You are on page 1of 5

Energy Conversion and Management 52 (2011) 789–793

Contents lists available at ScienceDirect

Energy Conversion and Management


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

Technical Note

Empirical correlating equations for predicting the effective thermal conductivity


and dynamic viscosity of nanofluids
Massimo Corcione *
Dipartimento di Fisica Tecnica, Sapienza University of Rome, via Eudossiana 18, 00184 Rome, Italy

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

Article history: In this paper, two empirical correlations for predicting the effective thermal conductivity and dynamic
Received 6 November 2009 viscosity of nanofluids, based on a high number of experimental data available in the literature, are pro-
Received in revised form 11 June 2010 posed and discussed. It is found that, given the nanoparticle material and the base fluid, the ratio between
Accepted 29 June 2010
the thermal conductivities of the nanofluid and the pure base liquid increases as the nanoparticle volume
Available online 3 August 2010
fraction and the temperature are increased, and the nanoparticle diameter is decreased. Additionally, also
the ratio between the dynamic viscosities of the nanofluid and the pure base liquid increases as the nano-
Keywords:
particle volume fraction is increased, and the nanoparticle diameter is decreased, being practically inde-
Nanofluids
Thermal conductivity
pendent of temperature. The ease of application of the equations proposed, and their wide regions of
Dynamic viscosity validity (the ranges of the nanoparticle diameter, volume fraction and temperature are 10–150 nm,
Empirical correlations 0.002–0.09 and 294–324 K for the thermal conductivity data, and 25–200 nm, 0.0001–0.071 and 293–
323 K for the dynamic viscosity data), make such equations useful by the engineering point of view,
for both numerical simulation purposes and thermal design tasks.
Ó 2010 Elsevier Ltd. All rights reserved.

1. Introduction of a number of new theoretical models for the evaluation of the


effective thermal conductivity of nanofluids, that basically account
Nanofluids are a new type of heat transfer fluids obtained by for the effects of the phenomena occurring at the solid/liquid inter-
suspending nano-sized particles into a base liquid. The term nano- face and/or the micro-mixing convection caused by the Brownian
fluid was coined in 1995 by Choi [1], who showed that the uniform motion of the nanoparticles, which is the case of the models pro-
dispersion of low concentrations of nanoparticles into a traditional posed by Yu and Choi [9], Xue [10], Kumar et al. [11], Koo and
liquid such as water, oil, and ethylene glycol, could noticeably im- Kleinstreuer [12], Jang and Choi [13,14], Xie et al. [15], Patel et al.
prove its thermal performance. Since then, owing to the potential [16], Ren et al. [17], Prasher et al. [18,19], Leong et al. [20], Xuan
of their impact upon several industrial sectors, nanofluids have at- et al. [21], Prakash and Giannelis [22], and Murshed et al. [23].
tracted the interest of an increasing number of scientists, as clearly However, these models show large discrepancies among each
reflected by the significant research effort dedicated to this topic, other, which clearly represents a restriction to their applicability.
whose main discoveries are summarized in the recent review-pa- Moreover, most of these models include empirical constants of pro-
pers written by Wang and Mujumdar [2], Trisaksri and Wongwises portionality whose values were often determined on the basis of a
[3], Daungthongsuk and Wongwises [4], and Murshed et al. [5]. limited number of experimental data, or were not clearly defined.
One of the major outcomes emerging from a thorough analysis Indeed, the thermal conductivity enhancement is not the only
of the available literature is that in most cases the models originally noteworthy effect originating from the suspension of nanoparticles
developed for composites and mixtures with micro-sized and milli- into a base fluid. In fact, a contemporary growth in dynamic viscos-
sized inclusions – namely, the models developed by Maxwell [6], ity occurs, which could be a serious limitation, either in terms of an
Hamilton and Crosser [7], and Davis [8] – fail dramatically in pre- exaggerated pressure drop increase in forced convection applica-
dicting the anomalously increased thermal conductivity of nano- tions, or in terms of a drastic fluid motion decrease in natural con-
particle suspensions (at least when the nanofluid temperature is vection situations. Accordingly, the possibility of calculating the
one or some tens degrees higher than ‘‘room” temperature), likely effective dynamic viscosity of a nanoparticle suspension seems
because such traditional models include only the effect of the crucial to establish if its use is actually advantageous with respect
nanoparticle concentration. This has motivated the development to the pure base liquid. In spite of this, leaving aside the theories
developed time ago for traditional colloid dispersions by Einstein
* Tel.: +39 06 44 58 54 43; fax: +39 06 48 80 120.
[24,25], Brinkman [26], Lundgren [27], and Batchelor [28], whose
E-mail address: massimo.corcione@uniroma1.it predictions typically under-estimate the dynamic viscosity of

0196-8904/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.enconman.2010.06.072
790 M. Corcione / Energy Conversion and Management 52 (2011) 789–793

Nomenclature

cf specific heat of the base fluid, J kg1 K1 Re Brownian-motion Reynolds number defined in Eq. (2)
D Brownian diffusion coefficient = kbT/(3plf dp), m2 s1 T temperature, K
df equivalent diameter of a base fluid molecule, m Tfr freezing point of the base fluid, K
dp diameter of the nanoparticle, m uB Brownian velocity of the nanoparticle, m s1
kb Boltzmann’s constant = 1.38066  1023 J K1
keff effective thermal conductivity of the nanofluid, Greek symbols
W m1 K1 u volume fraction of the nanoparticles
keff(M) effective thermal conductivity of the nanofluid accord- leff effective dynamic viscosity of the nanofluid, N m2 s
ing to Maxwell, W m1 K1 leff(B) effective dynamic viscosity of the nanofluid according to
kf thermal conductivity of the base fluid, W m1 K1 Brinkman, N m2 s
kp thermal conductivity of the nanoparticle, W m1 K1 lf dynamic viscosity of the base fluid, N m2 s
M molecular weight of the base fluid, kg mol1 qf mass density of the base fluid, kg m3
N Avogadro number = 6.022  1023 mol1 sD time required for a nanoparticle to move by a distance
Pr Prandtl number of the base fluid = cf lf/kf equal to its diameter, s

nanoparticle suspensions, only few models have recently been pro- dence, or the investigation procedure was not properly described
posed for describing the rheological behaviour of nanofluids, such in detail, or specific chemical dispersants/surfactants were used
as those developed by Koo [29], and Masoumi et al. [30]. Actually, in experiments, which could have significantly altered the ther-
as these models contain empirical correction factors based on an mo-mechanical behaviour of the suspension. Interestingly, after
extremely small number of experimental data, their region of these ‘‘spurious” results have been laid aside, which has brought
validity is someway limited. to the selection of data listed above, the degree of spreading among
Framed in this general background, the aim of the present paper the data has significantly reduced.
is to introduce and discuss two easy-to-apply empirical correlating By way of regression analysis, the following mean empirical
equations for predicting the effective thermal conductivity and dy- correlation with a 1.86% standard deviation of error is produced
namic viscosity of nanofluids, that, matching sufficiently well a (see Fig. 1):
high number of experimental data available in the literature,  10  0:03
keff T kp
may be usefully employed for numerical simulation purposes ¼ 1 þ 4:4Re0:4 Pr0:66 u0:66 ; ð1Þ
and thermal engineering design tasks. kf T fr kf
where Re is the nanoparticle Reynolds number, Pr is the Prandtl
2. Correlation for the effective thermal conductivity number of the base liquid, T is the nanofluid temperature, Tfr is
the freezing point of the base liquid, kp is the nanoparticle thermal
The equation proposed for the nanofluid effective thermal con- conductivity, and u is the volume fraction of the suspended
ductivity, keff, normalized by the thermal conductivity of the base nanoparticles.
fluid, kf, is derived from a wide variety of experimental data rela- In more detail, the nanoparticle Reynolds number is defined as
tive to nanofluids consisting of alumina, copper oxide, titania and qf uB dp
copper nanoparticles with a diameter in the range between Re ¼ ; ð2Þ
10 nm and 150 nm, suspended in water or ethylene glycol (EG).
lf
These data are extracted from the following sources: Masuda
Correlation data of the dimensionless effective thermal conductivity

et al. [31] for TiO2(27 nm) + H2O; Pak and Cho [32] for 1.5
TiO2(27 nm) + H2O; Lee et al. [33] for CuO(23.6 nm) + H2O, Water-based nanofluids.
Experimental data by Masuda et al. [31],
CuO(23.6 nm) + EG, Al2O3(38.4 nm) + H2O, and Al2O3(38.4 nm) + Pak and Cho [32], Lee et al. [33], Das et al. [35],
EG; Eastman et al. [34] for Cu(10 nm) + EG; Das et al. [35] for Chon et al. [36], Chon and Kihm [37],
CuO(28.6 nm) + H2O, and Al2O3(38.4 nm) + H2O; Chon et al. [36] 1.4 Xuan et al. [21], Murshed et al. [38],
Mintsa et al. [39], and
for Al2O3(47 nm) + H2O; Chon and Kihm [37] for Al2O3(47 nm Duangthongsuk and Wongwises [40].
and 150 nm) + H2O; Xuan et al. [21] for TiO2(27 nm) + H2O;
Murshed et al. [38] for Al2O3(80 nm) + H2O, and Al2O3(80 nm) +
1.3
EG; Mintsa et al. [39] for CuO(29 nm) + H2O; and Duangthongsuk
and Wongwises [40] for TiO2(21 nm) + H2O. The nanoparticle
volume fractions are in the range from 0.002 to 0.09, whereas
temperatures lie between 294 K and 324 K. 1.2
It seems worth pointing out that a certain dispersion of the
experimental data reported by different authors for the same type
of nanofluid is unavoidable. Indeed, in some cases the discrepan-
cies among the data available in the literature is of the order of 1.1 EG-based nanofluids.
50% or more, which may be ascribed to the different measurement Experimental data by
techniques used in experiments, as well as to the different degrees Lee et al. [33], Eastman et al. [34],
and Murshed et al. [38].
of dispersion/agglomeration obtained for the suspended nanopar-
ticles, and the accuracy of evaluation of their shape and size. In this 1.0
regard, it must be said that, in deriving the correlation proposed 1.0 1.1 1.2 1.3 1.4 1.5
Experimental data of the dimensionless effective thermal conductivity
here, some data-sets found in the literature have been discarded
because either the data were in a too sharp contrast with the main Fig. 1. Comparison of the empirical correlation Eq. (1) with the experimental data
body of the available results without any convincing physical evi- from literature.
M. Corcione / Energy Conversion and Management 52 (2011) 789–793 791

where qf and lf are the mass density and the dynamic viscosity of 3. Correlation for the effective dynamic viscosity
the base fluid, respectively, and dp and uB are the nanoparticle diam-
eter and mean Brownian velocity, respectively. Assuming absence As for the thermal conductivity, also the equation proposed for
of agglomeration, the nanoparticle Brownian velocity uB is calcu- the nanofluid effective dynamic viscosity, leff, normalized by the
lated as the ratio between dp and the time sD required to cover such dynamic viscosity of the base liquid, lf, is derived from a wide
distance, that, according to Keblinski et al. [41], is selection of experimental data available in the literature. These
2 3
data, relative to nanofluids consisting of alumina, titania, silica
dp plf dp and copper nanoparticles with a diameter ranging between
sD ¼ ¼ ; ð3Þ
6D 2kb T 25 nm and 200 nm, suspended in water, ethylene glycol (EG), pro-
pylene glycol (PG) or ethanol (Eth), are taken out of the following
where D is the Einstein diffusion coefficient and kb is the Boltz- sources: Masuda et al. [31] for TiO2(27 nm) + H2O, Pak and Cho
mann’s constant. Hence [32] for TiO2(27 nm) + H2O, Wang et al. [42] for Al2O3(28 nm) +
2kb T H2O, Das and co-workers [43,44] for Al2O3(38 nm) + H2O, Prasher
uB ¼ : ð4Þ et al. [45] for Al2O3(27 nm, 40 nm, and 50 nm) + PG, He et al. [46]
plf d2p for TiO2(95 nm) + H2O, Chen et al. [47,48] for TiO2(25 nm) + EG,
If we substitute Eq. (4) in Eq. (2), we obtain Chevalier et al. [49] for SiO2(35 nm, 94 nm, and 190 nm) + Eth,
Lee et al. [50] for Al2O3(30 nm) + H2O, and Garg et al. [51] for
2qf kb T Cu(200 nm) + EG. The nanoparticle volume fractions are in the
Re ¼ : ð5Þ
pl2f dp range from 0.0001 to 0.071. Temperatures lie between 293 K and
333 K.
Note that in the preceding equations all the physical properties The best-fit of the selected data enumerated above results in the
are calculated at the nanofluid temperature T. following mean empirical correlation with a 1.84% standard devia-
It is apparent that, once the nanoparticle material and the base tion of error (see Fig. 3):
fluid are assigned, the dimensionless effective thermal conductiv-
ity of the nanofluid, keff/kf, increases as u and T increase, and dp de-
leff 1
¼ ; ð7Þ
creases. Moreover, keff/kf depends marginally on the solid–liquid lf 1  34:87ðdp =df Þ0:3 u1:03
combination (as denoted by the small exponent of the particle–
where df is the equivalent diameter of a base fluid molecule, given
fluid thermal conductivity ratio), which, among other reasons,
by
may be due to the fact that most experimental data used to derive
!1=3
Eq. (1) are related to nanofluids with suspended oxide nanoparti- 6M
cles, whose thermal conductivities have same order of magnitude. df ¼ 0:1 ; ð8Þ
Npqf 0
In Fig. 2, the distributions of keff/kf vs. u that emerge from Eq. (1)
for Al2O3 + H2O, with dp and T as parameters, are compared with in which M is the molecular weight of the base fluid, N is the Avo-
the predictions of the Maxwell theory [6]: gadro number, and qf0 is the mass density of the base fluid calcu-
lated at temperature T0 = 293 K.
keff ðMÞ kp þ 2kf  2uðkf  kp Þ
¼ : ð6Þ It may be observed that, once the base fluid is assigned, the
kf kp þ 2kf þ uðkf  kp Þ
dimensionless effective dynamic viscosity of the nanofluid, leff/lf,
It may be noticed that the traditional Maxwell theory largely increases as dp decreases and u increases. Another interesting fea-
fails when applied to nanofluids. In fact, the Maxwell equation ture is that, within the approximation of Eq. (7), leff/lf is indepen-
tends either to under-estimate or to over-estimate the value of keff, dent of both the solid–liquid combination and the temperature, at
according as the nanoparticle diameter is small or large, respec-
tively, and the temperature of the suspension is high or low,
2.4
respectively. However, since the use of tiny nanoparticles is the
Correlation data of the dimensionless effective dynamic viscosity

Ethanol-based nanofluids.
rule, the adoption of Eq. (6) usually implies a more or less severe Experimental data by Chevalier et al. [49].
under-estimation of the nanofluid effective thermal conductivity. 2.2
Water-based nanofluids.
Experimental data by Masuda et al.
1.6 [31], Pak and Cho [32], Wang et al. [42],
2.0 Das et al. [43], Putra et al. [44],
Dimensionless effective thermal conductivity

dp = 20 nm, T = 324 K
He et al. [46], and Lee et al. [50].
1.5 dp = 100 nm, T = 324 K
1.8
dp = 20 nm, T = 294 K
1.4 dp = 100 nm, T = 294 K
1.6 PG-based nanofluids.
Maxwell equation
1.3 Experimental data by
Prasher et al. [45].
1.4
1.2
EG-based nanofluids.
Experimental data by
1.1 1.2 Chen et al. [47, 48],
and Garg et al.[51].
Al2O3 + H2O
1.0 1.0
0 0.01 0.02 0.03 0.04 0.05 0.06 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4
Volume fraction Experimental data of the dimensionless effective dynamic viscosity

Fig. 2. Distributions of keff(M)/kf and keff/kf vs. u for Al2O3 + H2O, with dp and T as Fig. 3. Comparison of the empirical correlation Eq. (7) with the experimental data
parameters. from literature.
792 M. Corcione / Energy Conversion and Management 52 (2011) 789–793

2.2 [2] Wang X-Q, Mujumdar AS. Heat transfer characteristics of nanofluids: a review.
d p = 25 nm water-based nanofluids Int J Therm Sci 2007;46:1–19.
Dimensionless effective dynamic viscosity

[3] Trisaksri V, Wongwises S. Critical review of heat transfer characteristics of


2.0 d p = 50 nm nanofluids. Renew Sust Energy Rev 2007;11:512–23.
[4] Daungthongsuk W, Wongwises S. A critical review of convective heat transfer
d p = 100 nm in nanofluids. Renew Sust Energy Rev 2007;11:797–817.
1.8 [5] Murshed SMS, Leong KC, Yang C. Thermophysical and electrokinetic properties
d p = 200 nm
of nanofluids – a critical review. Appl Therm Eng 2008;28:2109–25.
Brinkman equation [6] Maxwell JC. A treatise on electricity and magnetism. 3rd ed. New York: Dover;
1.6 1954.
[7] Hamilton RL, Crosser OK. Thermal conductivity of heterogeneous two
component systems. Ind Eng Chem Fund 1962;1:187–91.
1.4 [8] Davis RH. The effective thermal conductivity of a composite material with
spherical inclusions. Int J Thermophys 1986;7:609–20.
[9] Yu W, Choi SUS. The role of interfacial layers in the enhanced thermal
1.2 conductivity of nanofluids: a renovated Maxwell model. J Nanopart Res
2003;5:167–71.
[10] Xue Q-Z. Model for effective thermal conductivity of nanofluids. Phys Lett A
2003;307:313–7.
1.0
0 0.01 0.02 0.03 0.04 0.05 0.06 [11] Kumar DH, Patel HE, Kumar VRR, Sundararajan T, Pradeep T, Das SK. Model for
volume fraction φ heat conduction in nanofluids. Phys Rev Lett 2004;93:144301.
[12] Koo J, Kleinstreuer C. A new thermal conductivity model for nanofluids. J
Fig. 4. Distributions of leff(B)/lf and leff/lf vs. u for water-based nanofluids, with dp Nanopart Res 2004;6:577–88.
[13] Jang SP, Choi SUS. Role of Brownian motion in the enhanced thermal
as a parameter.
conductivity of nanofluids. Appl Phys Lett 2004;84:4316–8.
[14] Jang SP, Choi SUS. Effects of various parameters on nanofluid thermal
conductivity. J Heat Trans 2007;129:617–23.
least for particle volume fractions not too high and temperatures [15] Xie H, Fujii M, Zhang X. Effect of interfacial nanolayer on the effective thermal
not too far from ‘‘room” temperature (obviously, the evaluation conductivity of nanoparticle-fluid mixture. Int J Heat Mass Trans
2005;48:2926–32.
of the effective dynamic viscosity leff at a given temperature T re- [16] Patel HE, Sundararajan T, Pradeep T, Dasgupta A, Dasgupta N, Das SK. A micro-
quires that lf is calculated at such temperature T). convection model for the thermal conductivity of nanofluids. Pramana – J Phys
In Fig. 4, the distributions of leff/lf vs. u that emerge from Eq. 2005;65:863–9.
[17] Ren Y, Xie H, Cai A. Effective thermal conductivity of nanofluids containing
(7) for different values of dp are compared with the predictions spherical nanoparticles. J Phys D: Appl Phys 2005;38:3958–61.
of the Brinkman equation [26]: [18] Prasher R, Bhattacharya P, Phelan PE. Thermal conductivity of nanoscale
colloidal solutions (nanofluids). Phys Rev Lett 2005;94:025901.
leff ðBÞ 1 [19] Prasher R, Bhattacharya P, Phelan PE. Brownian motion-based convective–
¼ : ð9Þ
lf ð1  uÞ2:5 conductive model for the effective thermal conductivity of nanofluids. J Heat
Trans 2006;128:588–95.
[20] Leong KC, Yang C, Murshed SMS. A model for the thermal conductivity of
As observed earlier for the Maxwell theory, also the Brinkman
nanofluids – the effect of interfacial layer. J Nanopart Res 2006;8:245–54.
equation largely fails when applied to nanofluids, with a percent- [21] Xuan Y, Li Q, Zhang X, Fujii M. Stochastic thermal transport of nanoparticle
age error that increases as the nanoparticle diameter decreases. suspensions. J Appl Phys 2006;100:043507.
[22] Prakash M, Giannelis EP. Mechanism of heat transport in nanofluids. J
Computer-Aided Mater Des 2007;14:109–17.
4. Conclusions [23] Murshed SMS, Leong KC, Yang C. A combined model for the effective thermal
conductivity of nanofluids. Appl Therm Eng 2009;29:2477–83.
[24] Einstein A. Eine neue bestimmung der molekul-dimension (A new
In sum, in this paper two empirical equations for predicting the determination of the molecular dimensions). Ann Phys 1906;19:289–306.
effective thermal conductivity and dynamic viscosity of nanofluids, [25] Einstein A. Berichtigung zu meiner arbeit: Eine neue bestimmung der
based on a high number of experimental data available in the liter- molekul-dimension (Correction of my work: a new determination of the
molecular dimensions). Ann Phys 1911;34:591–2.
ature, have been proposed and discussed. According to the best-fit [26] Brinkman HC. The viscosity of concentrated suspensions and solutions. J Chem
of the literature data, once the nanoparticle material and the base Phys 1952;20:571.
fluid are assigned, the ratio between the thermal conductivities of [27] Lundgren T. Slow flow through stationary random beds and suspensions of
spheres. J Fluid Mech 1972;51:273–99.
the nanofluid and the pure base liquid increases as the nanoparti- [28] Batchelor G. The effect of Brownian motion on the bulk stress in a suspension
cle volume fraction and the temperature are increased, and the of spherical particles. J Fluid Mech 1977;83:97–117.
nanoparticle diameter is decreased. Moreover, also the ratio be- [29] Koo J. Computational nanofluid flow and heat transfer analyses applied to
micro-systems. Dissertation Thesis, North Carolina State University, Rayleigh,
tween the dynamic viscosities of the nanofluid and the pure base
NC; 2005.
liquid increases as the nanoparticle volume fraction is increased, [30] Masoumi N, Sohrabi N, Behzadmehr A. A new model for calculating the
and the nanoparticle diameter is decreased, being substantially effective viscosity of nanofluids. J Phys D: Appl Phys 2009;42:055501.
independent of temperature; additionally, the solid–liquid combi- [31] Masuda H, Ebata A, Teramae K, Hishinuma N. Alteration of thermal
conductivity and viscosity of liquid by dispersing ultra-fine particles
nation seems to have no evident effect on the rheological behav- (Dispersion of c-Al2O3, SiO2, and TiO2 ultra-fine particles). Netsu Bussei
iour of nanofluids. The comparison between the predictions 1993;4:227–33.
emerging from the correlations proposed for the dimensionless [32] Pak BC, Cho YI. Hydrodynamic and heat transfer study of dispersed fluids with
submicron metallic oxide particles. Exp Heat Trans 1998;11:151–70.
thermal conductivity and dynamic viscosity and the Maxwell and [33] Lee S, Choi SUS, Li S, Eastman JA. Measuring thermal conductivity of fluids
Brinkman equations, very often used in numerical studies on nano- containing oxide nanoparticles. J Heat Trans 1999;121:280–9.
fluids, shows that the traditional theories fail abundantly when [34] Eastman JA, Choi SUS, Li S, Yu W, Thompson LJ. Anomalously increased
effective thermal conductivity of ethylene glycol-based nanofluids containing
employed for nanofluids. Owing to their sufficiently wide regions copper nanoparticles. Appl Phys Lett 2001;78:718–20.
of applicability, the correlations developed are believed to repre- [35] Das SK, Putra N, Thiesen P, Roetzel W. Temperature dependence of thermal
sent an useful engineering tool for analysis and thermal design conductivity enhancement for nanofluids. J Heat Trans 2003;125:567–74.
[36] Chon CH, Kihm KD, Lee SP, Choi SUS. Empirical correlation finding the role of
applications. temperature and particle size for nanofluid (Al2O3) thermal conductivity
enhancement. Appl Phys Lett 2005;87:153107.
References [37] Chon CH, Kihm KD. Thermal conductivity enhancement of nanofluids by
Brownian motion. J Heat Trans 2005;127:810.
[38] Murshed SMS, Leong KC, Yang C. Investigations of thermal conductivity and
[1] Choi SUS. Enhancing thermal conductivity of fluids with nanoparticles. In:
viscosity of nanofluids. Int J Therm Sci 2008;47:560–8.
Siginer DA, Wang HP, editors. Developments and applications of non-
[39] Mintsa HA, Roy G, Nguyen CT, Doucet D. New temperature dependent thermal
Newtonian flows, vols. 31/MD–66. New York, NY, FED: ASME Publ.; 1995. p.
conductivity data for water-based nanofluids. Int J Therm Sci 2009;48:363–71.
99–105.
M. Corcione / Energy Conversion and Management 52 (2011) 789–793 793

[40] Duangthongsuk W, Wongwises S. Measurement of temperature-dependent [46] He Y, Jin Y, Chen H, Ding Y, Cang D, Lu H. Heat transfer and flow behaviour of
thermal conductivity and viscosity of TiO2-water nanofluids. Exp Therm Fluid aqueous suspensions of TiO2 nanoparticles (nanofluids) flowing upward
Sci 2009;33:706–14. through a vertical pipe. Int J Heat Mass Trans 2007;50:2272–81.
[41] Keblinski P, Phillpot SR, Choi SUS, Eastman JA. Mechanisms of heat flow in [47] Chen H, Ding Y, Tan C. Rheological behaviour of nanofluids. New J Phys
suspensions of nano-sized particles (nanofluids). Int J Heat Mass Trans 2007;9:367.
2002;45:855–63. [48] Chen H, Ding Y, He Y, Tan C. Rheological behaviour of ethylene glycol based
[42] Wang X, Xu X, Choi SUS. Thermal conductivity of nanoparticle-fluid mixture. J titania nanofluids. Chem Phys Lett 2007;444:333–7.
Thermophys Heat Trans 1999;13:474–80. [49] Chevalier J, Tillement O, Ayela F. Rheological properties of nanofluids flowing
[43] Das SK, Putra N, Roetzel W. Pool boiling characteristics of nano-fluids. Int J through microchannels. Appl Phys Lett 2007;91:233103.
Heat Mass Trans 2003;46:851–62. [50] Lee J-H, Hwang KS, Jang SP, Lee BH, Kim JH, Choi SUS, et al. Effective viscosities
[44] Putra N, Roetzel W, Das SK. Natural convection of nano-fluids. Heat Mass Trans and thermal conductivities of aqueous nanofluids containing low volume
2003;39:775–84. concentrations of Al2O3 nanoparticles. Int J Heat Mass Trans 2008;51:2651–6.
[45] Prasher R, Song D, Wang J, Phelan P. Measurements of nanofluid viscosity [51] Garg J, Poudel B, Chiesa M, Gordon JB, Ma JJ, Wang JB, et al. Enhanced thermal
and its implications for thermal applications. Appl Phys Lett 2006;89: conductivity and viscosity of copper nanoparticles in ethylene glycol
133108. nanofluid. J Appl Phys 2008;103:074301.

You might also like