwa, Structural, dielectric dispersion and relaxation, and
The authors declare that they have no known competing finan- optical properties of multiphase semicrystalline PEO/PMMA/ZnO
cial interests or personal relationships that could have appeared nanocomposites, Compos Interface 28 (8) (2021) 827–842.
[24] H. Li, L. Wang, Y. He, Y. Hu, J. Zhu, B. Jiang, Experimental investigation of
to influence the work reported in this paper. thermal conductivity and viscosity of ethylene glycol based ZnO nanofluids,
Appl. Therm. Eng. 88 (2015) 363–368.
[25] H. Kaya, K. Arslan, N. Eltugral, Experimental investigation of thermal
Acknowledgements performance of an evacuated U-tube solar collector with ZnO/ethylene
glycol-pure water nanofluids, Renew. Energ. 122 (2018) 329–338.
The University Grants Commission, New Delhi, is gratefully [26] M. Gupta, V. Singh, R. Kumar, Z. Said, A review on thermophysical properties of
nanofluids and heat transfer applications, Renew. Sust. Energ. Rev. 74 (2017)
acknowledged for the experimental facilities through SAP DRS-II 638–670.
Project Grant (No. F.530/12/DRS-II/2016(SAP-I)). [27] A.I. Khan, A.V. Arasu, A review of influence of nanoparticle synthesis and
geometrical parameters on thermophysical properties and stability of
nanofluids, Therm. Sci. Eng. Prog. 11 (2019) 334–364.
References [28] K.S. Suganthi, K.S. Rajan, Metal oxide nanofluids: review of formulation,
thermo-physical properties, mechanisms, and heat transfer performance,
Renew. Sust. Energ. Rev. 76 (2017) 226–255.
[1] Ü. Özgür, Y.I. Alivov, C. Liu, A. Teke, M.A. Reshchikov, S. Doğan, V. Avrutin, S.-J.
[29] F. Garoosi, Presenting two new empirical models for calculating the effective
Cho, H. Morkoç, A comprehensive review of ZnO materials and devices, J. Appl.
dynamic viscosity and thermal conductivity of nanofluids, Powder Technol.
Phys. 98 (2005) 041301.
366 (2020) 788–820.
[2] A.B. Djurišić, X. Chen, Y.H. Leung, A.M.C. Ng, ZnO nanostructures: growth,
[30] A.K. Verma, D. Singh, S. Singh, R.R. Yadav, Surfactant-free synthesis and
properties and applications, J. Mater. Chem. 22 (2012) 6526–6535.
experimental analysis of Mn-doped ZnO–glycerol nanofluids: an ultrasonic
[3] Q. Zhang, C.S. Dandeneau, X. Zhou, G. Cao, ZnO nanostructures for dye-
and thermal study, Appl. Phys. A 125 (2019) 253.
sensitized solar cells, Adv. Mater. 21 (41) (2009) 4087–4108.
[31] T. Zhang, C. Liu, Y. Gu, F. Jérôme, Glycerol in energy transportation: a state-of-
[4] J. Theerthagiri, S. Salla, R.A. Senthil, P. Nithyadharseni, A. Madankumar, P.
the-art review, Green Chem. 23 (2021) 7865–7889.
Arunachalam, T. Maiyalagan, H.-S. Kim, A review on ZnO nanostructured
[32] A.F. Cristino, I.A.S. Matias, D.E.N. Bastos, R.G. dos Santos, A.P.C. Ribeiro, L.M.D.R.
materials: energy, environmental and biological applications, Nanotechnology
S. Martins, Glycerol role in nano oxides synthesis and catalysis, Catalysts 10
30 (2019) 392001.
(2020) 1406.
[5] C.B. Ong, L.Y. Ng, A.W. Mohammad, A review of ZnO nanoparticles as solar
[33] C.A. Schwengber, H.J. Alves, R.A. Schaffner, F.A. da Silva, R. Sequinel, V.R. Bach,
photocatalysts: synthesis, mechanisms and applications, Renew. Sust. Energ.
R.J. Ferracin, Overview of glycerol reforming for hydrogen production, Renew.
Rev. 81 (2018) 536–551.
Sustain. Energy Rev. 58 (2016) 259–266.
[6] S.U. Ilyas, M. Narahari, J.T.Y. Theng, R. Pendyala, Experimental evaluation of
[34] S. Goyal, N.B. Hernández, E.W. Cochran, An update on the future prospects of
dispersion behavior, rheology and thermal analysis of functionalized zinc
glycerol polymers, Polym. Int. 70 (7) (2021) 911–917.
oxide-paraffin oil nanofluids, J. Mol. Liq. 294 (2019) 111613.
[35] M. Sharifpur, N. Tshimanga, J.P. Meyer, O. Manca, Experimental investigation
[7] Z.B. Siddique, S. Basu, P. Basak, Dielectric behavior of natural ester based
and model development for thermal conductivity of a-Al2O3-glycerol
mineral oil blend dispersed with TiO2 and ZnO nanoparticles as insulating fluid
nanofluids, Int. Commun. Heat Mass Transf. 85 (2017) 12–22.
for transformers, J. Mol. Liq. 339 (2021) 116825.

13
R.J. Sengwa, M. Saraswat and P. Dhatarwal Journal of Molecular Liquids 355 (2022) 118925

[36] N. Tshimanga, M. Sharifpur, J.P. Meyer, Experimental investigation and model [57] V. Manjula, T.V. Prasad, K. Balakrishna, K.C.J. Raju, T. Vishwam, Influence of
development for thermal conductivity of glycerol–MgO nanofluids, Heat hydrogen bond networks in glycerol/N-methyl-2-pyrrolidone mixtures
Transf. Eng. 37 (18) (2016) 1538–1553. studied by dielectric relaxation spectroscopy, J. Mol. Struct. 1227 (2021)
[37] N.E. Hjerrild, J.A. Scott, R. Amal, R.A. Taylor, Exploring the effects of heat and 129703.
UV exposure on glycerol-based Ag-SiO2 nanofluids for PV/T applications, [58] R.J. Sengwa, V. Khatri, S. Choudhary, S. Sankhla, Temperature dependent static
Renew, Energ. 120 (2018) 266–274. dielectric constant and viscosity behaviour of glycerol–amide binary mixtures:
[38] S. Akilu, A.T. Baheta, A.A. Minea, K.V. Sharma, Rheology and thermal characterization of dominant complex structures in dielectric polarization and
conductivity of non-porous silica (SiO2) in viscous glycerol and ethylene viscous flow processes, J. Mol. Liq. 154 (2010) 117–123.
glycol based nanofluids, Int. Commun. Heat Mass Transf. 88 (2017) 245–253. _
[59] G. Zyła, J. Fal, P. Estellé, Thermophysical and dielectric profiles of ethylene
[39] J.P. Meyer, S.A. Adio, M. Sharifpur, P.N. Nwosu, The viscosity of nanofluids: a glycol based titanium nitride (TiN–EG) nanofluids with various size of
review of the theoretical, empirical, and numerical models, Heat Transf. Eng. particles, Int. J. Heat Mass Transf. 113 (2017) 1189–1199.
37 (2016) 387–421. _
[60] J. Jadzyn, J. Świergiel, Electric relaxational effects induced by displacement
_
[40] J. Sobczak, J.P. Vallejo, J. Traciak, S. Hamze, J. Fal, P. Estellé, L. Lugo, G. Zyła, current in dielectric materials, Ind. Eng. Chem. Res. 51 (2) (2012) 807–813.
Thermophysical profile of ethylene glycol based nanofluids containing two [61] R.J. Sengwa, S. Choudhary, P. Dhatarwal, Characterization of relaxation
types of carbon black nanoparticles with different specific surface areas, J. Mol. processes over static permittivity frequency regime and compliance of the
Liq. 326 (2021) 115255. Stokes-Einstein-Nernst relation in propylene carbonate, J. Mol. Liq. 225 (2017)
[41] I. Wole-Osho, E.C. Okonkwo, D. Kavaz, S. Abbasoglu, An experimental 42–49.
investigation into the effect of particle mixture ratio on specific heat [62] H.P. Vankar, V.A. Rana, Electrode polarization and ionic conduction relaxation
capacity and dynamic viscosity of Al2O3-ZnO hybrid nanofluids, Powder in mixtures of 3-bromoanisole and 1-propanol in the frequency range of 20 Hz
Technol. 363 (2020) 699–716. to 2 MHz at different temperatures, J. Mol. Liq. 254 (2018) 216–225.
[42] T.S. Krishnakumar, A. Sheeba, V. Mahesh, M.J. Prakash, Heat transfer studies on [63] R.J. Sengwa, P. Dhatarwal, S. Choudhary, Static permittivities, viscosities,
ethylene glycol/water nanofluid containing TiO2 nanoparticles, Int. J. Refrig. refractive indices and electrical conductivities of the binary mixtures of
102 (2019) 55–61. acetonitrile with poly(ethylene glycol)-200 at temperatures 288.15–318.15 K,
[43] S.A. Adam, X. Ju, Z. Zhang, J. Lin, M.M. Abd El-Samie, C. Xu, Effect of J. Mol. Liq. 271 (2018) 128–135.
temperature on the stability and optical properties of SiO2-water nanofluids [64] S. Choudhary, P. Dhatarwal, R.J. Sengwa, Characterization of conductivity
for hybrid photovoltaic/thermal applications, Appl. Therm. Eng. 175 (2020) relaxation processes induced by charge dynamics and hydrogen-bond
115394. molecular interactions in binary mixtures of propylene carbonate with
[44] I. Carrillo-Berdugo, P. Estellé, E. Sani, L. Mercatelli, R. Grau-Crespo, D. Zorrilla, J. acetonitrile, J. Mol. Liq. 231 (2017) 491–498.
Navas, Optical and Transport Properties of Metal-Oil Nanofluids for Thermal [65] J.P. Vallejo, G. Zyła,_ L. Ansia, J. Fal, J. Traciak, L. Lugo, Thermophysical,
Solar Industry: Experimental Characterization, Performance Assessment, and rheological and electrical properties of mono and hybrid TiB2/B4C nanofluids
Molecular Dynamics Insights, ACS Sustainable Chem. Eng. 9 (2021) 4194–4205. based on a propylene glycol: water mixture, Powder Technol. 395 (2022) 391–
[45] M.H. Esfe, M.H. Kamyab, M. Valadkhani, Application of nanofluids and fluids in 399.
photovoltaic thermal system: an updated review, Sol. Energy 199 (2020) 796– _
[66] G. Zyła, J. Fal, S. Bikić, M. Wanic, Ethylene glycol based silicon nitride
818. nanofluids: an experimental study on their thermophysical, electrical and
_
[46] D.D. Rosa, M. Wanic, J. Fal, G. Zyła, L. Mercatelli, E. Sani, Optical and dielectric optical properties, Physica E Low Dimens. Syst. Nanostruct. 104 (2018) 82–90.
properties of ethylene glycol-based nanofluids containing nanodiamonds with [67] M. Kole, T.K. Dey, Thermophysical and pool boiling characteristics of ZnO-
various purities, Powder Technol. 356 (2019) 508–516. ethylene glycol nanofluids, Int. J. Therm. Sci. 62 (2012) 61–70.
[47] B. Chakraborty, K.Y. Raj, A.K. Pradhan, B. Chatterjee, S. Chakravorti, S. Dalai, [68] Y. Li, J. Fernández-Seara, K. Du, Á.Á. Pardiñas, L.L. Latas, W. Jiang, Experimental
Investigation of dielectric properties of TiO2 and Al2O3 nanofluids by frequency investigation on heat transfer and pressure drop of ZnO/ethylene glycol-water
domain spectroscopy at different temperatures, J. Mol. Liq. 330 (2021) 115642. nanofluids in transition flow, Appl. Therm. Eng. 93 (2016) 537–548.
[48] M.F. Coelho, M.A. Rivas, E.M. Nogueira, T.P. Iglesias, Permittivity of (40 nm and [69] S. Mukherjee, S. Jana, P.C. Mishra, P. Chaudhuri, S. Chakrabarty, Experimental
80 nm) alumina nanofluids in ethylene glycol at different temperatures, J. investigation on thermo-physical properties and subcooled flow boiling
Chem. Thermodyn. 158 (2021) 106423. performance of Al2O3/water nanofluids in a horizontal tube, Int. J. Therm.
[49] T.P. Iglesias, J.C.R. Reis, On the definition of excess electrical conductivity, J. Sci. 159 (2021) 106581.
Mol. Liq. 344 (2021) 117764. [70] M.N. Rashin, J. Hemalatha, A novel ultrasonic approach to determine thermal
[50] R.J. Sengwa, S. Choudhary, P. Dhatarwal, Dielectric and electrical behaviour conductivity in CuO–ethylene glycol nanofluids, J. Mol. Liq. 197 (2014) 257–
over the static permittivity frequency regime, the refractive indices and 262.
viscosities of PC–PEG binary mixtures, J. Mol. Liq. 252 (2018) 339–350. [71] D. Che˛cińska-Majak, K. Klimaszewski, M. Stańczyk, A. Bald, R.J. Sengwa, S.
[51] R.J. Sengwa, P. Dhatarwal, S. Choudhary, A comparative study of different Choudhary, Static permittivity, density, speed of sound, and refractive index of
metal oxide nanoparticles dispersed PVDF/PEO blend matrix-based advanced 2-propoxyethanol mixtures with water in a wide temperature range, J. Chem.
multifunctional nanodielectrics for flexible electronic devices, Mater. Today Thermodyn. 102 (2016) 164–177.
Commun. 25 (2020) 101380. [72] F. Hevia, V. Alonso, J.A. González, L.F. Sanz, I.G. de la Fuente, J.C. Cobos, Density,
[52] R.J. Sengwa, Comparative dielectric study of mono, di and trihydric alcohols, speed of sound, refractive index and relative permittivity of methanol, propan-
Indian J. Pure Appl. Phys. 41 (2003) 295–300. 1-ol or pentan-1-ol+ aniline liquid mixtures. Application of the Kirkwood-
_
[53] J. Jadzyn, J. Świergiel, Molecular understanding of the viscosity variety within Fröhlich model, J. Mol. Liq. 322 (2021) 114988.
self-assembled hydroxyl liquids, J. Mol. Liq. 293 (2019) 111472. [73] B. Mahmoud, H.P. Rice, L. Mortimer, M. Fairweather, J. Peakall, D. Harbottle,
[54] T. Kusukawa, G. Niwa, T. Sasaki, R. Oosawa, W. Himeno, M. Kato, Observation Acoustic method for determination of the thermal properties of nanofluids,
of a hydrogen-bonded 3D structure of crystalline glycerol, Bull. Chem. Soc. Jpn. Ind. Eng. Chem. Res. 58 (42) (2019) 19719–19731.
86 (3) (2013) 351–353. [74] D. Belhadj, A. Negadi, P. Venkatesu, I. Bahadur, L. Negadi, Density, speed of
[55] S. Hazra, M. Michael, T. Nandi, Investigations on optical and photo-thermal sound, refractive index and related derived/excess properties of binary
conversion characteristics of BN-EG and BN/CB-EG hybrid nanofluids for mixtures (furfural + dimethyl sulfoxide), (furfural + acetonitrile) and
applications in direct absorption solar collectors, Sol. Energy Mater. Sol. Cells (furfural + sulfolane) at different temperatures, J. Mol. Liq. 330 (2021) 115436.
230 (2021) 111245. [75] M. Kumar, M.A. Khan, C.P. Yadav, D.K. Pandey, D. Singh, Ultrasonic
[56] X. Li, Y. Li, L. Kong, F. Li, C. Wang, Density, viscosity, surface tension, excess characterization of binary mixture of 2, 3-dichloroaniline and polyethylene
properties and molecular interaction of diethylene glycol (1) + 1,2- glycols, J. Chem. Thermodyn. 161 (2021) 106557.
propanediamine (2) at atmospheric pressure and T = 293.15 K  318.15 K, J.
Mol. Liq. 345 (2022) 117703.

14

You might also like