Sunteți pe pagina 1din 8

ARTICLE

pubs.acs.org/IECR

Effect of Nanoconvection Caused by Brownian Motion on the


Enhancement of Thermal Conductivity in Nanofluids
Reza Azizian,† Elham Doroodchi,‡ and Behdad Moghtaderi†,*

Centre for Energy and ‡Centre for Advanced Particle Processing Chemical Engineering, School of Engineering Faculty of Engineering
and Built Environment, The University of Newcastle, Callaghan, NSW 2308, Australia

ABSTRACT: Nanofluids have attracted considerable attention in recent years as effective working fluids for heat transfer
applications. This is not surprising given that nanofluids which are essentially suspensions of nanoparticles in a base fluid, exhibit
higher thermal conductivity than conventional heat transfer fluids. The mechanisms responsible for such anomalous enhancement
in thermal conductivity are still not well understood despite many experimental and theoretical investigations carried out on the
subject. In this study, experiments were carried out on 70 nm alumina-in-water nanofluids with volume concentrations up to 13% in
an attempt to develop a clearer understanding of the different mechanisms that could be responsible for enhancement of thermal
conductivity in nanofluids. A set of experiments also were conducted on titanium dioxide-in-water nanofluids at volume
concentrations of up to 5% to consider the effect of material on nanoconvection due to Brownian motion. Our findings indicate
that nanoconvection caused by Brownian motion is the dominant mechanism responsible for the observed enhancements in thermal
conductivity of nanofluids.

’ INTRODUCTION liquid. Yu and Choi5 proposed that the liquid nanolayer acts as a
The performance, compactness and cost of heat-exchange thermal bridge between a nanoparticle and base fluid, and this is
operations in various industrial applications (e.g., power genera- the key to enhancing the thermal conductivity of the nanofluids.
tion, petrochemical, fuel/chemical production, air-conditioning, Following the pioneering work of Yu and Choi5 many other
transportation, MEMS and microelectronics) are often limited studies820 were conducted on various aspects of nanolayering
by the low thermal conductivities of conventional heat transfer and its effect on thermal conductivity enhancement of the nanofluids.
fluids such as water, mineral oil, and ethylene glycol. Driven by Leong’s model15 in particular showed extremely good agreement with
the need to overcome these limitations, particularly in emerging experimental data. This led many to believe that the nanolayering
fields such as MEMS (microelectro-mechanical systems), devel- phenomenon alone is responsible for anomalous enhancements
opment of high performance heat transfer fluids has been a highly in thermal conductivity of nanofluids. However, it was shown
active area of research over the past decade. Early studies focused later16,17 that much of the agreement between Leong’s model
on enhancing the thermal conductivity of conventional heat and experimental data was due to an incorrect derivation.
transfer fluids by employing suspensions of particulate solids Another important mechanism that comes into play when the
with high thermal conductivities in the base fluid. The early studies, particles get smaller is the nature of heat transport inside the
however, used suspensions of millimeter or micrometer sized nanoparticle. While for large particles the heat conduction
particles, which, despite some degree of enhancement experi- expressed by the Fourier Law adequately explains the nature of
enced problems such as poor suspension stability and hence heat transfer, the ballistic heat transport becomes important
clogging that are particularly detrimental for systems using mini when the temperature gradient becomes large or sample size is
and/or microchannels. To overcome these problems later stud- smaller than the mean free path of the phonons.21 In the case of
ies shifted their focus to the so-called nanofluids which are conductive heat transport phonons go through the material in
essentially engineered colloids containing nanoparticles (typically random directions while in the case of the ballistic heat transport
less than 100 nm) of metals, metal oxides, and carbon nanotubes all phonons jump directly from lower surface to the upper
suspended at low volume fraction (typically 4% or less) in a base surface. According to the Debye theory, the mean free path of
fluid such as water or organic liquids. the phonon is given by1,21
Four mechanisms have been identified in the literature175 as
10aTm
being responsible for heat transfer enhancement in nanofluids l¼ ð1Þ
under stationary conditions, namely, (i) molecular-level layering γT
of the liquid at the liquid/particle interface, (ii) the nature of heat
transport inside the nanoparticles, (iii) aggregation and cluster-
Special Issue: Nigam Issue
ing, and (iv) the Brownian motion of nanoparticles within the
base fluid. Henderson and Swol3 were the first to have proposed Received: May 24, 2011
that liquid molecules close to a solid surface form a layered Accepted: June 30, 2011
structure. Keblinski and co-workers showed1 that the atomic Revised: June 27, 2011
structure of this liquid layer is more ordered than that of the base Published: June 30, 2011

r 2011 American Chemical Society 1782 dx.doi.org/10.1021/ie201110k | Ind. Eng. Chem. Res. 2012, 51, 1782–1789
Industrial & Engineering Chemistry Research ARTICLE

Table 1. Comparison of Diffusion Time Scales for Water


Based Nanofluids with 5 nm Nanoparticles25
phenomena formula time-scale (s)

Brownian motion 3
4πμr /(6kBT) 3.8  108
nanoconvection 2
4r /v 3.11  1011
heat diffusion (conduction) 4r2CpF/(6kf) 2.83  1011

where a is the lattice constant, Tm is melting point, T is


temperature, and γ is the Gruneisen parameter.22 On the basis
of eq 1 the mean free path of the phonons for alumina at 293 K is
about 40 nm which is much smaller than the particle size typically
Figure 1. Particle distribution measured with the dynamic light scatter-
used in nanofluid applications. As a result, the ballistic heat ing (DLS) method.
transport cannot fully explain the observed enhancement in the
thermal conductivity of such fluids. Given the above background, it is clear that despite the recent
Aggregation is another mechanism that could be responsible progress the mechanisms responsible for enhancement of ther-
for the enhancement of heat transport in nanofluids. It has been mal conductivity in nanofluids are not well understood and,
shown recently1 that the clustering of the nanoparticles into hence, heat exchange processes in which nanofluids are used
percolating patterns creates a path with a lower thermal resistance cannot be accurately optimized. The main difficulty in under-
for heat flow, in turn, leading to the higher thermal conductivity standing such mechanisms is the impact of potential mechanisms
in nanofluids. The effective thermal conductivity increases rapidly on each other and their interactive nature. The study which is
with an initial increase of the aggregate size. This enhancement is reported here was an attempt to address this shortcoming at least
mainly due to the increased size of the backbone that promotes in part. Specifically, we investigated the effect of nanoconvection
rapid conduction across the cluster. Since the backbone is primarily in isolation from ballistic heat transport and formation of aggregates.
responsible for the thermal conductivity enhancements, bringing For this purpose relatively large nanoparticles and high solid
more particles to it will increase the enhancement rather than volume fractions were employed to avoid aggregate formations
dead ends.20 Aggregation/clustering tend to impact upon other (see the next section for details).
mechanisms responsible for thermal conductivity enhancement
of nanofluids, particularly the Brownian motion.
The Brownian motion gives rise to particleparticle collision ’ EXPERIMENTAL SECTION
inside the base fluid and, hence, is one of the potential mecha-
Nanofluid samples were prepared by making stable colloidal
nisms for heat transport in nanofluids.1 The particle velocity due
suspensions of alumina and titanium dioxide nanoparticles in
to Brownian motion can be estimated from
water using a two-step procedure. In the first step alumina (Sigma
rffiffiffiffiffiffiffiffiffiffi Aldrich), and titanium dioxide (Nanostructured & Amorphous
3kB T Materials, Inc.) nanoparticles were added to water at solid volume
Vp ¼ ð2Þ
M fractions of up to 13% and 5%, respectively. The average particle
sizes reported by the manufacturer for alumina and titanium
where kB is a Boltzmann constant, T is the temperature, and M is dioxide nanoparticles were 50 and 10 nm, respectively. The
the mass of particle. It has been shown by the order of magnitude desired nanofluids were then made by dispersing nanoparticles in
analysis1 that for nanoparticles (size range between 5 and the waterparticle slurry using an ultrasonic stirrer (Sonicator
100 nm), the magnitude of Vp is not high enough to make the 3000, Unisonics Pty, Ltd.) for 10 min. The particle size distribu-
transport of heat due to Brownian motion faster than that of the tion within nanofluid samples were measured by a dynamic light
diffusional heat transport (i.e., heat conduction) (see Table 1). scattering (DLS) method using a Zetasizer-Nanoseries instru-
However, as shown recently by several research groups,2325 the ment. DLS measurements indicated that the nominal particle size
nanoscale convection caused in the base fluid by particles within nanofluid samples was generally about 70 nm for alumina
colliding due to Brownian motion can reach a characteristic time and 30 nm for titanium dioxide nanoparticles (see Figure 1 for
quite comparable with that of the diffusional heat transport. example).
Prasher and co-workers for instance24 showed through an order Measurements of the thermal conductivity were performed in
of magnitude analysis that the time required for the effect of a conventional “transient hot wire, THW” apparatus20,5360
nanoconvection to be felt in a distance equal to the nanoparticle which has been used in the past for similar measurements with
diameter is almost the same as the time-scale for the heat errors reported as low as 0.2%.59 The THW apparatus operates
diffusion. The same behavior was theoretically shown to exist based on the principle that if a linear heat source with a constant
even for particles as small as 5 nm25 (see Table 1). In addition to and uniform output (i.e., hot wire) is placed in a liquid, the
the above pioneering works many other theoretical studies have thermal conductivity of the liquid can be derived from the
been carried out on the impact of Brownian motion on resulting temperature change over a known time interval.59
nanofluids.2639 These studies overwhelmingly indicate that The THW apparatus employed in this study was developed in
the Brownian motion by itself cannot explain the enhancement 2007 by one of our students60 and consists of a THW cell and a
of thermal conductivity in nanofluids, while the nanoconvection Wheatstone bridge circuit (see Figure 2). The cell comprises a
due to Brownian motion could be the potential mechanism for test tube and an electrical wire (see Figure 3). The wire was
heat transfer enhancement in nanofluids although further experi- coated with Teflon to minimize the attraction of charged alumina
mental proof is needed. particles toward the wire. Deionized water was used as the base
1783 dx.doi.org/10.1021/ie201110k |Ind. Eng. Chem. Res. 2012, 51, 1782–1789
Industrial & Engineering Chemistry Research ARTICLE

bar throughout the experimental campaign. Each experiment was


repeated three times to confirm reproducibility and ensure that
collected data are statistically consistent.

’ RESULTS AND DISCUSSION


One of the key objectives of this study was to minimize the
impact of mechanisms such as nanolayering, ballistic heat trans-
port, and aggregation on thermal conductivity enhancement in
nanofluids so that the nanoconvection mechanism can be studied
in relative isolation. To this end, we focused on using relatively
large nanoparticles particularly in the case of alumina for which as
noted earlier from DLS measurements the nominal particle size
of nanofluid samples was about 70 nm. According to Ashcroft22
and based on eq 1 the mean free path of the phonons for alumina
at 293 K is approximately 40 nm which is much smaller that the
nominal particle size of 70 nm in our nanofluid samples. Hence,
the impact of ballistic heat transport on the thermal conductivity
enhancement can be considered negligible here.
Moreover, the use of relatively large particles (e.g., 70 nm used
in this study) reduces the likelihood of aggregate formation.
From a systematic series of laboratory experiments we have
observed that small particles (e20 nm) generally tend to form
large aggregates, whereas particles with large diameters exhibit
less tendency for clustering and usually form much smaller
aggregates, if any. Similar observations have been reported in
the open literature and two possible explanations have been
provided for the observed behavior,50,7173 namely, (i) the influ-
ence of attractive surface forces which become more significant as
the particle diameter decreases,72 and (ii) the increased chance of
particleparticle collision when at a fixed solid volume fraction
the size of particles is reduced. The increased collision frequency
combined with strong attractive forces allows clusters to over-
come the relevant energy barriers and grow to relatively large
aggregates. The impact of surface forces and collision on aggregate
formation can be best expressed in terms of the isoelectric point
(IEP) which is essentially the pH at which a given molecule or
surface carries no net electrical charge (i.e., the negative and
positive charges are equal). At or near IEP the electrostatic
Figure 2. (a) Schematic representation of the thermal hot wire (THW) repulsion between particles is no longer sufficient to keep the
setup, (b) picture of the University of Newcastle THW system. particles apart and as a result aggregates can rapidly grow. Therefore,
for a given suspension the aggregate reaches its maximum size at
IEP. In contrast, at pH levels away from IEP the tendency for
clustering is quite weak and, hence, if aggregates are formed their
physical dimensions will be significantly small. The IEP for
suspensions is not unique and is a function of several parameters.
Typically for alumina nanofluids (without any other additives
such as surfactants, etc.) the IEP is between 8 and 9 (see for
example ref 72 and ref 73).
Figure 4 shows the suspension pH of alumina nanofluid
samples measured in our laboratory. Evidently, the suspension
pH varies as the solid volume fraction is increased. Specifically,
for volume fractions smaller than 2% the suspension pH initially
falls to values as low as 6 but it gradually recovers until a pH value
of close to 7 is reached. A similar trend has been reported by
Timofeeva and co-workers.72 Interestingly, the values of pH
shown in Figure 4 indicate that some of level of dissociation has
taken place in the nanofluid samples where OH ions have
Figure 3. THW setup (a) test cell assembly; (b) Teflon-coated electrical wire. possibly formed aluminum hydroxides, in turn, increasing the
concentration of surface charge-determining H+ ions in the
fluid in all experiments and as well as calibrations. The thermal suspension. In any case, as illustrated in Figure 4, the suspension
conductivity measurements were carried out at 293.15 K and 1 pH for nanofluid samples employed in our study is far below the
1784 dx.doi.org/10.1021/ie201110k |Ind. Eng. Chem. Res. 2012, 51, 1782–1789
Industrial & Engineering Chemistry Research ARTICLE

Figure 4. Measurements of suspension pH as a function of solid volume


fraction.

Figure 6. Comparison of present experimental data with the literature


data for solid volume fractions between 0 and 6% (a) alumina-in-water;
(b) titanium dioxide-in-water.

Figure 5. Enhancement in thermal conductivity for alumina- and


titanium dioxide-based nanofluids over an extended range of solid
volume fractions.

pH values corresponding to IEP of typical alumina suspensions.


As such, the likelihood of aggregate formation in our nanofluids
must have been minimal indicating that the selected experi-
mental parameters were effective in isolating nanoconvection
mechanism from clustering and aggregation.
Let us focus now on validating the thermal conductivity data
obtained in this study against those reported by other research-
ers. Figure 5 shows the scattered plot of thermal conductivity
enhancement of nanofluid samples (i.e., keff/kf) versus the Figure 7. Comparison of present experimental data for alumina with
volume fraction of nanoparticles (i.e., solid volume fraction). the literature data for solid volume fractions between 0 and 18%.
The plot has been determined by comparing the thermal conduc-
tivity of the base water and those of nanofluid samples measured The data presented in Figure 5 are in good agreement with the
using the THW setup described previously. As can be seen, the data reported in the open literature for alumina and titanium
accuracy of data reported in Figure 5 is excellent with uncertain- oxide based nanofluids.30,55,6265,70,74,75 For instance as Figure 6
ties ranging from (1 to (4% for a 95% confidence level. The panels a and b illustrate, for solid volume fractions of up to 6%
relationship between the thermal conductivity enhancement and our measurements for alumina and titanium dioxide show little
solid volume fraction shown in Figure 5 for alumina and titanium discrepancy with the data reported by references 7,19,30,57,
dioxide generally follow a linear trend between the solid volume 65,71,74, and 75. Slightly higher discrepancies are observed when
fractions of 0% and 5%. Beyond 5% concentration the alumina our data set is compared with those reported by Eastman and co-
plot follows a nonlinear trend approaching an asymptotic value of workers71 for alumina, although both data sets show a similar
1.2 at solid volume concentrations greater than 12%. This linear trend (see Figure 6a). Likewise, a comparison of our
asymptotic value can be considered as an upper limit for thermal measurements with those of Minsta62 over a wider range of solid
conductivity enhancement at least for alumina cases studied here. volume fractions (between 0 and 13%) for alumina confirms that
Because of the nonlinear behavior the thermal conductivity there is a nonlinear relationship between the solid volume
enhancement becomes smaller in differential terms as the volume fraction and the thermal conductivity enhancement at volume
fraction is increased. The apparent shift from the linear to the fractions greater than 5% (Figure 7).
nonlinear behavior for alumina appears to take place at or near a Now let us change our attention to nanoconvection and its
solid volume fraction of 6%. impact on thermal conductivity enhancement. Given the technical
1785 dx.doi.org/10.1021/ie201110k |Ind. Eng. Chem. Res. 2012, 51, 1782–1789
Industrial & Engineering Chemistry Research ARTICLE

difficulties associated with the direct measurement of nanocon-


vection, let us examine the influence of nanoconvection in an
indirect manner, that is, by comparing the measurements shown
in Figure 5 against several well know models/correlations of
which one incorporates the effect of nanoconvection. These are
the classical HamiltonCrosser,66 and Burggeman11,61,67 models as
well as the correlations recently proposed by Li and Peterson68
and Chon et al.69 Among these the Chon’s correlation takes into
account the effect of nanoconvection by incorporating the
particle velocity associated with the Brownian motion (see the
following paragraphs for more details).
The Hamilton and Crosser model66 is an extension of the
classical Maxwell theory to nonspherical particles. The model
predicts the effective thermal conductivity of a nanofluid from eq 3:
" #
kef f kp þ ðn  1Þkf  ðn  1Þðkf  kp Þϕ
¼ ð3Þ
kf kp þ ðn  1Þkf þ ðkf  kp Þϕ

where keff, kf, and kp are the thermal conductivities of the nanofluid,
base fluid, and particle respectively, ϕ is the solid volume fraction,
and n is an empirical shape factor given by n = 3/ψ where ψ is the
particle sphericity.
The classical model of Bruggeman11,61,67 combines the effec-
tive medium and fractal theories and determines the effective
thermal conductivity of the nanofluid from eq 4:
Figure 8. Comparison of present experimental data with several
kef f 1 1pffiffiffiffi models/correlations reported in the literature (a) for alumina-in-water;
¼ ½ð3ϕ  1Þkp þ ð2  3ϕÞkf  þ Δ ð4Þ (b) titanum dioxide-in-water.
kf 4kf 4
where Δ is given by eq 5:
"  2  # Ff V p d p
Δ ¼ ð3ϕ  1Þ
kp2 2
þ ð2  3ϕÞ þ 2ð2 þ 9ϕ  9ϕ Þ
2 kp Re  ð9Þ
kf kf μ

ð5Þ where μ is dynamic viscosity of the base fluid, Ff is density of the


base fluid, R is thermal diffusivity, dp is particle diameter, and Vp
The correlation recently proposed by Li and Peterson68 is based is a characteristic particle velocity. As mentioned earlier, Chon’s
on the linear regression analysis of the temperature and solid correlation takes into account the effect of indirect Brownian
volume fraction and has been specifically developed for predict- motion (nanoconvection). This is achieved by assuming that Vp
ing the thermal conductivity of alumina based nanofluids. The is equivalent to the Brownian motion velocity of the particles
model is given the following expression:68 defined as
kef f kb T
¼ 1 þ 0:764481ϕ Vp  ð10Þ
kf 3πμdp lf
þ 0:018688867T  0:462147175 ð6Þ where kb is the Boltzmann constant (1.3807  1023 J/K), and lf
where T is the nanofluid temperature. is the mean free path of base fluid molecules calculated from
Similarly, Chon and co-workers recently proposed69 a semi- 1
empirical correlation based on the Buckingham theory and linear lf ¼ pffiffiffi πdf 2 ð11Þ
2m
regression analysis of the experimental measurements of thermal
conductivity of alumina-based nanofluids: where m is a molecular number density (note that eq 10 is a
slightly different version of eq 2).
kef f
¼1 Figure 8a illustrates the comparison between the present
kf alumina experimental data and predictions of the above-mentioned
!0:3690   models/correlations. As discussed before, while the experimental
df kp 0:7476 0:9955 1:2321 data collected in this study follow a general nonlinear trend, at
þ 64:7ðϕ 0:7460
Þ Pr Re
dp kf solid volume fractions between 0 and 5% the trend is linear. Over
this range of solid volume fractions titanium dioxide also shows a
ð7Þ linear relationship between the thermal conductivity enhance-
where Pr and Re denote the Prandtl and Reynolds numbers, ment and solid volume fraction. A change of slope for alumina
respectively (see eqs 8 and 9). (i.e., the switch over from linear to nonlinear trend) occurs
approximately at the solid volume fraction of 6%. Due to this
μ
Pr  ð8Þ distinct change of slope which has also been reported by several
Ff R other researchers,62 none of the models/correlations studied
1786 dx.doi.org/10.1021/ie201110k |Ind. Eng. Chem. Res. 2012, 51, 1782–1789
Industrial & Engineering Chemistry Research ARTICLE

here can fully predict the present experimental data. However, aggregation/clustering, and ballistic heat transport. Measure-
some of these models/correlations can accurately match some ments were compared with a range of experimental data reported
part of the experimental data (see Figure 8 a). in the literature to verify their validity. Measurements were also
One of the first observations from Figure 8a is that the compared with several well known models/correlations in an
correlation by Li and Peterson68 does not show any enhance- attempt to investigate the role of nanoconvection in an indirect
ment in thermal conductivity (i.e., all predictions < 1) and under manner. Experimental data presented in this paper clearly show
predicts the experimental data by a relatively large margin. This that over an extended range of solid volume fractions (i.e.,
can be partly assigned to the fact that Li’s correlation68 was 013%) the relation between the particle loading and thermal
derived based on experiments over the temperature range conductivity enhancement is not linear and follows a parabolic
between 27 and 36 °C, whereas the experimental data collected profile. More importantly, the comparison between the experi-
as part of this study were obtained at room temperature. Given mental data and predictions suggests that nanoconvection be-
that Li’s correlation (i.e., eq 6) is essentially a line, it can be comes more pronounced at solid volume fractions greater than
correlated with present experimental data only if the constant 6%. However, this value which corresponds to a nominal particle
0.462147175 in eq 6 is replaced by a new constant which size of 70 nm is not unique and varies depending on the nominal
represents an increase of about 19% (i.e., 0.37378). size of nanoparticles with which the nanofluid has been prepared.
Another interesting observation from Figure 8a is that the
effective medium theories such as the HamiltonCrosser model ’ AUTHOR INFORMATION
and Bruggeman models are generally accurate at solid volume
fractions less than 6%. But these models overpredict the thermal Corresponding Author
conductivity enhancement at higher solid volume fractions (see *Tel.: +61 2 4985-4411. Fax: +61 2 4921-6893. E-mail: Behdad.
Figure 8a). This observation is in good agreement with a recent Moghtaderi@newcastle.edu.au.
benchmark study by Buongiorno et al.,70 in which more than 30
organizations worldwide measured the thermal conductivity of
identical samples by using different experimental methods. ’ ACKNOWLEDGMENT
As far as Chon’s correlation69 is concerned, its predictions are The authors wish to acknowledge the financial support of the
quite accurate at solid volume fractions greater than 6% rather University of Newcastle (Australia), Granite Power Pty, Ltd., and
than at lower values, as evident from Figure 8a. The experimental the Australian Research Council through the ARC-Linkage
data appear to follow the HamiltonCrosser model for low Grant LP100200871 (2010-2012).
particle loadings before switching to follow the Chon’s correla-
tion. Given that Chon’s correlation does consider the nanocon- ’ NOMENCLATURE
vection mechanism and noting that the impact of mechanisms
such as nanolayering, ballistic heat transport and aggregation Variables
have been minimized in our experiments, the close agreement a = lattice constant
between the experimental data and predictions at solid volume cp = specific heat capacity
fractions between 6 and 13% suggest that over this range the d = diameter
nanoconvection must be a key contributing factor in the ob- h = thickness of interfacial layer
served thermal conductivity enhancement. This can be partly due k = thermal conductivity
to higher particle loadings at this range of solid volume fraction kb = Boltzman’s Constant (1.3805  1023 J/K)
which inevitably create larger and stronger nanoscale chaotic l = mean free path
structures. m = molecular density
The behavior shown in Figure 8b by titanium dioxide is M = mass of the particle
slightly different to that exhibited by alumina in Figure 8a. As n = empirical shape factor
can be seen from Figure 8b because of their smaller average sizes Pr = Prandtl number
the effect of nanoconvection is more pronounced for titanium r = radius
dioxide particles. As such, unlike the alumina case (Figure 8a) Re = Reynolds number
Chon’s correlation69 can predict the experimental data for T = temperature
titanium dioxide even for volume fractions less than 5%. This t = time
observation is very interesting because the empirical correlation S = surface area per unit volume
by Chon and co-workers was originally proposed for predicting V = velocity
the thermal conductivity enhancement of alumina rather than
titanium oxide. Greek Symbols
r = thermal diffusivity
ϕ = particle volume fraction
’ SUMMARY AND CONCLUSIONS μ = dynamics viscosity
A set of experiments were conducted on 70 nm alumina-in- υ = kinematic viscosity
water nanofluids at volume concentrations of up to 13%, and γ = gruneisen parameter
30 nm titanium dioxide-in-water up to 5%. The emphasis of the F = density
study was on nanoscale convection caused by the Brownian ψ = sphericity
motion and its impact on thermal conductivity enhancement
in nanofluids. Relatively large diameter alumina nanoparticles ’ SUBSCRIPT AND SUPERSCRIPT
were used in the preparation of nanofluid samples to eliminate eff = effective
other heat transport enhancing mechanisms such nanolayering, f = fluid (liquid)
1787 dx.doi.org/10.1021/ie201110k |Ind. Eng. Chem. Res. 2012, 51, 1782–1789
Industrial & Engineering Chemistry Research ARTICLE

l = interfacial layer (nanolayer) (23) Jang, S. P.; Choi, S. U. S. Role of Brownian motion in the
m = melting point enhanced thermal conductivity of nanofluids. Appl. Phys. Lett. 2004,
p = particle (solid) 84, 4316.
(24) Prasher, R.; Bhattacharya, P.; Phelan, P. E. Thermal conductivity of
nanoscale colloidal solutions (nanofluids). Phy. Rev. Lett. 2005, 94, 1.
(25) Azizian, M. R.; Aybar, H. S.; Okutucu, T. Effect of nanoconvec-
’ REFERENCES tion due to Brownian motion on thermal conductivity of nanofluids,
(1) Keblinski, P.; Phillpot, S. R.; Choi, S. U. S.; Eastman, J. A. Proc. 7th IASME/WSEAS Int. Conf. Heat Transfer, Therm. Eng. Environ.
Mechanisms of heat flow in suspensions of nano-sized particles (HTE ’09) 2009, 53.
(nanofluids). Int. J. Heat Mass Transfer 2002, 45, 855. (26) Jang, S. P.; Choi, S. U. S. Effects of various parameters on
(2) Eastman, J. A.; Phillpot, S. R.; Choi, S. U. S.; Keblinski, P. nanofluid thermal conductivity. J. Heat Transfer 2007, 129, 618.
Thermal transport in nanofluids. Annu. Rev. Mater. Res. 2004, 34, (27) Koo, J.; Kleinstreuer, C. A new thermal conductivity model for
219. nanofluids. J. Nanopart. Res. 2004, 6, 577.
(3) Henderson, J. R.; Swol, F. V. On the interface between a fluid and (28) Koo, J.; Kleinstreuer, C. Impact analysis of nanoparticle motion
a planar wall: Theory and simulations of a hard sphere fluid at a hard wall. mechanisms on the thermal conductivity of nanofluids. Int. Commun.
Mol. Phys. 1984, 51, 991. Heat Mass Transfer 2005, 32, 1111.
(4) Yu, C. J.; Richter, A. G.; Datta, A.; Durbin, M. K.; Dutta, P. (29) Shukla, R. K.; Dhir, V. K. Effect of Brownian motion on thermal
Molecular layering in a liquid on a solid substrate: An X-ray reflectivity conductivity of nanofluids. J. Heat Transfer 2008, 130, 042406.
study. Phys. B 2000, 283, 27. (30) Das, S. K.; Putra, N.; Thiesen, P.; Roetzel, W. Temperature
(5) Yu, W.; Choi, S. U. S. The role of interfacial layers in the dependence of thermal conductivity enhancement for nanofluids. J. Heat
enhanced thermal conductivity of nanofluids: A renovated Maxwell Transfer 2003, 125, 567.
model. J. Nanopart. Res. 2003, 5, 167. (31) Beck, M. P.; Sun, T.; Teja, S. A. The thermal conductivity of
(6) Lee, S.; Choi, S. U. S.; Li, S.; Eastman, J. A. Measuring thermal alumina nanoparticles dispersed in ethylene glycol. Fluid Phase Equilib.
conductivity of fluids containing oxide nanoparticles. J. Heat Transfer. 2007, 260, 275.
1999, 121, 281. (32) Evans, W.; Fish, J.; Keblinski, P. Role of Brownian motion
(7) Masuda, H.; Ebata, A.; Teramae, K.; Hishinuma, N. Alternation hydrodynamics on nanofluid thermal conductivity. Appl. Phys. Lett.
of thermal conductivity and viscosity of liquid by dispersing ultrafine 2006, 88, 093116.
particles (dispersions of γ-Al2O3, SiO2, and TiO2 ultrafine particles). (33) Putnam, S. A.; Cahill, D. G.; Braun, P. V.; Ge, Z.; Shimmin,
Netsu Bussei (Jpn) 1993, 4, 227. R. G. Thermal conductivity of nanoparticle suspensions. J. Appl. Phys.
(8) Chandrasekar, M.; Suresh, S. A review on the mechanisms of 2006, 99, 084308.
heat transport in nanofluids. Heat Transfer Eng. 2009, 30, 1136. (34) Nie, C.; Marlow, W. H.; Hassan, Y. A. Discussion of proposed
(9) Xue, Q.; -Z. Model for effective thermal conductivity of nano- mechanisms of thermal conductivity enhancement in nanofluids. Int.
fluids. Phys. Lett. A 2003, 307–313. J. Heat Mass Transfer 2008, 51, 1342.
(10) Yu, W.; Choi, S. U. S. The role of interfacial layers in the (35) Kumar, D. H.; Patel, H. E.; Kumar, V. R. R.; Sundararajan, T.;
enhanced thermal conductivity of nanofluids: A renovated Hamilton Pradeep, T.; Das, S. K. Model for heat conduction in nanofluids. Phys.
Crosser model. J. Nanopart. Res. 2004, 6, 355. Rev. Lett. 2004, 93, 144301.
(11) Wang, B. -X.; Zhou, I. P.; Peng, X. F. A fractal model for (36) Keblinski, P.; Cahill, D. G. Comments on model for heat
predicting the effective thermal conductivity of liquid with suspension of conduction in nanofluids. Phys. Rev. Lett. 2005, 95, 209401.
nanoparticles. Int. J. Heat Mass Transfer 2003, 46, 2665. (37) Das, S. K.; Sundararajan, T.; Pradeep, T.; Patel, H. E. Reply to
(12) Xue, L.; Keblinski, P.; Phillpot, S. R.; Choi, S. U. S.; Eastman, “comments on 'model for heat conduction in nanofluids'”. Phys. Rev. Lett.
J. A. Effect of liquid layering at the liquidsolid interface on thermal 2005, 95, 209402.
transport. Int. J. Heat Mass Transfer 2004, 47, 4277. (38) Bastea, S. Comment on “model for heat conduction in nano-
(13) Xue, Q.; Xu, W. M. A model of thermal conductivity of fluids”. Phys. Rev. Lett. 2005, 95, 019401.
nanofluids with interfacial shells. Mater. Chem. Phys. 2005, 90, 298. (39) Das, S. K.; Sundararajan, T.; Pradeep, T.; Patel, H. E. Reply to
(14) Xie, H.; Fujii, M.; Zhang, X. Effect of interfacial nanolayer on “Comment on 'model for heat conduction in nanofluids'”. Phys. Rev. Lett.
the effective thermal conductivity of nanoparticlefluid mixture. Int. 2005, 95, 019402.
J. Heat Mass Transfer 2005, 48, 2926. (40) Philip, J.; Shima, P. D.; Raj, B. Nanofluid with tunable thermal
(15) Leong, K. C.; Yang, C.; Murshed, S. M. S. A model for the properties. Appl. Phys. Lett. 2008, 92, 043108.
thermal conductivity of nanofluidsThe effect of interfacial layer. (41) Shima, P. D.; Philip, J.; Raj, B. Role of microconvection induced
J. Nanopart. Res. 2006, 8, 245. by Brownian motion of nanoparticles in the enhanced thermal con-
(16) Doroodchi, E.; Evans, T. M.; Moghtaderi, B. Comments on the ductivity of stable nanofluids. Appl. Phys. Lett. 2009, 94, 223101.
effect of liquid layering on the thermal conductivity of nanofluids. (42) Philip, J.; Shima, P. D.; Raj, B. Enhancement of thermal
J. Nanopart. Res. 2009, 11, 1501. conductivity in magnetite-based nanofluid due to chainlike structures.
(17) Azizian, M. R.; Doroodchi, E.; Moghtaderi, B. The role of liquid Appl. Phys. Lett. 2007, 91, 203108.
layering on the enhancement of thermal conductivity in nanofluids. 14th (43) Philip, J; Shima, P. D.; Raj, B. Nanofluid with tuneable thermal
ASME Int. Heat Transfer Conf. 2010, 6, 659. properties. Appl. Phys. Lett. 2008, 92, 043108.
(18) Tillman, P.; Hill, J. M. Determination of nanolayer thickness for (44) Philip, J.; Laskar, J. M.; Raj, B. Magnetic field induced extinction
a nanofluid. Int. Commun. Heat Mass. 2007, 34, 399. of light in a suspension of Fe3O4 nanoparticles. Appl. Phys. Lett. 2008,
(19) Lee, D. Thermophysical properties of interfacial layer in nano- 92, 221911.
fluids. Langmuir. 2007, 23, 6011. (45) Wen, D.; Lin, G.; Vafaei, S.; Zhang, K. Review of nanofluids for
(20) Evans, W.; Prasher, R.; Fish, J.; Meakin, P.; Phelan, P.; heat transfer applications. Particuology 2009, 7, 141.
Keblinski, P. Effect of aggregation and interfacial thermal resistance on (46) Karthikeyan, N. R.; Philip, J.; Raj, B. Effect of clustering on the
thermal conductivity of nanocomposites and colloidal nanofluids. Int. thermal conductivity of nanofluids. Mater. Chem. Phys. 2008, 109, 50.
J. Heat Mass Transfer 2008, 14, 1431. (47) Shih, W. H.; Shih, W. Y.; Kim, S. I.; Liu, J.; Aksay, I. A. Scaling
(21) Geiger, G. H.; Poirier, D. R. Transport Phenomena in Metallurgy; behavior of the elastic properties of colloidal gels. Phys. Rev. A 1990,
Addison-Wesley: Reading, MA, 1973. 42, 4772.
(22) Ashcroft, N. W.; Mermin, N. D. Solid State Physics; Holt (48) Xuan, Y.; Li, Q.; Hu, W. Aggregation structure and thermal
Rinehart and Winston: New York, 1976. conductivity of nanofluids. AICHE. J. 2003, 49, 1038.

1788 dx.doi.org/10.1021/ie201110k |Ind. Eng. Chem. Res. 2012, 51, 1782–1789


Industrial & Engineering Chemistry Research ARTICLE

(49) Xu, J.; Yu, B. M.; Yun, M. J. Effect of clusters on thermal (73) Peng, Z.; Doroodchi, E.; Evans, G. DEM simulation of aggrega-
conductivity in nanofluids. Chin. Phys. Lett. 2006, 23, 2819. tion of suspended nanoparticles. Powder Technol. 2010, 204, 91–102.
(50) Prasher, R.; Phelan, P. E.; Bhattacharya, P. Effect of aggregation (74) Wang, Z. L.; Tang, D. W.; Liu, S.; Zheng, X. H.; Araki, N.
kinetics on the thermal conductivity of nanoscale colloidal solutions Thermal-Conductivity and Thermal-Diffusivity Measurements of Na-
(nanofluid). Nano Lett. 2006, 6, 1529. nofluids by 3ω Method and Mechanism Analysis of Heat Transport. Int.
(51) Feng, Y.; Yu, B.; Xu, P.; Zou, M. The effective thermal J. Thermophys. 2007, 28, 1255.
conductivity of nanofluids based on the nanolayer and the aggregation (75) Turgut, A.; Tavman, I.; Chirtoc, M.; Schuchmann, H. P.;
of nanoparticles. J. Phys. D. Appl. Phys. 2007, 40, 3164. Sauter, C.; Tavman, S. Thermal Conductivity and Viscosity Measure-
(52) Li, Y. H.; Qu, W.; Feng, J. C. Temperature dependence of ments of Water-Based TiO2 Nanofludis. Int. J. Thermophys. 2009,
thermal conductivity of nanofluids. Chin. Phys. Lett. 2008, 25, 3319. 30, 1213.
(53) Hong, K. S.; Hong, T. K.; Yang, H. S. Thermal conductivity of
Fe nanofluids depending on the cluster size of nanoparticles. Appl. Phys.
Lett. 2006, 88, 1.
(54) Xuan, Y.; Li, Q. Heat transfer enhancement of nanofluids. Int.
J. Heat Mass Transfer 2000, 21, 58.
(55) Hong, T. K.; Yang, H. S.; Choi, C. J. Study of the enhancement
thermal conductivity of Fe nanofluids. J. Appl. Phys. 2005, 97, 1.
(56) Lo, C. H.; Tsung, T. T.; Chen, L. C.; Su, C. H.; Lin, H. M.
Fabrication of copper oxide nanofluid using submerged arc nanoparti-
cles synthesis system (SANSS). J. Nanopart. Res. 2005, 7, 313.
(57) Zhang, X.; Gu, H.; Fujii, M. Effective thermal conductivity and
thermal diffusivity of nanofluids containing spherical and cylindrical
nanoparticles. J. Appl. Phys. 2006, 100, 1.
(58) Li, Y.; Zhou, J.; Tung, S.; Schneider, E.; Xi, S. A review on
development of nanofluid preparation and characterization. Powder
Technol. 2009, 196, 89.
(59) Healy, J. J.; Groot, J. J.; Kestin, J. The theory of the transient hot-
wire method for measuring thermal conductivity. Phys. B+C 1976, 82, 392.
(60) Allen, D. J. Enhanced Thermal Conductivity of Nanofluids.
Chemical Engineering Thesis, The University of Newcastle in Australia,
2007.
(61) Murshed, S. M. S.; Leong, K. C.; Yang, C. Enhanced thermal
conductivity of TiO2-water based nanofluid. Int. J. Therm. Sci. 2005,
44, 367.
(62) Mintsa, H. A.; Roy, G.; Nguyen, C. T.; Doucet, D. New
temperature-dependent thermal conductivity data for water-based na-
nofluids. Int. J. Therm. Sci. 2009, 48, 363.
(63) Gao, L.; Zhou, X. F. Differential effective medium theory for
thermal conductivity in nanofluids. Phys. Lett. A 2006, 348, 355.
(64) Prasher, R.; Bhattacharya, P.; Phelan, P. E. Brownian-motion-
based convective-conductive model for the effective thermal conductiv-
ity of nanofluids. J. Heat Transfer 2006, 128, 588.
(65) Wang, X.; Xu, W.; Choi, S. U. S. Thermal conductivity of
nanoparticlefluid mixture. J. Thermophys. Heat Transfer 1999, 4, 474.
(66) Hamilton, R. L.; Crosser, O. K. Thermal conductivity of
heterogeneous two component systems. Ind. Eng. Chem. Fundam. 1962,
1, 187.
(67) Ding, Y. L.; Wen, D. S.; Williams, R. A. Nanofluids for heat
transfer intensification—Where are we and where should we go? 6th Int.
Symp. Heat Transfer 2004, 66.
(68) Li, C. H.; Peterson, G. P. Experimental investigation of
temperature and volume fraction variations on the effective thermal
conductivity of nanoparticle suspensions (nanofluids). J. Appl. Phys.
2006, 99, 1.
(69) Chon, C. H.; Kihm, K. D.; Lee, S. P.; Choi, S. U. S. Empirical
correlation finding the role of temperature and particle size for nanofluid
(Al2O3) thermal conductivity enhancement. Appl. Phys. Lett. 2005,
87, 153107.
(70) Buongiorno, J.; Venerus, D. C.; Prabhat, N.; McKrell, T.;A
benchmark study on the thermal conductivity of nanofluids. J. Appl. Phys.
2009, 106, 094312.
(71) Eastman, J. A.; Choi, S. U. S.; Li, S.; Thompson, L. J.; Lee, S.
Enhanced thermal conductivity through the development of nanofluids.
Symp. Nanophase Nanocomp. Mater. II 1997, 3.
(72) Timofeeva, E. V.; Gavrilov, A. N.; McCloskey, J. M.; Tolmachev,
Y. V.; Sprunt, S.; Lopatina, L. M.; Selinger, J. V. Thermal conductivity and
particle agglomeration in alumina nanofluids: Experimental and theory.
Phys. Rev. E 2007, 76, 061203.

1789 dx.doi.org/10.1021/ie201110k |Ind. Eng. Chem. Res. 2012, 51, 1782–1789

S-ar putea să vă placă și