Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

This article was downloaded by: [University of Arizona]

On: 19 January 2013, At: 23:49


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Experimental Heat Transfer: A Journal of Thermal


Energy Generation, Transport, Storage, and Conversion
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/ueht20

HYDRODYNAMIC AND HEAT TRANSFER STUDY OF


DISPERSED FLUIDS WITH SUBMICRON METALLIC OXIDE
PARTICLES
a b
Bock Choon Pak & Young I. Cho
a
Division of Mechanical Engineering, Chonbuk National University, Chonju, Chonbuk, Korea
b
Department of Mechanical Engineering & Mechanics, Drexel University, Philadelphia,
Pennsylvania, USA
Version of record first published: 23 Apr 2007.

To cite this article: Bock Choon Pak & Young I. Cho (1998): HYDRODYNAMIC AND HEAT TRANSFER STUDY OF DISPERSED FLUIDS
WITH SUBMICRON METALLIC OXIDE PARTICLES, Experimental Heat Transfer: A Journal of Thermal Energy Generation, Transport,
Storage, and Conversion, 11:2, 151-170

To link to this article: http://dx.doi.org/10.1080/08916159808946559

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to
anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents
will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should
be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims,
proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in
connection with or arising out of the use of this material.
HYDRODYNAMIC AND HEAT TRANSFER
STUDY OF DISPERSED FLUIDS WITH
SUBMICRON METALLIC OXIDE PARTICLES

Bock Choon Pak


Division ofMechanical Engineering, Chonbuk National University,
Chonju, Chonbuk; Korea

Young I. Cho
Downloaded by [University of Arizona] at 23:49 19 January 2013

Department of Mechanical Engineering & Mechanics, Drexel University,


Philadelphia, Pennsylvania, USA

a
Turbulent friction and heat transfer behaviors of disp.rs.d fluids .e., ultrafin. metallic
oxide particles suspended in water) in a circular pipe were investigated experimentally.
Viscosity measurements were also conducted using a Brookfield rotating viscometer. Two
different metallic oxide particles, y-alumina CAl,O,) and titanium dioxide (TiO,). with
mean diameters of 13 and 27 nm, respectively, were used as suspended particles. The
Reynolds and Prandtl numbers varied in the ranges 10'_10 5 and 6.5-12.3, respectively.
The viscosiJies of the dispersedfluids wiJh ,..Al,O, and TiO, particles at a 10% volume
concentration were approximately 200 and 3 times greater than thai of water, respectively.
These viscosity results were significantly larger than the predictions from the classical theory
of suspension rheology. Darcy friction factors for the dispersed fluids of the volume
concentration ranging from J% to 3% coincided well with KLJys' correlation for turbulent
flow of a single-phase fluid. The Nusselt number of the dispersedfluids for fully developed
turbulent flow increased with increasing volume concemmtion as weU as the Reynolds
number. However. it wasfound that the convective heat transfer coefficientof the dispersed
fluid at a volume concentration of 3% was 12% smaller than tluu of pure water when
compared under the condition of constant average velocity. Therefore, better selection of
particles having higher thermal conductivity and larger size is recommended in order
to utilize dispersed fluids as a working medium to enhance heat transfer performance. A
new correlation for the turbulent connective heat transfer for dilute dispersed fluids
with submicron metallic oxide particles is given by the following equation: Nu =
0.021 ReO.s Pr O.5•

Responding to the need for more efficient heat transfer systems, researchers
and engineers have introduced various heat transfer enhancement techniques since
the mid-1950s. The exponential increase in the number of research articles devoted

Received 24 December 1997; accepted 5 January 1998.


The authors acknowledge the financial support of The Electrical Engineering & Science
Research Institute, Korea, under Grant 96-G-17. The authors also acknowledge emeritus Professor
Masuda at Tohoku University, Japan, for useful discussions on the dispersion scheme of submicron
metallic oxide particles in liquid.
Address correspondence to Prof. Bock Choon Pak, Division of Mechanical Engineering, Chon-
buk National University, Chonju, Chonbuk, 560-756, Korea. E-mail: bcpak@moak.chonbuk.ac.kr

Experimental Heat Transfer, 11:151-170, 1998


Copyright © 1998 Taylor & Francis
0891.6152/98 $12.00 + .00 151
152 B. C. PAK AND Y. I. CHO

NOMENCLATURE
c specific heat, kJ Ikg K 1] viscosity, mPa . s
D inside tube diameter, m <\> concentration, %
f Darcy friction factor

"
k
convective heat transfer coefficient,
W/m'K
thermal conductivity. W 1m K
Subscripts

b bulk
Nu Nusselt number f fluid
Pr Prandtl number m mass
q" heat flux, W 1m' p particle
x axial distance from an inlet of heat r relative
transfer lest section, m v volume
p density, kg/m' w water
Downloaded by [University of Arizona] at 23:49 19 January 2013

10 this subject thus far manifests the importance of heat transfer enhancement
technology. Among the many articles, several recent review articles [1-3] and
handbook chapters [4, 5], covering a variety of methods for augmenting heat
transfer, are available in the literature.
Generally, enhancement techniques can be categorized as passive and active
methods, depending on whether or not an external power is required [3]. The
passive schemes, requiring no direct application of external power, include the use
of extended surfaces, swirl flow devices, and additives for liquids and gases. On the
other hand, the active methods consist of the vibration of the heated surface and
fluid, the injection or suction of fluid, etc. The application of the passive methods
in a liquid flow system includes the use of solid particles, gas bubbles, and
phase-change materials in carrier flow.
Ahuja [6, 7] performed experiments on the augmentation of heat transport in
the laminar flow of polystyrene suspensions. His experiments were based on the
overall heat transfer measurements in a shell-and-tube heat exchanger, employing
the Graetz entrance flow conditions. The results showed remarkable enhancements
of the Nusselt number and heat exchanger effectiveness when polystyrene spheres
were added to a single-phase liquid. Sohn and Chen [8, 9] reported the effective
thermal conductivity of a dispersed two-phase mixture in a low-velocity and
Couette flow condition. When the particle Peclet number was sufficiently high, the
thermal conductivity increased with increasing shear rate.
Recently, a new technique utilizing phase-change particles was introduced to
increase the convective heat transfer coefficient in a turbulent pipe flow in district
cooling and thermal control systems [10-13]. In particular, Choi et al. [13] used
tetradecane (i.e., phase-change materials, 0.4 mm in diameter when frozen) and
successfully demonstrated that the turbulent heat transfer could be significantly
enhanced in a circular pipe flow. A small amount of a surfactant (less than 300
ppm) could maintain the dispersion of solidifying tetradecane particles in a carrier
fluid, thus allowing the mixture slurry to be pumped without clogging the piping
system.
Masuda et al. [14] investigated the possibility of controlling thermophysical
properties of fluids by dispersing metallic oxide particles whose thermophysical
DISPERSED FLUIDS WITH SUBMICRON METALLIC OXIDE PARTICLES 153

properties were significantly different from those of carrier fluids. They demon-
strated that the effective thermal conductivity of the mixtures of liquid and
submicron metallic oxide particles (such as y-Al20 3 and Ti0 2 ) increased consider-
ably from the value of the liquid. They showed that the increases in the thermal
conductivity were approximately 32% and 11%, respectively, for y-Al 2 0 3 and Ti0 2
particles in water at a volume concentration of 4.3%. Hence, the dispersed fluids
might be used as a fluid medium for convective heat transfer enhancement as long
as the increase in the pressure drop due to the addition of the particles is not
significantly large.
To verify this concept in practical heat transfer problems, the fluid dynamics
and heat transfer characteristics of these dispersed fluids should be clearly under-
stood. The objective of the present study is to investigate the hydrodynamic and
Downloaded by [University of Arizona] at 23:49 19 January 2013

convective heat _transfer characteristics of dispersed fluids containing submicron


metallic oxide particles in turbulent pipe flow.

EXPERIMENTAL SETUP AND INSTRUMENTATION


A schematic illustration of the flow loop built in a recirculating mode is
shown in Figure 1. The flow loop includes a stainless steel reservoir tank (0.2 rrr'), a
pump, a bypass line, a surge tank, a calming chamber, a hydrodynamic entry
section, a heat transfer test section, a mixing chamber, and a fluid collection
apparatus for measuring and calibrating flow rates. The hydrodynamic entry
section and the heat transfer section were manufactured using a seamless, stainless
steel tube (type 304), of which the inside diameter and the total length were 1.066
ern and 480 ern, respectively. The hydrodynamic entry section in the current study
was long enough (i.e., x /D = 157) to accomplish fully developed flow at the

Inputs from 16 Thenmocouples and 2 Pressuretaps NO Converter


I -I I
I 1_______ I[ L
L_. HYDRODYNAMld ---:...-:...-:..-:,-----j
1~ ENTRYSECTlON! HEAT TRANSFER TEST SECTION r _J
(XIO"57) (XIO' 330) _J TI
res

Flow Control Valve

Chiller
Collecting
Reservoir

Mono Pump
Digital Weighing balance

Figure I. Schematic diagram of the experimental setup for pressure drop and heat transfer.
154 B. C. PAK AND Y. I. CHO

entrance of the heat transfer test section. The remaining parts of the flow loop,
except for the hydrodynamic entry and heat transfer test section, were constructed
from a welded, stainless steel tube with an outer diameter of 2.52 cm.
Fourteen thermocouples (Tvtype) and two pressure taps were mounted along
the heat transfer test section for temperature and pressure drop measurements,
respectively. Specifically, at two different locations, eight additional thermocouples
were installed circumferentially to check circumferential temperature variations
during tests. The thermocouples were cemented with a high-thermal-conductivity
epoxy to the outside surface of the heat transfer test tube. The thermocouples were
calibrated by the freezing and boiling points of distilled water before they were
attached to the test section. The pressure drop measurements were performed with
a Validyne DPI03-18 pressure transducer. Prior to the experiments, the pressure
Downloaded by [University of Arizona] at 23:49 19 January 2013

transducer was carefully calibrated with an inclined mercury manometer accurate


to 0.5 mmHg. The pressure drop measurements were conducted without heating
the test section prior to the heat transfer experiments, to minimize the tempera-
ture effect on the hydrodynamic behavior of test fluids. All the temperature and
pressure data were recorded directly into an IBM-compatible PC through a
16-channel, 16-bit, high-resolution data acquisition system.
To obtain a constant-heat-flux boundary condition, the heat transfer test
section was heated electrically by a constant DC power supply capable of delivering
a maximum of 15 kW at 15 V. In order to minimize parasitic heat loss resulting
from axial heat conduction, the test section was isolated both electrically and
thermally from its upstream and downstream sections by nylon bushings. To
minimize heat loss to the surroundings, the heat transfer section, as well as the
calming and mixing chambers, were wrapped with 7.6-cm-thick fiberglass insulation
blankets. Also, a rapid chiller, of which the cooling capability was 54,000 kJ /h, was
utilized, so that the bulk temperature of the test fluids could be kept constant at
the inlet of the test section.
The essential parameters measured during the test include outer wall temper-
atures along the heat transfer test section at 14 axial locations, bulk temperatures,
mass flow rates, electric power inputs, and pressure drops along the test section. In
particular, the inlet and outlet bulk temperatures were measured in both the
calming and mixing chambers.
The flow rate was controlled with two valves-one at the main flow loop and
the other at a bypass line (see Figure 1). Flow rates were measured directly by
weighing test fluids exiting from the outlet of the flow loop. A 2-m-long piece of
flexible vinyl tubing was connected to the end of the flow loop in such a way that
the flow rate measurement procedure would not change the flow rate itself, a
phenomenon that is not a trivial problem if a three-way valve is used.

PREPARATION OF DISPERSED FLUIDS AND


PROPERTY MEASUREMENTS
Submicron y-alumina (y-Al z0 3 ) and titanium dioxide (TiO z) particles were
chosen for the dispersed fluids used in the present study. The y-alumina is an
amphoteric substance, i.e., it is dissolvable in either acids or strong hydroxide bases
[15]. Heating y-alumina above 373 K converts it to a-alumina. Finely divided
DISPERSED FLUIDS WITH SUBMICRON METALLIC OXIDE PARTICLES 155

Table I. Physical properties of sub micron metallic oxide particles

Mean Thermal
Submicron diameter Density Specific surface conductivity Specific heat
particles (nrn)" (kgyrrr') area (m'/g)b (W/m K) (kJ/kg K)'

y-AJ,03 13 3,880 100 ± 15 36.0 0.773


TiO, (Anatase) 27 4,175 56.1 8.4 0.692

Q Data from the manufacturers.


b Data for 0-2% porosity, at 300 K.
c Data from [16], at 298 K.

alumina has a very high surface area, a property that makes it useful as a support
Downloaded by [University of Arizona] at 23:49 19 January 2013

for catalysts and as a dehydrating agent. Titanium is the ninth most abundant
element in the earth's crust and is well known for its unusual resistance to
corrosion under ordinary conditions. Titanium dioxide (Ti0 2 ) , also an amphoteric
substance, is the most important titanium compound; it is a bright white substance
used as a pigment in paint, paper, and many other products.
Distilled water was used for the suspending liquid medium. The mean
diameters of y-A1 2 0 3 and Ti0 2 particles were 13 and 27 nm, respectively. The
other physical properties of the particles are shown in Table 1, and micropho-
tographs observed with a magnification of 120,00 are shown in Figure 2. Both types
of particles have irregular, grainlike shapes with an aspect ratio of approximately 1.
Since the density difference between metallic oxide particles and water is so
great, a method to prevent the settling of the metallic oxide particles in water was

(a) y-AIP3 partiles (b) noz particles (Anatase type)

Figure 2. Microphotographs of submicron metallic particles. A magnification


of 120.000 was used.
t56 B. C. PAK AND Y. I. eHO

imperative. Two kinds of techniques could be applied [14]: one was to use the
electrostatic repulsion forces between particle surfaces, and the other was to use
surfactants. Since the latter method was found to cause significant changes in the
thermophysical properties of dispersed fluids [14], the former technique was used
in the present study. The principle and preparation procedure of dispersed fluids
are summarized briefly below.
When the submicro metallic oxide particle comes into contact with water, a
hydroxyl radical ( - OH) is formed at the surface of the metallic oxide particle. The
overall behavior of the particle-water interaction depends on whether the water is
acidic or alkalic. For example, the particle surface in acidic water (i.e., having
excess hydrogen ions) has a positive charge, since a hydrogen ion from the acidic
water is combined with a hydroxyl radical at the surface of the metallic particle. On
Downloaded by [University of Arizona] at 23:49 19 January 2013

the other hand, the surface of a metallic oxide particle in an alkalic water (i.e.,
having excess hydroxyl radicals) has a negative charge, since a hydrogen ion is
removed from the hydroxyl radical at the particle surface. Equations (1) and (2)
show these two reactions in acidic and alkalic water in general, respectively:

M - OH + H+-> MOH~ (l)

(2)

where M indicates a metal cation. In summary, alumina particles exhibit basic


properties in the presence of strong acids and acidic properties in the presence of
strong hydroxide bases. The reactions of the two compounds with hydrogen and
hydroxide ions are summarized below [15].

Compound Reaction with H+ produces Reaction with OH - produces

AI,O, or AI(OH), [Al(OH)4] or AIO:


TiO, or Ti(OH)4 [Ti(OH)4J'- or TiOj-

However, at a certain pH value, the mixture reaches an equipotential (or


equivalence) point, at which the numbers of positive ions (MOHi) and negative
ions (MO-) are exactly the same. In other words, if the pH of a dispersed fluid is
near the equipotential value, it will not be easy to disperse fine particles in water.
The equipotential point depends strongly on the type of metallic oxide particle.
From a series of experiments, the authors found that y-A1 2 0 3 and Ti0 2 particles
can be well dispersed homogeneously at pH values of 3 and 10, respectively. By
using these pH values, a so-called electric double layer is formed at the surface of
the submicron metallic oxide particle, and subsequently, the particle is suspended
almost indefinitely in water without forming a cluster due to the repulsive force
between particles.
In preparing dispersed fluids for a desired mass concentration, the required
amount of particles weighed by a precision balance was mixed with a carrier fluid,
and the pH value of the dispersed fluid was adjusted with an extremely small
amount of hydrochloric acid or sodium hydroxide, depending on the type of
particles. After stirring with a high-speed mixer at a speed of 10,000 rpm for
DISPERSED FLUIDS WITH SUBMICRON METALLIC OXIDE PARTICLES 157

approximately 2 h, the dispersed fluid was allowed to settle for 5-6 days. Any
possible sedimentation was checked carefully by visual observation and by measur-
ing the densities of collected samples from the middle and bottom of a reservoir.
The authors could not find any evidence of irreversible settling due to density
difference at the above-mentioned pH values (i.e., pH = 3 for y-A1 z0 3 and
pH = 10 for TiO z particles).
The viscosities of the dispersed fluids were measured over a wide range of
shear rate by a Brookfield rotating viscometer with cone-and-plate geometry. The
fundamental principle and overall procedure for measuring viscosity using a
cone-and-plate rotating viscometer can be found elsewhere [17). In order to
examine the effect of the volume concentration on the non-Newtonian behavior of
the viscosity of dispersed fluids, the volume concentration was varied from 1% to
Downloaded by [University of Arizona] at 23:49 19 January 2013

10%. The density of the mixture was measured by using hydrometers. The effective
thermal conductivities of dispersed fluids reported by Masuda et al. [14) were used
in the data reduction of the present study.

RESULTS OF VISCOSITY MEASUREMENTS


In the investigation of the rheology of suspensions, the volume concentration,
defined as the fraction of space of the total suspension occupied by the suspended
material, is often used instead of mass concentration. The suspension rheology
depends greatly on the hydrodynamic forces, which act on the surface of the
particles or aggregates of particles. Hence, the volume concentration, not the mass
concentration, is often used in defining concentration. However, it is much more
difficult to make a dispersed liquid precisely at a desired volume concentration
than at a desired mass concentration. Therefore, in the present study, the volume
concentration was determined from the mass concentration at the dispersed fluid
by the following equation:

(3)

where Pp and Pw are the densities of metallic oxide particles and water, respec-
tively, and <lJu and <lJm are the volume and mass concentrations (%) of the
dispersed fluid, respectively. Once the volume concentration is determined, the
density of the dispersed liquid could be determined from

(4)

Figure 3 gives the comparison of density values between experimentally


measured values using hydrometers at 298 K and those calculated from Eq. (4) at
various volume concentrations. The maximum deviation was 0.6% at a volume
concentration of 31.6%. Thus, the authors concluded that the above-mentioned
procedure to determine the density of a dispersed fluid was valid.
In the viscosity measurement with the Brookfield cone-and-plate viscometer,
three different cone angles of 0.8°, 1.565°, and 3.0° were used, depending on the
158 B. C. PAK AND Y. I. CRO

1160
0 y·AI,o, (calculated)
• y·AI,o, (experiment) ~
1120 D TiO, (calculated)
~
E
en
• TiO, (experiment)
~ ~ 0.6%
a.
1080 II
.?;o
'iii D

0
c:
'" •
1040 !l

Downloaded by [University of Arizona] at 23:49 19 January 2013

1000
0 2 3 4 5

Volume concentration. <1>, (%)

Figure 3. Density versus volume concentration for the dispersed fluids.


Calculated results were obtained using Eq. (4).

concentration of the test fluids. Prior to measuring the viscosities of the test fluids,
calibration runs for the viscometer were conducted with distilled water and a
standard viscosity fluid (4.3 mPa s at 298 K) supplied by the manufacturer of the
Brookfield viscometer. All the experiments were carried out at a constant tempera-
ture of 298 K. Figure 4 shows the viscosity results for water and a standard viscosity
sample, of which the average deviation was approximately 2% for all of the
calibration runs.

Standard sample
----~----;!r----~-;!r--1a~---8r-------

.?;o 10 0 ·•·•·•....·• .. ·········.··········111····..........·..·•·•···•···•.•...•.


'§ Water

:>'" 0 pH= 3
t:. pH= 6
+ pH = 12

Shear rate, y (1/s)

Figure 4. Calibration runs for the Brookfield viscometer and pH


effects on viscosity.
DISPERSED FLUIDS WITH SUBMICRON METALLIC OXIDE PARTICLES 159

Figure 4 also shows the influence of pH on viscosity. The viscosity measure-


ments of the same test fluids at three different pH values, 3, 6, and 12, were
conducted to examine the possible effect on the viscosity of adding a small amount
of hydrochloric acid or sodium hydroxide in preparation of the dispersed fluids.
The current experimental results, as shown in Figure 4, revealed no influence of
pH on the viscosity of single-phase fluids, as expected. The results in Figure 4 are
provided as a reference so that one can compare the effect of pH on viscosity with
and without metallic oxide particles.
Figures 5-8 show viscosity results for the two dispersed fluids used in the
present study. Figures 5 and 6 present the effect of volume concentration on the
viscosities of y-A1 20 3 and Ti0 2 dispersed fluids, respectively. It is worthy of note
that the range of shear rates was different for the two fluids due to limitations of
Downloaded by [University of Arizona] at 23:49 19 January 2013

the Brookfield viscometer in measuring torque. In general, the viscosities of the


two dispersed fluids increased with increasing volume concentration. Furthermore,
the viscosity of the dispersed liquid with y-A1 20 3 particles depicted the shear-
thinning behavior at approximately a 3% volume concentration. On the other
hand, the dispersed liquid with Ti0 2 particles showed shear-thinning behavior at a
10% volume concentration. Ippolito [18) reported the influence of sodium hydrox-
ide on the viscosity of aqueous bentonite suspensions. He reported that flocculated
systems showed the shear-thinning characteristics of non-Newtonian viscosity in
general and very high viscosities at a low shear rate, often giving the impression of
a yield stress.
Figure 7 presents the relative viscosity as a function of the volume concentra-
tion. The relative viscosity is usually defined as the ratio of the viscosity of a fluid
with suspended particles to that of the fluid alone. In the case of the dispersed
fluid with Ti0 2 particles shown in Figure 7, the relative viscosity of 10% volume
concentration, for example, is approximately 3, while that of the disperse system

• ..... •
..c
0 <1>\1 = 1,00%
<1>\1= 3,00%
<I>, = 5,00%
• <1>\1= 7,00%
• <I>, =10.00%

~
.~
• ••
n
.. ..
'" 10' Cc n n
5 Cc
.. .... .. . .... .
0 00 0 0 00 0

10' 10' 10 3

Shear rate. y (l/s)

Figure 5. Viscosity versus shear rate for the dispersed fluids with
y-Al,O, particles at different volume concentrations.
160 B. C. PAK AND Y. I. CHO

10'
0 <1>, = 0.99%

" 0
5.00%
<1>v= 3.16%
cf)v=

"'
oi
0..
.s
• <1>,
<1>, = 8.00%
• = 10.00%
I=' • •• • • •• •
~
'iii
8<Jl • •• • • •• •
5 0 00 0 0 00 0

" "" "


Downloaded by [University of Arizona] at 23:49 19 January 2013

0 00 0
10°
10' 10 2
10 3

Shear rate, y (1/s)


Figure 6. Viscosity versus shear rate for the dispersed fluids with
Ti0 2 particles at different volume concentrations.

with y-Al z0 3 particles at the same volume concentration is approximalely 200.


Present results in Figure 7 indicate significantly larger values than the values
predicted from the analytical theory by Batchelor [19] for dilute fluid dispersed
with solid spheres, which can be given as

7), = 1 + 2.5<1>" + 6.2<1>~ (5)

10 3
0 Y - 19.2
Y = 115.2
" y = 384.0
0

"" 10 2 --- Batchelor


~0
u
<Jl
's
Ql
>
~ 10'
Qj
~

2 4 6 8 10 12

Volume concentration, <1>, (%)

Figure 7. Relative viscosity versus volume concentration at different


shear rates.
DISPERSED FLUIDS WITH SUBMICRON METALLIC OXIDE PARTICLES 161

where 71, is the relative viscosity and <l>v is the volume concentration (%) divided by
100. Of note is that Batchelor's equation covers the present volume concentration
range (i.e., 1-10%). The uncertainty in Batchelor's equation is in the determina-
tion of the constant coefficient of the third term on the right-hand side of Eq. (5).
In particular, the coefficient value reported varies from 5 to 15. Even if one uses
the maximum coefficient 15, it accounts for no more than a 40% increase in
viscosity over the continuous liquid phase. The significant increase in the viscosity
of the dispersed fluids observed in the present study may be attributed primarily to
the viscoelectric effect, which has to do with the fact that the effective particle
dimension is larger than its radius and equal to the Debye length [20]. Rutgers [21]
also pointed out that the sphere size should have a profound influence on viscosity;
i.e., a much larger relative viscosity would have been observed if the sphere
Downloaded by [University of Arizona] at 23:49 19 January 2013

diameter was smaller than 0.5 JLm. Furthermore, the irregular shape of the
particles with a rough surface may have resulted in the increase of a specific
surface area, thus increasing the viscosity of the dispersed fluid. Figure 7 also
shows that the relative viscosity increases with decreasing shear rate at a constant
volume concentration, especially in the case of the dispersed fluid with y-A1 20 3
particles. The results agree qualitatively with previous experimental investigations
[22, 23], which presented the relative viscosity for monodisperse lattices at zero and
infinite shear rates.
Figures Sa and Sb present the temperature-dependent behavior of viscosity
for the y-A1 2 0 3 and Ti0 2 dispersed fluids, respectively. For comparison, the
corresponding curves for distilled water are shown as solid lines in the figures. The
viscosities of the dispersed fluids decreased asymptotically with increasing tempera-
ture. The rate of decrease of the viscosity with respect to temperature became
greater as the volume concentration increased for the two dispersed fluids, but
more conspicuously for the y-A1 2 0 3 fluid. The data from the curve fittings for
temperature-dependent viscosity for each volume concentration were used for the
data reduction procedure in the subsequent pressure drop and heat transfer
experiments to take into account the temperature effect on viscosity.

PRESSURE DROP AND HEAT TRANSFER EXPERIMENTS


System calibration runs were carried out with distilled water to check the
validity of the experimental apparatus and overall experimental procedure before
the experiments with dispersed fluids were performed. First, accuracy tests in heat
transfer measurements were carried out by comparing the electric energy supplied
to the stainless steel tube with the thermal energy removed by the fluid. The
energy balance ratio, E.B.R., is defined as

thermal energy removed by fluid


E.B.R. = --,-----,----'----,,-.,....,---'-------:-
electric energy supplied by power supply

Figure 9 shows the energy balance ratio, which was calculated by measuring
the bulk mean temperatures of water at the inlet and outlet of the test section, and
162 B. C. PAK AND Y. I. CHO

(a)
4

Iwater + r-AI,o,1 o <1>. = 1.34%


-6 <1>. = 2.78%
(i) 3
l1i - water
a,
.s 6
6
"" 2 6

~ 0
6
8
II)
0
6
5 6
0 6
0
0
0
Downloaded by [University of Arizona] at 23:49 19 January 2013

0
10 20 30 40 50 60 70 80
Temperature (0C)
(b)
1.4
IJ
water + TiO, 0 <1>. = 0.99%
I 6 <1>. = 2.04%
~ 1.2 IJ
II)
6 IJ <1>.=3.16%
l1i
c, 0 6
- water
.s 1.0
IJ

""
~ 0.8
8
II)

5
0.6

0.4
10 20 30 40 50 60 70
Temperature (0C)

Figure 8. Temperature effects on viscosity of the dispersed fluids: (a)


with y-AJ,O,: (b) with TiO, particles. The viscosity data shown were
measured at a shear rate of 395 s - I.

the voltage drop and current over the test section. The energy balance ratio was
within approximately ± 2% of unity. The uniform heat flux condition was ensured
by checking thermocouple readings at four circumferentially different points at two
axial locations (i.e., x/D = 105 and 210) along the test section. Several measure-
ments at different Reynolds numbers and heat inputs showed that the cirumferen-
tial temperature variations of the test tube were negligible.
The aecuracy of the pressure measurements was checked by measuring the
pressure drop along the test section with water. The Darcy friction factors were
calculated and compared with those of Blasius [24] and Kays [25]. Figure lOa shows
the Darcy friction coefficients obtained from the present study. The deviation of
the measured Darcy friction coefficients from the predicted values with both
DISPERSED FLUIDS WITH SUBMICRON METALLIC OXIDE PARTICLES 163

1.10

1.05

ci o ~
a:i 1.00
uj o o

0.95
Downloaded by [University of Arizona] at 23:49 19 January 2013

0.90
0 2 4 6 6 10

Electric heat input (kW)

Figure 9. Energy balance ratio (E.B.R.) of the heating system.

Blasius' and Kays' correlations was less than 2% over the range of the Reynolds
number.
The accuracy of the temperature measurements was also tested by comparing
the measured Nusselt numbers at the fully developed region of the test section to
the predicted values from the Dittus-Boelter [26] and Notter-Rouse equations [27].

Dittus-Boelter: Nu = 0.023 Reo.s PrO.4 (6)

Notter-Rouse: Nu = 5 + 0.015 Re o.s 56 PrO. 347 (7)

As depicted in Figure lOb, excellent agreement was observed for water between
the present results and these well-established correlations. The convective heat
transfer coefficient and the corresponding Nusselt number are usually calculated
using the following equations:

q"
h= (8)
Tw - Tb

hD
Nu=- (9)
kf

where hand Nu are the heat transfer coefficient and the Nusselt number,
respectively, and D and k , are the inside tube diameter and thermal conductivity
of the fluid, respectively. More detailed information on the data reduction proce-
dure can be found elsewhere [28].
The difference between inlet and outlet temperatures, as well as between wall
and bulk temperatures, should be small in order to minimize the effect of
temperature-dependent viscosity on the overall heat transfer coefficient. The
164 B. C. PAK AND Y. I. CHO

o Present Study
Blasius
Kays
Downloaded by [University of Arizona] at 23:49 19 January 2013

Reynolds number, Re
(b)
103
0 Present study
::J
z --- Dittus and Boelter
- Notter and Rouse
t;;
D
E
::J
C 10 2
""gj
Q)

::J
Z

10' 3
10 10' 10'

Reynolds number. Re
Figure 10. (a) Comparison of Darcy friction factors obtained with
distilled water in the present study with the well-known correlations (i.e.,
Blasius [24] and Kays [25]). (b) Comparison of the Nusselt numbers
obtained with distilled water in the present study with those calculated
from Dittus and Boelter [26Jand Notter and Rouse [27].

maximum temperature difference between the inlet and outlet was less than lOoe,
and the inlet temperature of the fluid was kept at 200 e for all runs. The local
Reynolds and Prandtl numbers were calculated with the thermophysical properties
based on local mean bulk temperature. The Nusselt number in the thermally fully
developed region was determined by averaging local Nusselt numbers over the heat
transfer test section. As shown in Figure lOb, the measured Nusselt numbers for
water are between the two values predicted by the well-established correlations
[26, 27].
It is essential in multi phase phenomena to define clearly the nondimensional
parameters such as Reynolds number and Prandtl number. Table 2 shows repre-
sentative values of effective thermal conductivity and specific heat for dispersed
fluids at 300 K. It is of note that the specific heat of dispersed fluids was calculated
DISPERSED FLUIDS WITH SUBMICRON METALLIC OXIDE PARTICLES 165

Table 2. Representative data for effective thermal conductivity and specific heat of dispersed fluids
with submicron metallic oxide particles at 300 K [14]

Volume Effective thermal Specific heat


Dispersed fluid concentration (%) conductivity (W 1m K) (klykg K) Pr a

y-A120 3 + water 1.34 0.675 4.133 8.51


2.78 0.745 4.084 12.33
4.33 0.810
TiO, + water 0.99 0.634 4.145 6.54
2.04 0.647 4.110 7.00
3.16 0.662 4.070 7.40
4.35 0.679
Water 0 0.613 4.179 5.83
Downloaded by [University of Arizona] at 23:49 19 January 2013

a Calculated with the effective thermal conductivity for each volume concentration.

using the following equation based on the volume fraction, since there were no
data available in the literature.

(10)

The Prandtl number of dispersed fluids was determined with these values by the
usual definition.
Figure 11 and 12 present the Darcy friction factors and heat transfer
coefficients as a function of the Reynolds number for the dilute dispersion fluids
with submicron y-Al Z0 3 and TiO z particles, respectively. The dispersed fluids of
dilute concentrations (i.e., less than 3% volume concentration) were used in
pressure drop and heat transfer measurements from a practical point of view.

0r-AI;O, (<I>, - 1.34%)


•r-Ar,o, (<I>, =2.78%)
A TiO, (<I>, =0.99%)
0 TiO, (<I>, =2.04%)
+ TiO, (<I>, =3.16%)
1 - Kays

-- ~

10·2 3
10 10' 10'

Reynolds number, Re

Figure 1 t. Darcy friction factors of dispersed fluids.


166 B. C. PAK AND Y. I. CHO

(a)
10'
SZ
1:
~
s:
C
Q)
'0
IE
Q) 10'
8 =
y-AI,o,(l\>, 1.34%)
~

~ =
y-AI,o,(l\>, 2.78%)
III
c: no, (l\>, 0.99%)
=
~ no, (l\>, 2.04%)
=
ro
Q)
0 no, (l\>, 3.16%)
=
Downloaded by [University of Arizona] at 23:49 19 January 2013

J: Dittus-Boelter
10 3
10' 10'
Reynolds number, Re

(b)
10 3

:>
z
G;
.c
E
:>
c: 0 =
y-AI,O,(l\>, 1.34%)

1z
:>
..•
.
=
y-AI,O,(l\>, 2.78%)
no, (l\>, 0.99%)
=
no, (l\>, 2.04%)
=
0 =
TiO, (l\>, 3.16%)
Dittus-Boelter
10'
10' 10'

Reynolds number, Re
Figure 12. (a) Heat transfer coefficient versus Reynolds number. (b)
Corresponding Nusselt number versus Reynolds number for dispersed
fluids with both y-A1,03 and TiO, particles. Solid lines represent the
data for a single-phase fluid.

Kays' correlation for the turbulent flow of a single-phase fluid is also shown
as a solid line in Figure 11 to facilitate comparison. As shown in the figure, the
friction coefficients for dispersed fluids coincide well with Kays' prediction regard-
less of the volume concentration, implying that pumping power can be estimated
with a well-known correlation when the volume concentration in water is less than
or close to 3%.
The deviation from the predicted values by Kays' correlation was within 3%
over the range of Reynolds numbers. The effect of particle size on the friction
DISPERSED FLUIDS WITH SUBMICRON METALLIC OXIDE PARTICLES 167

factor was not seen with the two different particle sizes (i.e., 13 and 27 nrn) as long
as the volume concentration was less than 3%.
Ahuja [6] reported that the Darcy friction factor for polystyrene suspensions
of 50 ILm in diameter at 20% volume concentration was approximately 7.5% less
(i.e., f = 59 IRe) than the corresponding value calculated from the fully developed
laminar flow correlation (i.e., f = 64/Re), an observation that can be attributed to
the so-called e-phenomenon [29]. However, the present results did not show such a
decrease in the friction factor, which may be attributed to the fact that the particle
size in the present study was much smaller than that used by Ahuja, the volume
concentration was low, and the flow in the present study was turbulent.
Figures 12a and 12b present turbulent heat transfer coefficients and the
. corresponding Nusselt numbers for dilute dispersed fluids. The Dittus-Boelter
Downloaded by [University of Arizona] at 23:49 19 January 2013

correlation for a single-phase fluid (i.e., zero concentration) is also shown as a solid
line for comparison. As shown in the figure, the convective heat transfer coefficient
increased with increasing Reynolds number and volume concentration. For a fixed
Reynolds number, the enhancement in the convective heat transfer coefficient due
to the addition of submicron y-A1 20 3 particles was approximately 45% at the
volume concentration of 1.34%, and approximately 75% at the concentration of
2.78%. The heat transfer enhancement for the Ti0 2 dispersed fluids was less than
that for the y-A1 20 3 fluids at the same concentration. When one compares the
percentage increase in the heat transfer coefficient and that in the effective
thermal conductivity of dispersed fluids (i.e., see Figure 12a and Table 2), the
former is substantially greater than the latter. The authors attribute this to
enhanced mixing caused by submicron particles near the walls.
Figure 13 shows a plot of Nusselt numbers versus Reynolds numbers. From
this plot, the authors have obtained a new correlation, the solid line in Figure 13,

10'
+ water (<I>, = 0 %)
0 y-AIP, (<I>, = 1.34%)
• y-AIP, (<I>, = 2.78%)
l>. TiO, (<I>, = 0.99%)

:t-Q.
• TiO, (<I>, = 2.04%)
c TiO, (<I>, = 3.16%)

--
Z
::J
10'

10' '-::- ~ __ ~_~ ~~~.......J

10' 10'

Reynolds number. Re

Figure 13. Dimensionless representation of heat transfer measure-


ments for the dispersed tluids.
168 B. C. PAK AND Y. J. CHO

which can be described by the following relationship:


Nu = 0.021 Reo.s Pro.s (10)
The above correlation was obtained by curve-fitting all the data for the dispersed
fluids including water. It is interesting to note that the exponent 0.8 in the
correlation of the Reynolds number occurs in a wide range of physical situations
involving turbulent flows. The present correlation again finds that the exponent 0.8
is also the best choice for dispersed fluids with submicron particles. The exponent
of the Prandtl number in the new correlation for dispersed fluids is 0.5 as shown in
Eq. (to), whereas the exponent of the Prandtl number for single-phase fluids is 0.4
or 0.347 [i.e., see Eqs. (6) and (7)], suggesting that the Nusselt number for the
dispersed fluids is slightly more dependent on the Prandtl number than that for the
Downloaded by [University of Arizona] at 23:49 19 January 2013

single-phase fluids. Note that the maximum deviation in the curve-fitting was 4.8%,
and the ranges of the Reynolds and Prandtl numbers tested were 104-10 5 and
6.54-12.33, respectively.
However, one needs to examine the pumping penalty due to the increased
viscosity of dispersed fluids. The dispersed fluid of 2.78% with y-AJ Z0 3 particles
will be used as an example. The viscosity of the dispersed fluid is approximately 3
mPa s, i.e., 300% larger than that of single-phase water. Assuming that the flow
rate and thus average velocity are fixed, the Reynolds number of the dispersed
fluid becomes 36.2% of that of water, considering the change in density of the
dispersed fluid. As a result, the friction factor calculated by Blasius' correlation for
the dispersed fluid is 28% greater than that of water. Since the pressure drop is
directly proportional to the friction factor and density when the flow velocity
remains constant, the pumping penalty due to the use of 2.78% y-AJ Z0 3 particles
is estimated to be 31%. Similarly, when the corresponding heat transfer is com-
pared under the condition of constant average velocity, the convective heat transfer
coefficient of the dispersed fluid is found to be 12% smaller than that of pure
water, because the Reynolds number of the dispersed fluid is 36.2% of that of
water, and the heat transfer coefficient is proportional to Re.o.8 This result is
attributed primarily to the significant increase in viscosity due to the extremely
small size of the particles and the relatively small increase in the effective thermal
conductivity of the dispersed fluid. Therefore, it is concluded that in order to
utilize dispersed fluids as a working medium to enhance heat transfer performance,
better selection of particles, i.e., selecting particles that have higher thermal
conductivity and larger size, is a crucial point.

SUMMARY AND CONCLUSIONS


Turbulent friction and heat transfer behaviors of dispersed fluids with submi-
cron y-AJ Z0 3 and TiO z particles were investigated experimentally in a circular
pipe. Viscosity measurements of those fluids were also conducted by using a
Brookfield viscometer. Important findings are summarized briefly below.

I. The viscosity for the dispersed fluids with y-AJ Z03 and TiO z particles
showed shear-thinning behavior at or above volume concentration of 3%
and 10%, respectively.
DISPERSED FLUIDS WITH SUBM/CRON METALLIC OXIDE PARTICLES /69

2. At the 10% volume concentration, the relative viscosity for the dispersed
fluid with y-A1 20) particles was approximately 200, while it was approxi-
mately 3 for dispersed fluids with Ti0 2 particles. These viscosity results
are significantly larger than the predictions from the classical theory of
suspension rheology. The submicron size of metallic particles in dispersed
fluids plays a significant role in the determination of viscosity.
3. The Darcy friction factors for dilute dispersion fluids used in the present
study coincided well with the values predicted from Kays' correlation for
turbulent flow of a single-phase fluid. Due to the increase in the viscosity
of dispersed fluids, there is an additional pumping penalty of approxi-
mately 30% at a volume concentration of 3%.
4. The Nusselt number for the dispersed fluids increased with increasing
Downloaded by [University of Arizona] at 23:49 19 January 2013

volume concentration as well as Reynolds number. However, it was found


that the convective heat transfer coefficient of the dispersed fluid was 12%
smaller than that of pure water when compared under the condition of
constant average velocity. Therefore, better selection of particles having
higher thermal conductivity and larger size is recommended in order to
utilize dispersed fluids as a working medium to enhance heat transfer
performance.
5. Under the experimental ranges of volume concentration (0-3%), the
Reynolds number 00 4-10 5 ) , and the Prandtl number (6.54-12.33) in the
present study, a new correlation for the dispersed fluids y-A1 20) and Ti0 2
particles is suggested as Nu = 0.021 Reo.s PrO. 5 .

REFERENCES
1. A. E. Bergles, Survey and Evaluation of Techniques to Augment Convective Heat and
Mass Transfer, in U. Grigull and E. Hahne (eds.), Progress in Heat and Mass Transfer
vol. 1, pp. 331-424, Pergamon Press, New York, 1969.
2. W. Nakayama, Enhancement of Heat Transfer, Heat Transfer 1982, Proc. 7th Int. Heat
Transfer Conf., vol. 1, pp. 223-240, Hemisphere, Washington, DC, 1982.
3. A. E. Bergles, Some Perspectives on Enhanced Heat Transfer-Second-Generation
Heat Transfer Technology, Trans. ASME, J. Heat Transfer, vol. 110, pp. 1082-1096,
1988.
4. A. E. Bergles, Augmentation of Heat Transfer, in Heat Exchanges Design Handbook, vol.
2, pp. 2.5.11-1-12, Hemisphere, Washington, DC, 1983.
5. A. E. Bergles, Techniques to Augment Heat Transfer, in Handbook of Heat Transfer
Applications, pp. 3-1-80, McGraw-Hill, New York, 1985.
6. A. S. Ahuja, Augmentation of Heat Transport in Laminar Flow of Polystyrene Suspen-
sions. 1. Experiments and Results, J. Appl. Phys., vol. 46, no. 8, pp. 3408-3416, 1975.
7. A. S. Ahuja, Thermal Design of a Heat Exchanger Employing Laminar Flowof Particle
Suspensions, Int. J. Heat Mass Transfer, vol. 25, no. 5, pp. 725-728, 1982.
8. C. W. Shon and M. M. Chen, Microconvective Thermal Conductivity in Disperse
Two-Phase Mixture as observed in a Laminar Flow, Trans. ASME, J. Heat Transfer, vol.
104, pp. 47-51, 1984.
9. C. W. Shon and M. M. Chen, Heat Transfer Enhancement in Laminar Slurry Pipe Flows
with Power Law Thermal Conductivity, Trans. ASME, J. Heat Transfer, vol. 106, pp.
539-542, 1984.
170 B. C. PAK AND Y. 1. CHD

10. K. E. Kasza and M. M. Chen, Development of Enhanced Heat Transfer/Transport/


Storage Slurries for Thermal-System Improvement, Argonne Natl. Lab. ANL-82-50,
1982.
11. B. C. Pak, Y. I. Cho, and H. G. Lorsch, Use of Advanced Law Temperature Heat
Transfer Fluid for District Cooling Systems, Int. District Heating and Cooling System
Conference, Virginia Beach, VA, vol. 85, no. 269, pp. 368-376, 21-24 June 1989.
12. P. Charunyakorn, S. Sengupta, and S. K. Roy, Forced Convection Heat Transfer in
Microencapsulated Phase Change Material Slurries: Flow in Circular Ducts, Int. J. Heat
Mass Transfer, vol. 34, no. 3, pp. 819-833, 1991.
13. E. Choi, Y. I. Cho, and H. G. Lersch, Forced Convection Heat Transfer with Phase-
Change Material Slurries: Turbulent Flow in a Circular Tube, Int. J. Heat Mass
Transfer, vol. 37, pp. 207-215,1994.
14. H. Masuda, A. Ebata, K. Teramae, and N. Hishinuma, Alteration of Thermal Conductiv-
Downloaded by [University of Arizona] at 23:49 19 January 2013

ity and Viscosity of Liquid by Dispersing Ultra-fine Particles (Dispersion of 1'-Al Z03 ,
SiO z and TiO z Ultra-Fine Particles), Netsu Bussei (Japan), vol. 4, no. 4, pp. 227-233,
1993.
15. J. C. Bailar, T. Moeller, J. Kleinberg, C. Guss, M. E. Castellion, and C. Metz, Chemistry,
3d ed., pp. 397-398, Harcourt Bruce Jovanovich, New York, 1989.
16. The Thermophysical Properties Research Center (TPRC), Thermophysical Properties of
Mattcr-Specific Heat (Nonmetallic Solids), Vol. 5, Purdue University, 1970.
17. H. A. Barnes, J. F. Hutton, and K. Walters, An Introduction to Rheology, pp. 25-35,
Elsevier, Amsterdam, 1989.
18. M. Ippolito, Rheology of Dispersed Systems-Influence of NaCI on the Viscous Proper-
ties of Aqueous Bentonite Suspensions, in G. Astarita et al. (eds.), Rheology, Vol. 2:
Fluids, pp. 651-658, Plenum Press, New York, 1980.
19. G. K. Batchelor, The Effect of Brownian Motion on the Bulk Stress in a Suspension of
Spherical Particles, J. Fluid Mech., vol. 83, part 1, pp, 97-117,1977.
20. Ronald F. Probstein, Physiochemical Hydrodynamics, pp. 185-190, Butterworths, 1989.
21. I. R. Rutgers, Relative Viscosity and Concentration, Rheol. Acta, vol. 2, pp. 305-438,
1962.
22. I. M. Krieger, Rheology of Monodisperse Lattices, Adv. Colloid Interface Sci., vol. 3, pp.
111-136,1972.
23. C. G. de Kruif, E. M. F. van Ievsel, A. Vrij, and W. B. Russel, Hard Sphere Colloidal
Dispersions: Viscosity as a Function of Shear Rate and Volume Fraction, J. Chern.
Phys., vol. 83, pp. 4717-4725, 1985.
24. H. Schlichting, Boundary Layer Theory, 7th cd., pp. 596-600, McGraw-Hill, New York,
1979.
25. W. M. Kays and M. E. Crawford, Convective Heat and Mass Transfer, 3d ed., pp.
244-249, McGraw-Hili, New York, 1993.
26. F. W. Dittus and L. M. K. Boelter, Heat Transfer in Automobile Radiators of the
Tubular Type, Uniu. Calif. Publ. in Eng., vol. 11, pp. 443-461,1930.
27. R. H. Notter and M. W. Rouse, A Solution to the Turbulent Graetz Problem-III. Fully
Developed Region Heat Transfer Rates, Chern. Eng. Sci., vol. 27, pp. 2073-2093, 1972.
28. B. C. Pak, Y. I. Cho, and S. U. S. Choi, A Study of Turbulent Heat Transfer in a
Sudden-Expansion Pipe with Drag-Reducing Viscoelastic Fluid, Int. J. Heat Mass
Transfer, vol. 34, no. 4/5, pp. 1195-1208, 1991.
29. H. L. Goldsmith and S. G. Mason, in F. R. Eirich (ed.), Rheology: Theory and
Applications, vol. 4, chap. 2, Academic Press, New York, 1967.

You might also like