Вы находитесь на странице: 1из 9

Chapter 8

Thermal Conductivity of Nanofluids


Pawel Keblinski

Nanofluids (colloidal suspensions of solid nanoparticles) sparked excitement as well


as controversy. In particular, a number of researches reported dramatic increases of
thermal conductivity with small nanoparticle loading, while others showed moderate increases consistent with the effective medium theories on well-dispersed conductive spheres. Here we discuss potential mechanisms that were put forward in
order to understand nanoflouid thermal conductivity and demonstrate that particle
aggregation is currently the only physically reasonable mechanism that can explain
the majority of the experimental data.

8.1 Introduction
Cooling is one of the most important technical challenges facing numerous diverse
industries, including microelectronics, transportation, solid-state lighting, and manufacturing. Developments driving the increased thermal loads that require advances
in cooling include faster speeds (in the multi-GHz range) and smaller features (to
< 100 nm) for microelectronic devices, higher power engines, and brighter optical devices. The conventional method for increasing heat dissipation is to increase
the area available for exchanging heat with a heat transfer fluid, but this approach
requires an undesirable increase in the size of thermal management system. There
is therefore an urgent need for new and innovative coolants with improved performance. About a decade ago a novel concept of nanofluids, i.e., heat transfer fluids
containing suspensions of nanoparticles, was proposed as a means of meeting these
challenges [1].
Nanofluids are solidliquid composite materials consisting of solid nanoparticles
or nanofibers, with sizes typically on the order of 1100 nm, suspended in a liquid.
Nanofluids have attracted great interest recently due to reports of greatly enhanced
thermal properties at low volume fractions. For example, a small amount (less
than 1% volume fraction) of copper nanoparticles and carbon nanotubes dispersed

S. Volz (ed.), Thermal Nanosystems and Nanomaterials, Topics in Applied Physics, 118
c Springer-Verlag Berlin Heidelberg 2009
DOI 10.1007/978-3-642-04258-4 8, 

213

214

Pawel Keblinski

in ethylene glycol and oil, respectively, was reported to increase the inherently poor
thermal conductivity of the liquid by 40% and 150%, respectively [25]. Conventional suspensions of well-dispersed particles require high concentrations (> 10%)
of particles to achieve such enhancement; problems of rheology and stability, which
are amplified at high concentrations, preclude the widespread use of conventional
slurries as heat transfer fluids. In several cases, the observed enhancement in thermal conductivity of nanofluids is orders-of-magnitude larger than predicted by wellestablished theories of dispersed particles [612].

8.2 Excitement, Controversy, and New Physics


The large thermal conductivity enhancements reported by experiment led to excitement but also to controversy. The origin of the excitement was that the measured
thermal conductivity was often much larger than that predicted by well-established
effective medium theories under the assumption of well-dispersed particles. For a
system of well dispersed particles, in the limit of low particle volume fraction , and
with much higher particle conductivity NP than fluid conductivity fluid , the effective medium (Maxwell) theory for vanishing interfacial thermal resistance leads to
the relation [13, 14]
NF
= 1 + 3 .
(8.1)
fluid
Equation (8.1) predicts only moderate thermal transport increase, independently of
the conductivity of the filler. As already discussed, many experiments reported much
larger increases.
This mismatch between predicted and observed values led not only to excitement
but also to controversy, since in some experiments, the measured thermal conductivity was not very large and consistent with the prediction of (8.1) [1519]. This
initially led to the belief that some experiments had to be wrong, particularly considering that different research groups often obtained different results for what were
presumably the same nanofluids.
Further excitement and controversy were associated with potential mechanisms
put forward to explain the anomalous experimental observations. This discussion
was initiated by Keblinski et al. [20]. In the following we will review the key mechanism analyzed and discussed in the literature.

8.2.1 Brownian Motion


Motivated by the unusual thermal conductivity enhancements, a number of researchers promoted the concept of Brownian motion induced micro- or nanoconvection [2123]. They argued in different ways that each Brownian particle generates a
long-range velocity field in the surrounding fluid, akin to that present around a par-

8 Thermal Conductivity of Nanofluids

215

ticle moving with a constant velocity, which decays approximately as the inverse of
the distance from the particle center. The ability of large volumes of fluid dragged
by the nanoparticles to carry substantial amounts of heat was credited for the large
thermal conductivity increases in nanofluids.
A key weakness of this argument is that the thermal diffusivity DT of the base
fluid, which measures the rate of the heat flow via thermal conduction, is several orders of magnitude larger than nanoparticle diffusivity DNP , which measures the rate
of mass motion due to nanoparticle diffusion, whence the magnitude of possible
nanoconvection effects is negligible [24]. Furthermore, the velocity field around a
Brownian particle is much shorter-ranged than that around a particle moving with a
constant velocity [25]. The low estimates for the contribution to thermal transport of
Brownian motion were supported by the results of molecular dynamics simulations
[24, 26], and by the results of experimental measurements on well-dispersed spherical nanoparticle fluids [1619, 27], all showing thermal conductivity enhancements
(positive and negative) that agree with (8.1). In a direct experimental investigation,
the density effects associated with the postulated nanoconvection [23] were experimentally tested with lighter silica and Teflon particles, and were shown to be
incorrect [28]. The nanoconvection velocities were further shown to be of the order of thermophoretic velocities, which for most nanofluids were insignificant (as
low as 109 m/s). Even for sub-nanometer clusters, as evidenced from molecular
dynamics simulations, the thermophoretic velocities are exceedingly small, and this
effectively precludes a discernible contribution to the nanofluid thermal conductivity from any conceivable nanoconvection mechanism [29].

8.2.2 Interfacial Liquid Layer


Considering that the molecular structure of liquid at the solid interface is more ordered, possibilities of larger thermal conductivity of this ordered liquid layer and
tunneling of heat carrying phonons from one particle to another were put forward
[20]. The subsequent molecular dynamics simulation work concluded that those
mechanisms do not contribute significantly to heat transfer [30]. For strong solid
liquid interactions, typical of those in nanofluids with metallic nanoparticles, a percolating network of amorphous-like fluid structures can emerge which can facilitate
additional thermal conduction paths [31, 32]. However, a discernible increase in
thermal conductivity is possible only for exceedingly small colloidal particles (limited to few tens of atoms). Experiments have shown that the structured interfacial
fluid layers are limited to a few molecular spacings from a solid surface [33], which
makes them inadequate to influence the thermal transport in common nanofluids.
Furthermore, interpretation of the cooling rates of Au nanoparticles suspended in
water and organic solvents does not appear to require any unusual thermophysical
properties of the surrounding liquid to explain the experimental results [34].

216

Pawel Keblinski

8.2.3 Interfacial Thermal Resistance


As described above, the liquid layering does not generate thermal conductivity increases. There is, however, a well-known interfacial effect degrading thermal conductivity: the thermal interfacial resistance Rk , most commonly defined via its inverse, the interfacial conductance G = 1/Rk . G is related to the heat flux JQ and the
temperature drop T at the interface via
JQ = GT .

(8.2)

A simple measure of the relative importance of the interfacial resistance in the overall heat flow in a composite can be obtained from the equivalent thickness h, defined
as the distance over which the temperature drop is the same as at the interface. This
thickness is given by the ratio of the fluid thermal conductivity fluid to the interfacial conductance, i.e., h = fluid /G.
The interfacial thermal resistance effect on well-dispersed spherical particle composites can be estimated by effective medium theory to be [15]

1
NF
,
1 = 3
fluid
+2

(8.3)

where is the ratio of the particle radius to the equivalent matrix thickness h. According to (8.3), when the particle radius becomes equal to the equivalent matrix
thickness ( = 1), there is no enhancement at all, while for larger interfacial resistance ( < 1), the addition of particles decreases the thermal conductivity of the
composite. This effect can explain the decrease of thermal conductivity of nanofluids below the base fluid value [18, 19] and much lower than expected thermal conductivity increases of carbon nanotube nanofluids and carbon nanotube polymer
composites [35, 36].

8.2.4 Near Field Radiation


Somewhat less popular was the idea of anomalously high radiative energy transfers
between nanoparticles [37]. Since the classical radiation theory does not predict
radiative transfer of significance for nanofluids, molecular dynamics simulations of
thermal energy exchange between two silica nanoparticles were carried out, resulting in the prediction of enormous thermal conductance due to near-field interactions
when particles were closely separated [37]. However, the thermal conductance obtained from molecular dynamics was significantly larger than that obtained under
the assumption that all the thermal energy is exchanged between two nanoparticles
within a single atomic vibration period. Further analysis demonstrated that nearfield interactions do not affect the nanofluid thermal conductivity [38].

8 Thermal Conductivity of Nanofluids

217

8.2.5 Particle Clustering


Maxwells expression, the limiting case of which is given by (8.1), corresponds to
the situation of well-dispersed nanoparticles. Such structure is the least efficient
from the point of view of thermal transport enhancement since conductive particles
are separated from each other by low conductivity liquid. Aggregation of particles
into sparse clusters, or ideally into linear chains, leads to extended and highly conductive paths for the heat flow. In fact, there is a well established understanding of
the impact of morphology on thermal conductivity of a composite. Depending on
the morphology, the conductivity can vary greatly, even at the same volume fraction
of the components. It is generally accepted that the conductivity has to fall between
so-called HashinShtrikman (HS) bounds, obtained under an effective medium analysis [39] without any restriction on the volume fraction. The HS bounds are given
by




3 (NP fluid )
3(1 )(NP fluid )
NF 1
fluid 1+
NP .
3fluid + (1 )(NP fluid )
3NP (NP fluid)
(8.4)
In the inequality given by (8.4), the lower (Maxwell) bound corresponds physically
to a set of well-dispersed nanoparticles in a fluid matrix, while the upper limit corresponds to large pockets of fluid separated by linked, chain-forming or clustered
nanoparticles. From the point of view of circuit analysis, the lower HS limit lies
closer to conductors connected in series, while the upper limit lies closer to conductors in a parallel mode. The HS bounds do not give a precise mechanism of
thermal conductance, but set the most restrictive limits based on knowledge of the
volume fraction alone. It is relatively well-known that a large number of experimental data on solid composites, and to a lesser extent the data on liquid mixtures,
fall between the HS bounds [29]. An unbiased estimate (favoring neither series nor
parallel modes) predicts thermal conductivity values that lie between the upper and
lower HS bounds [40].
The evidence from scanning electron microscopy (SEM) points to the existence
of partial clustering and chain-like linear aggregation [1012, 41]. Viscosity data
on nanofluids has also shown an anomalous increase as compared to the Einstein
model for well dispersed particles [42]. A large increase in the viscosity is another
indication of aggregation in the nanofluids. Interestingly, for a carbon nanotube suspension, an effective medium theory accounting for the very high aspect ratio of the
nanotubes [14], predicts thermal conductivity enhancements that are in fact well
above the reported values. This was attributed to a significant interfacial resistance
to heat flow between the carbon nanotubes and the fluid [35].

218

Pawel Keblinski

8.3 Discussion
Clear experimental evidence for extensive clustering observed in nanofluids indicates that the expected thermal conductivity can fall within a wide range bounded
by (8.4) and strongly depends on the actual aggregation. Recently, using a multilevel effective medium theory, it was demonstrated that the thermal conductivity of
nanofluids can be significantly enhanced by the aggregation of nanoparticles into
chain-like clusters, and this enhancement can be quite dramatic for large, but sparse
clusters [43]. Predictions of this aggregation model were in excellent agreement
with detailed numerical calculations on model nanofluids involving fractal clusters [43]. Thus, allowing for clustering effects dramatically broadens the thermal
conductivity range that is consistent with the effective medium theory [29]. Quite
strikingly, it was demonstrated that the vast majority of experimental results fall
within these bounds, supporting the classical nature of thermal conduction behavior
in nanofluids [44].
The increase in the relative thermal conductivity of nanofluids with temperature [4] is another example of anomalous behavior that cannot be explained on the
assumption of well-dispersed particles. Similarly, the increase in the thermal conductivity with decreasing particle size cannot be explained if the particles are well
dispersed. The probability of aggregation increases with increasing temperature and
decreasing particle size [45]. Therefore the apparent contradictions between experiment and theory, such as particle size effects, can be resolved by weighing in the
ability of nanoparticles to form linear clusters. Furthermore, the temperature dependence is not as strong as it was previously believed to be, with recent experiments
showing a similar variation for both nanofluids and the base fluid [19, 29]. This
implies that the mechanism for increase in the thermal conductivity of water (presumably from the hydrogen bonded structures) is partly responsible for the thermal
conductivity increase in nanofluids as well. Conversely, it is reasonable to expect a
decrease in the nanofluid thermal conductivity for a base fluid that has a negative
change in conductivity with increasing temperature.
It remains a challenge to accurately identify and manipulate the cluster configuration to modify the thermal transport properties of a nanofluid. The two characterization techniques, DLS and SEM, have limitations in assessing the structure of
nanoparticles. DLS measurements are limited to dilute suspensions ( < 1%) for
most nanofluids, while SEM imaging can be performed only after drying the base
fluid. While the science of making well-dispersed colloids has reached a fair level
of maturity, attempts to generate targeted nanoparticle configurations are still in an
evolving phase.
The fact that significant aggregation is required to obtain substantial increases in
thermal transport has an important consequence for the potential application of such
fluids in flow-based cooling, which is the most important benefit from the technological point of view. It is well known that aggregation into sparse but large clusters
increases fluid viscosity. Such increases become very dramatic when the aggregates
start to touch each other, which can occur at very low volume fractions, as low as
0.2% [46]. Therefore, the same aggregate structures that are most beneficial to the

8 Thermal Conductivity of Nanofluids

219

thermal transport are also the most detrimental to the fluid flow characteristics. Future research should therefore address the issue of optimizing nanofluid structure
with the best combination of thermal conductivity and viscosity.

References
1. Choi, S.U.S.: Enhancing thermal conductivity of fluids with nanoparticles. In: Siginer, D.A.,
Wang, H.P., and Div., F.E. (Eds.) Developments and Applications of Non-Newtonian Flows,
American Society of Mechanical Engineers, New York (1995) pp. 99105
2. Choi, S.U.S., Zhang, Z.G., Yu, W., Lockwood, F.E., and Grulke, E.A.: Anomalous thermal
conductivity enhancement in nanotube suspensions. Appl. Phys. Lett. 79, 22522254 (2001)
3. Eastman, J.A., Choi, S.U.S., Li, S., Yu, W., and Thompson, L.J.: Anomalously increased effective thermal conductivities of ethylene glycol-based nanofluids containing copper nanoparticles. Appl. Phys. Lett. 78, 718720 (2001)
4. Das, S.K., Putra, N., Thiesen, P., and Roetzel, W.: Temperature dependence of thermal conductivity enhancement for nanofluids. J. Heat Transfer 125, 567574 (2003)
5. Patel, H.E., Das, S.K., Sundararajan, T., Nair, A.S., George, B., and Pradeep, T.: Thermal
conductivities of naked and monolayer protected metal nanoparticle based nanofluids: Manifestation of anomalous enhancement and chemical effects. Appl. Phys. Lett. 83, 29312933
(2003)
6. Chopkar, M., Das, P.K., and Manna, I.: Synthesis and characterization of nanofluid for advanced heat transfer applications. Scripta Materialia 55, 549552 (2006)
7. Chopkar, M., Kumar, S., Bhandari, D.R., Das, P.K., and Manna, I.: Development and characterization of Al2 Cu and Ag2 Al nanoparticle dispersed water and ethylene glycol based
nanofluid. Materials Science and Engineering: B 139, 141148 (2007)
8. Hong, K.S., Hong, T.-K., and Yang, H.-S.: Thermal conductivity of Fe nanofluids depending
on the cluster size of nanoparticles. Appl. Phys. Lett. 88, 031901 (2006)
9. Kang, H.U., Kim, S.H., and Oh, J.M.: Estimation of thermal conductivity of nanofluid using
experimental effective particle volume. Experimental Heat Transfer 19, 181191 (2006)
10. Murshed, S.M.S., Leong, K.C., and Yang, C.: Enhanced thermal conductivity of TiO2 water
based nanofluids Int. J. Therm. Sci. 44, 367373 (2005)
11. Zhu, H., Zhang, C., Liu, S., Tang, Y., and Yin, Y.: Effects of nanoparticle clustering and alignment on thermal conductivities of Fe3 O4 aqueous nanofluids. Appl. Phys. Lett. 89, 023123
(2006)
12. Zhu, H.T., Zhang, C.Y., Tang, Y.M., and Wang, J.X.: Novel synthesis and thermal conductivity
of CuO nanofluid. J. Phys. Chem. C 111, 16461650 (2007)
13. Maxwell, J.C.: A Treatise on Electricity and Magnetism, II edn. Clarendon, Oxford (1881)
14. Nan, C.-W., Birringer, R., Clarke, D.R., and Gleiter, H.: Effective thermal conductivity of particulate composites with interfacial thermal resistance. J. Appl. Phys. 81, 66926699 (1997)
15. Keblinski, P., Eastman, J.A., and Cahill, D.G.: Nanofluids for thermal transport. Materials
Today 8, 3644 (2005)
16. Putnam, S.A., Cahill, D.G., Braun, P.V., Ge, Z., and Shimmin, R.G.: Thermal conductivity of
nanoparticle suspensions. J. Appl. Phys. 99, 084308 (2006)
17. Venerus, D.C., Kabadi, M.S., Lee, S., and Perez-Luna, V.: Study of thermal transport in
nanoparticle suspensions using forced Rayleigh scattering. J. Appl. Phys. 100, 094310 (2006)
18. Zhang, X., Gu, H., and Fujii, M.: Effective thermal conductivity and thermal diffusivity of
nanofluids containing spherical and cylindrical nanoparticles. J. Appl. Phys. 100, 044325
(2006)
19. Zhang, X., Gu, H., and Fujii, M.: Experimental study on the effective thermal conductivity
and thermal diffusivity of nanofluids. Int. J. Thermophysics 27, 569580 (2006)

220

Pawel Keblinski

20. Keblinski, P., Phillpot, S.R., Choi, S.U.S., and Eastman, J.A.: Mechanisms of heat flow in
suspensions of nano-sized particles (nanofluids). Int. J. Heat and Mass Transfer 45, 855863
(2002)
21. Jang, S.P., and Choi, S.U.S.: Role of Brownian motion in the enhanced thermal conductivity
of nanofluids. Appl. Phys. Lett. 84, 43164318 (2004)
22. Koo, J., and Kleinstreuer, C.: A new thermal conductivity model for nanofluids. J. Nanoparticle Res. 6, 577588 (2004)
23. Prasher, R., Bhattacharya, P., and Phelan, P.E.: Thermal conductivity of nanoscale colloidal
solutions (nanofluids). Phys. Rev. Lett. 94, 025901 (2005)
24. Evans, W., Fish, J., and Keblinski, P.: Role of Brownian motion hydrodynamics on nanofluid
thermal conductivity. Appl. Phys. Lett. 88, 093116 (2006)
25. Keblinski, P., and Thomin: Hydrodynamic field around a Brownian particle, Phys. Rev. E 73,
Rapid Communication 010502 (2006)
26. Vladkov, M., and Barrat, J.-L.: Modeling transient absorption and thermal conductivity in a
simple nanofluid. Nano Lett. 6, 12241228 (2006)
27. Rusconi, R., Rodari, E., and Piazza, R.: Optical measurements of the thermal properties of
nanofluids. Appl. Phys. Lett 89, 261916 (2006)
28. Eapen, J., Williams, W.C., Buongiorno, J., Hu, L.-W., Yip, S., Rusconi, R., and Piazza, R.:
Mean-field versus microconvection effects in nanofluid thermal conduction. Phys. Rev. Lett.
99, 095901 (2007)
29. Eapen, J., Buongiorno, J., Hu, L.-W., Yip, S., Rusconi, R., and Piazza, R.: Mean-field bounds
and the classical nature of thermal conduction in nanofluids. Manuscript under preparation
(2007)
30. Xue, L., Keblinski, P., Phillpot, S.R., Choi, S.U.S., and Eastman, J.A.: Effect of liquid layering
at the liquidsolid interface on thermal transport. Int. J. Heat and Mass Transfer 47, 42774283
(2004)
31. Eapen, J., Li, J., and Yip, S.: Beyond the Maxwell limit: Thermal conduction in nanofluids
with percolating fluid structures. In press, Phys. Rev. E. arXiv:0707.2164v1 (2007)
32. Eapen, J., Li, J., and Yip, S.: Mechanism of thermal transport in dilute nanocolloids. Phys.
Rev. Lett. 98, 028302 (2007)
33. Yu, C.-J., Richter, A.G., Kmetko, J., Dugan, S.W., Datta, A., and Dutta, P.: Structure of interfacial liquids: X-ray scattering studies. Phys. Rev. E 63, 021205 (2001)
34. Wilson, O.M., Hu, X., Cahill, D.G., and Braun, P.V.: Colloidal metal particles as probes of
nanoscale thermal transport in fluids. Phys. Rev. B 66, 224301 (2002)
35. Huxtable, S.T., Cahill, D.G., Shenogin, S., Xue, L., Ozisik, R., Barone, P., Usrey, M., Strano,
M.S., Siddons, G., Shim, M., and Keblinski, P.: Interfacial heat flow in carbon nanotube suspension. Nature Materials 2 (2003)
36. Shenogin, S,. Xu, L., Ozisik, R., Cahill, D., and Keblinski, P.: Role of thermal boundary resistance on the heat flow in carbon-nanotube composites. J. Appl. Phys. 95, 81368144 (2004)
37. Domingues, G., Volz, S., Joulain, K., and Greffet, J.-J.: Heat transfer between two nanoparticles through near field interaction. Phys. Rev. Lett. 94, 085901 (2005)
38. Ben-Abdallah, P.: Heat transfer through near-field interactions in nanofluids. Appl. Phys. Lett
89, 113117 (2006)
39. Hashin, Z., and Shtrikman, S.: A variational approach to the theory of the effective magnetic
permeability of multiphase materials. J. Appl. Phys. 33, 3125 (1962)
40. Landauer, R.: The electrical resistance of binary metallic mixtures. J. Appl. Phys. 23, 779784
(1952)
41. Kim, S.H., Choi, S.R., and Kim, D.: Thermal conductivity of metal-xxide nanofluids: Particle
size dependence and effect of laser irradiation. J. Heat Transfer 129, 298307 (2007)
42. Prasher, R., Song, D., Wang, J., and Phelan, P.: Measurements of nanofluid viscosity and its
implications for thermal applications. Appl. Phys. Lett. 89, 133108 (2006)
43. Prasher, R., Evans, W., Meakin, P., Fish, J., Phelan, P., and Keblinski, P.: Effect of aggregation
on thermal conduction in colloidal nanofluids. Appl. Phys. Lett. 89, 143119 (2006)

8 Thermal Conductivity of Nanofluids

221

44. Keblinski, P., Prasher, R. and Eapen, J.: Thermal conductance of nanofluids: Is the controversy over? J. Nanopart. Res. Published online: www.springerlink.com/content/
5w710jn586w38v25/ (2008)
45. Prasher, R., Phelan, P.E., and Bhattacharya, P.: Effect of aggregation kinetics on the thermal
conductivity of nanoscale colloidal solutions (nanofluid). Nano Lett. 6, 15291534 (2006)
46. Kwak, K., and Kim, C.: Viscosity and thermal conductivity of copper oxide nanofluid dispersed in ethylene glycol. KoreaAustralian Rheology Journal 17, 3540 (2005)
47. Eastman, J.A., Choi, S.U.S., Li, S., Thompson, L.J., and Lee, S.: Enhanced thermal conductivity through the development of nanofluids, 311. Materials Research Society (MRS): Fall
Meeting, Boston, USA (1997)
48. Every, A.G., Tzou, Y., Hasselmanan, D.P.H., and Raj, R.: The effect of particle size on the
thermal conductivity of ZnS/diamond composites. Acta Metallurgica et Materialia 40, 123
129 (1992)
49. Hong, T.K., Yang, H.S., and Choi, C.J.: Study of the enhanced thermal conductivity of Fe
nanofluids. J. Appl. Phys. 97, 064311 (2005)
50. Hwang, Y., Lee, J.K., Lee, C.H., Jung, Y.M., Cheong, S.I., Lee, C.G., Ku, B.C., and Jang,
S.P.: Stability and thermal conductivity characteristics of nanofluids. Thermochimica Acta
455, 7074 (2007)
51. Lee, S., Choi, S.U.S., Li, S., and Eastman, J.A.: Measuring thermal conductivity of fluids
containing oxide nanoparticles. J. Heat Transfer 121, 280289 (1999)
52. Li, C.H., and Peterson, G.P.: Experimental investigation of temperature and volume fraction
variations on the effective thermal conductivity of nanoparticle suspensions (nanofluids). J.
Appl. Phys. 99, 084314 (2006)
53. Masuda, H., Ebata, A., Teramae, K., and Hishinuma, N.: Alteration of thermal conductivity
and viscosity of liquid by dispersing ultra-fine particles (dispersion of -Al2 O3 , SiO2 , and
TiO2 ultra-fine particles). Netsu Bussei (Japan) 7, 227233 (1993)
54. Murshed, S.M.S., Leong, K.C., and Yang, C.: Determination of the effective thermal diffusivity of nanofluids by the double hot-wire technique. J. Phys. D Appl. Phys. 39, 53165322
(2006)
55. Shaikh, S., Lafdi, K., and Ponnappan, R.: Thermal conductivity improvement in carbon
nanoparticle doped PAO oil: An experimental study. J. Appl. Phys. 101, 064302 (2007)
56. Wen, D., and Ding, Y.: Effective thermal conductivity of aqueous suspensions of carbon nanotubes (carbon nanotube nanofluids). J. Thermophys. Heat Transfer 18, 481485 (2004)
57. Wen, D., and Ding, Y.: Experimental investigation into convective heat transfer of nanofluids
at the entrance region under laminar flow conditions. Int. J. Heat and Mass Transfer 47, 5181
5188 (2004)
58. Wen, D., and Ding, Y.: Natural convective heat transfer of suspensions of titanium dioxide
nanoparticles (nanofluids). IEEE Trans. Nanotech. 5, 220227 (2006)

Вам также может понравиться