Thermal Conductivity of Nanofluids

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 moder- ate increases consistent with the effective medium theories on well-dispersed con- ductive 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 man- ufacturing. 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 opti- cal 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 perfor- mance. 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 solid–liquid composite materials consisting of solid nanoparticles or nanofibers, with sizes typically on the order of 1–100 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 213 DOI 10.1007/978-3-642-04258-4 8, Springer-Verlag Berlin Heidelberg 2009

description

Thermal Conductivity of Nanofluids

Transcript of Thermal Conductivity of Nanofluids

Page 1: Thermal Conductivity of Nanofluids

Chapter 8Thermal Conductivity of Nanofluids

Pawel Keblinski

Nanofluids (colloidal suspensions of solid nanoparticles) sparked excitement as wellas controversy. In particular, a number of researches reported dramatic increases ofthermal conductivity with small nanoparticle loading, while others showed moder-ate increases consistent with the effective medium theories on well-dispersed con-ductive spheres. Here we discuss potential mechanisms that were put forward inorder to understand nanoflouid thermal conductivity and demonstrate that particleaggregation is currently the only physically reasonable mechanism that can explainthe majority of the experimental data.

8.1 Introduction

Cooling is one of the most important technical challenges facing numerous diverseindustries, including microelectronics, transportation, solid-state lighting, and man-ufacturing. Developments driving the increased thermal loads that require advancesin cooling include faster speeds (in the multi-GHz range) and smaller features (to< 100 nm) for microelectronic devices, higher power engines, and brighter opti-cal devices. The conventional method for increasing heat dissipation is to increasethe area available for exchanging heat with a heat transfer fluid, but this approachrequires an undesirable increase in the size of thermal management system. Thereis therefore an urgent need for new and innovative coolants with improved perfor-mance. About a decade ago a novel concept of nanofluids, i.e., heat transfer fluidscontaining suspensions of nanoparticles, was proposed as a means of meeting thesechallenges [1].

Nanofluids are solid–liquid composite materials consisting of solid nanoparticlesor nanofibers, with sizes typically on the order of 1–100 nm, suspended in a liquid.Nanofluids have attracted great interest recently due to reports of greatly enhancedthermal properties at low volume fractions. For example, a small amount (lessthan 1% volume fraction) of copper nanoparticles and carbon nanotubes dispersed

S. Volz (ed.), Thermal Nanosystems and Nanomaterials, Topics in Applied Physics, 118c

213DOI 10.1007/978-3-642-04258-4 8, © Springer-Verlag Berlin Heidelberg 2009

Page 2: Thermal Conductivity of Nanofluids

214 Pawel Keblinski

in ethylene glycol and oil, respectively, was reported to increase the inherently poorthermal conductivity of the liquid by 40% and 150%, respectively [2–5]. Conven-tional suspensions of well-dispersed particles require high concentrations (> 10%)of particles to achieve such enhancement; problems of rheology and stability, whichare amplified at high concentrations, preclude the widespread use of conventionalslurries as heat transfer fluids. In several cases, the observed enhancement in ther-mal conductivity of nanofluids is orders-of-magnitude larger than predicted by well-established theories of dispersed particles [6–12].

8.2 Excitement, Controversy, and New Physics

The large thermal conductivity enhancements reported by experiment led to excite-ment but also to controversy. The origin of the excitement was that the measuredthermal conductivity was often much larger than that predicted by well-establishedeffective medium theories under the assumption of well-dispersed particles. For asystem of well dispersed particles, in the limit of low particle volume fraction φ , andwith much higher particle conductivityΛNP than fluid conductivityΛfluid, the effec-tive medium (Maxwell) theory for vanishing interfacial thermal resistance leads tothe relation [13, 14]

ΛNF

Λfluid= 1 + 3φ . (8.1)

Equation (8.1) predicts only moderate thermal transport increase, independently ofthe conductivity of the filler. As already discussed, many experiments reported muchlarger increases.

This mismatch between predicted and observed values led not only to excitementbut also to controversy, since in some experiments, the measured thermal conduc-tivity was not very large and consistent with the prediction of (8.1) [15–19]. Thisinitially led to the belief that some experiments had to be wrong, particularly con-sidering that different research groups often obtained different results for what werepresumably the same nanofluids.

Further excitement and controversy were associated with potential mechanismsput forward to explain the ‘anomalous’ experimental observations. This discussionwas initiated by Keblinski et al. [20]. In the following we will review the key mech-anism analyzed and discussed in the literature.

8.2.1 Brownian Motion

Motivated by the ‘unusual’ thermal conductivity enhancements, a number of resear-chers promoted the concept of Brownian motion induced micro- or nanoconvec-tion [21–23]. They argued in different ways that each Brownian particle generates along-range velocity field in the surrounding fluid, akin to that present around a par-

Page 3: Thermal Conductivity of Nanofluids

8 Thermal Conductivity of Nanofluids 215

ticle moving with a constant velocity, which decays approximately as the inverse ofthe distance from the particle center. The ability of large volumes of fluid draggedby the nanoparticles to carry substantial amounts of heat was credited for the largethermal conductivity increases in nanofluids.

A key weakness of this argument is that the thermal diffusivity DT of the basefluid, which measures the rate of the heat flow via thermal conduction, is several or-ders of magnitude larger than nanoparticle diffusivity DNP, which measures the rateof mass motion due to nanoparticle diffusion, whence the magnitude of possiblenanoconvection effects is negligible [24]. Furthermore, the velocity field around aBrownian particle is much shorter-ranged than that around a particle moving with aconstant velocity [25]. The low estimates for the contribution to thermal transport ofBrownian motion were supported by the results of molecular dynamics simulations[24,26], and by the results of experimental measurements on well-dispersed spheri-cal nanoparticle fluids [16–19,27], 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 ex-perimentally tested with lighter silica and Teflon particles, and were shown to beincorrect [28]. The nanoconvection velocities were further shown to be of the or-der of thermophoretic velocities, which for most nanofluids were insignificant (aslow as 10−9 m/s). Even for sub-nanometer clusters, as evidenced from moleculardynamics simulations, the thermophoretic velocities are exceedingly small, and thiseffectively precludes a discernible contribution to the nanofluid thermal conductiv-ity 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 or-dered, 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 thosemechanisms do not contribute significantly to heat transfer [30]. For strong solid–liquid interactions, typical of those in nanofluids with metallic nanoparticles, a per-colating network of amorphous-like fluid structures can emerge which can facilitateadditional thermal conduction paths [31, 32]. However, a discernible increase inthermal conductivity is possible only for exceedingly small colloidal particles (lim-ited to few tens of atoms). Experiments have shown that the structured interfacialfluid layers are limited to a few molecular spacings from a solid surface [33], whichmakes them inadequate to influence the thermal transport in common nanofluids.Furthermore, interpretation of the cooling rates of Au nanoparticles suspended inwater and organic solvents does not appear to require any unusual thermophysicalproperties of the surrounding liquid to explain the experimental results [34].

Page 4: Thermal Conductivity of Nanofluids

216 Pawel Keblinski

8.2.3 Interfacial Thermal Resistance

As described above, the liquid layering does not generate thermal conductivity in-creases. There is, however, a well-known interfacial effect degrading thermal con-ductivity: the thermal interfacial resistance Rk, most commonly defined via its in-verse, the interfacial conductance G = 1/Rk. G is related to the heat flux JQ and thetemperature drop ΔT at the interface via

JQ = GΔT . (8.2)

A simple measure of the relative importance of the interfacial resistance in the over-all heat flow in a composite can be obtained from the equivalent thickness h, definedas the distance over which the temperature drop is the same as at the interface. Thisthickness is given by the ratio of the fluid thermal conductivity Λfluid to the interfa-cial conductance, i.e., h =Λfluid/G.

The interfacial thermal resistance effect on well-dispersed spherical particle com-posites can be estimated by effective medium theory to be [15]

ΛNF

Λfluid−1 = 3φ

γ−1γ+ 2

, (8.3)

where γ is the ratio of the particle radius to the equivalent matrix thickness h. Ac-cording to (8.3), when the particle radius becomes equal to the equivalent matrixthickness (γ = 1), there is no enhancement at all, while for larger interfacial resis-tance (γ < 1), the addition of particles decreases the thermal conductivity of thecomposite. This effect can explain the decrease of thermal conductivity of nanoflu-ids below the base fluid value [18, 19] and much lower than expected thermal con-ductivity increases of carbon nanotube nanofluids and carbon nanotube polymercomposites [35, 36].

8.2.4 Near Field Radiation

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

Page 5: Thermal Conductivity of Nanofluids

8 Thermal Conductivity of Nanofluids 217

8.2.5 Particle Clustering

Maxwell’s expression, the limiting case of which is given by (8.1), corresponds tothe situation of well-dispersed nanoparticles. Such structure is the least efficientfrom the point of view of thermal transport enhancement since conductive particlesare separated from each other by low conductivity liquid. Aggregation of particlesinto sparse clusters, or ideally into linear chains, leads to extended and highly con-ductive paths for the heat flow. In fact, there is a well established understanding ofthe impact of morphology on thermal conductivity of a composite. Depending onthe morphology, the conductivity can vary greatly, even at the same volume fractionof the components. It is generally accepted that the conductivity has to fall betweenso-called Hashin–Shtrikman (HS) bounds, obtained under an effective medium anal-ysis [39] without any restriction on the volume fraction. The HS bounds are givenby

Λfluid

[1+

3φ(ΛNP−Λfluid)3Λfluid +(1−φ)(ΛNP−Λfluid)

]≤ΛNF ≤

[1− 3(1−φ)(ΛNP−Λfluid)

3ΛNP−φ(ΛNP−Λfluid)

]ΛNP .

(8.4)In the inequality given by (8.4), the lower (Maxwell) bound corresponds physicallyto a set of well-dispersed nanoparticles in a fluid matrix, while the upper limit cor-responds to large pockets of fluid separated by linked, chain-forming or clusterednanoparticles. From the point of view of circuit analysis, the lower HS limit liescloser to conductors connected in series, while the upper limit lies closer to con-ductors in a parallel mode. The HS bounds do not give a precise mechanism ofthermal conductance, but set the most restrictive limits based on knowledge of thevolume fraction alone. It is relatively well-known that a large number of experi-mental 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 norparallel modes) predicts thermal conductivity values that lie between the upper andlower HS bounds [40].

The evidence from scanning electron microscopy (SEM) points to the existenceof partial clustering and chain-like linear aggregation [10–12, 41]. Viscosity dataon nanofluids has also shown an anomalous increase as compared to the Einsteinmodel for well dispersed particles [42]. A large increase in the viscosity is anotherindication of aggregation in the nanofluids. Interestingly, for a carbon nanotube sus-pension, an effective medium theory accounting for the very high aspect ratio of thenanotubes [14], predicts thermal conductivity enhancements that are in fact wellabove the reported values. This was attributed to a significant interfacial resistanceto heat flow between the carbon nanotubes and the fluid [35].

Page 6: Thermal Conductivity of Nanofluids

218 Pawel Keblinski

8.3 Discussion

Clear experimental evidence for extensive clustering observed in nanofluids indi-cates that the expected thermal conductivity can fall within a wide range boundedby (8.4) and strongly depends on the actual aggregation. Recently, using a multi-level effective medium theory, it was demonstrated that the thermal conductivity ofnanofluids can be significantly enhanced by the aggregation of nanoparticles intochain-like clusters, and this enhancement can be quite dramatic for large, but sparseclusters [43]. Predictions of this aggregation model were in excellent agreementwith detailed numerical calculations on model nanofluids involving fractal clus-ters [43]. Thus, allowing for clustering effects dramatically broadens the thermalconductivity range that is consistent with the effective medium theory [29]. Quitestrikingly, it was demonstrated that the vast majority of experimental results fallwithin these bounds, supporting the classical nature of thermal conduction behaviorin nanofluids [44].

The increase in the relative thermal conductivity of nanofluids with tempera-ture [4] is another example of anomalous behavior that cannot be explained on theassumption of well-dispersed particles. Similarly, the increase in the thermal con-ductivity with decreasing particle size cannot be explained if the particles are welldispersed. The probability of aggregation increases with increasing temperature anddecreasing particle size [45]. Therefore the apparent contradictions between exper-iment and theory, such as particle size effects, can be resolved by weighing in theability of nanoparticles to form linear clusters. Furthermore, the temperature depen-dence is not as strong as it was previously believed to be, with recent experimentsshowing a similar variation for both nanofluids and the base fluid [19, 29]. Thisimplies that the mechanism for increase in the thermal conductivity of water (pre-sumably from the hydrogen bonded structures) is partly responsible for the thermalconductivity increase in nanofluids as well. Conversely, it is reasonable to expect adecrease in the nanofluid thermal conductivity for a base fluid that has a negativechange in conductivity with increasing temperature.

It remains a challenge to accurately identify and manipulate the cluster config-uration to modify the thermal transport properties of a nanofluid. The two charac-terization techniques, DLS and SEM, have limitations in assessing the structure ofnanoparticles. DLS measurements are limited to dilute suspensions (φ < 1%) formost nanofluids, while SEM imaging can be performed only after drying the basefluid. While the science of making well-dispersed colloids has reached a fair levelof maturity, attempts to generate targeted nanoparticle configurations are still in anevolving phase.

The fact that significant aggregation is required to obtain substantial increases inthermal transport has an important consequence for the potential application of suchfluids in flow-based cooling, which is the most important benefit from the techno-logical point of view. It is well known that aggregation into sparse but large clustersincreases fluid viscosity. Such increases become very dramatic when the aggregatesstart to touch each other, which can occur at very low volume fractions, as low as0.2% [46]. Therefore, the same aggregate structures that are most beneficial to the

Page 7: Thermal Conductivity of Nanofluids

8 Thermal Conductivity of Nanofluids 219

thermal transport are also the most detrimental to the fluid flow characteristics. Fu-ture research should therefore address the issue of optimizing nanofluid structurewith 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. 99–105

2. Choi, S.U.S., Zhang, Z.G., Yu, W., Lockwood, F.E., and Grulke, E.A.: Anomalous thermalconductivity enhancement in nanotube suspensions. Appl. Phys. Lett. 79, 2252–2254 (2001)

3. Eastman, J.A., Choi, S.U.S., Li, S., Yu, W., and Thompson, L.J.: Anomalously increased ef-fective thermal conductivities of ethylene glycol-based nanofluids containing copper nanopar-ticles. Appl. Phys. Lett. 78, 718–720 (2001)

4. Das, S.K., Putra, N., Thiesen, P., and Roetzel, W.: Temperature dependence of thermal con-ductivity enhancement for nanofluids. J. Heat Transfer 125, 567–574 (2003)

5. Patel, H.E., Das, S.K., Sundararajan, T., Nair, A.S., George, B., and Pradeep, T.: Thermalconductivities of naked and monolayer protected metal nanoparticle based nanofluids: Man-ifestation of anomalous enhancement and chemical effects. Appl. Phys. Lett. 83, 2931–2933(2003)

6. Chopkar, M., Das, P.K., and Manna, I.: Synthesis and characterization of nanofluid for ad-vanced heat transfer applications. Scripta Materialia 55, 549–552 (2006)

7. Chopkar, M., Kumar, S., Bhandari, D.R., Das, P.K., and Manna, I.: Development and char-acterization of Al2Cu and Ag2Al nanoparticle dispersed water and ethylene glycol basednanofluid. Materials Science and Engineering: B 139, 141–148 (2007)

8. Hong, K.S., Hong, T.-K., and Yang, H.-S.: Thermal conductivity of Fe nanofluids dependingon 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 usingexperimental effective particle volume. Experimental Heat Transfer 19, 181–191 (2006)

10. Murshed, S.M.S., Leong, K.C., and Yang, C.: Enhanced thermal conductivity of TiO2–waterbased nanofluids Int. J. Therm. Sci. 44, 367–373 (2005)

11. Zhu, H., Zhang, C., Liu, S., Tang, Y., and Yin, Y.: Effects of nanoparticle clustering and align-ment on thermal conductivities of Fe3O4 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 conductivityof CuO nanofluid. J. Phys. Chem. C 111, 1646–1650 (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 par-

ticulate composites with interfacial thermal resistance. J. Appl. Phys. 81, 6692–6699 (1997)15. Keblinski, P., Eastman, J.A., and Cahill, D.G.: Nanofluids for thermal transport. Materials

Today 8, 36–44 (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 conductivityand thermal diffusivity of nanofluids. Int. J. Thermophysics 27, 569–580 (2006)

Page 8: Thermal Conductivity of Nanofluids

220 Pawel Keblinski

20. Keblinski, P., Phillpot, S.R., Choi, S.U.S., and Eastman, J.A.: Mechanisms of heat flow insuspensions of nano-sized particles (nanofluids). Int. J. Heat and Mass Transfer 45, 855–863(2002)

21. Jang, S.P., and Choi, S.U.S.: Role of Brownian motion in the enhanced thermal conductivityof nanofluids. Appl. Phys. Lett. 84, 4316–4318 (2004)

22. Koo, J., and Kleinstreuer, C.: A new thermal conductivity model for nanofluids. J. Nanoparti-cle Res. 6, 577–588 (2004)

23. Prasher, R., Bhattacharya, P., and Phelan, P.E.: Thermal conductivity of nanoscale colloidalsolutions (nanofluids). Phys. Rev. Lett. 94, 025901 (2005)

24. Evans, W., Fish, J., and Keblinski, P.: Role of Brownian motion hydrodynamics on nanofluidthermal 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 asimple nanofluid. Nano Lett. 6, 1224–1228 (2006)

27. Rusconi, R., Rodari, E., and Piazza, R.: Optical measurements of the thermal properties ofnanofluids. 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 boundsand 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 layeringat the liquid–solid interface on thermal transport. Int. J. Heat and Mass Transfer 47, 4277–4283(2004)

31. Eapen, J., Li, J., and Yip, S.: Beyond the Maxwell limit: Thermal conduction in nanofluidswith 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 inter-facial 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 ofnanoscale 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 sus-pension. Nature Materials 2 (2003)

36. Shenogin, S,. Xu, L., Ozisik, R., Cahill, D., and Keblinski, P.: Role of thermal boundary resis-tance on the heat flow in carbon-nanotube composites. J. Appl. Phys. 95, 8136–8144 (2004)

37. Domingues, G., Volz, S., Joulain, K., and Greffet, J.-J.: Heat transfer between two nanoparti-cles through near field interaction. Phys. Rev. Lett. 94, 085901 (2005)

38. Ben-Abdallah, P.: Heat transfer through near-field interactions in nanofluids. Appl. Phys. Lett89, 113117 (2006)

39. Hashin, Z., and Shtrikman, S.: A variational approach to the theory of the effective magneticpermeability of multiphase materials. J. Appl. Phys. 33, 3125 (1962)

40. Landauer, R.: The electrical resistance of binary metallic mixtures. J. Appl. Phys. 23, 779–784(1952)

41. Kim, S.H., Choi, S.R., and Kim, D.: Thermal conductivity of metal-xxide nanofluids: Particlesize dependence and effect of laser irradiation. J. Heat Transfer 129, 298–307 (2007)

42. Prasher, R., Song, D., Wang, J., and Phelan, P.: Measurements of nanofluid viscosity and itsimplications 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 aggregationon thermal conduction in colloidal nanofluids. Appl. Phys. Lett. 89, 143119 (2006)

Page 9: Thermal Conductivity of Nanofluids

8 Thermal Conductivity of Nanofluids 221

44. Keblinski, P., Prasher, R. and Eapen, J.: Thermal conductance of nanofluids: Is the contro-versy 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 thermalconductivity of nanoscale colloidal solutions (nanofluid). Nano Lett. 6, 1529–1534 (2006)

46. Kwak, K., and Kim, C.: Viscosity and thermal conductivity of copper oxide nanofluid dis-persed in ethylene glycol. Korea–Australian Rheology Journal 17, 35–40 (2005)

47. Eastman, J.A., Choi, S.U.S., Li, S., Thompson, L.J., and Lee, S.: Enhanced thermal conduc-tivity through the development of nanofluids, 3–11. Materials Research Society (MRS): FallMeeting, Boston, USA (1997)

48. Every, A.G., Tzou, Y., Hasselmanan, D.P.H., and Raj, R.: The effect of particle size on thethermal 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 Fenanofluids. 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 Acta455, 70–74 (2007)

51. Lee, S., Choi, S.U.S., Li, S., and Eastman, J.A.: Measuring thermal conductivity of fluidscontaining oxide nanoparticles. J. Heat Transfer 121, 280–289 (1999)

52. Li, C.H., and Peterson, G.P.: Experimental investigation of temperature and volume fractionvariations 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 conductivityand viscosity of liquid by dispersing ultra-fine particles (dispersion of γ-Al2O3, SiO2, andTiO2 ultra-fine particles). Netsu Bussei (Japan) 7, 227–233 (1993)

54. Murshed, S.M.S., Leong, K.C., and Yang, C.: Determination of the effective thermal diffu-sivity of nanofluids by the double hot-wire technique. J. Phys. D Appl. Phys. 39, 5316–5322(2006)

55. Shaikh, S., Lafdi, K., and Ponnappan, R.: Thermal conductivity improvement in carbonnanoparticle 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 nan-otubes (carbon nanotube nanofluids). J. Thermophys. Heat Transfer 18, 481–485 (2004)

57. Wen, D., and Ding, Y.: Experimental investigation into convective heat transfer of nanofluidsat 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 dioxidenanoparticles (nanofluids). IEEE Trans. Nanotech. 5, 220–227 (2006)