Sei sulla pagina 1di 10

S.

Lee
Department of Mechanical Engineering,
Kyonggi University,
Suwon, Kyonggi-Do Korea Measuring Thermal Conductivity
e-mail: shinpyo@kuic.kyonggi.ackr
of Fluids Containing Oxide
S. U.-S. Choi
Energy Technology Division,
e-mail: choi@anl.gov
Nanoparticles
Mem. ASME
Oxide nanofluids were produced and their thermal conductivities were measured by a
transient hot-wire method. The experimental results show that these nanofluids, contain-
S. Li ing a small amount of nanoparticles, have substantially higher thermal conductivities than
Materials Science Division the same liquids without nanoparticles. Comparisons between experiments and the Ham-
ilton and Crosser model show that the model can predict the thermal conductivity of
nanofluids containing large agglomerated Al203 particles. However, the model appears to
J. A. Eastman be inadequate for nanofluids containing CuO particles. This suggests that not only
Materials Science Division particle shape but size is considered to be dominant in enhancing the thermal conductivity
of nanofluids.
Argonne National Laboratory,
Argonne, IL 60439

1 Introduction plications in a number of areas, including biotechnology, nano-


electronic devices, scientific instruments, and transportation (Ash-
Traditional heat transfer fluids, such as water, oil, and ethylene
ley, 1994; Rohrer, 1996).
glycol mixture are inherently poor heat transfer fluids. There is a
Modern nanotechnology provides great opportunities to process
strong need to develop advanced heat transfer fluids, with signif-
and produce materials with average crystallite sizes below 50 nm.
icantly higher thermal conductivities and improved heat transfer
Recognizing an opportunity to apply this emerging nanotechnol-
characteristics than are presently available. Despite considerable
ogy to established thermal energy engineering, Argonne has de-
previous research and development focusing on industrial heat
veloped the concept of a new class of heat transfer fluids called
transfer requirements, major improvements in heat transfer capa-
"nanofluids," which transfer heat more efficiently than conven-
bilities have been held back because of a fundamental limit in the
tional fluids (Choi, 1995). Nanofluids are engineered by suspend-
thermal conductivity of conventional fluids.
ing ultrafine metallic or nonmetallic particles of nanometer dimen-
It is well known that metals in solid form have thermal conduc- sions in traditional heat transfer fluids such as water, engine oil,
tivities that are higher than those of fluids by orders of magnitude. and ethylene glycol.
For example, the thermal conductivity of copper at room temper- Argonne has already produced nanofluids and conducted proof-
ature is about 700 times greater than that of water and about 3000 of-concept tests (Eastman et al, 1997). In particular, it was dem-
times greater than that of engine oil (Touloukian and Ho, 1970). onstrated that oxide nanoparticles, such as A1203 and CuO, have
Even oxides such as alumina (A1203), which are good thermal excellent dispersion properties in water, oil, and ethylene glycol
insulators compared to metals such as copper, have thermal con- and form stable suspensions (i.e., significant settling does not
ductivities more than an order-of-magnitude larger than water. occur in static suspensions even after weeks or months).
Therefore, fluids containing suspended solid particles are expected Nanofluids are expected to exhibit superior properties relative to
to display significantly enhanced thermal conductivities relative to those of conventional heat transfer fluids and fluids containing
those of conventional heat transfer fluids. micrometer-sized particles. Because heat transfer takes place at the
In fact, numerous theoretical and experimental studies of the surface of the particle, it is desirable to use particles with a large
effective thermal conductivity of suspensions that contain solid total surface area. The surface-area-to-volume ratio is 1000 times
particles have been conducted since Maxwell's theoretical work larger for particles with a 10-nm diameter than for particles with a
was published more than 100 years ago (Maxwell, 1881). How- lO-jitm diameter. The much larger surface areas of nanoparticles
ever, all of the studies on thermal conductivity of suspensions have relative to those of conventional particles should not only improve
been confined to those produced with millimeter or micrometer- heat transfer capabilities, but also increase the stability of the
sized particles. Until now, researchers had no way to prevent solid suspensions. Nanoparticles offer extremely large total surface ar-
particles from eventually settling out of suspension. The lack of eas and therefore have great potential for application in heat
stability of suspensions that involve coarse-grained particles is transfer.
undoubtedly a primary reason why fluids with dispersed coarse- The use of the conventional millimeter and micrometer-sized
grained particles have not been previously commercialized. particles in heat transfer fluids in practical devices is greatly
We are on the verge of a new scientific and technological era, limited by the tendency of such particles to settle rapidly and to
the standard of which is the nanometer (billionths of a meter). clog mini and microchannels. However, nanoparticles appear to be
Initially sustained by progress in miniaturization, this new devel- ideally suited for applications in which fluids flow through small
opment has helped form a highly interdisciplinary science and passages, because the nanoparticles are stable and small enough
engineering community. Nanotechnology is expected to have ap- not to clog flow passages. This will open the possibility of using
nanoparticles in microchannels for many envisioned applications.
Successful employment of nanofluids will support the current
Contributed by the Heat Transfer Division for publication in the JOURNAL OF
HEAT TRANSFER. Manuscript received by the Heat Transfer Division, July 7, 1998;
trend toward component miniaturization by enabling the design of
revision received, Jan. 25, 1999. Keywords: Conduction, Enhancement, Heat smaller and lighter heat exchanger systems. One such application
Transfer, Nanoscale, Two-Phase. Associate Technical Editor: A. Majumdar. is in the next generation of cooling systems for mirrors and

280 / Vol. 121, MAY 1999 Copyright 1999 by ASME Transactions of the ASME

Downloaded 07 Sep 2009 to 169.229.32.136. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
monochromators in high-intensity X-ray sources such as Ar- Table 1 Average particle diameters
gonne's Advanced Photon Source. Because the X-ray beam creates
tremendous heat, cooling rates of 2000-3000 W/cm2 must be CuO A1203
achievable. An advanced cooling process that employs microchan- Number-weighted particle diameter (nm) 18.6 0.6 24.4 1.0
nels with nanofiuids could provide more efficient cooling than any Area-weighted particle diameter (nm) 23.6 1.0 38.4 2.0
existing approach (Lee and Choi, 1996).
Another advantage of nanofiuids over conventional suspensions
of coarse-grained particles in fluid systems is in their expected
improved abrasion-related properties. Because coarse-grained par- other processing techniques include the ability to produce particles
ticles are relatively large, and thus have significant mass, they can under cleaner conditions. For example, chemical techniques often
abrade surfaces that they contact. This results in shortened life- result in undesirable surface coatings on particles.
times of such components as water pumps and bearings. In con- If powders are produced by gas condensation, some agglomer-
trast, the small sizes and masses of nanoparticles would impart ation of individual particles occurs. It is well known, however, that
little kinetic energy in collisions with component surfaces and thus these agglomerates require little energy to fracture into smaller
would be expected to produce little or no damage. A recent paper constituents, and thus it is possible that even agglomerated nanoc-
(Hu and Dong, 1998) shows that titanium oxide nanoparticles in rystalline powders can be successfully dispersed into fluids and
oil, unlike conventional particles, reduce the friction coefficient result in good properties.
and increase resistance to wear. Therefore, it is possible that In this study, oxide nanofiuids were produced by a two-step
nanoparticles would actually improve the lubricating properties of method, in which first oxide nanoparticles are prepared, followed
heat transfer fluids. by a second step in which the powders are dispersed into the base
fluids in a mixing chamber. A1203 and CuO nanoparticles pro-
The primary objective of the present study is to investigate
duced by gas condensation (Nanophase Technologies Corp., Bun-
experimentally the thermal conductivity behavior of oxide
Ridge, IL) were used as the oxide nanoparticles, and were dis-
nanofiuids with low particle concentrations (1-5 vol.%).
persed in water or ethylene glycol. A polyethylene container (3.78
L) was used as a mixing chamber and was shaken thoroughly to
2 Production and Characterization of Nanofiuids ensure a homogeneous suspension.
Nanofiuids are produced by dispersing nanometer-scale solid Nanoparticles were characterized before and after dispersion in
particles into liquids such as water, ethylene glycol, or oils. Mod- liquids. Characterization of particle size before dispersion in liq-
ern fabrication technology provides great opportunities to process uids was carried out using transmission electron microscopy tech-
materials at micrometer and nanometer scales. "Nanostructured" niques. Both CuO and A1203 particles exhibited a log-normal size
or "nanocrystalline" materials are comprised of individual crystal- distribution, as expected for particles produced by the gas-
lites that are typically <50 nm in size. Bulk materials formed by condensation process. The particle distributions were fit to a log-
the consolidation of large numbers of nanocrystallites often exhibit normal function and both number and area-weighted values were
greatly modified properties compared to those of conventional determined for the average particle diameters. The number-
coarser-grained materials (for a review, see Gleiter, 1989). The weighted size distribution weights all particles equally. The area-
change in behavior is due to the large fraction of atoms in a weighted size distribution weights the particles according to their
nanostructured material that are located within only a few atomic surface areas. Hence, the area weighting will emphasize larger
spacings of one or more crystallite surfaces (called "grain bound- particles more than small ones. Since heat transfer is a surface
aries" in a bulk consolidated nanostructured material). Noncon- phenomena, the area-weighted average is expected to produce the
solidated nanocrystalline powders also show modified properties more relevant values. The values that come from the fits are shown
due to their large free surface areas (e.g., catalytic activity is in Table 1.
enhanced significantly using nanocrystalline powders). One interesting and important thing is that we observe the CuO
Much progress has been made recently in the processing of particles to be smaller than the A1203 particles, which is consistent
nanocrystalline materials. Current nanocrystalline materials tech- with our observation in this study that CuO-nanofluids exhibit
nology can produce large quantities of powders with average better thermal conductivity values than Al203-nanofluids.
particle sizes in the 10-100 nm range. Several processing routes Transmission electron microscopy of nanoparticles distilled
can be used to produce nanocrystalline materials, such as gas- from nanofluid solutions was used to characterize the effects of
condensation, mechanical attrition, or chemical precipitation tech- solutions on agglomeration behavior. Transmission electron mi-
niques (for a review of the many possible processing techniques, crographs showing the particle size and morphologies of agglom-
see Gleiter (1989)). Gas-condensation processing (Granqvist and erated A1203 and CuO powders are seen in Fig. 1. It can be seen
Buhrman, 1976; Kimoto et al., 1963), which was used in the that both A1203 and CuO nanoparticles agglomerate to form much
current experiments, is the most common method used currently in larger particles than individual grains before dispersion. Some
laboratory production of nanostructured materials. agglomerates are as large as 100 nm. In addition to CuO exhibiting
Briefly, gas-condensation processing involves the vaporization a smaller grain size than A1203, the CuO agglomerate sizes are
of metallic or nonmetallic precursor species in the presence of a smaller than those of A1203. No attempt was made to break up the
controlled gas pressure (typically a few Torr of an inert gas such agglomerated nanoparticles because even agglomerated nanopar-
as helium). Collisions between the vapor and the inert gas result in ticles were successfully dispersed into liquids and formed stable
the condensation of nanometer-sized particles of the precursor suspensions.
material, which are then collected and used in either powder or To verify systematically the effects of volume fraction and
bulk consolidated form. Advantages of gas-condensation over conductivities of solid phase and liquid on the thermal conductiv-

Nomenclature
a = radius R = resistance Subscripts
g = Euler's constant t = time p = particle
k = thermal conductivity T = temperature
o = base fluid
K = thermal diffusivity a = particle volume fraction
ref = reference temperature
n = shape factor 8T = temperature difference
q = electric power i/; = sphericity

Journal of Heat Transfer MAY 1999, Vol. 121 / 281

Downloaded 07 Sep 2009 to 169.229.32.136. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
transient hot-wire method were made by Kestin and Wakeham
(1978), Roder (1981), and Johns et al. (1988). A hot-wire system
involves a wire suspended symmetrically in a liquid in a vertical
cylindrical container. The wire serves both as heating element and
as thermometer. Almost without exception, platinum is the wire of
choice.
The mathematical model that one attempts to approximate is
that of an infinite-line source of heat suspended vertically in an
infinite medium. The method is called transient because the power
is applied abruptly and briefly. The working equation is based on
a specific solution of Fourier's law and can be found in standard
text (Carslaw and Jaeger, 1959):
q l AK \
T
-^=4^lnl?Cfj' (1)

where T(t) is the temperature of the wire in the fluid at time t, TKf
is the temperature of the cell, q is the applied electric power, k is
the thermal conductivity, K is the thermal diffusivity of the fluid,
a is the radius of the wire, and In C = g, where g is Euler's
constant.
The relationship given by Eq. (1) implies a straight line for a
plot of 8T versus In (t). In practice, systematic deviations occur at
both short and long times. However, for each experimental mea-
surement, there is a range of times over which Eq. (1) is valid, that
is, the relationship between ST and In (0 is linear. The slope of the
ST versus In (i) relationship is obtained over the valid range, i.e.,
between times /, and t2, and, using the applied power, we calcu-
lated thermal conductivity from

where T2 7\ is the temperature rise of the wire between times


ti and t2. From the temperature coefficient of the wire's resistance,
the temperature rise of the wire can be determined by the change
in its electrical resistance as the experiment progresses.
Despite the advantages of the transient hot-wire method, it is
impossible to measure the thermal conductivity of electrically
conducting fluids because current flows through the liquids, the
Fig. 1(A) heat generation of the wire becomes ambiguous, and polarization
Fig. 1 Transmission electron micrographs showing the particle size
occurs on the wire's surface. This method is thus normally re-
and morphology of the agglomerated (a) Al 2 0 3 and (b) CuO powders stricted to electrically nonconducting fluids such as noble gases
used in this study and organic liquids.
Only a few attempts have thus far been made to expand the
transient hot-wire method to measure electrically conducting liq-
uids. Nagasaka and Nagashima (1981) used a platinum wire (di-
ity of nanofluids, four oxide nanofluids with different volume
ameter = 40 /urn) coated with a thin electrical insulation layer
fractions were produced. These four systems consisted of CuO in
(thickness = 7.5 (am) to measure the thermal conductivity of an
water, CuO in ethylene glycol, A1203 in water, and A1203 in
NaCl solution, and they analyzed the effects on the thermal con-
ethylene glycol, with particle loadings up to a maximum of 5 ductivity measurement due to this thin insulation layer. Because
vol.%. nanofluids are likely to be electrically conducting (metallic nano-
particles and the suspending fluid such as water are electrically
3 Experiments conducting materials), the ordinary transient hot-wire technique
To measure the thermal conductivity of nanofluids, we used a cannot be used. Therefore, Nagasaka and Nagashima's method has
transient hot-wire method. Because our nanofluids are electrically been adopted in this experiment.
conductive, this caused difficulty in applying the ordinary transient
hot-wire technique. A new hot-wire cell and electrical system have 3.2 Experimental Apparatus and Procedure. A transient
been designed for our experiment according to the method pro- hot-wire cell was designed and constructed specifically for the
posed by Nagasaka and Nagashima (1981). measurement of the thermal conductivities of nanofluids. The
experimental apparatus and the electrical circuit used in this study
3.1 Transient Hot-Wire Method. In this study, a transient are shown schematically in Fig. 2. In the Wheatstone bridge, Rw is
hot-wire method has been adopted because recent advances in the resistance of the hot wire, /?, is a 1 kil potentiometer, R2 is a
electronic techniques have helped to establish this method as one 1 kCl resistor, R, is a 15 il resistor, and RA is a 20 il dummy
of the most accurate ways to determine fluid thermal conductivity. resistor used to stabilize the DC power supply. Adjusting the
The advantage of this method lies first in its almost complete resistance of the potentiometer R, allows the offset voltage from
elimination of the effects of natural convection, whose unwanted the Wheatstone bridge to be canceled out and thus the high-voltage
presence presents problems for measurements made with a steady- gain of the analog-to-digital (A/D) converter can be used.
state apparatus. In addition, the method is very fast relative to Platinum is used for the hot wire because its resistance/
steady-state techniques. temperature relationship is well known over a wide temperature
The major expositions of both theory and application of the range. In the resistance thermometer grade of Pt wire, the wire

282 / Vol. 121, MAY 1999 Transactions of the AS ME

Downloaded 07 Sep 2009 to 169.229.32.136. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
DC Power
Supply

PC for Data Storage


Nanofluids in
Cylinder
Wheatstone Bridge
Fig. 2 Schematic diagram of transient hot-wire apparatus for measuring thermal
conductivities of nanofluids (A/0 = analog-to-digital)

diameters available are 25.4 (xm, 50.8 /xm, and 76.2 ftm. A water and ethylene glycol in the temperature range of 290 K to 310
76.2-/xm diameter Pt wire was used because the other two wires K and at atmospheric pressure.
were too fragile to be coated with electrical insulation, and we Typical temperature rises of the wire in water or ethylene glycol
believed that the ratios of wire diameter to thickness of the insu- are shown in Fig. 3. Thirty data points from 3 to 6 s were used to
lation layer were too low for the other two wires. The wire was calculate the slope of the temperature rise. We can see that while
welded to rigid copper supporters, with one side of a supporter the electric power applied to the wire for water and ethylene glycol
connected to a rubber string so the wire tension could be adjusted is the same, the slopes of the temperature rises are different, which
to hold the wire straight. means that the thermal conductivities of water and ethylene glycol
The wire and welded spots were coated with an epoxy adhesive are different, as expected. Because the thermal conductivity of
which has excellent electrical insulation and heat conduction. The water is greater than that of ethylene glycol, the slope for water is
thickness of the coating on the wire is estimated to be less than 10 lower than that for ethylene glycol.
/xm. Because the specific resistance of the 76.2 /xm wire is low, a Figure 4 shows the measured thermal conductivities derived
very long wire (the length is 180 mm, and the resistance at 20C from Fig. (3) using Eq. (2), along with literature data for water and
is 4.594 Cl) was used to provide the high resistance required for a ethylene glycol (Touloukian and Ho, 1970). To determine the
hot-wire sensor. We did not use a two-wire compensation system uncertainty in the thermal conductivity measurement, we used Eq.
(Roder, 1981) to eliminate effects due to axial conduction at both (2) as the data reduction equation and performed an uncertainty
ends of the wire because identical geometry of the welded points analysis based on the presentations of Abernethy et al. (1985) and
and diameter of the coated hot wires cannot be guaranteed. Instead, Coleman and Steele (1989). The applied electric power was mea-
the long hot wire (wire length-to-diameter ratio is ^2300) was sured with an uncertainty of 0.26 percent and the wire temperature
used as a hot-wire sensor to minimize end conduction loss. A 500 variation with time was measured with an uncertainty of 1.4
mL polypropylene cylinder with inner diameter of 50 mm and percent. The uncertainty in the thermal conductivity for water at
length of 340 mm was used as the nanofluid container; the outside 304 K was 1.43 percent. The calibration data in Fig. 4 show that
of the cylinder is thermally insulated with glass fiber. The hot-wire our thermal conductivity measurement system and procedure is
sensor system is located at the center of the cylinder and the accurate to 1.5 percent in the temperature range tested, which is in
alignment to the direction of gravity is adjusted by the string excellent agreement with the results of the uncertainty analysis.
system described above. Because the enhanced thermal conductivity of nanofluids is ex-
Switching the power supply from R4 to the Wheatstone bridge pected to be >1.5 percent of the base fluids at the minimum
initiates the voltage change in the hot wire, and this varying volume fraction of 0.01, this accuracy is considered adequate for
voltage over time is recorded by the A/D converter with resolution quantitatively determining the thermal conductivity of nanofluids.
of 1.53 mV at a sampling rate of ten times per second (compared
4.2 Measurement of Thermal Conductivity of A1203 and
to Nagasaka and Nagashima's six times per second for NaCl
CuO Nanofluids. The thermal conductivity measurements were
solution). From this measured voltage variation and Ohm's law
made at room temperature and no attempt was made to maintain
applied to the electric circuit shown in Fig. 2, the resistance change
the temperature of nanofluids at a constant temperature because the
of the wire and the heating current through the wire can be
fluctuations of the room temperature in the laboratory were very
calculated. Finally, the temperature variation of the wire can be
small.
calculated by the temperature-resistance relationship of the Pt
wire. A value of 0.0039092/C (Bentley, 1984) has been used for Figure 5 shows the measured thermal conductivity of four oxide
the resistance temperature coefficient of the Pt wire. A linear nanofluids as a function of nanoparticle volume fraction. The
curve-fit of the experimental data in the log scale was performed results show that nanofluids, containing only a small amount of
by using a commercial software. nanoparticles, have substantially higher thermal conductivities
than the same liquids without nanoparticles. For the copper oxide/
ethylene glycol system, thermal conductivity can be enhanced by
4 Results and Discussion more than 20 percent at a volume fraction of 0.04 (4 vol.%).
Our experimental results clearly show that the thermal conduc-
4.1 Calibration With Base Fluids. The base fluids used in tivity ratios increase almost linearly with volume fraction, but with
this study as the suspending liquids are deionized water and different rates of increase for each system. The present experimen-
ethylene glycol. To establish the accuracy of our thermal conduc- tal data also show that the thermal conductivity of nanofluids
tivity measurements, calibration experiments were performed for depends on the thermal conductivities of both the base fluids and

Journal of Heat Transfer MAY 1999, Vol. 121 / 283

Downloaded 07 Sep 2009 to 169.229.32.136. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
u

u
3
2
I

10
Time, t (s)

Fig. 3 Typical temperature rises versus logarithm of time

particles. For nanofluids using the same nanoparticles, the conduc- tive thermal conductivity of suspensions containing spherical par-
tivity ratio increases of ethylene glycol nanofluid systems are ticles increases with the volume fraction of the solid particles. For
always higher than those of water nanofluid systems. For nanoflu- nonspherical particles, it is known that the thermal conductivity of
ids using the same liquid, the conductivity ratio of the CuO system suspensions depends not only on the volume fraction of the par-
is always higher than that of the A1203 system. We believe that ticles, but also on the shape of the dispersed particles.
CuO results in better thermal conductivity values than A1203 Hamilton and Crosser (1962) developed an elaborate model for
primarily because the CuO particles are smaller than the A1203 the effective thermal conductivity of two-component mixtures as a
particles as shown in Table 1. In addition, it is possible that the function of the conductivity of the pure materials, the composition
processing of the CuO nanoparticles could have resulted in incom- of the mixture, and the shape of the dispersed particles. The
plete oxidation, which would then be manifested by the presence conductivity of two-component mixtures can be calculated as
of a small amount of unreacted high thermal conductivity Cu being follows:
present in addition to the CuO.
4.3 Comparison With Hamilton and Crosser Model. Be- kp+ (n - l)k0 - (re - \)a(k0 - kp)
k = kQ (3)
cause of the absence of a theory for the thermal conductivity of k + (n - l)k0 + a(k0 - k)
nanofluids, an existing model for the solid/liquid system has been
used to compare the predicted values with the measured thermal where k is the mixture thermal conductivity, kQ is the liquid
conductivity of nanofluids. Maxwell's model shows that the effec- thermal conductivity, kp is the thermal conductivity of solid par-

0.7

0.6

|
0.5 - water
ethylene glycol
o water: measured
0.4 ethylene glycol: measured

13

0.3

0.2
290 295 300 305 310

Temperature, T (K)

Fig. 4 Validation of present transient hot-wire method measuring thermal conduc-


tivity of base fluids; water, ethylene glycol

284 / Vol. 121, MAY 1999 Transactions of the ASME

Downloaded 07 Sep 2009 to 169.229.32.136. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
1.50
- e water + A1 2 0 3
- a water + CuO
- ethylene glycol + A1 2 0 3
1.40 - ethylene glycol + CuO

1.30

a
a 1.20
o
u

1.10

1.00
0 0.01 0.02 0.03 0.04 0.05 0.06

Volume fraction

Fig. 5 Enhanced thermal conductivity of oxide nanofluids system

tides, a is the particle volume fraction, and n is the empirical significant in enhancing the thermal conductivity of nanofluids, as
shape factor given by shown in Fig. 6(c-d) for CuO with krlka < 4.
= 3/^, (4) 4.4 Comparison With Experiments of Masuda et al. To
our knowledge no experimental studies have been carried out on
where ty is sphericity, defined as the ratio of the surface area of a the thermal conductivity of CuO nanofluids. Direct comparison
sphere (with a volume equal to that of the particle) to the surface between the present experimental data and those of other investi-
area of the particle. This model shows that nonspherical shapes gators is possible only for water/Al203 nanofluids. Recently, Ma-
will increase conductivity above that of spheres. suda et al. (1993) showed experimentally that A1203 particles at a
Although Eq. (3) is highly nonlinear, if the thermal conductivity volume fraction of 4.3 percent can increase the thermal conduc-
ratio of solid phase to liquid phase, kplkQ : 1 and the volume tivity of water by 30 percent.
fraction is low, Eq. (3) can be linearized as follows: For comparison, their experimental data and present work for
k/k0" 1 + net. (5) water/Al203 nanofluids are shown in Fig. 7. The thermal conduc-
tivity ratios in our present experiments are lower than those of
Applying the Hamilton and Crosser model to aluminum oxide Masuda et al. by more than 20%. This is not surprising, since the
and copper oxide nanoparticles in water and ethylene glycol, we mean diameter of A1203 particles used in the experiments of
estimated the thermal conductivity ratios for two values of n (6 for Masuda et al. is 13 nm, which in the present experiments is 38 nm.
cylinders and 3 for spheres). The present experimental thermal Because surface-area-to-volume ratio is three times larger for
conductivities of oxide nanofluids are compared with the predicted particles with a 13-nm diameter than for particles with a 38 nm
results of Hamilton and Cross using Eq. (3). Comparisons of diameter, a greater improvement in effective thermal conductivity
increase of the thermal conductivity ratios predicted by the is expected.
Hamilton-Crosser model with results from the present experiments Another reason for the significant differences is that Masuda
are shown in Fig. 6(a) for water/Al203 nanofluids, Fig. 6(b) for et al. used the electrostatic repulsion technique and a high-
ethylene glycol/Al203 nanofluids, Fig. 6(c) for water/CuO speed shearing disperser (up to *20,000 r/min). These tech-
nanofluids, and Fig. 6(d) for ethylene glycol/CuO nanofluids. niques might change the morphology of the nanoparticles,
For A1203 nanofluids, it is seen from Fig. 6(a) and Fig. 6(b) that which was emphasized in the Hamilton-Crosser model as a
the predicted thermal conductivity ratios for spheres (n = 3) are shape factor n. Mainly because we did not use such techniques,
in good agreement with the present experiments. This is consistent nanoparticles in our systems were agglomerated and thus be-
with the TEM micrographs which show that A1203 particles are came larger and more spherical than those used by Masuda et al.
quite close to perfect spheres. The thermal conductivity of A1203 In fact, Fig. 7 clearly shows that the measured data by Masuda
nanofluids can be predicted satisfactorily by the Hamilton-Cross et al. agree better with the estimated thermal conductivity from
model because these nanofluids contain large agglomerated spher- the Hamilton and Crosser model, assuming the shape factor of
ical particles. n = 6 (cylinders), and that our data agree better with the model
Figures 6(c) and 6(0^ show the measured thermal conductivity predictions with the shape factor of n = 3 (spheres). However,
of nanofluids containing CuO, along with the model predictions. It it is important to note that even the model profile for nonspheri-
is seen that the agreement between the present experiments and cal particles begins to diverge from the experimental data of
model predictions is not good. The measured thermal conductivity Masuda et al. at a very low volume fraction. This strongly
ratios for CuO nanofluids are substantially larger than the suggests that not only particle shape but also size is considered
Hamilton-Crosser model predictions. It is possible that a small to be dominant in enhancing the thermal conductivity of
quantity of metallic Cu in the CuO material used in these exper- nanofluids. Therefore, it is reasonable to expect that the model
iments increased the thermal conductivity of CuO nanofluids. will fail to predict the thermal conductivity of nanofluids con-
In the Hamilton-Crosser model, the thermal conductivity of taining particles smaller than 13 nm.
suspensions depends strongly on particle shape when the ratio of Because heat transfer between the particles and the fluid takes
the conductivities of the two phases kp/k0 is high as shown in Fig. place at the particle-fluid interface, our expectation is that if we
6(a-b) for alumina. When kp/k0 = 0(1), particle shape is not have more interfacial area, then we would expect heat transfer to

Journal of Heat Transfer MAY 1999, Vol. 121 / 285

Downloaded 07 Sep 2009 to 169.229.32.136. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
1.50
O present experiments
Hamilton-Crosser model: spheres
1.40 - Hamilton-Crosser model: cylinders

1.30

1.20

1.10 O O

1.00
0.01 0.02 0.03 0.04 0.05 0.06
Volume fraction
Fig. 6(a)

1.50
O present experiments
Hamilton-Crosser model: spheres
1.40 - Hamilton-Crosser model: cylinders

1.30

1.20

1.10
H

1.00
0.01 0.02 0.03 0.04 0.05 0.06
Volume fraction

Fig. 6(b)

Fig. 6 Comparison of increase of thermal conductivity ratio between Hamilton-Crosser model and present experiment (a) water/AI2Oa nanofluids, (b)
ethylene glycol/AI 2 0 3 nanofluids, (c) water/CuO nanofluids, and (d) ethylene glycol/CuO nanofluids

be more efficient and thus quicker. If the behavior was just a interest in the present study, the phonon mean-free path can be
simple rule of mixtures, where all that matter is the volume of estimated using an equation that Debye derived (Geiger and
particles, then particle surface area wouldn't matter. However, our Poirier, 1973). Calculations show that for oxides such as A1203,
thermal conductivity data indicate this not to be the case. There- the estimated phonon mean-free path is comparable to the particle
fore, it is desirable to use particles with a large-surface-area-to- sizes we used. Therefore, the intrinsic thermal conductivity of
volume ratio. Previous theoretical work by Hamilton and Crasser oxide nanoparticles may be reduced compared to bulk oxides due
(1962) focused on the possible effects of increasing particle sur- to the size effect. Supporting this idea, the thermal conductivity of
face area by controlling particle shapes to be nonspherical. How- yttria-stabilized zirconia (YSZ) thin films with controlled nanoc-
ever, improvement in surface area per particle volume attainable rystalline grain sizes was recently measured (Soyez et al., 1998).
by this strategy is limited. As particle size decreases, the surface At room temperature, approximately a factor of two reduction in
area of the particle decreases as the square of the length dimension, thermal conductivity of 10-nm grain-sized YSZ compared to that
while the volume decreases as the cube of the length dimension. of the bulk material is seen.
Because of this "square/cube law," the surface-area-to-volume All existing theories, models, and correlations for thermal
ratio of nanoparticles is three orders of magnitude greater than that conductivity (Maxwell, 1873; Hamilton and Crosser, 1962;
of microparticles. Therefore, a much more dramatic improvement Hashin and Shtrikman, 1962; Jeffrey, 1973; Jackson, 1975;
in effective thermal conductivity is expected as a result of decreas- Davis, 1986; Bonnecaze and Brady, 1991; Lu and Lin, 1996)
ing the particle size than can be obtained by altering the particle were developed for fluids with relatively large particles and are
shapes of larger particles. Because all previous studies of the very limited and often contradictory when applied to fluids
thermal conductivity of suspensions have been confined to those containing nanoparticles because, in these theories, the effec-
containing millimeter or micrometer-sized particles, existing mod- tive thermal conductivity of liquid/particle suspensions depends
els appear to be inadequate for nanofluids containing particles as only on the volume fraction and shape of the suspended parti-
small as 10 nm. cles, not on particle size.
It is well known that, in the microscale regime, the thermal There is a large body of literature that addresses the effect of
conductivity of a thin film material is much less than its bulk value, the boundary resistance between the particle and the fluid on the
due to the scattering of the primary carriers of energy (phonon thermal conductivity of composites and suspensions (Ben-
and/or electron) at its boundary (Flik and Tien, 1990; Majumdar, veniste, 1987; Chiew and Glandt, 1987; Concalves and Kolodz-
1998). For electrical insulators such as the oxide materials of iej, 1993; Auriault and Ene, 1994; Ni et al., 1997). However,

286 / Vol. 121, MAY 1999 Transactions of the ASME

Downloaded 07 Sep 2009 to 169.229.32.136. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
1.50
O present experiments

1.40 - _ Hamilton-Crosser model: cylinders

1.30

a 1.20
o

O
1.10 O
S o
0
1.00 O '" ' ' '
0.01 0.02 0.03 0.04 0.05 0.06
Volume fraction
Fig. 6(c)

1.50
O present experiments
Hamilton-Crosser model: spheres
1.40 Hamilton-Crosser model: cylinders

1.30

1.20

1.10

1.00
0.01 0.02 0.03 0.04 0.05 0.06
Volume fraction
Fig. 6(d)

these models predict decreased thermal conductivity of nanoflu- size, in contrast to the predictions based on boundary resistance.
ids due to increased surface boundary resistance when particle Also, these models are confined to conventional suspensions
size decreases. Comparison of our A1203 in water results with containing millimeter or micrometer-sized particles. Nanofluids
those of Masuda et al. (1993) clearly indicate that the effective appear to be quite different from conventional suspensions. The
conductivity for this system increases with decreasing particle much larger surface areas of nanophase powders relative to

O present experiments
Masuda et al.
1.40 Hamilton-Crosser model: cylinders
o
a

1.30
y
y
S*

D
1.20

*****-" ^^-*"
D ** ^ "^
1.10 ^"" ^^^To
" "
** ' ^-r-"""*^
. - - ^ - " * ^ - - '

*"* _ - P T " ^ ^

LOOP
0.01 0.02 0.03 0.04 0.05 0.06
Volume fraction

Fig. 7 Comparison of present experimental thermal conductivity ratios with experi-


ments of Masuda et al. and model predictions for water/AI 2 0 3 nanofluids

Journal of Heat Transfer MAY 1999, Vol. 121 / 287

Downloaded 07 Sep 2009 to 169.229.32.136. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
those of conventional powders not only markedly improve Ahuja, A. S 1975, "Augmentation of Heat Transport in Laminar Flow of Poly-
conduction heat transfer capabilities, but also increase the sta- styrene Suspensions. 1. Experiments and Results," J. Appl. Phys., Vol. 46, No. 8, pp.
3408-3416.
bility of suspensions. Only recently have nanoparticles of 50 Ashley, S 1994, "Small-scale Structure Yields Big Property Payoffs," Mechanical
nm or less become available to investigators. We need to Engineering, Vol. 116, No. 2, pp. 52-57.
develop a more comprehensive theory in the future to explain Auriault, J.-L., and Ene, H. I 1994, "Macroscopic Modelling of Heat Transfer in
the behavior of nanofluids. It appears that the thermal conduc- Composites with Interfacial Thermal Barrier," Int. J. Heat Mass Transfer, Vol. 37,
No. 18, pp. 2885-2892.
tivity of nanofluids is very much surface area and structure Bentley, J. P., 1984, "Temperature Sensor Characteristics and Measurement Sys-
dependent. Any new models of nanofluid thermal conductivity tem Design," J. Phys. E: Set Instrum., Vol. 17, pp. 430-439.
should include the surface area and structure dependent behav- Benveniste, Y 1987, "Effective Thermal Conductivity of Composites with a
ior as well as the size effect and boundary resistance. Thermal Contact Resistance between the Constituents," J. Appl. Phys., Vol. 61, No.
16, pp. 2840-2843.
Bonnecaze, R. R., and Brady, J. F., 1991, "The Effective Conductivity of Random
5 Conclusions and Future Research Plan Suspensions of Spherical Particles," Proc. R. Soc. Lond., Vol. A432, pp. 445-465.
Carslaw, H. S., and Jaeger, i. C , 1959, Conduction of Heat in Solids, 2nd Ed.,
To experimentally investigate the thermal conductivity behavior Oxford University Press, New York, p. 510.
of dilute nanofluids, we measured the thermal conductivity of four Chiew, Y. C , and Glandt, E. D 1987, "Effective Conductivity of Dispersions: The
oxide nanofluids (A1203 in water, A1203 in ethylene glycol, CuO in Effect of Resistance at the Particle Surfaces," Chem. Eng. Sci., Vol. 42, No. 11, pp.
water, and CuO in ethylene glycol) by a transient hot-wire method. 2677-2685.
Choi, U. S., 1995, "Enhancing Thermal Conductivity of Fluids with Nanopar-
The present experimental results show that nanofluids, contain- ticles," Developments and Applications of Non-Newtonian Flows, D. A. Siginer and
ing only a small amount of nanoparticles, have substantially higher H. P. Wang, eds., FED-Vol. 231/MD-Vol. 66, ASME, New York, pp. 99-105.
thermal conductivities than the same liquids without nanoparticles. Coleman, H. W., and Steele, W. G., 1989, Experimentation and Uncertainty
For the copper oxide/ethylene glycol system, thermal conductivity Analysis for Engineers, John Wiley and Sons, New York.
Concalves, L. C. C , and Kolodziej, J. A., 1993, "Determination of Effective
can be enhanced by more than 20 percent at a volume fraction of Thermal Conductivity in Fibrous Composites with Imperfect Thermal Contact be-
0.04 (4 vol.%). In the low-volume fraction range tested (up to tween Constituents," Int. Comm. Heat Mass Transfer, Vol. 20, pp. 111-121.
0.05), the thermal conductivity ratios increase almost linearly with Davis, R. H., 1986, "The Effective Thermal Conductivity of a Composite Material
volume fraction, but with different rates of increase for each with Spherical Inclusions," International Journal ofThermophysics, Vol. 7, No. 3, pp.
system. The present experimental data also show that the thermal 609-620.
Eastman, J. A., Choi, U. S Li, S., Thompson, L. J., and Lee, S., 1997, "Enhanced
conductivity of nanofluids depends on the thermal conductivities Thermal Conductivity through the Development of Nanofluids," Proceedings of the
of both the base fluids and particles: For nanofluids using the same Symposium on Nanophase and Nanocomposite Materials II, Vol. 457, Materials
nanoparticles, the conductivity ratio increases of ethylene glycol Research Society, Boston, pp. 3-11.
nanofluid systems are always higher than those of water nanofluid Flik, M. I., and Tien, C. L., 1990, "Size Effect on the Thermal Conductivity of
High-r c Thin-Film Superconductors," ASME JOURNAL OF HEAT TRANSFER, Vol. 112,
systems; For nanofluids using the same liquid, the conductivity pp. 872-881.
ratio of the CuO system is always higher than that of the A1203 Geiger, G. H., and Poirier, D. R., 1973, Transport Phenomena in Metallurgy,
system. Addison-Wesley, Reading, MA, p. 190.
Gleiter, H 1989, "Nanocrystalline Materials," Prog. Mater. Sci, Vol. 33, pp.
Comparisons between the measured thermal conductivity of 223-315.
nanofluids and the predicted values of a model developed by Granqvist, C. G., and Buhrman, R. A., 1976, "Ultrarine Metal Particles," J. Appl.
Hamilton and Crosser (1962) were made. The Hamilton-Crosser Phys., Vol. 47, p. 2200.
model is capable of predicting the thermal conductivity of nanoflu- Hamilton, R. L., and Crosser, O. K., 1962, "Thermal Conductivity of Hetero-
geneous Two-Component Systems," / & EC Fundamentals, Vol. 1, No. 3, pp.
ids containing large agglomerated A1203 particles. However, the 187-191.
model appears to be inadequate for nanofluids containing CuO Hashin, Z., and Shtrikman, S., 1962, "A Variational Approach to the Theory of the
particles. Further work is required to clarify the reasons for this Effective Magnetic Permeability of Multiphase Materials," J. Appl. Phys., Vol. 33,
discrepancy. No. 10, pp. 3125-3131.
Hu, Z. S., and Dong, J. X., 1998, "Study on Antiwear and Reducing Friction
To clearly understand the mechanisms of thermal conductivity Additive of Nanometer Titanium Oxide," WEAR, Vol. 216, pp. 92-96.
enhancement of nanofluids, further research is needed. One of the Jackson, D. J., 1975, Classical Electrodynamics, 2nd Ed., John Wiley and Sons,
major differences between experimental data and the theoretical London.
model is that the rate of increase of thermal conductivity depends Jeffrey, D. J., 1973, "Conduction Through a Random Suspension of Spheres,"
strongly on particle size. Therefore, not only the effect of particle Proc. R. Soc. Lond, Vol. A335, pp. 355-367.
Johns, A. I., Scott, A. C, Watson, J. T. R and Ferguson, D 1988, "Measurement
shape, which was considered by Hamilton and Crosser (1978), but of the Thermal Conductivity of Gases by the Transient Hot Wire Method," Phil.
more importantly, the effect of particle size should be investigated. Trans. R. Soc. Lond, Vol. A325, pp. 295-356.
We need to develop a more comprehensive theory in the future to Kestin, J., and Wakeham, W. A., 1978, "A Contribution to the Theory of the
explain the complex behavior of nanofluids. Transient Hot-wire Technique for Thermal Conductivity Measurement," Physica,
Vol. 92A, pp. 102-116.
In our study, the thermal conductivity of stationary nanofluids Kimoto, K., Kamilaya, Y., Nonoyama, M., and Uyeda, R., 1963, "An Electron
was considered. Several investigators have reported augmentation Microscope Study on Fine Metal Particles Prepared by Evaporation in Argon Gas at
of the effective thermal conductivity of suspensions with Low Pressure," Jpn. J. Appl. Phys., Vol. 2, p. 702.
millimeter-sized polystyrene particles under laminar flow (Ahuja Lee, S. P., and Choi, U. S., 1996, "Application of Metallic Nanoparticle Suspen-
sions in Advanced Cooling Systems," Recent Advances in Solids/Structures and
1975; Sohn and Chen 1981). Therefore, we expect that the effec- Application of Metallic Materials, Y. Kwon, D. Davis, and H. Chung, eds., PVP-Vol.
tive thermal conductivity of nanofluids under flow conditions 342/MD-Vol. 72, ASME, New York, pp. 227-234.
might be higher than that seen in the present experimental results. Lu, S., and Lin, H 1996, "Effective Conductivity of Composites Containing
Hence, heat transfer tests to assess the thermal performance of Aligned Spherical Inclusions of Finite Conductivity," J. Appl. Phys., Vol. 79, No. 9,
pp. 6761-6769.
nanofluids under controlled flow conditions are currently being Majumdar, A., 1998, "Microscale Energy Transport in Solids," Microscale Energy
conducted. Transport, C. L. Tien, A. Majumdar, and F. Gerner, eds., Taylor & Francis, Wash-
ington, DC.
Masuda, H., Ebata, A., Teramae, K., and Hishinuma, N., 1993, "Alteration of
Acknowledgments Thermal Conductivity and Viscosity of Liquid by Dispersing Ultra-fine Particles
This work was supported by the U.S. Department of Energy, (Dispersion of 7-Al 2 0 3 , Si0 2 , and Ti0 2 Ultra-fine particles)," Netsu Bussei (Japan),
Vol. 4, No. 4, pp. 227-233.
BES-DMS under Contract W-31-109-Eng-38, and by a grant from Maxwell, J. C , 1881, "A Treatise on Electricity and Magnetism," 2nd Ed., Vol. 1,
Argonne National Laboratory's Coordinating Council for Science Clarendon Press, Oxford, U.K., p. 435.
and Technology. Nagasaka, Y., and Nagashima, A., 1981, "Absolute Measurement of the Thermal
Conductivity of Electrically Conducting Liquids by the Transient Hot-wire Method,"
J. Phys. E: Sci. Instrum., Vol. 14, pp. 1435-1440.
References Ni, F Gu, G. Q and Chen, K. M., 1997, "Effective Thermal Conductivity of
Abemethy, R. B., Benedict, R. P., and Dowdell, R. B 1985, "ASME Measurement Nonlinear Composite Media with Contact Resistance," Int. J. Heat Mass Transfer,
Uncertainty," ASME Journal of Fluids Engineering, Vol. 107, pp. 161-164. Vol. 40, pp. 943-949.

288 / Vol. 121, MAY 1999 Transactions of the ASME

Downloaded 07 Sep 2009 to 169.229.32.136. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Roder, H. M., 1981, "A Transient Hot Wire Thermal Conductivity Apparatus for Dispersed Two-Phase Mixtures as Observed in a Low Velocity Couette Flow Exper-
Fluids," Journal of Research of the National Bureau of Standards, Vol. 86, No, 5, pp. iment," ASME JOURNAL OF HEAT TRANSFER, Vol. 103, pp. 47-51.
457-493. Soyez, G Eastman, J. A., DiMelfi, R. J., and Thompson, L. J., 1998, private
Rohrer, H., 1996, "The Nanoworld: Chances and Challenges," Microelectronic communication.
Engineering, Vol. 32, No. 1-4, pp. 5-14. Touloukian, Y. S and Ho, C. Y., eds 1970 to 1977, Thermal Properties of Matter,
Sohn, C. W and Chen, M. M 1981, "Microconvective Thermal Conductivity in The TPRC Data Series, Plenum Press, New York.

Journal of Heat Transfer MAY 1999, Vol. 121 / 289


All in-text references underlined in blue are linked to publications on ResearchGate, letting you access and read them immediately.
Downloaded 07 Sep 2009 to 169.229.32.136. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Potrebbero piacerti anche