Ahmad
Amiri
*a,
S. N.
Kazi
*a,
Mehdi
Shanbedi
*b,
Mohd Nashrul
Mohd Zubir
a,
Hooman
Yarmand
a and
B. T.
Chew
a
aDepartment of Mechanical Engineering, University of Malaya, Kuala Lumpur, Malaysia. E-mail: ahm.amiri@gmail.com; salimnewaz@um.edu.my
bDepartment of Chemical Engineering, Faculty of Engineering, Ferdowsi University of Mashhad, Mashhad, Iran. E-mail: mehdi.shanbedi@stu-mail.um.ac.ir
First published on 30th November 2015
In a one-pot microwave-assisted method, multi-walled carbon nanotubes (MWNT) are functionalized with hexylamine (HA). Based on FT-IR, TGA-DTG, Raman, EDS, CHNS/O and TEM results, the functionalization of MWNT with HA was confirmed . The effect of microwave-assisted functionalized MWNT with HA charged transformer oil at different concentrations was experimentally investigated for electrical, thermal and rheological properties. According to the breakdown voltage, flash point, density, electrical conductivity, thermal parameters and viscosity results of the synthesized transformer oil-based coolants they could be an appropriate alternative for different transformers operating at a nominal voltage less than 170 kV.
In the field of transformer technology, a higher thermal conductivity of TO means a higher heat transfer rate and smaller equipment size, implying a longer lifetime and performance.2 In order to reach the abovementioned purpose, an innovative idea was to improve dielectric coolants.4 These type of fluids are synthesized by adding different nanostructures to TO, with the aim of increasing the insulating and thermal properties of oil.4 A number of studies demonstrated that about 5% growth in the thermal conductivity of TO can result in considerable cost saving to the electrical power generation and transmission distribution infrastructure.5 Despite the idea about the fluids including solid particles with higher thermal conductivities than base fluids being reported a century ago, the innovative concept for synthesizing coolants was introduced by Dr Choi and his team in 1995.6
Recently, researchers have studied the performance of transformer oil-based coolant nanofluids of different nanoparticles such as magnetite nanoparticles, Al2O3.7–9 The results commonly suggested that a nanoparticle-based transformer oil coolant can be utilized to increase the cooling of a power transformer’s core. It is obvious that the thermal conductivity of the nanostructures may play an important role in the suspension’s thermal conductivity. In the field of coolants, a higher extent of thermal conductivity of the nanoparticles means higher thermal conductivity of the coolants as well as the higher heat transfer coefficient.10–12
Among different nanostructures, carbon nanotubes (CNT) with unique thermo-physical, electrical, and mechanical properties can be selected as promising materials.13–16 On the other hand, various applications of CNTs cannot be fully realized due to weak dispersion in solvents and feeble interactions with other materials, which result from strong intertube van der Waals interactions.17,18 To increase the interactivity of CNTs, surface functionalization was suggested as an efficient and common approach. Therefore, different functionalities such as carboxylic groups,2,19 aminoacids17,20,21 etc. have been used for various applications. Aravind et al.19 functionalized CNTs with carboxylic groups to enhance their water dispersibility. Also, Zeinali et al.2 experimentally investigated the effects of carboxylated CNT on TO. Eight effective parameters were investigated and their results reported a drop in the breakdown voltage of TO when the carboxylated CNTs were applied.2
It is obvious that the acidity of transformer oil is a destructive property. According to a previous study,6 the acidity of TO increases with an increase of water content in the oil and it becomes more soluble to the oil, which ultimately deteriorates its insulation property.22,23 So, despite the use of carboxylic groups on CNT surfaces it can also be used to increase dispersibility in different media, thus it cannot be a good choice for TO as an additive. In order to realize excellent dispersion of CNTs in different media, a plethora of carboxylic groups must be decorated on the surface of the CNTs.24 This process first produces numerous defects on the CNT surface and then destroys the main structure creating promising properties with the shortening of their lengths.24
Choi et al.25 synthesized and investigated the dispersions of Al2O3 and AlN nanoparticles in TO with an oleic acid surfactant. In order to synthesize a new kind of transformer oil-based coolant, they applied ceramic nanoparticles to enhance the thermal conductivity of TO and the electric insulation property. An 8% improvement in the thermal conductivity and 20% in the overall heat transfer coefficient were observed at an AlN volume concentration of 0.5%. In addition, the viscosity of the transformer oil-based coolant plays a key role in the performance of other required instruments such as the pump and it has a direct connection with required power through the cooling systems.
Chen et al.26 studied the influence of different basefluids such as silicone oil, glycerol, and water and they observed Newtonian behavior at various nanoparticle concentrations and temperatures. Ahmadi et al.27 investigated the influence of CNT on engine oil coolants and their viscosities at various concentrations. They observed about 13% and 3.3% increases in the flash point and pour point of coolants from 0.1 weight% of CNT over base oil.
In the current study, multi-walled carbon nanotubes (MWNTs) were first functionalized with hexylamine (HA) in one-pot by applying a green procedure under microwave irradiation. In order to realize good dispersibility in TO, a decrease in defects and the removal of the acidic property of TO, the proposed method had no acid treatment phase, which is the general step in previous studies. The functionalization phase was investigated via different characterization methods. In the application phase, the HA functionalized MWNT (MWNT–HA) was added to the basefluid and its properties such as, breakdown voltage, density, flash point, pour point, and transformer performance were investigated at two weight concentrations.
To synthesize the MWNT–HA based transformer oil coolant of different weight concentrations, the HA treated MWNT (MWNT–HA) was sonicated with known amounts of TO for 30 min. Due to the good dispersibility of HA in the oil phase, the MWNT–HA could achieve an appropriate dispersion in TO media. The easily-miscible hexylamine molecules are responsible for better dispersion of MWNT–HA. The synthesized transformer oil-based coolants of different weight concentrations, such as 0.001% and 0.005% were prepared. Some essential properties of TO such as viscosity, density, flash point, pour point, breakdown voltage, electrical conductivity, and thermal conductivity are shown in Table S1 (ESI†).
A wire cylindrical heating element was employed to heat the oil and coolants. 4 PT-100 thermocouples with the accuracy of ±0.1 °C were installed in the reservoir of fluid to measure the average temperature of the fluid. Moreover, four thermocouples were fixed at 4 different sides of the transformer’s wall to calculate the average temperature of the walls. Meanwhile, at an actual distance of 5 cm of the bigger wall, a 55 W blower was employed vertically to cool the big wall and introduce forced convection heat transfer.
The ammeter, thermocouples, and the voltmeter had highest precision data of 0.001 A, 0.1 °C, and 0.1 V, respectively. The uncertainty of the experiments were calculated using the Holman method32 and the maximum uncertainty in calculating the heat transfer coefficient and Nusselt number were less than 2.1 and 5.4%, respectively.
Raman spectroscopy (Renishaw confocal spectrometer at 514 nm), thermogravimetric analysis, TGA, (TGA-167 50 Shimadzu), and transmission electron microscopy, TEM, (HT7700, High-Contrast/High-Resolution Digital TEM) were used to analyze samples. Regarding TGA, the mass loss of samples was measured at a heating rate of 10 °C min−1 in air. The preparation of TEM samples consisted of several steps, such as the sonication of MWNTs in ethanol, dropping the sample on a lacey carbon grid followed by drying under vacuum. Electrical conductivity of the samples was measured using an electrical conductivity meter by using two coaxial electrodes, which are the inner electrode and outer electrode. To conduct the measurement, a specific amount of sample was injected between the two electrodes. When the voltage (V) was applied to the electrodes, there would be a current (I) flowing through the sample. Then, the electrical conductivity of the sample was measured using eqn (1)33
![]() | (1) |
Parameters (h & Nu) were obtained from eqn (2)–(4):
First, the rate of input power was calculated by eqn (2).
Q = VI | (2) |
h = Q/A(Th − Tc) | (3) |
The amount of Nu can be calculated from eqn (4)
Nu = hL/k | (4) |
Meanwhile, the breakdown voltage, flash point, pours point, density, electrical conductivity, thermal conductivity and viscosity were obtained experimentally. A Brookfield LVDV-III rheometer was used to measure the amounts of viscosity. A KD2 thermal analyzer (Decagon Devices, Inc., USA) was used to determine the thermal conductivity of coolants and TO at various temperatures. Also, the thermal conductivity reported in this study is the average of 4 replicated measurements with the error value less than 0.0016.
As discussed above, the covalent functionalization has been performed by the formation of a semi-stable diazonium ion which initiated a radical reaction between the nanotubes and HA chains. FT-IR spectroscopy has been used as one of the best ways to obtain evidence, which is illustrated in Fig. 2 panel (a). In contrast to the pristine sample, the FT-IR spectrum of MWNT–HA demonstrates a couple of peaks at the ranges of 2850–3000 cm−1, indicating the presence of C–H stretching vibrations.35 Also, the peaks at 1396 cm−1 and 1460 cm−1, arise from the bending vibrations of the CH2 group.35 Formation of the aforementioned peaks after functionalization along with the lack of amine peaks can depict the diazonium reaction between the MWNTs and HA.29,34 More importantly, the peak at 1492 cm−1 was associated with the stretching vibrations of CC, which is due to disruption of aromatic π-electrons on the MWNTs surface.
![]() | ||
Fig. 2 (a) FT-IR spectra, (b) Raman spectra, (c) thermogravimetric analysis (TGA) and derivative thermogravimetric (DTG) curves of the pristine and MWNT–HA, (d–f) TEM images of MWNT–HA. |
Raman characterization is a strong measurement for analyzing structure and sp2 and sp3 hybridized carbon atoms in carbon-based materials and functionalization by following alterations in holes.36–39 The Raman spectra of the pristine MWNT, and MWNT–HA are presented in Fig. 2 panel (b). While the pristine MWNTs are weak in terms of D band intensity, the fairly strong D band in the MWNT–HA sample can be seen at 1343 cm−1. The ratio of the intensities of the D-band to that of the G-band (ID/IG) was considered to be the amount of disordered carbon (sp3-hybridized carbon) relative to graphitic carbon (sp2-hybridized carbon).39 In functionalization studies of MWNT, the higher intensity ratio of ID/IG indicates a higher disruption of aromatic π–π electrons, implying partial damage of the graphitic carbon produced by covalent functionalization.39,40 The ID/IG ratio of MWNT–HA (0.91) is relatively higher than that of pristine MWNTs (0.49), which confirmed the successful functionalization via a diazonium reaction under microwave irradiation. A significant increase in ID/IG can also confirm that the present method is completely successful for functionalization of MWNT without acid-treatment phase.
As further evidence, thermogravimetric analysis was conducted to investigate functionalization of MWNTs with HA. TGA is a technique of thermal analysis in which alterations in the structure of materials are measured as a function of temperature. Also, TGA presents results about chemical functionalization of MWNTs with different functional groups.41,42
Fig. 2 panel (c) presents the TGA and derivative thermogravimetric (DTG) curves of the pristine MWNTs and MWNT–HA. It can be seen that the TGA results of the pristine sample illustrate no mass loss up to 500 °C. However, there is an obvious weight loss in the temperature range of 50–150 °C in the HA-MWNT curve. This mass loss was attributed to the functionality of HA as an unstable organic part on the surface of the MWNTs. Also, it is obvious that the DTG curve of the pristine MWNTs shows no phase of degradation up to 500 °C and a noticeable step of mass loss in the temperature range of 500–730 °C is attributed to the bulk degradation temperatures of the main graphitic structures. In addition to the first step of mass loss, the first step of degradation in the DTG curve of the MWNT–HA in the temperature range of 50–150 °C is associated with the bulk degradation temperatures of HA as an unstable organic loaded on the surface of the MWNTs.
The degree of functionalization of modified MWNTs is generally calculated by means of thermogravimetric analysis (TGA). The functionalization degree of MWNT–HA was calculated in order to assess the functionalization yield of the proposed method. Eqn (5) is used to measure the functionalization degree from TGA results:43
![]() | (5) |
Overall, TGA or DTG, FT-IR and Raman results are in good agreement with each other regarding functionalization of MWNTs with HA. Utilizing a microwave method and semi-stable diazonium ions simultaneously provided a very effective route to attack the main graphitic structures of MWNTs.
Element symbol | Pristine MWNTs | MWNT–HA |
---|---|---|
C | 93.1 ± 0.7 | 90.2 ± 0.9 |
O | 2.9 ± 1.1 | 3.2 ± 0.5 |
N | 0.0 | 0.1 ± 0.9 |
Sample | Chemical composition (wt%) | |||
---|---|---|---|---|
C | N | O | H | |
Pristine MWNT | 90.23 | 0.00 | 2.54 | 1.09 |
MWNT–HA | 89.85 | 0.09 | 2.41 | 3.28 |
To check whether the decrease in signal intensity over time is just due to reversible particle settling or gradual irreversible particle aggregation, Fig. 4 panel (c) was plotted. The UV-vis spectrum of MWNT–HA for weight concentrations of 0.001 and 0.005 obtained by performing UV measurements for 30 days on suspensions mechanically shaken every day. The sonication was performed for 5 min at room temperature with a power rating of 390 W. The measurement was again carried out at the peak wavelength of each material to trace the alteration in intensity which can be further used to describe the suspension stability at both weight fractions of 0.001% and 0.005%. It can be seen that both colloidal mixtures show a lower than 1% decrease in the relative concentration as time progressed, indicating that the colloidal suspensions have remained stable for 1 month without irreversible particle aggregation.
The particle size distribution change was analyzed using the dynamic light scattering (DLS) method to check the aggregate size over time. For measurements, the samples were transferred into the folded capillary cell (polycarbonate with gold plated electrodes) for investigation of particle size distribution using a Zetasizer Nano (Malvern Instruments Ltd., United Kingdom) at 25 °C.
Fig. 5 shows graphs of the particle size distributions in the nanofluids measured using the DLS method. These graphs present the size distribution of the particles. Fig. 5 panels (a–f) show the DLS results of MWNT–HA. Although the difference is insignificant, the overall distribution patterns show that the distribution is found to move to a slightly larger particle size when the distribution of the final day is compared to those of previous days, i.e., the particles are insignificantly aggregated. Also, a comparison of particle size distribution of pristine MWNTs (Fig. 5g) and MWNT–HA (Fig. 5a–f) presented a big difference. While the pristine MWNTs completely degraded after 24 h to reach a particle size distribution of 3033 nm, the MWNT–HA material shows a particle size distribution around 400 nm for 35 day, indicating a stable colloidal system in TO media. Therefore, the existence of MWNT–HA in TO provides a great benefit as MWNT–HA provides a stable colloidal system which can enhance the heat transfer properties.
![]() | ||
Fig. 5 Average particle size distribution after (a) 1 day, (b) 7 days, (c) 14 days, (d) 21 days, (e) 28 days, (f) 35 days for MWNT–HA and (g) 1 day for pristine MWNTs. |
Table 3 demonstrates the particle size distributions and zeta potential for pristine MWNT/TO and MWNT–HA/TO colloids over a period of 35 days. Firstly, MWNT–HA/TO shows no big aggregation and coagulation at both concentrations. It can be seen that the particle size distribution for MWNT–HA/TO nanofluids shows a gradual increase in the overall hydrodynamic size, which substantiates the formation of small aggregation, which is in agreement with the UV-vis results. On the other hand, pristine MWNTs mostly settled after 24 h. The above results confirm the critical role of HA molecules which act as a covalent stabilizer to prevent rapid colloidal instability associated with the increase in graphitic domain within the carbon-based structure.
Label in Fig. 5 | Sample | Time (day) | Average particle size distributions (nm) | Polydispersity index (PDI) | Zeta potential (mV) | Mobility (µmcm V−1 s−1) |
---|---|---|---|---|---|---|
a | MWNT–HA | 1 | 367.4 | 0.297 | −52.3 | −4.097 |
b | MWNT–HA | 7 | 407.3 | 0.222 | −49.4 | −3.870 |
c | MWNT–HA | 14 | 401.5 | 0.259 | −46.9 | −3.678 |
d | MWNT–HA | 21 | 411.0 | 0.239 | −38.6 | −3.024 |
e | MWNT–HA | 28 | 400.9 | 0.245 | −46.6 | −3.655 |
f | MWNT–HA | 35 | 389.3 | 0.251 | −50.4 | −3.951 |
g | Pristine MWNTs | 1 | 3033.0 | 0.655 | −23.6 | −1.853 |
According to stabilization theory, the electrostatic repulsions between particles increases if the zeta potential has a high absolute value which then leads to good stability of the suspensions. Particles with a high surface charge tend not to agglomerate, since contact is opposed. Typically accepted zeta-potential values are summarized in Table S2.†44
Table 3 also shows the zeta-potential and the polydispersity index (PDI) for pristine MWNTs and MWNT–HA/TO at their natural pH. The zeta-potential and polydispersity index (PDI) are commonly utilized as an index of the magnitude of electrostatic interaction between colloidal particles and thus can be considered as a measure of the colloidal stability of the solution. According to Table S2,† the zeta potential must be as large as possible (positively or negatively) to make a common repulsive force between the particles.45 It can be seen that after functionalization with HA, MWNT–HA shows a more negatively charged zeta-potential and is around −45 mV over a period of 35 days. The zeta potential results of MWNT–HA in TO media suggest an appropriate stability over a period of 35 days at 25 °C. Indeed, the zeta-potential gradually show some fluctuations over a period of 35 days, despite remaining mostly stable with time.45
Fig. 7 shows the thermal conductivity plot of MWNT–HA/TO coolant as a function of temperature and concentration. Two different weight concentrations of 0.001% and 0.005% are considered and the variation of thermal conductivity with concentration and temperature are studied. To avoid a significant increase in effective viscosity and electrical behavior, MWNT–HA at low concentrations is considered in the present study. Fig. 7 clearly demonstrates that the thermal conductivity of MWNT–HA/TO coolant is higher than that of pure TO. It can also be seen that the thermal conductivity of the TO decreases as the temperature increases, where the MWNT–HA/TO coolants at both concentrations show an upward trends. In a similar study, Beheshti et al.2 showed that the thermal conductivity of MWNT-COOH/TO coolants increased up to 60 °C and it decreased considerably at higher temperatures. This drop may be attributed to the settlement of MWNT, which can result from the disruption of carboxyl groups on the MWNT surface at high temperature. In contrast with a previous study,10 the problem of accumulation at 60 °C is completely eradicated by the functionalization of MWNT with HA. On the other hand, some researchers in the field of heat transfer concluded that the nanoparticles are governed by the continuous irregular reciprocating motion and phenomena such as Brownian motion, thermophoresis and diffusiophoresis which could influence thermal conductivity of the coolant.4 Above all, it is a big success that the trend of thermal conductivity illustrates a rising trend over the temperature range of 30–80 °C.
![]() | ||
Fig. 7 Thermal conductivity plot of pure TO and MWNT–HA/TO coolant at different weight concentrations. |
On the other hand, the obtained results confirm that temperature plays a key role in increasing the thermal conductivity of MWNT–HA/TO coolant, which cannot be neglected. For coolants loaded with MWNTs, the vital role in heat transfer enhancement is played by the nanoparticles.50,51 Liquid molecules generate layers around the MWNTs, thereby altering the local ordering of the liquid layer at the interface region. Thus, the liquid layer at the interface would exhibit a higher thermal conductivity compared to the base-fluid. Differences in thermal conductivity enhancement may be attributable to differences of the thermal boundary resistance around the nanoparticles occurring for different base fluids.52 The higher slope in thermal conductivity of MWNT–HA/TO coolants with higher concentration is due to a higher rate of formation of the surface nanolayers.10 In addition, the role of Brownian motion of particles in nanofluids may be an important parameter in determining the thermal conductivity enhancement and is also an important factor, in particular for cases with significant change in viscosity with temperature, which is certainly the case for TO.52 According to recent studies,19,53,54 increasing the local ordering of the liquid layers at the interface region is one of the main reasons for thermal conductivity enhancement. Thus, the resistance for heat removal is decreased by thinning of the thermal boundary layer via MWNT loading in basefluids and preparing nanolayers with higher thermal conductivity than the bulk liquid.
A summary of experimental studies on the thermal conductivity of TO-based nanofluids is listed in Table S3.† It is worth mentioning that most of the recent studies focused on suspensions with high nanoparticle concentration, which increase the possibility of sediment. Compared with recent studies,2 the present work reaches higher enhancement in thermal conductivity under similar experimental conditions. For example, Patel et al.55 obtained 10, 11.5, 14 and 17% enhancements in the thermal conductivities of TO-based nanofluids by the addition of Al2O3 at 20, 30, 40 and 50 °C, respectively. Such enhancements are obtained for 3% Al2O3 loading, while we work at very low weight fractions of 0.001 and 0.005. In fact, the strategy of the present study is the utilization of low concentration of highly-soluble MWNT–HA in TO to avoid sediment, since the bulk fluid in the transformer is motionless.
Natural convection heat transfer occurred by the buoyancy force, which results from the density deviations formed by temperature variations in the layers of fluid. This type of heat transfer generates in the absence of an external source for the movement of the bulk fluid. According to a previous study,56 when a fluid is in contact with a surface with higher temperature, the molecules of the fluid separate and form a layer of lower density. As a result, the layers of fluid with lower density are displaced by the cooler layers of fluid, thus the cooler layers of fluid sink. It is noteworthy that the nanostructure concentrations and thermal conductivity can also influence natural convection in coolants.2
The natural convection heat transfer coefficient and Nusselt number of pure TO and MWNT–HA/TO coolants versus input power are illustrated in Fig. 8 panels (a) and (b), respectively. It can be seen that similar upward trends are obvious in the presence of the pure TO and MWNT–HA altered coolants in both figures. It is obvious that the heat transfer coefficient is a function of temperature. According to the results, the input power increases with an increase of the average temperature of the bulk fluid in the transformer, which is implying a higher heat transfer coefficient. As the thermal conductivity of the coolants at different temperatures is applied for calculation, the evaluated Nusselt number results show some weak fluctuations.
Forced convection is a mechanism of heat transfer in which fluid motion is created via an external source such as a fan and pump. The amount of forced convection heat transfer coefficient and Nusselt number versus input power are respectively shown in Fig. 8 panels (c) and (d) for pure oil and MWNT–HA/TO coolants. The natural convection heat transfer coefficient and Nusselt number increase with an increase of input power and concentration, the forced convection heat transfer also shows similar results of enhancement with an increase of power and concentration. Surprisingly, the amount of thermal conductivity, heat transfer coefficient and/or Nusselt number show significant enhancements in the presence of MWNT–HA, which is attributed to the good dispersibility of MWNT in TO as one of the reasons.
The enhancement of convective heat transfer can be attributed to a reduction of the thermal boundary layer thickness. Aravind et al.50,57 showed that carbon nanomaterials such as MWNTs have a tendency to decrease the thermal boundary layer thickness, which lowers the difference between temperatures of the bulk fluid and wall in the transformer and subsequently enhances the convective heat transfer coefficient. To clarify this, the heat transfer coefficient can be approximately modeled as k/δt, where δt represents the thermal boundary layer thickness. So, to increase the heat transfer coefficient, either k should be increased or δt should be decreased or both. As a future comparison, Table S4† shows a comparison of the natural convection and forced convection heat transfer coefficients of nanofluids with different nanoparticle loading. The convective heat transfer coefficient of the MWNT–HA/TO nanofluid was compared with those of other nanoparticle-based TO nanofluids reported in the literature. Obviously, the present samples had a relatively high natural convection and forced convection heat transfer coefficients in comparison to other nanoparticles loading in TO at similar weight fraction, in particular, as compared with MWNT-COOH/TO nanofluids2 under similar experimental conditions. Our results show natural convection and forced convection heat transfer coefficient enhancements of 23 and 28% at weight fraction of 0.005, which demonstrate a significant increase at a very low concentration.
Breakdown voltage of pure TO with and without sonication and MWNT–HA/TO at both weight concentrations of 0.001 and 0.005 are plotted in Fig. 9 and the mean breakdown voltages with standard deviation are listed in Table 4. In order to measure the dielectric breakdown voltage with different weight fractions of MWNT–HA in the colloidal suspension, the experiment was repeated 60 times for each of the weight concentrations according to the ASTM D-92 standard using the dielectric breakdown measurement device, Mugger’s automatic laboratory oil tester. All experiments were performed using mushroom electrodes set at a gap of 1 mm and at room temperature. The dielectric breakdown voltage of all samples was measured 1 day after preparing. The most influential factor affecting the performance of the dielectric strength of the transformer oil is the degradation caused by water and other contaminants,59 which with the sonication process could also introduce moisture to the oil. Fig. 9 panels (a) and (b) show the dielectric breakdown voltage of pure oil with and without 30 min sonication (time for performing the diazonium reaction) to investigate the effect of the sonication process on TO performance. Results suggest that the sonication has an insignificant effect on the dielectric breakdown voltage (∼0.4 kV), which is obtained by doing just 30 min sonication using the Bioruptor (4 s “ON”, 4 s “OFF”).
![]() | ||
Fig. 9 The dielectric breakdown voltage of pure oil (a) with and (b) without 30 min sonication and (c) MWNT–HA/TO at (c) 0.001 and (d) 0.005. |
Samples | Mean breakdown voltage (kV) | S.D. (kV) |
---|---|---|
Pure TO | 58.03166667 | 1.137174 |
Sonicated pure TO | 57.65333333 | 1.311039 |
MWNT–HA/TO (0.001%) | 54.09833333 | 1.60521 |
MWNT–HA/TO (0.005%) | 52.06666667 | 2.47283 |
Fig. 9 panels (c) and (d) show the experimental results with the various weight concentrations of MWNT–HA. Generally, increasing the MWNT–HA concentration leads to a mild decrease of the dielectric breakdown voltage. Again, the mean breakdown voltages and standard deviations for the above-mentioned samples are given in Table 4. It was found that breakdown voltages level-off with the rising test number, no upward or downward trend was obtained in experiments, which owes to the effective energy control of the test equipment. With the increase of weight concentration of MWNT–HA in TO from 0.001 to 0.005%, the dielectric strength decreases from 54.098 kV to 52.066 kV. The standard deviations of both samples are larger than that of transformer oil, indicating that the breakdown voltage of the samples is less predictable. The MWNT–HA loading in TO media leads to a small drop in the breakdown voltage, which is mainly caused by the MWNTs.
A seta semi-automatic cleveland open cup flash point tester was used to measure the flash point. The flash points of the pure TO with and without sonication and also MWNT–HA/TO at two different weight fractions were measured using the basis of American Society for Testing and Materials (ASTM) D-92.61 The trend of changes of flash point as a function of MWNT–HA concentration is shown in Fig. 11. Firstly, the flash point was tested for the pure oil with and without sonication to investigate the effect of sonication. Variation of the flash point in the presence of sonication and lack of sonication is shown in Fig. 11, indicating an increase of around one degree in the volatility of the TO when performing sonication for 30 min. Also, the variation of the flash point as a function of MWNT–HA concentration is indicated with an increase in the volatility of the TO with MWNT addition. It is obvious that a higher flash point temperature is required for safer handling of TO. The rate of increase in the flash point of the MWNT–HA/TO coolant at the 0.001 weight fraction with respect to the TO is 4.69%, and the highest amount of increase is related to the 0.005 weigh fraction sample, which is 7.86%.
Similar to other oil-based nanofluids, the rheological behavior of MWNT–HA/TO (Fig. 12a) shows two typical characteristics: (i) an enhancement of viscosity with increasing weight fraction of MWNT–HA, and (ii) a decrease in viscosity with increasing temperature, which is due to weakening of the intermolecular forces of the fluid itself. Fig. 12b illustrates the amount of enhancement in the viscosity with MWNT–HA loading. It shows the enhancement in the MWNT–HA/oil viscosity is less than 10% for both weight fractions and all temperatures. It can be concluded that the enhancement in the MWNT–HA/oil viscosity is insignificant with an increase in weight concentration, which can be attributed to the low weight fraction of the MWNT–HA in the oil. In addition, in agreement with Ko et al.,62 the viscosity decreases with an increase in temperature.
The viscosity of pure TO and MWNT–HA/TO coolant as a function of the shear rate for different concentrations are presented in Fig. S1 (ESI†). The effective viscosities of pure transformer oil and treated coolants were measured for the shear rate range of 35–235 s−1. Firstly, MWNT–HA/TO and pure TO have a linear shear stress/shear rate relationship, which can be categorized as a Newtonian fluid.2,63 It may be seen for all temperatures as the viscosity of the TO is independent of the shear strain rate, indicating Newtonian behavior. Also, after the addition of MWNT–HA, the samples showed Newtonian behavior.
We believe that since TO exhibits Newtonian behavior, it dominates the rheological properties and the whole mixture behaves like a Newtonian fluid with low concentrations of MWNT–HA. Due to some fluctuations and the significant error in measurement, we were not able to identify the exact amount of viscosity for a shear rate of around zero. Based on some recent research,2,63,64 TO is categorized as a Newtonian fluid.
Nu | Nusselt number |
T c | Average temperature of walls, °C |
H | Heat transfer coefficient, W m−2 K−1 |
K | Thermal conductivity, W m−1 K−1 |
L | Length, m |
T h | Average temperature of oil, °C |
Q | Input power, W |
A | Heat transfer area |
V | Voltage, V |
I | Current, A |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra17687e |
This journal is © The Royal Society of Chemistry 2015 |