Next Article in Journal
Intelligent Identification of MoS2 Nanostructures with Hyperspectral Imaging by 3D-CNN
Next Article in Special Issue
Nanofluid Development Using Silver Nanoparticles and Organic-Luminescent Molecules for Solar-Thermal and Hybrid Photovoltaic-Thermal Applications
Previous Article in Journal
Sulfonated Pentablock Copolymer Membranes and Graphene Oxide Addition for Efficient Removal of Metal Ions from Water
Previous Article in Special Issue
Tribological Behavior of Nanolubricants Based on Coated Magnetic Nanoparticles and Trimethylolpropane Trioleate Base Oil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermal Conductivity Performance of 2D h-BN/MoS2/-Hybrid Nanostructures Used on Natural and Synthetic Esters

by
Jaime Taha-Tijerina
1,*,
Hélio Ribeiro
2,3,
Karla Aviña
1,
Juan Manuel Martínez
1,
Anna Paula Godoy
2,
Josué Marciano de Oliveira Cremonezzi
2,
Milene Adriane Luciano
3,
Marcos Antônio Gimenes Benega
2,
Ricardo Jorge Espanhol Andrade
2,
Guilhermino José Macedo Fechine
2,
Ganguli Babu
4 and
Samuel Castro
4
1
Departamento de Ingeniería, Universidad de Monterrey, Av. Ignacio Morones Prieto 4500 Pte., San Pedro Garza García, NL 66238, Mexico
2
Mackenzie Institute for Research in Graphene and Nanotechnologies-MackGraphe, Universidade Presbiteriana Mackenzie, Rua da Consolacão, 896, Sao Paulo, SP CEP 01302-907, Brazil
3
Departamento de Química, Universidade Federal de Minas Gerais, Av. Antonio Carlos, 6627, Pampulha, Belo Horizonte, MG CEP 31270-901, Brazil
4
Materials Science and NanoEngineering, Rice University, 6100 Main St., Houston, TX 77005, USA
*
Author to whom correspondence should be addressed.
Submission received: 19 May 2020 / Revised: 5 June 2020 / Accepted: 9 June 2020 / Published: 12 June 2020

Abstract

:
In this paper, the thermal conductivity behavior of synthetic and natural esters reinforced with 2D nanostructures—single hexagonal boron nitride (h-BN), single molybdenum disulfide (MoS2), and hybrid h-BN/MOS2—were studied and compared to each other. As a basis for the synthesis of nanofluids, three biodegradable insulating lubricants were used: FR3TM and VG-100 were used as natural esters and MIDEL 7131 as a synthetic ester. Two-dimensional nanosheets of h-BN, MoS2, and their hybrid nanofillers (50/50 ratio percent) were incorporated into matrix lubricants without surfactants or additives. Nanofluids were prepared at 0.01, 0.05, 0.10, 0.15, and 0.25 weight percent of filler fraction. The experimental results revealed improvements in thermal conductivity in the range of 20–32% at 323 K with the addition of 2D nanostructures, and a synergistic behavior was observed for the hybrid h-BN/MoS2 nanostructures.

Graphical Abstract

1. Introduction

For over a century, fossil fuels have been the major contributor to the energy sector. Challenges to mitigate pollution, global warming, and more will be critical. This, together with the imminent scarcity of oil reserves, increases in oil prices, rises in lubricant disposal costs, as examples, will promote the development of alternative sources of energy as well as novel technologies to attend the current and future world energy needs with sustainability and efficiency [1,2,3].
Petroleum-based fluids and lubricants are applied as lubricants and dielectric and coolant material in electrical or electronic devices such as power systems (transformers), machinery, and the automotive and biomedical industries, among others [4,5,6,7]. Paraffinic and naphthenic oils are examples of mediums used for these applications [8]. Nevertheless, these materials exacerbate environmental issues due to their eco-toxicity, incomplete biodegradability, and highly probable carcinogenic characteristics [9,10]. Some of the main limitations to be solved regarding these applied fluids are how to dispose of them and prevent products to have a significant negative impact on health and environmental-related consequences, i.e., skin and respiratory diseases or in case of a spillage [11,12,13,14].
More often, biodegradable synthetic products are used in environmentally-sensitive areas such as electric, electronic, automotive, agricultural, and biomedical industries [4,5,6,7,15,16,17]. The increased environmental awareness is a primary driving force for the novel technological developments in these fields [11,18,19]. Synthetic and natural esters are an alternate to petroleum-based oils because they have non-toxic attributes and are renewable and environmentally-friendly. Moreover, synthetic esters are highly resistant to oxidation [20,21] and have high fire points. It has been proven that these fluids reduce hydrocarbon and carbon monoxide emission levels, which is the reason why their use has been increased recently in the industry. Furthermore, they possess great characteristics such as excellent lubricity, high thermal conductivity, high flash and fire points, and biodegradability, as compared to petroleum-based fluids. Nevertheless, despite specific environmental benefits of natural or vegetable fluids, their characteristics do not always accomplish the required properties [22].
Due to their inherent risks, such as fire and pollution (caused by leakage and spills), natural alternatives derived from triglycerides in vegetable oils and animal fat are being sought [23,24,25,26]. Although a great sustainable feedstock for novel materials, these are susceptible to variations in their compositions due to different processing, crop conditions, and climate and environment, among others [27]. In general, the natural esters are presented as fatty acids, containing the ester function and a long aliphatic chain that dictate some chemical and mechanical properties of the oils [28]. The oils based on saturated fatty acids are more chemically stable and very viscous. Unsaturated (mono, di, tri) can be much less viscous but are prone to oxidation [28], which leads to an increase in viscosity and consequently affect thermal, tribological, and other properties of the oil [29].
In industrial and thermal systems, heat dissipation is a paramount phenomenon. Devices, machines, and equipment miniaturization design require better performance (thermal dissipation efficiency) of fluids and lubricants due to environmental challenges and global market competition [22,30]. Several heat transfer fluids have been under research with the aid of nanoreinforcements to improve their performance [4,5,31,32,33,34]. In recent years, environmental awareness has gained increased attention and is a driving force towards the novel technology and innovation developments in diverse fields. Biodegradability has become a key design factor and request for products. In this sense, a good candidate to replace petroleum-based oils are synthetic or natural esters based on vegetable oils such as sunflower, coconut, corn, soybean, rapeseed, olive, palm oil, and others [18,19,30,35,36]. These materials are an alternative to mineral oils because they have non-toxic properties and are renewable and environmentally-friendly.
Two-dimensional nanomaterials appear as an attractive alternative due to the high surface area they possess for heat transfer. Among these 2D nanomaterials are hexagonal boron nitride (h-BN) and molybdenum disulfide (MoS2), which are suitable for development of high-performance composites applied in several industrial and technological sectors due to their exceptional electrical, thermal, and mechanical properties [37,38,39]. In this sense, hexagonal boron nitride (h-BN) is investigated in this study, since it is applied to several substrates to reduce viscosity and increase thermal conductivity [7,24,37,38,39,40,41]. The capability of h-BN to improve the thermal conductivity of solid nanocomposites based on epoxy and polyurethane was previously reported by Ribeiro et al. [37,38,39]. Nonetheless, the application of h-BN to mineral oils, natural esters, and other fluids, is being studied [5,7,26,42]. Taha-Tijerina et al. [7] showed that h-BN was able to improve the thermal conductivity of a mineral oil, and—differently to graphene—functioned as an electric insulator rather than a conductor. Such phenomena is due to the ability of some nanoparticles, such as F3O4 [25], h-BN [25], Al2O3 [43], SPION [44], CuO [43], and BaTiO3 [45] to trap moving electrons. Du et al. [42] applied boron nitride (BN) to a vegetable oil and found that the thermal conductivity was increased in comparison with the pure oil. The use of BN allowed the lubricant to be less viscous at higher temperatures which further enhanced the temperature dissipation. Salehirad and Nikje [5] reported similar results applying h-BN to a mineral oil: increased thermal conductivity, lower viscosity, and better insulation capability when compared to the pure mineral oil. Li et al. [46] developed ethylene glycol/boron nitride nanofluids, varying the size of the nanoparticles. It was demonstrated that nanofluids containing larger nanoparticles have higher improvements in thermal conductivity than smaller nanoparticles. Ilhan et al. [47] reported an enhancement of thermal conductivity for h-BN nanofluids of up to 26%, 22% and 16% at 3 vol.% for water, water/ethylene glycol mixture, and ethylene-glycol-based nanofluids, respectively.
Molybdenum disulfide (MoS2) involves very optimal thermal and chemical stabilities. It demonstrates applications as a catalyst and lubricant due to its unique properties like anisotropy, photocorrosion resistance and chemical inertness [48,49]. MoS2 nanostructures have good thermo-physical features and great anti-friction properties, making it another promising material as reinforcement for the cooling purposes of nanofluids. Diverse factors such as the composition and loading of nanoparticles, suspension stability, base fluid composition, nanofluids preparation method, surface modifier, and surfactant application may influence the thermo-physical properties of nanofluids, including viscosity and thermal transport [50,51,52,53,54,55]. Su et al. [52] investigated stability and thermal transport behavior of MoS2, water-based, and oil-based nanofluids at various concentrations (0.01 wt.% to 0.5 wt.%). It was observed that by increasing the filler fraction, the thermal conductivity increased. Furthermore, the thermal conductivity improvement in MoS2 and water-based nanofluids was higher than for oil-based nanofluids, which according to Su et al. was due to a better dispersion and suspension stability achieved on water-based fluid. Zeng et al. found that thermal conductivity of dibenzyl-toluene-based MoS2 nanofluids increased from 17.5 to 37.5% as the temperature increased from 40 to 180 °C [53].
In this work, the effect of incorporation and homogeneous dispersion of single h-BN, MoS2, and hybrid h-BN/MoS2 mixture at various concentrations (by weight) within synthetic and natural esters, aims to improve their thermal conductivity performance, which is determined at various evaluating temperatures (up to 323 K). The contributions of this work rely on the aspect of biodegradable materials suitable for industrial applications, such as energy and metal-mechanic field. The nanofluids with only h-BN or MoS2 nanofiller components were compared with their hybrid h-BN/MoS2 combination.

2. Materials and Methods

In our study, synthetic esters Midel 7131 (M&I Materials-Manchester, UK) and natural esters—Envirotemp® FR3™ (Cargill Industrial Specialties–Minneapolis, MN, USA) and VG-100 (Prolec GE International, Apodaca, México) [56] (Table 1)—were used as base materials to develop different nanofluids with h-BN, MoS2, and h-BN/MoS2 nanostructures (Table 2) at various concentrations of 0.01, 0.05, 0.10, 0.15, and 0.25 wt.%. The nanofluids with the hybrid composition h-BN/MoS2 were used at a mass ratio of 1:1 between h-BN and MoS2. Nanofillers were obtained by wet exfoliation in tetrahydrofuran (THF), using probe ultrasound for 5 h at an amplitude of 35%. The exfoliated material was filtered and dried in a vacuum oven for 24 h at 80 °C. To prepare the nanofluids, a two-step method to homogeneously disperse the 2D nanostructures within the esters was used. Glass containers were prepared with each set of fluids at different filler fractions: 0.01, 0.05, 0.10, 0.15, and 0.25 wt.% of the nanostructures for each set of fluid. Extended water bath sonication (~4–5 h) was used (Branson ultrasonic homogenizer model 5510-Danbury, CT, USA, 40 kHz). To avoid the nanostructures’ agglomeration and quick sedimentation, the temperature of the water bath was maintained constant at room temperature (24 °C). Samples were maintained on a drawer for at least 2 weeks without significant sedimentation. Experimental evaluations were performed after 3 days of sample preparation. This process assists in obtaining the stable homogeneous nanofluids, which are further evaluated.
Nanomaterials’ morphology and thickness were analyzed by atomic force microscopy (AFM) using an Asylum Research MFP-3D-AS (Belo Horizonte, MG, Brazil), operated in contact mode. An Olympus AC240TS probe (Belo Horizonte, MG, Brazil) and a 70 kHz resonant frequency were used. Specimens were prepared by ultrasonically dispersing the 2D nanostructures in isopropanol (IPA) and water mixture. This was followed by the deposition of dispersion drops on mica, which was then dried. A Quanta™ 200 FEG-FEI microscope (Belo Horizonte, MG, Brazil), operated under a vacuum with a 30.0 kV accelerating voltage, was used to obtain SEM images. The 2D nanostructures were prepared by dispersing the material in IPA/water mixture using an ultrasonic water bath (2 h); subsequently, the sample was cast onto a silicon substrate and evaporated in air at room temperature. The chemical composition of the 2D nanostructures was investigated (EDS) using a FEI QUANTA™ 200 microscope (Belo Horizonte, MG, Brazil). To evaluate the homogeneity of the mixture between them, EDS mapping was performed.
SEM was also performed to characterize the microstructure of the h-BN/MoS2 mixture. Figure 1a–d shows SEM images of h-BN, MoS2, and h-BN/MoS2 nanosheets with different morphologies and sizes. Figure 1a shows a typical SEM image of a large-size h-BN sample, which consists of thin and crumpled sheets with a lateral dimension of about ~5–20 µm. Figure 1b shows the morphology of MoS2. Figure 1c,d shows a typical SEM image of the h-BN/MoS2 mixture, which keeps its features between its single components. Finally, Figure 1e–f presents the AFM image of h-BN/MoS2 and its corresponding height profile (thickness approximately 2.3 nm).
The h-BN/MoS2 mixture can be also confirmed by EDS maps obtained from the SEM image (Figure 2a), which displays the homogeneity of the mixture of h-BN and MoS2. Elements B, N, Mo, and S are uniformly distributed in the whole area according to Figure 2b–e.
Raman spectra of the h-BN nanostructures were obtained by a Raman Confocal Spectrometer WITec, model Alpha 300R (Sao Paulo, SP, Brazil), using a laser (532 nm) as the excitation source. The MoS2 nanosheets on the Raman spectrum showed the main characteristic bands of this nanostructure (Figure 3a). The first one, in 382 cm−1, is attributed to E2g vibrational mode, and the second (in 410 cm−1) is related to A1g mode. On the other hand, these modes correspond to the in-plane vibrations of sulfur atoms in one direction and molybdenum atoms in another one, and to out-of-plane vibrations (A1g) of sulfur atoms. For the sample containing h-BN, a spectrum with only one intense band in 1372 cm−1, in relation to the E2g vibrational mode, was observed. In relation to the mixture of these nanofillers, the same three bands from MoS2 and h-BN nanosheets were presented, without another additional band or significative shifting from original positions related from the original nanomaterials. [37,63].
X-ray diffraction (XRD) evaluation was performed using a Shimadzu XRD-7000 system (Sao Paulo, SP, Brazil) employing Cu Kα radiation (λ = 1.5418 Å). The XRD patterns were obtained in the 2θ range between 4° and 80° at a scanning rate of 4° min−1. These results were found to be fully matched with a JCPDS 37–1492 card. The XRD spectrum of MoS2 presented diffraction peaks in equal to 14.4°, 32.8°, 39.6°, 49.8°, and 60.3° related to (002), (100), (103), (105), and (110) planes, respectively. The diffractogram of h-BN depicted peaks in equal to 26.7° and 41.7° from the (002) and (100) planes (Figure 3b) [38]. The mixture presented the same peaks related to both nanofillers, without additional peaks or significative shifting from the original MoS2 or h-BN diffraction spectra.
Thermal conductivity measurements were carried out on nanofluids at various nanostructure concentrations according to the transient hot-wire (THW) technique, with a KD2 Pro device. A temperature-dependence scan was also performed, maintaining a thermal equilibrium for at least 10 min before each set of evaluations. A minimum of 7 measurements were taken for each set of experiments to report average values with standard deviation as error bars.

3. Results

Improvements in thermal conductivity in synthetic and natural ester systems have been obtained for different types of nanofillers. Figure 4 depicts the temperature-dependent thermal conductivity performance of Midel 7131 nanofluids with h-BN, MoS2, and h-BN/MoS2 at various filler fractions. The synthetic ester did not show significant temperature dependency (less than 1% at 50 °C): actually, a decrement in thermal conductivity was observed as temperature was increased, similar to other authors’ findings [64,65]. In general, the thermal conductivity of the evaluated nanofluids gradually increased for all the fillers and concentrations studied as temperature was also increased.
Figure 4a depicts the effect of h-BN on the thermal conductivity of Midel 7131. As filler fraction and temperature is increased, thermal conductivity also increases. For instance, at 323 K, improvements of 12, 14, 15 and 19% are obtained at 0.01, 0.05, 0.10, and 0.15 wt.%, respectively. A maximum enhancement of 24.2% in thermal conductivity was observed at 0.25 wt.% at 323 K. A similar improvement trend was observed for the nanofluids with MoS2 addition (Figure 4b), showing higher enhancement than h-BN reinforcement. Here, the improvement was 16, 19, 20.5, 23 and 26.5% at 0.01, 0.05, 0.10, 0.15, and 0.25 wt.%, respectively. Thermal conductivity showed significant improvement when hybrid nanostructures were applied as reinforcement. In this case, the incorporation of h-BN/MoS2 nanofiller, contributed to improvement, was 20, 22, 24.5, 27 and 32% at 0.01, 0.05, 0.10, 0.15, and 0.25 wt.%, respectively (Figure 4c).
Natural ester systems are shown in Figure 5 (FR3) and Figure 6 (VG-100). These two green fluids displayed very similar levels of improvement for all fillers. Figure 5a and Figure 6a show the effect of h-BN-filler fraction and temperature rise, reaching a maximum improvement of ~20% in thermal conductivity at 0.25 wt.% at 323 K.
The addition of MoS2 within natural esters showed an improvement trend similar to h-BN (Figure 5b and Figure 6b). In this case, thermal conductivity enhancement of 10, 14, 16, 19 and 21% at 0.01, 0.05, 0.10, 0.15, and 0.25 wt.% was observed for FR3 fluid (Figure 5b). Moreover, VG-100 nanofluids showed a slightly higher enhancement, reaching a maximum of 23% at 0.25 wt.% at 323 K (Figure 6b).
The incorporation of the h-BN/MoS2 nanofiller to natural esters resulted in significant improvements in thermal conductivity. Achieving enhancements of 13, 19, 24, 27 and 30% at 0.01, 0.05, 0.10, 0.15, and 0.25 wt.% was observed for FR3 fluid (Figure 5c). Moreover, enhancements of 16, 21, 24, 30 and 32% at 0.01, 0.05, 0.10, 0.15, and 0.25 wt.%, respectively (Figure 6c), were observed for VG-100 nanofluids.
The incorporation of 2D nanostructures significantly improves the performance of the nanofluids. h-BN contributes to enhance thermal conductivity and MoS2 also contributes to the fluid’s reinforcement and may help to obtain an homogenous dispersion of the nanofillers due to its well-known lubricant properties [66,67]. It is suggested that due to the low filler fraction used, the observed enhancement on thermal conductivity is due to molecular interactions (collisions) between the fluid and the 2D nanostructures [7,14,37,53,68,69]. However, observing thermal conductivity behavior during the temperature-dependent evaluations indicates that this increase is based on the percolation mechanism as well as the contribution of the Brownian motion of the sheet-like nanostructures [70,71,72]. As the concentration of 2D nanostructures is increased in the synthetic or natural esters, the chance of phonons to get scattered in the contiguous nanostructures increases proportionally, which leads to enhanced contact conductance [73]. Consequently, thermal conduction channels are formed, which may increase the thermal conductivity due to the percolation mechanism.
The heat transfer between colliding nanostructures may increase the thermal conductivity of the nanofluids. Particularly, a higher temperature corresponds to a more intense Brownian motion [38]. Thus, at 323 K, the thermal conductivity of nanofluids is more apparent than at the other evaluating temperatures. Moreover, liquid layering at the nanostructures/fluid can also contribute to the improvement of thermal conductivity behavior [74,75,76].
In the hybrid nanofluids, it was observed that at room temperature, the effect of coupling h-BN and MoS2 was similar as a single component reinforcement, either h-BN or MoS2. Moreover, as evaluating temperature increased, the value of thermal conductivity was higher than the single component nanofluids [77,78]. h-BN contributed to the increase of the thermal conductivity as well as MoS2, promoting the dispersion of the nanostructures due to their lubricant properties [38,79].

4. Conclusions

Thermal transport phenomena in nanofluids was influenced by physical and chemical characteristics of the base fluid, as well as interactions with the reinforcement nanostructures. Environmentally-friendly nanofluids based on single h-BN, single MoS2, and hybrid h-BN/MoS2 nanostructures were developed at various filler fractions, and temperature-dependence evaluations were performed for thermal conductivity. In general, all nanofluids showed a temperature-dependence in thermal conductivity performance, indicating the role of the interaction of nanostructures among synthetic and natural esters. The addition and homogeneous dispersion of these 2D nanomaterials within conventional esters showed significant positive results on the effective thermal conductivity performance. Thermal conductivity improved in the range of 21–23% for single reinforcement of h-BN or MoS2 on natural esters and 23–27% for synthetic esters at 323 K, respectively. Furthermore, for the hybrid h-BN/MoS2 nanofluids, an improvement of 30–32% was observed for natural esters and 32% for synthetic esters at 323 K, respectively. Hence, the hybrid h-BN/MoS2 nanostructures dispersed in conventional esters induced a high thermal conductivity, suggesting that the hybrid h-BN/MoS2 have highly desirable multifunctional features for advanced materials applications.
This research showed the potential of an interesting synergistic effect resulting from the contribution of two 2D nanostructures for synthetic and natural esters. Increased environmental awareness is the main driving force for the development of novel technologies, such as the use of biodegradable fluids and lubricants in environmentally-sensitive areas, which have great potential to succeed in industrial applications.

Author Contributions

Conceptualization, J.T.-T., H.R. and S.C.; methodology, J.T.-T. and H.R.; formal analysis and investigation, J.T.-T., A.P.G., J.M.d.O.C., M.A.L., K.A., J.M.M., R.A., M.A.G.B., S.C. and R.J.E.A.; resources, J.T.-T., H.R., R.A., G.B. and R.J.E.A.; data curation, J.T.-T., H.R., K.A. and J.M.M.; writing—original draft preparation, J.T.-T., H.R, A.P.G., J.M.d.O.C.; writing—review and editing, J.T.-T., H.R., G.B., S.C.; G.J.M.F., G.G.S., K.A. and R.J.E.A.; visualization, J.T.-T. and H.R.; supervision, J.T.-T. and H.R.; project administration, J.T.-T. and H.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

Authors would like to thank to Glaura Silva for support given in this research. Authors would like to acknowledge the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - Brasil (CAPES) - PRINT 88887.310339/2018-00 and the Universidad de Monterrey for their support given to develop this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, R. The challenges and opportunities of nanofluids. In Proceedings of the 2019 2nd International Conference on Electrical Materials and Power Equipment (ICEMPE), Guangzhou, China, 7–10 April 2019; pp. 110–114. [Google Scholar]
  2. Amin, D.; Walvekar, R.; Khalid, M.; Vaka, M.; Mujawar, N.M.; Gupta, T.C.S.M. Recent Progress and Challenges in Transformer Oil Nanofluid Development: A Review on Thermal and Electrical Properties. IEEE Access 2019, 7, 151422–151438. [Google Scholar] [CrossRef]
  3. Rafiq, M.; Lv, Y.Z.; Zhou, Y.; Ma, K.B.; Wang, W.; Li, C.R.; Wang, Q. Use of vegetable oils as transformer oils—A review. Renew. Sustain. Energy Rev. 2015, 52, 308–324. [Google Scholar] [CrossRef]
  4. Wong, K.V.; De Leon, O. Applications of nanofluids: Current and future. Adv. Mech. Eng. 2010, 2, 519659. [Google Scholar] [CrossRef] [Green Version]
  5. Salehirad, M.; Nikje, M.M.A. Properties of modified hexagonal boron nitride as stable nanofluids for thermal management applications. Russ. J. Appl. Chem. 2019, 92, 78–86. [Google Scholar] [CrossRef]
  6. Fofana, I. 50 years in the development of insulating liquids. IEEE Electr. Insul. Mag. 2013, 29, 13–25. [Google Scholar] [CrossRef]
  7. Taha-Tijerina, J.; Narayanan, T.N.; Gao, G.; Rohde, M.; Tsentalovich, D.A.; Pasquali, M.; Ajayan, P.M. Electrically insulating thermal nano-oils using 2D fillers. ACS Nano 2012, 6, 1214–1420. [Google Scholar] [CrossRef]
  8. Rouabeh, J.; M’barki, L.; Hammami, A.; Jallouli, I.; Driss, A. Studies of different types of insulating oils and their mixtures as an alternative to mineral oil for cooling power transformers. Heliyon 2019, 5, e01159. [Google Scholar] [CrossRef] [Green Version]
  9. Madanhire, I.; Mbohwa, C.; Madanhire, I.; Mbohwa, C. Development of lubricants. In Mitigating Environmental Impact of Petroleum Lubricants; Madanhire, I., Mbohwa, C., Eds.; Springer: Cham, Switzerland, 2016; pp. 85–101. [Google Scholar]
  10. Nowak, P.; Kucharska, K.; Kamiński, M. Ecological and health effects of lubricant oils emitted into the environment. Int. J. Environ. Res. Public Health 2019, 16, 3002. [Google Scholar] [CrossRef] [Green Version]
  11. Krolczyk, G.M.; Maruda, R.W.; Krolczyk, J.B.; Wojciechowski, S.; Mia, M.; Nieslony, P.; Budzike, G. Ecological trends in machining as a key factor in sustainable production—A review. J. Clean. Prod. 2019, 218, 601–615. [Google Scholar] [CrossRef]
  12. Benedicto, E.; Carou, D.; Rubio, E.M. Technical, economic and environmental review of the Lubrication/Cooling systems used in machining processes. Procedia Eng. 2017, 184, 99–116. [Google Scholar] [CrossRef]
  13. Peña-Parás, L.; Maldonado-Cortés, D.; Taha-Tijerina, J. Eco-friendly nanoparticle additives for lubricants and their tribological characterization. In Handbook of Ecomaterials; Torres, L.M., Kharissova, O.V., Kharisov, B.I., Eds.; Springer Nature: Cham, Switzerland, 2019; pp. 1–21. [Google Scholar]
  14. Taha-Tijerina, J.; Narayanan, T.N.; Avali, S.; Ajayan, P.M. 2D structures-based energy management nanofluids. In Proceedings of the ASME International Mechanical Engineering Congress and Exposition, Houston, TX, USA, 9–15 November 2012. IMECE 2012-87890. [Google Scholar]
  15. Huang, Z.; Li, J.; Yao, W.; Wang, F.; Wan, F.; Tan, Y.; Mehmood, M.A. Electrical and thermal properties of insulating oil-based nanofluids: A comprehensive overview. IET Nanodielectr. 2019, 2, 27–40. [Google Scholar] [CrossRef]
  16. Taha-Tijerina, J.J. Thermal transport and challenges on nanofluids performance. In Microfluidics and Nanofluidics; Kandelousi, M.S., Ed.; IntechOpen: Rijeka, Croatia, 2018; pp. 215–256. [Google Scholar]
  17. Nagendramma, P.; Kaul, S. Development of ecofriendly/biodegradable lubricants: An overview. Renew. Sustain. Energy Rev. 2012, 16, 764–774. [Google Scholar] [CrossRef]
  18. Heikal, E.K.; Elmelawy, M.S.; Khalil, S.A.; Elbasuny, N.M. Manufacturing of environment friendly biolubricants from vegetable oils. Egypt. J. Pet. 2017, 26, 53–55. [Google Scholar] [CrossRef] [Green Version]
  19. Woma, T.Y.; Lawal, S.A.; Abdulrahman, A.S.; Olutoye, M.A.; Ojapah, M.M. Vegetable oil based lubricants: Challenges and prospects. Tribology 2019, 14, 60–70. [Google Scholar] [CrossRef] [Green Version]
  20. Abdalla, H.S.; Patel, S. The performance and oxidation stability of sustainable metalworking fluid derived from vegetable extracts. Proc. Inst. Mech. Eng. B 2006, 220, 2027–2040. [Google Scholar] [CrossRef]
  21. Petlyuk, A.M.; Adams, R.J. Oxidation stability and tribological behavior of vegetable oil hydraulic fluids. Tribol. Trans. 2004, 47, 182–187. [Google Scholar] [CrossRef]
  22. Taha-Tijerina, J.; Aviña, K.; Diabb, J.M. Tribological and Thermal Transport Performance of SiO2-Based Natural Lubricants. Lubricants 2019, 7, 71. [Google Scholar] [CrossRef] [Green Version]
  23. Wilhelm, H.M.; Tulio, L.; Jasinski, R.; Almeida, G. Aging markers for in-service natural ester-based insulating fluids. IEEE Trans. Dielectr. Electr. Insul. 2011, 18, 714–719. [Google Scholar] [CrossRef]
  24. Ghani, A.S.; Muhamad, N.A.; Noorden, Z.A.; Zainuddin, H.; Bakar, A.; Talib, M.A. Methods for improving the workability of natural ester insulating oils in power transformer applications: A review. Electr. Power Syst. Res. 2018, 163, 655–667. [Google Scholar] [CrossRef]
  25. Yao, W.; Wang, W.; Sun, P.; Mehmood, M.A.; Huang, Z. Effect of h-BN and Fe3O4 nanoparticles on streamer propagation and dissipation in vegetable oil based nanofluids. AIP Adv. 2019, 9, 085306. [Google Scholar] [CrossRef] [Green Version]
  26. Maharana, M.; Baruah, N.; Nayak, S.K.; Meher, N.; Iyer, P.K. Condition assessment of aged ester-based nanofluid through physicochemical and spectroscopic measurement. IEEE Trans. Instrum. Meas. 2019, 68, 4853–4863. [Google Scholar] [CrossRef]
  27. Hoekman, S.K.; Broch, A.; Robbins, C.; Ceniceros, E.; Natarajan, M. Review of biodiesel composition, properties, and specifications. Renew. Sustain. Energy Rev. 2012, 16, 143–169. [Google Scholar] [CrossRef]
  28. Eberhardt, R.; Muhr, H.M.; Lick, W.; Baumann, F.; Pukel, G. Comparison of alternative insulating fluids. In Annual Report Conference on Electrical Insulation and Dielectric Phenomena, Quebec, QC, Canada; IEEE: Piscataway, NJ, USA, 2008; pp. 591–593. [Google Scholar]
  29. IEEE Power Engineering Society. C57.147-2018—IEEE Guide for Acceptance and Maintenance of Natural Ester Insulating Liquid in Transformers. 2018. Available online: https://0-standards-ieee-org.brum.beds.ac.uk/standard/C57_147-2018.html (accessed on 3 March 2020).
  30. Srikant, R.R.; Rao, P.N. Use of vegetable-based cutting fluids for sustainable machining. In Sustainable Machining; Davim, J.P., Ed.; Springer: Cham, Switzerland, 2017; pp. 31–46. [Google Scholar]
  31. Azmi, W.H.; Sharma, K.V.; Mamat, R.; Najafi, G.; Mohamad, M.S. The enhancement of effective thermal conductivity and effective dynamic viscosity of nanofluids—A review. Renew. Sustain. Energy Rev. 2016, 53, 1046–1058. [Google Scholar] [CrossRef]
  32. Taha-Tijerina, J.J.; Narayanan, T.N.; Tiwary, C.S.; Lozano, K.; Chipara, M.; Ajayan, P.M. Nanodiamond-based thermal fluids. ACS Appl. Mater. Interfaces 2014, 6, 4778–4785. [Google Scholar] [CrossRef] [PubMed]
  33. Sarviya, R.M.; Fuskele, V. Review on Thermal Conductivity of Nanofluids. Mater. Today Proc. 2017, 4, 4022–4031. [Google Scholar] [CrossRef]
  34. Leong, K.Y.; Ku, A.K.Z.; Ong, H.C.; Ghazali, M.J.; Baharum, A. Synthesis and thermal conductivity characteristic of hybrid nanofluids—A review. Renew. Sustain. Energy Rev. 2017, 75, 868–878. [Google Scholar] [CrossRef]
  35. Koshy, C.P.; Rajendrakumar, P.K.; Thottackkad, M.V. Evaluation of the tribological and thermo-physical properties of coconut oil added with MoS2 nanoparticles at elevated temperatures. Wear 2015, 330–331, 288–308. [Google Scholar] [CrossRef]
  36. Padmini, R.; Vamsi Krishna, P.; Rao, K.M. Effectiveness of vegetable oil based nanofluids as potential cutting fluids in turning AISI 1040 steel. Tribol. Int. 2016, 94, 490–501. [Google Scholar] [CrossRef]
  37. Ribeiro, H.; Trigueiro, J.P.C.; Lopes, M.C.; Pedrotti, J.J.; Woellner, C.F.; Silva, W.M.; Silva, G.G.; Ajayan, P.M. Enhanced thermal conductivity and mechanical properties of hybrid MoS2/h-BN polyurethane nanocomposites. J. Appl. Polym. Sci. 2018, 135, 46560. [Google Scholar] [CrossRef]
  38. Ribeiro, H.; Trigueiro, J.P.C.; Silva, W.M.; Woellner, C.F.; Owuor, P.S.; Chipara, A.C.; Lopes, M.C.; Tiwary, C.S.; Pedrotti, J.J.; Villegas, R.; et al. Hybrid MoS2/h-BN nanofillers as synergic heat dissipation and reinforcement additives in epoxy nanocomposites. ACS Appl. Mater. Interfaces 2019, 11, 24485–24492. [Google Scholar] [CrossRef]
  39. Ribeiro, H.; Trigueiro, J.P.C.; Woellner, C.F.; Pedrotti, J.J.; Miquita, D.R.; Silva, W.M.; Lopes, M.C.; Fechinee, G.J.M.; Luciano, M.A.; Silva, G.G.; et al. Higher thermal conductivity and mechanical enhancements in hybrid 2D polymer nanocomposites. Polym. Test. 2020, 87, 106510. [Google Scholar] [CrossRef]
  40. Charoo, M.S.; Wani, M.F. Tribological properties of h-BN nanoparticles as lubricant additive on cylinder liner and piston ring. Lubr. Sci. 2017, 29, 241–254. [Google Scholar] [CrossRef]
  41. Ribeiro, H.; Trigueiro, J.P.C.; Owuor, P.S.; Machado, L.D.; Woellner, C.F.; Pedrotti, J.J.; Jaques, Y.M.; Kosolwattana, S.; Chipara, A.; Silva, W.M.; et al. Hybrid 2D nanostructures for mechanical reinforcement and thermal conductivity enhancement in polymer composites. Compos. Sci. Technol. 2018, 159, 103–110. [Google Scholar] [CrossRef]
  42. Du, B.X.; Li, X.L.; Li, J.; Tao, X.Y. Effects of BN nanoparticles on thermal conductivity and breakdown strength of vegetable oil. In Proceedings of the 2015 IEEE 11th International Conference on the Properties and Applications of Dielectric Materials (ICPADM), Sydney, Australia, 19–22 July 2015; pp. 476–479. [Google Scholar]
  43. Mohamad, N.A.; Azis, N.; Jasni, J.; Ab Kadir, M.Z.A.; Yunus, R.; Yaakub, Z. Impact of Fe3O4, CuO and Al2O3 on the AC breakdown voltage of palm oil and coconut oil in the presence of CTAB. Energies 2019, 12, 1605. [Google Scholar] [CrossRef] [Green Version]
  44. Kurimský, J.; Rajňák, M.; Cimbala, R.; Rajnič, J.; Timko, M.; Kopčanský, P. Effect of magnetic nanoparticles on partial discharges in transformer oil. J. Magn. Magn. Mater. 2020, 496, 165923. [Google Scholar]
  45. Ghoneim, S.S.M.; Sabiha, N.A.; Hessien, M.M.; Alahmadi, A. Evaluation of dielectric breakdown strength of transformer oil with BaTiO3 and NiFe2O4 nanoparticles. Electr. Eng. 2019, 101, 369–377. [Google Scholar] [CrossRef]
  46. Li, Y.; Zhou, J.; Luo, Z.; Tung, S.; Schneider, E.; Wu, J.; Li, X. Investigation on two abnormal phenomena about thermal conductivity enhancement of BN/EG nanofluids. Nanoscale Res. Lett. 2011, 6, 443. [Google Scholar] [CrossRef] [Green Version]
  47. Ilhan, B.; Kurt, M.; Ertürk, H. Experimental investigation of heat transfer enhancement and viscosity change of hBN nanofluids. Exp. Therm. Fluid Sci. 2016, 77, 272–283. [Google Scholar] [CrossRef]
  48. Voiry, D.; Salehi, M.; Silva, R.; Fujita, T.; Chen, M.; Asefa, T.; Shenoy, V.B.; Eda, G.; Chhowalla, M. Conducting MoS2 nanosheets as catalysts for hydrogen evolution reaction. Nano Lett. 2013, 13, 6222–6227. [Google Scholar] [CrossRef]
  49. Liu, G.; Robertson, A.W.; Li, M.M.J.; Kuo, W.C.H.; Darby, M.T.; Muhieddine, M.H.; Lin, Y.C.; Suenaga, K.; Stamatakis, M.; Warner, J.H.; et al. MoS2 monolayer catalyst doped with isolated Co atoms for the hydrodeoxygenation reaction. Nat. Chem. 2017, 9, 810–816. [Google Scholar] [CrossRef]
  50. Gupta, M.; Singh, V.; Kumar, R.; Said, Z. A review on thermophysical properties of nanofluids and heat transfer applications. Renew. Sustain. Energy Rev. 2017, 74, 638–670. [Google Scholar] [CrossRef]
  51. Babita; Sharma, S.K.; Gupta, S.M. Preparation and evaluation of stable nanofluids for heat transfer application: A review. Exp. Therm. Fluid Sci. 2016, 79, 202–212. [Google Scholar] [CrossRef]
  52. Su, Y.; Gong, L.; Li, B.; Chen, D. An experimental investigation on thermal properties of molybdenum disulfide nanofluids. In Proceedings of the 2015 International Conference on Materials, Environmental and Biological Engineering, Guilin, China, 28–30 March 2015; pp. 881–885. [Google Scholar]
  53. Zeng, Y.-X.; Zhong, X.-W.; Liu, Z.-Q.; Chen, S.; Li, N. Preparation and enhancement of thermal conductivity of heat transfer oil-based MoS2 nanofluids. J. Nanomater. 2013, 2013, 1–6. [Google Scholar]
  54. Ellahi, R.; Sait, S.M.; Shehzad, N.; Ayaz, Z. A hybrid investigation on numerical and analytical solutions of electro-magnetohydrodynamics flow of nanofluid through porous media with entropy generation. Int. J. Numer. Methods Heat Fluid Flow. 2019, 30, 834–854. [Google Scholar] [CrossRef]
  55. Aboalhamayie, A.; Festa, L.; Ghamari, M. Evaporation rate of colloidal droplets of jet fuel and carbon-based nanoparticles: Effect of thermal conductivity. Nanomaterials 2019, 9, 1297. [Google Scholar] [CrossRef] [Green Version]
  56. GE Prolec—VG-100: Fluid Natural Ester-based Fluid for Transformers. Monterrey. Available online: http://prolecge.com/wp-content/uploads/2017/10/VG-100_Technical_Information.pdf (accessed on 10 March 2020).
  57. M&I Materials Ltd. MIDEL 7131 Technical Datasheet No 14, Thermal Properties. Available online: https://www.midel.com/app/uploads/2018/09/MIDEL_7131_Thermal_Properties.pdf (accessed on 28 February 2020).
  58. Cargil Inc. Envirotemp FR3TM fluid Formulated for Performance. Minnesota, USA. Available online: https://www.cargill.com/doc/1432076501275/envirotemp-fr3-brochure.pdf (accessed on 28 February 2020).
  59. Rodríguez-Díaz, J.; Contreras, J.E.; Montes, J.A. Aplicación de un éster natural base soya en transformadores. In Proceedings of the 2018 IEEE 38th Central America and Panama Convention (CONCAPAN XXXVIII), San Salvador, El Salvador, 7–9 November 2018. [Google Scholar]
  60. Cai, Q.; Scullion, D.; Gan, W.; Falin, A.; Zhang, S.; Watanabe, K.; Taniguchi, T.; Chen, Y.; Santos, E.J.G.; Li, L.H. High thermal conductivity of high-quality monolayer boron nitride and its thermal expansion. Sci. Adv. 2019, 5, eaav0129. [Google Scholar] [CrossRef] [Green Version]
  61. Gandi, A.N.; Schwingenschlögl, U. Thermal conductivity of bulk and monolayer MoS2. Europhys. Lett. 2016, 113, 36002. [Google Scholar] [CrossRef]
  62. Krishnamoorthy, A.; Rajak, P.; Norouzzadeh, P.; Singh, D.J.; Kalia, R.K.; Nakano, A.; Vashishta, P. Thermal conductivity of MoS2 monolayers from molecular dynamics simulations. AIP Adv. 2019, 9, 035042. [Google Scholar] [CrossRef] [Green Version]
  63. Li, H.; Zhang, Q.; Yap, C.C.R.; Tay, B.K.; Edwin, T.H.T.; Olivier, A.; Baillargeat, D. From Bulk to Monolayer MoS2: Evolution of Raman Scattering. Adv. Funct. Mater. 2012, 22, 1385–1390. [Google Scholar] [CrossRef]
  64. Nadolny, Z.; Dombek, G.; Nadolny, Z.; Dombek, G. Electro-insulating nanofluids based on synthetic ester and TiO2 or C60 nanoparticles in power transformer. Energies 2018, 11, 1953. [Google Scholar] [CrossRef] [Green Version]
  65. Salama, M.M.M.; Mansour, D.E.A.; Daghrah, M.; Abdelkasoud, S.M.; Abbas, A.A. Thermal performance of transformers filled with environmentally friendly oils under various loading conditions. Int. J. Electr. Power Energy Syst. 2020, 118, 105743. [Google Scholar] [CrossRef]
  66. Bartels, T.; Bock, W.; Braun, J.; Busch, C.; Buss, W.; Dresel, W.; Freiler, C.; Harpescheid, M.; Heckler, R.-P.; Hörner, D.; et al. Lubricants and Lubrication. In Ullmann’s Encyclopedia of Industrial Chemistry; Bohnet, M., Ed.; Wiley-VCH Verlag GmbH & Co.: Weinheim, Germany, 2003. [Google Scholar]
  67. Song, I.; Park, C.; Choi, H.C. Synthesis and properties of molybdenum disulphide: From bulk to atomic layers. RSC Adv. 2015, 5, 7495–7514. [Google Scholar] [CrossRef] [Green Version]
  68. Zhou, H.; Zhu, J.; Liu, Z.; Yan, Z.; Fan, X.; Lin, J.; Wang, G.; Yan, Q.; Yu, T.; Ajayan, P.M.; et al. High thermal conductivity of suspended few-layer hexagonal boron nitride sheets. Nano Res. 2014, 7, 1232–1240. [Google Scholar] [CrossRef]
  69. Yan, R.; Simpson, J.R.; Bertolazzi, S.; Brivio, J.; Watson, M.; Wu, X.; Kis, A.; Luo, T.; Walker, A.R.H.; Xing, H.G. Thermal conductivity of monolayer molybdenum disulfide obtained from temperature-dependent raman spectroscopy. ACS Nano 2014, 8, 986–993. [Google Scholar] [CrossRef]
  70. Lee, S.L.; Saidur, R.; Sabri, M.F.M.; Min, T.K. Molecular dynamic simulation: Studying the effects of Brownian motion and induced micro-convection in nanofluids. Numer. Heat Transf. Part A Appl. 2016, 69, 643–658. [Google Scholar] [CrossRef]
  71. Peña-Parás, L.; García-Pineda, P.; Maldonado-Cortés, D.; Garza, G.T.; Taha-Tijerina, J. Temperature dependence of the extreme-pressure behavior of CuO and TiO2 nanoparticle additives in metal-forming polymeric lubricants. Ind. Lubr. Tribol. 2017, 69, 730–737. [Google Scholar] [CrossRef]
  72. Teruel, M.; Aguilar, T.; Martínez-Merino, P.; Carrillo-Berdugo, I.; Gallardo-Bernal, J.J.; Gómez-Villarejo, R.; Alcántara, R.; Fernández-Lorenzo, C.; Navas, J. 2D MoSe2-based nanofluids prepared by liquid phase exfoliation for heat transfer applications in concentrating solar power. Sol. Energy Mater. Sol. Cells. 2019, 200, 109972. [Google Scholar] [CrossRef]
  73. Shin, S.; Chen, R. Thermal transport measurements of nanostructures using suspended micro-devices. Nanoscale Energy Transp. 2020, 12-1, 12–33. [Google Scholar]
  74. Taha-Tijerina, J.; Peña-Parás, L.; Maldonado-Cortés, D. 2D-Based Nanofluids: Materials Evaluation and Performance. In Two-Dimensional Materials—Synthesis, Characterization and Potential Applications; Nayak, P.K., Ed.; IntechOpen: Rijeka, Croatia, 2016; pp. 153–198. [Google Scholar]
  75. Taha-Tijerina, J.; Cadena-de la Peña, N.; Cue-Sampedro, R.; Rivera-Solorio, C. Thermo-physical evaluation of dielectric mineral oil-based nitride and oxide nanofluids for thermal transport applications. J. Therm. Sci. Technol. 2019, 14, JTST0007. [Google Scholar] [CrossRef] [Green Version]
  76. Iqbal, S.M.; Raj, C.S.; Michael, J.J.; Irfan, A.M. A Comparative Investigation of Al2O3/H2O, SiO2/H2O and ZrO2/H2O Nanofluid for Heat Transfer Applications. Dig. J. Nanomater. Biostruct. 2017, 12, 255–264. [Google Scholar]
  77. Parto, K.; Pal, A.; Xie, X.; Cao, W.; Banerjee, K. Interfacial thermal conductivity of 2D layered materials: An atomistic approach. In Proceedings of the 2018 IEEE International Electron Devices Meeting (IEDM), San Francisco, CA, USA, 1–5 December 2018. [Google Scholar]
  78. Liu, Y.; Ong, Z.Y.; Wu, J.; Zhao, Y.; Watanabe, K.; Taniguchi, T.; Chi, D.; Zhang, G.; Thong, J.T.L.; Qiu, C.-W.; et al. Thermal conductance of the 2D MoS2/h-BN and graphene/h-BN interfaces. Sci. Rep. 2017, 7, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Sharma, V.; Kagdada, H.L.; Jha, P.K.; Śpiewak, P.; Kurzydłowski, K.J. Thermal transport properties of boron nitride based materials: A review. Renew. Sustain. Energy Rev. 2020, 120, 109622. [Google Scholar] [CrossRef]
Figure 1. SEM images of h-BN (a), MoS2 (b), and h-BN/MoS2 nanosheets (c,d). AFM image of the h-BN/MoS2 mixture and its height profile (e,f).
Figure 1. SEM images of h-BN (a), MoS2 (b), and h-BN/MoS2 nanosheets (c,d). AFM image of the h-BN/MoS2 mixture and its height profile (e,f).
Nanomaterials 10 01160 g001
Figure 2. SEM image of h-BN/MoS2 system (a) and its resulting EDS mapping of the elements B, N, Mo, and S on SiO2 grid (bf).
Figure 2. SEM image of h-BN/MoS2 system (a) and its resulting EDS mapping of the elements B, N, Mo, and S on SiO2 grid (bf).
Nanomaterials 10 01160 g002
Figure 3. (a) Raman and XRD spectra of h-BN, (b) MoS2 and the h-BN/MoS2 mixture.
Figure 3. (a) Raman and XRD spectra of h-BN, (b) MoS2 and the h-BN/MoS2 mixture.
Nanomaterials 10 01160 g003
Figure 4. Conductivity performance of Midel 7131: (a) h-BN, (b) MoS2, and (c) h-BN/MoS2 nanofluids under temperature-dependence evaluation (percentage of filler amount is mentioned).
Figure 4. Conductivity performance of Midel 7131: (a) h-BN, (b) MoS2, and (c) h-BN/MoS2 nanofluids under temperature-dependence evaluation (percentage of filler amount is mentioned).
Nanomaterials 10 01160 g004
Figure 5. Conductivity performance of FR3: (a) h-BN, (b) MoS2, and (c) h-BN/MoS2 nanofluids under temperature-dependence evaluation (percentage of filler amount is mentioned).
Figure 5. Conductivity performance of FR3: (a) h-BN, (b) MoS2, and (c) h-BN/MoS2 nanofluids under temperature-dependence evaluation (percentage of filler amount is mentioned).
Nanomaterials 10 01160 g005
Figure 6. Conductivity performance of VG-100: (a) h-BN, (b) MoS2, and (c) h-BN/MoS2 nanofluids under temperature-dependence evaluation (percentage of filler amount is mentioned).
Figure 6. Conductivity performance of VG-100: (a) h-BN, (b) MoS2, and (c) h-BN/MoS2 nanofluids under temperature-dependence evaluation (percentage of filler amount is mentioned).
Nanomaterials 10 01160 g006
Table 1. Material properties [56,57,58,59].
Table 1. Material properties [56,57,58,59].
General PropertiesMaterials
UnitsStandardEnvirotemp ® FR3 TMMidel 7131 VG-100
Density @ 25 °C g/cm3ASTM D9240.920.950.92
Kinematic viscosity @ 25 °C mm2/sASTM D445-53-
Kinematic viscosity @ 40 °C mm2/sASTM D4453329.531
Kinematic Viscosity @ 100 °Cmm2/sASTM D44585.36
Thermal conductivity @ 25 °C(W m−1 K−1)ASTM D78960.1670.147 -
Thermal conductivity @ 50 °C(W m−1 K−1)ASTM D7896-0.145-
Pour point (°C)ASTM D97−18 to −23−56−12 to −15
Table 2. Nanostructures characteristics at 298 K.
Table 2. Nanostructures characteristics at 298 K.
General Propertiesh-BNMoS2Units
Purity9899.98%
Density2.295.06gr/cm3
Particle size 5.0–10.0μm
Thermal conductivity300–500 [60]38–131 [61,62]W/m K

Share and Cite

MDPI and ACS Style

Taha-Tijerina, J.; Ribeiro, H.; Aviña, K.; Martínez, J.M.; Godoy, A.P.; Cremonezzi, J.M.d.O.; Luciano, M.A.; Gimenes Benega, M.A.; Andrade, R.J.E.; Fechine, G.J.M.; et al. Thermal Conductivity Performance of 2D h-BN/MoS2/-Hybrid Nanostructures Used on Natural and Synthetic Esters. Nanomaterials 2020, 10, 1160. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10061160

AMA Style

Taha-Tijerina J, Ribeiro H, Aviña K, Martínez JM, Godoy AP, Cremonezzi JMdO, Luciano MA, Gimenes Benega MA, Andrade RJE, Fechine GJM, et al. Thermal Conductivity Performance of 2D h-BN/MoS2/-Hybrid Nanostructures Used on Natural and Synthetic Esters. Nanomaterials. 2020; 10(6):1160. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10061160

Chicago/Turabian Style

Taha-Tijerina, Jaime, Hélio Ribeiro, Karla Aviña, Juan Manuel Martínez, Anna Paula Godoy, Josué Marciano de Oliveira Cremonezzi, Milene Adriane Luciano, Marcos Antônio Gimenes Benega, Ricardo Jorge Espanhol Andrade, Guilhermino José Macedo Fechine, and et al. 2020. "Thermal Conductivity Performance of 2D h-BN/MoS2/-Hybrid Nanostructures Used on Natural and Synthetic Esters" Nanomaterials 10, no. 6: 1160. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10061160

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop