experimental thermal and fluid sciencedli9/publications/etfs 2011 pao...convective thermal transport...

12
Laminar convective heat transfer of alumina-polyalphaolefin nanofluids containing spherical and non-spherical nanoparticles Leyuan Yu a , Dong Liu a,, Frank Botz b a Department of Mechanical Engineering, University of Houston, Houston, TX 77204, USA b METSS Corporation, Westerville, OH 43082, USA article info Article history: Received 19 May 2011 Received in revised form 27 September 2011 Accepted 12 October 2011 Available online 19 October 2011 Keywords: Nanofluids Convective heat transfer Pressure drop Thermal conductivity Viscosity Minichannel abstract As a potential candidate for advanced heat transfer fluid, nanofluids have been studied extensively in the literature. Past investigations were largely limited to thermophysical property measurements of aqueous nanofluids synthesized with spherical nanoparticles. No comprehensive information is yet available for convective thermal transport of nanofluids containing non-spherical particles, especially of the non- aqueous nanofluids. In this work, an experimental study was conducted to investigate the thermophys- ical properties and convective heat transfer of Al 2 O 3 -polyalphaolefin (PAO) nanofluids containing both spherical and rod-like nanoparticles. The effective viscosity and effective thermal conductivity of the nanofluids were measured and compared to predictions from several existing theories in the literature. It was found that, in addition to the particle volume fraction, other parameters, including the aspect ratio, the dispersion state and the aggregation of nanoparticles as well as the shear field, have significant impact on the effective properties of the nanofluids, especially of those containing non-spherical parti- cles. The pressure drop and convection heat transfer coefficient were also measured for the nanofluids in the laminar flow regime. Although established theoretical correlations provide satisfactory prediction of the friction factor and Nusselt number for nanofluids containing spherical nanoparticles, they fail for nanofluids containing rod-like nanoparticles. The results indicate that in a convective flow, the shear- induced alignment and orientational motion of the particles must be considered in order to correctly interpret the experimental data of the nanofluids containing non-spherical nanoparticles. Ó 2011 Elsevier Inc. All rights reserved. 1. Introduction Nanofluids have been studied extensively as a promising candi- date of advanced heat transfer fluids for thermal management applications in microelectronics, transportation, power and military systems [1]. While dramatic enhancement in thermal con- ductivity was reported in the early literature, most recent studies find the convective heat transfer of nanofluids is only mildly aug- mented and the measured thermal properties can be well corre- lated by the conventional effective medium theory (EMT) [2,3]. Wide discrepancies exist among different studies regarding how the dispersed nanoparticles alter the convective thermal transport of nanofluids. Moreover, the past investigations were conducted primarily for water-based nanofluids synthesized with spherical nanoparticles. Due to the simple geometry of these particles, it is reasonable to neglect the effects of the particle distribution and orientation on the effective rheological and thermal properties as well as on the convective heat transfer of the nanofluids. However, when non-spherical nanoparticles, such as carbon nanotubes (CNTs) and titanate nanotubes (TNTs), are used in formulating the nanofluids, these effects may no longer be ignored as the result of the hydrodynamic interactions between the particles and the surrounding fluid medium. Also, the use of aqueous nanofluids be- comes infeasible in many practical applications due to their limits in the dielectric property and the operating temperature range. Consequently, there is a need to systematically study the convec- tive thermal transport of non-aqueous nanofluids synthesized with non-spherical nanoparticles. Several experimental studies were performed to study the effective thermophysical properties and convective heat transfer of nanofluids containing non-spherical particles with high length-to-diameter ratio (aspect ratio). Cherkasova and Shan [4] studied the effects of particle aspect-ratio and dispersion state on the effective thermal conductivity of aqueous nanofluids with mul- ti-walled nanotubes (MWNTs). They found the effective thermal conductivity decreases with reduced MWNT aspect ratio, and the conductivity enhancement was primarily due to the presence of individualized long nanotubes rather than the bundles of aggre- gated low aspect ratio nanotubes. The measured conductivity data showed a good agreement with the EMT predictions, and no 0894-1777/$ - see front matter Ó 2011 Elsevier Inc. All rights reserved. doi:10.1016/j.expthermflusci.2011.10.005 Corresponding author. E-mail address: [email protected] (D. Liu). Experimental Thermal and Fluid Science 37 (2012) 72–83 Contents lists available at SciVerse ScienceDirect Experimental Thermal and Fluid Science journal homepage: www.elsevier.com/locate/etfs

Upload: others

Post on 14-Feb-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

  • Experimental Thermal and Fluid Science 37 (2012) 72–83

    Contents lists available at SciVerse ScienceDirect

    Experimental Thermal and Fluid Science

    journal homepage: www.elsevier .com/locate /et fs

    Laminar convective heat transfer of alumina-polyalphaolefin nanofluidscontaining spherical and non-spherical nanoparticles

    Leyuan Yu a, Dong Liu a,⇑, Frank Botz ba Department of Mechanical Engineering, University of Houston, Houston, TX 77204, USAb METSS Corporation, Westerville, OH 43082, USA

    a r t i c l e i n f o

    Article history:Received 19 May 2011Received in revised form 27 September2011Accepted 12 October 2011Available online 19 October 2011

    Keywords:NanofluidsConvective heat transferPressure dropThermal conductivityViscosityMinichannel

    0894-1777/$ - see front matter � 2011 Elsevier Inc. Adoi:10.1016/j.expthermflusci.2011.10.005

    ⇑ Corresponding author.E-mail address: [email protected] (D. Liu).

    a b s t r a c t

    As a potential candidate for advanced heat transfer fluid, nanofluids have been studied extensively in theliterature. Past investigations were largely limited to thermophysical property measurements of aqueousnanofluids synthesized with spherical nanoparticles. No comprehensive information is yet available forconvective thermal transport of nanofluids containing non-spherical particles, especially of the non-aqueous nanofluids. In this work, an experimental study was conducted to investigate the thermophys-ical properties and convective heat transfer of Al2O3-polyalphaolefin (PAO) nanofluids containing bothspherical and rod-like nanoparticles. The effective viscosity and effective thermal conductivity of thenanofluids were measured and compared to predictions from several existing theories in the literature.It was found that, in addition to the particle volume fraction, other parameters, including the aspect ratio,the dispersion state and the aggregation of nanoparticles as well as the shear field, have significantimpact on the effective properties of the nanofluids, especially of those containing non-spherical parti-cles. The pressure drop and convection heat transfer coefficient were also measured for the nanofluidsin the laminar flow regime. Although established theoretical correlations provide satisfactory predictionof the friction factor and Nusselt number for nanofluids containing spherical nanoparticles, they fail fornanofluids containing rod-like nanoparticles. The results indicate that in a convective flow, the shear-induced alignment and orientational motion of the particles must be considered in order to correctlyinterpret the experimental data of the nanofluids containing non-spherical nanoparticles.

    � 2011 Elsevier Inc. All rights reserved.

    1. Introduction

    Nanofluids have been studied extensively as a promising candi-date of advanced heat transfer fluids for thermal managementapplications in microelectronics, transportation, power andmilitary systems [1]. While dramatic enhancement in thermal con-ductivity was reported in the early literature, most recent studiesfind the convective heat transfer of nanofluids is only mildly aug-mented and the measured thermal properties can be well corre-lated by the conventional effective medium theory (EMT) [2,3].Wide discrepancies exist among different studies regarding howthe dispersed nanoparticles alter the convective thermal transportof nanofluids. Moreover, the past investigations were conductedprimarily for water-based nanofluids synthesized with sphericalnanoparticles. Due to the simple geometry of these particles, it isreasonable to neglect the effects of the particle distribution andorientation on the effective rheological and thermal properties aswell as on the convective heat transfer of the nanofluids. However,

    ll rights reserved.

    when non-spherical nanoparticles, such as carbon nanotubes(CNTs) and titanate nanotubes (TNTs), are used in formulatingthe nanofluids, these effects may no longer be ignored as the resultof the hydrodynamic interactions between the particles and thesurrounding fluid medium. Also, the use of aqueous nanofluids be-comes infeasible in many practical applications due to their limitsin the dielectric property and the operating temperature range.Consequently, there is a need to systematically study the convec-tive thermal transport of non-aqueous nanofluids synthesized withnon-spherical nanoparticles.

    Several experimental studies were performed to study theeffective thermophysical properties and convective heat transferof nanofluids containing non-spherical particles with highlength-to-diameter ratio (aspect ratio). Cherkasova and Shan [4]studied the effects of particle aspect-ratio and dispersion state onthe effective thermal conductivity of aqueous nanofluids with mul-ti-walled nanotubes (MWNTs). They found the effective thermalconductivity decreases with reduced MWNT aspect ratio, and theconductivity enhancement was primarily due to the presence ofindividualized long nanotubes rather than the bundles of aggre-gated low aspect ratio nanotubes. The measured conductivity datashowed a good agreement with the EMT predictions, and no

    http://dx.doi.org/10.1016/j.expthermflusci.2011.10.005mailto:[email protected]://dx.doi.org/10.1016/j.expthermflusci.2011.10.005http://www.sciencedirect.com/science/journal/08941777http://www.elsevier.com/locate/etfs

  • Nomenclature

    A channel cross-sectional area, m2

    f friction factorh heat transfer coefficient, W/m2 Kk thermal conductivity, W/m Kl length of nanorod, nmL length of test tube, md diameter of nanorods, nmD channel inner diameter, mCp specific heat, kJ/kg KM fractal indexNu Nusselt numberP pressure, PaPe Peclet numberPr Prandtl numberq00 heat flux, W/m2

    Q volumetric flow rate, m3/sr aspect ratio of nanorodRb interfacial resistance, m2 K/WRe Reynolds number

    T temperature, �Cu velocity, m/sx axial position, mx� inverse of Graetz number

    Greek symbols/ volume concentrationl viscosity, N s/m2

    q density, kg/m3

    Subscriptsf base fluidin inleto outer wallout outletp particler relativex localw wall

    L. Yu et al. / Experimental Thermal and Fluid Science 37 (2012) 72–83 73

    anomalous increase was observed. Ding and his co-workers mea-sured the rheological properties of ethylene–glycol (EG)-basednanofluids containing TNTs [5,6]. Strong shear thinning behaviorwas found at high particle volume fractions in the EG-TNT nanofl-uids. The initial high viscosity was attributed to the resistance thatarises when the Brownian rotation of the rod-like TNTs must beovercome for the nanotubes to align with the shear field, as wellas the higher effective particle volume fraction caused by the par-ticle aggregation than in a well-dispersed solution. Ding et al. [7]conducted convective heat transfer experiments of aqueous CNTnanofluids flowing through a circular tube. The authors found thatthe heat transfer enhancement was caused by the dynamic thermalconductivity of the nanofluids under shear conditions, which wasmuch higher than the value measured at static conditions. Theyalso suggested that the development of thermal boundary layermay be affected by shear thinning and the migration/aggregationof the long aspect ratio CNTs. Similar observations were made byChen et al. [8] in their study of convective heat transfer of water-TNT nanofluids. Yang et al. [2] explored the effects of the Reynoldsnumber, temperature, particle concentration on convective heattransfer of nanofluids through a circular channel. Two kinds ofnanofluids were formulated by dispersing disk-like graphite nano-particles in a commercial automatic transmission fluid (ATF) and amixture of synthetic base oils, respectively. The enhancement ofheat transfer coefficient was much less than what was predictedby the Seider–Tate correlation with the effective thermal conduc-tivity measured under static conditions. It was hypothesized thatthe shear-induced nanoparticle alignment disrupts the particle–particle interaction, which was assumed to be the main energypathway in the nanofluids, and results in the deterioration of con-vective heat transfer.

    Polyalphaolefins (PAO) are a group of synthetic engine oils thathave been used extensively as lubricants and coolants in variousmilitary and aerospace applications [9]. As a heat transfer fluid,their thermal performance and energy efficiency are seriously lim-ited by the low intrinsic thermal conductivity of PAO (0.132 W/m Kvs. 0.6 W/m K for water). Therefore, it is of great practical interestto develop PAO-based nanofluids with enhanced thermophysicalproperties, which can also serve as a good sample for the funda-mental study of convective thermal transport in non-aqueousnanofluids. Yang et al. [10] measured the thermal and rheologicalproperties of PAO-CNT nanofluids at different dispersant

    concentrations, dispersing energy levels and nanoparticle concen-trations. They found nanoparticle aggregration was the mostimportant factor contributing to the effective properties: Largeragglomerates resulted in both higher thermal conductivity andhigher viscosity. The thermal conductivity of nanofluids increasedwith the CNT aspect ratio until the CNTs were long enough to forma percolated network structure. Zhou et al. [11] investigated thedependence of viscosity on the shear-rate and the temperature inPAO-Al2O3 nanofluids. At low particle volume fraction (1 vol%and 3 vol% for nanospheres and 1 vol% for nanorods), the viscosityshowed a weak dependence on shear rate and the nanofluids canbe approximated as a Newtonian fluid. The relative viscosity (de-fined as lr = l/lf) was found to be independent of temperature,indicating the rheological properties of nanofluids were primarilydominated by the base fluid. Shaikh et al. [12] measured the ther-mal conductivity of three types of PAO nanofluids containing CNTs,exfoliated graphite (EXG) and heat treated nanofibers (HTTs),respectively. They observed that the thermal conductivityenhancement was the most significant for PAO-CNT nanofluids,followed by EXG and HTT. Nelson et al. [9] conducted convectiveheat transfer experiments of PAO nanofluids in a plain offset finheat exchanger. The nanofluids were synthesized with exfoliatedgraphite fibers. They showed the augmented Nusselt number wasdue to the precipitation of nanoparticles on the wall of the heat ex-changer which acted as nanoscale fins to enhance the convectiveheat transfer.

    The literature survey reveals that thermal transport of non-aqueous nanofluids containing non-spherical particles has notbeen well understood. Particularly, there is a clear call for asystematic investigation on the effects of the particle–fluid andparticle–particle interactions on the effective thermophysicalproperty and convective heat transfer characteristics of the nanofl-uids. In this work, an experimental study was conducted to explorethe single-phase forced convection of PAO-Al2O3 nanofluids con-taining both spherical and rod-like nanoparticles (nanorods)through a circular minichannel. The effective viscosity and thermalconductivity were measured for the two types of nanofluids, andthe data were compared to predictions from the effective mediumtheory. The pressure drop and convective heat transfer of the nano-fluids were characterized over the Reynolds number range of110–630. It was found that the aspect ratio, dispersion state andaggregation of the nanoparticles contribute significantly to the

  • 74 L. Yu et al. / Experimental Thermal and Fluid Science 37 (2012) 72–83

    effective thermophysical properties of nanofluids containingnon-spherical particles, and the shear-induced alignment andorientational motion of the non-spherical nanoparticles play animportant role in affecting the pressure drop and heat transfercharacteristics of the nanofluids.

    Fig. 1. TEM image of the rod-like Al2O3 nanoparticles.

    2. Experiments and methods

    2.1. Preparation of nanofluids

    The nanofluids were formulated by dispersing boehmitealumina nanoparticles in 2 centiStokes (cSt) PAO under ultrasoni-cation. Two types of nanofluids were prepared. The first type (re-ferred to as NF1) contains spherical nanoparticles, and thesecond type (referred to as NF2) contains nanorods. Due to thehydrophobic nature of PAO, special dispersants of minusculeamount were added to alleviate the aggregation of nanoparticlesto stabilize the nanofluids. All the samples were found stable with-out precipitation for at least 7 days during the experiment period.

    The diameter of the spherical nanoparticles contained in NF1was measured using a dynamic light scattering (DLS) instrument(Brookhaven Instrument BI-200SM), and was found to be about60 nm. The diameter and length of the nanorods contained inNF2 were obtained using transmission electron microscopy(TEM). Before the TEM measurement, the sample was preparedby diluting the nanofluid containing 0.65 vol% nanorods with pure2 cSt PAO at a ratio of 1:10 and sonicating the mixture for 4 h. Fig. 1shows one representative TEM image. The measured diameters ofthe individual nanorods (d) range from 5 nm to 11 nm, and thelengths (l) range from 80 nm to 106 nm. Taking the average mea-surements from 10 TEM images, the mean diameter and lengthof the nanorods were d = 7.0 nm and l = 85 nm, respectively, whichcorrespond to an average aspect ratio of r = l/d � 12.

    2.2. Viscosity and thermal conductivity measurements

    Both the effective viscosity and thermal conductivity of thenanofluids were measured under static conditions. The viscositywas measured using a capillary viscometer (Cannon-Ubbelohde9721-R53). The thermal conductivity was characterized by a ther-mal property analyzer (KD2 Pro, Decagon Devices). The manufac-turer specified measurement uncertainty was 5% for the thermalproperty analyzer, which was confirmed by reproducing the liter-ature values for thermal conductivity of deionized (DI) water.

    2.3. Convective heat transfer experiments

    The apparatus for the convective heat transfer experiments(shown in Fig. 2) has been reported in details in Ref. [13]. In brief,a gear pump (IDEX Micropump 67-GA-V21) was used to circulatethe nanofluid through the test loop. The flow rate was measuredby a turbine flowmeter (McMillan G111). A liquid–liquid heat ex-changer (Lytron LL520G14) was used together with an air-cooledchiller (Neslab MERLIN 25) to reduce the temperature of theheated nanofluid to room temperature before it flows back to thereservoir. Readings of the flow rate, temperature and pressuremeasurements were collected by a data requisition system (Agilent34970A) and processed in a computer.

    The test tube is a circular minichannel made of stainless steel,and it measures 1.09 mm in inner diameter (D) and 0.25 mm inwall thickness. The total tube length (L) is 306 mm. The test tubeis resistively heated by passing a DC current through it. The voltagedrop across the channel was measured by the data acquisition sys-tem, and the current was obtained by using an accurate shuntresistor. Six copper–constantan (T-type) thermocouples (Omega

    5TC-TT-T40-36) were attached to the outer wall of the channel at44 mm axial intervals (TC1 through TC6). The temperature read-ings from these thermocouples were extrapolated to yield the localtemperatures at the inner wall. To minimize the heat loss to theambient, the test tube was wrapped heavily with thermal insulat-ing materials. Two thermocouple probes (Omega TMT IN-020G-6)were employed to measure the fluid temperatures at the inlet andoutlet of the channel. Two absolute pressure transducers (OmegaPX319-050A5V and PX319-030A5V) were installed to measurethe pressure drop across the channel.

    In the pressure drop experiments, the flow rate was adjusted bya control valve. In each experiment run, the data were read into thedata acquisition system after the flow rate and the pressure signalsstabilized. The flow rate was then increased in small increment andthe procedure repeated. In the heat transfer experiments, thepower input to the test tube was maintained at a constant level.The flow rate was first set to the maximum value and graduallydecreased in subsequent experiments. In all experiments, eachsteady-state value was calculated as an average of 100 readingsfor all flow rate, pressure, temperature, and power measurements.

    3. Data reduction

    The pressure drop and the flow rate were measured to obtainthe Reynolds number, Re, and the Darcy friction factor, f, whichare defined as

    Re ¼ quD=l ð1Þ

    f ¼ ðDP=LÞDqu2=2

    ð2Þ

    In Eq. (2), DP is the pressure drop across the channel length, and iscalculated by subtracting the inlet and outlet pressure losses fromthe measured overall pressure drop [14,15].

    In the heat transfer experiments, a uniform heat flux boundarycondition is assumed on the channel wall. The wall heat flux is cal-culated from the sensible heat gain by the fluid as

    Q 00 ¼ qQCpðTout � TinÞ=A ð3Þwhere the fluid properties are evaluated at the mean temperature

    T ¼ ðTin þ ToutÞ=2 ð4Þ

  • Fig. 2. Schematic of the experimental apparatus.

    L. Yu et al. / Experimental Thermal and Fluid Science 37 (2012) 72–83 75

    The local convective heat transfer coefficient is defined as

    hxðxÞ ¼ q00=½TwðxÞ � TðxÞ� ð5Þ

    where the local wall temperature Tw(x) is extrapolated from thetemperature, Tw,o(x), measured at the outer wall of the channel, fol-lowing [16]

    TwðxÞ ¼ Tw;oðxÞ þqQCpðTout � TinÞ

    4pkwL�

    qQCpðTout � TinÞD2o2pkwðD2o � D

    2ÞLln

    DoD

    ð6Þ

    The local fluid temperature T(x) can be calculated from the en-ergy conservation

    TðxÞ ¼ Tin þ q00pDx=ðqQCpÞ ð7Þ

    Accordingly, the local Nusselt number is

    Nux ¼hx � D

    kð8Þ

    where the thermal conductivity is evaluated at the correspondingfluid temperatures.

    In the foregoing analysis (Eqs. (1)–(8)), the density and specificheat of nanofluids were estimated as follows

    qðTÞ ¼ / � qpðTÞ þ ð1� /Þ � qf ðTÞ ð9Þ

    CpðTÞ ¼/ � ðqpCp;pÞðTÞ þ ð1� /Þ � ðqf Cp;f ÞðTÞ

    qðTÞ ð10Þ

    and the thermal conductivity and viscosity were from the experi-mental measurements.

    4. Measurement uncertainties

    The measurement uncertainties for the temperature, flow rateand pressure drop were ±0.3 �C, 1% of full scale and 2% of full scale,respectively. A standard error analysis [17] revealed that theuncertainties in the reported friction factor and heat transfer coef-ficient were in the ranges of 5.2–13.8% and 1.6–3.5%, respectively.

    5. Results and discussion

    5.1. Viscosity

    Addition of particles, particularly anisotropic particles such asspheroids, rods or disks, into a base fluid results in an effectiveviscosity that is higher than the inherent viscosity of the base fluid.This is due to the Brownian and hydrodynamic motions of the dis-persing particles.

    For suspensions of spherical particles, the relative viscosity canbe expressed by the Batchelor equation as a function of nanoparti-cle volume fraction

    lr ¼llf¼ 1þ ½l�/þ kH/2 þ Oð/3Þ ð11Þ

    where the intrinsic viscosity is [l] = 2.5 and the Huggins coefficientis kH = 6.2 [18,19]. The first two terms at the RHS of Eq. (11) are re-lated to the particle diffusion; the third term arises in concentratedsuspensions, and the coefficient kH is very sensitive to the rheolog-ical structure of the suspension [20]. At infinite dilution, Eq. (11) re-duces to the well-known Einstein equation, lr = 1 + 2.5/.

    In static suspensions of rod-like particles, the impact of transla-tional Brownian motion of the particles is negligible on the viscos-ity, as compared to that due to the rotational Brownian motion. Assuch, the effective viscosity is mainly affected by the competitionbetween the shear force and the rotational Brownian motion, rep-resented by the rotational Peclet number

    Perot ¼_c

    Drð12Þ

    where _c is the shear rate, and Dr is the rotational diffusion constantof the particles. It is easy to see the low shear limit (as encounteredunder the static conditions) corresponds to Perot� 1. If the particleconcentration is sufficiently high, the particles will overlap andinteract hydrodynamically, and the rotational freedom is restricted.Recasting the particle volume concentration / in terms of the num-

    ber density v, it yields v ¼ /= p4 d2l

    � �for a dispersion of rod-like par-

    ticles with a diameter d and a length l. The minimum overlapconcentration is given by [21]

    v� ¼ 1l3

    ð13Þ

  • 76 L. Yu et al. / Experimental Thermal and Fluid Science 37 (2012) 72–83

    If the particle dispersion is infinitely dilute (v < 0.1v�), each particlecan rotate freely and the relative viscosity of the nanofluid is

    lr ¼ 1þ ½l�/ ð14Þ

    where [l] can be calculated from one of the following equations[22,23]

    ½l� ¼ 415

    r2

    ln rfor r � 1 ð15Þ

    or

    ½l� ¼ r2

    51

    3ðln 2r � 1:5Þ þ1

    ln 2r � 0:5

    � �þ 1:6 for r > 15 ð16Þ

    At higher shear rate, i.e., Perot P r3, the orientation distribution ofthe particles due to shear forces dominates over the Brownianmotion, and [l] can be estimated by the equation by Hinch and Leal[24]

    ½l� ¼ 0:315 rln r

    ð17Þ

    If v P 0.1v�, the rod–rod hydrodynamic interaction starts tocontribute to the effective viscosity of the suspension. For instance,0.1v� corresponds to a particle volume concentration of 0.055 vol%for nanorods with l = 85 nm and d = 7 nm in this work. When theparticle concentration falls in the range of 0.1v� < v < v�, the suspen-sion can still be considered dilute. The relative viscosity at lowshear limit (Perot� 1) is estimated from the Berry–Russel equation[25]

    lr ¼ 1þ ½l�/þ kH½l�2/2 ð18Þ

    where kH ¼ 25 1� 0:00142Pe2rot

    � �and [l] can be found from Eqs. (15)

    or (16). At higher particle concentrations, v� < v < (dl2)�1, or equiva-lently, 0.55 vol% < / < 6.5 vol% for the PAO-nanofluids in the presentstudy, the particle suspension is considered semi-dilute. In this con-centration range, the rotational volumes of adjacent rods will over-lap and the rods are entangled. The low-shear viscosity of suchparticle suspensions is given by [21]

    lr ¼ 1þr2

    15 ln r/þ 36r

    6

    5p2b ln r/3 ð19Þ

    where b is a numerical factor (b = 103–104). When the particle con-centration further increases to the maximum packing fraction, /m,the Krieger–Dougherty correlation [26] can be used to estimatethe relative viscosity

    lr ¼ ½1� ð/=/mÞ��½l�/m ð20Þ

    where /m = 5.4/r for r� 1 [27,28].In addition to the rotational interaction between particles, two

    more effects must be considered in the study of the effective vis-cosity of nanofluids: the particle aggregation and the shear flow.Most nanoparticles naturally adhere to each other, and the aggre-gation can only be suppressed to some degree with stabilizingagents. As a consequence, the nominal particle volume fraction, /, should be replaced by an effective volume fraction, /a, as sug-gested by Chen et al. [29,30]

    /a ¼ /aaa

    � �3�Mð21Þ

    where a is the primary nanoparticle size, aa is the effective size ofthe aggregates; and M is the fractal index. The value of M varies be-tween 1.6 and 2.3, and is approximately 1.8 for spherical nanopar-ticles. Since (aa/a) > 1, the particle aggregation will result in /a > /,which, in turn, leads to a higher viscosity of the nanofluids, as canbe manifested by Eqs. (11)–(20). When a shear flow is applied,the rod-like particles tend to align spontaneously with the flowfield. As the shear rate increases, the impact of the rotational

    motion of nanoparticles on viscosity will diminish. In fact, shearthinning behavior was observed in nanofluids of high particle vol-ume fraction under high shear conditions [29,30]. As a consequence,the effective viscosity of nanofluids in forced convective flow is ex-pected to be lower than the values measured at the staticconditions.

    In this work, the viscosity of nanofluids was measured at 25 �Cunder static conditions for particle volume fractions of 0.33, 0.49,0.65, and 1.3 vol%, respectively. A capillary viscometer was usedfor this purpose. The results are presented in Fig. 3 in terms ofthe relative viscosity. Fig. 3a shows the measured data of NF1and the comparison with theoretical predictions. The viscosity ofnanofluids clearly increases with the nanoparticle volume fraction.The original Batchelor equation (Eq. (11)) underpredicts the exper-imental data; but when used in combination with the effective vol-ume fraction calculated from Eq. (21), it provides a betterprediction of the viscosity data. In doing so, aa/a must be knownas a priori, however, it is employed as a fitting parameter due tothe difficulty in obtaining the actual values for each sample. It isfound that the selection of aa/a = 4.12 and M = 1.8 yields a goodagreement between the predictions and the measured data witha coefficient of determination of R2 = 0.9843. For simplicity, anempirical correlation is proposed for the nanofluids containingspherical nanoparticles

    lr ¼ 1þ 13:67/þ 185:42/2 for NF1 ð22Þ

    Fig. 3b illustrates the measured lr of NF2 as a variation of /. Ascompared to the results in Fig. 3a, the viscosity of NF2 is distinctlyhigher than that of NF1 at the same particle concentration and thedifference widens as the concentration increases. Similar observa-tions were reported by Chen et al. [8] and Zhou et al. [11]. The re-sults suggest the important role played by the particle geometryand aspect ratio in the rheological properties of nanofluids. Alsoshown in Fig. 3b are the predictions from the theoretical modelsdeveloped for suspensions of rod-like particles (Eqs. (14), (15),(18)–(20)). Unfortunately, they all underestimate the measure-ment data. Considering the aggregation effect, the effective volumefraction of the rod-like particles can be estimated from Eq. (21),where aa/a = 1.48 and M = 1.95 were chosen. With the effectivevolume fraction, the modified theoretical predictions are examinedagain in Fig. 3c, which shows the dilute suspension model yieldsthe best overall agreement (R2 = 0.9920) with the experimentaldata. An empirical correlation similar to Eq. (22) was proposedfor the nanofluids containing nanorods

    lr ¼ 1þ 27:29/þ 296:92/2 for NF2 ð23Þ

    It is noted that Eqs. (22) and (23) are valid for / 6 1.3%.The viscosities of nanofluids at elevated temperatures were not

    investigated in this work. They were estimated by

    lðTÞ ¼ lr � lf ðTÞ ð24Þ

    where the temperature-dependence of the viscosity of the basefluid (PAO) is obtained from [31]

    lf ðTÞ ¼ 10ð109:67=T3:923Þ � 0:7

    � � 10�6 � 1360� 4:56T þ 0:0157T2

    ��0:28 10�4T3 þ 0:174 10�7T4

    �ð25Þ

    5.2. Thermal conductivity

    For solid–liquid mixtures, the relative thermal conductivity canbe estimated by the Hamilton–Crosser model [32]

    kr ¼kkf¼ kp þ ðn� 1Þkf � ðn� 1Þðkf � kpÞ/

    kp þ ðn� 1Þkf þ ðkf � kpÞ/ð26Þ

  • 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.41.00

    1.05

    1.10

    1.15

    1.20

    Rb = 6.5 x 10-8 m2K/W

    k/k f

    Volume fraction φ (v%)

    Experiment (NF1)

    Hamilton-Crosser model (Eq. (26)) Maxwell-Garnett model (Eq. (28))

    (a)

    (b)

    0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.41.00

    1.05

    1.10

    1.15

    1.20

    Rb = 6.5 x 10-8 m2K/W

    R2 = 0.9962

    k/k f

    Volume fraction φ (v%)

    Experiment (NF1)

    Hamilton-Crosser model (Eqs. (21) & (26))(aa/a = 4.12, M = 1.8) Maxwell-Garnett model (Eqs. (21) & (28))(aa/a = 4.12, M = 1.8))

    Fig. 4. The relative thermal conductivity of NF1 as a function of nanoparticlevolume fraction.

    0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.41.0

    1.1

    1.2

    1.3

    1.4

    1.5

    1.6

    1.7

    R2 = 0.9843Rela

    tive

    visc

    osity

    μr

    Volume fraction φ (v%)

    Experiment (NF1) Batchelor model (Eq. (11)) Batchelor model (Eqs. (11) & (21)) (aa/a = 4.12, M = 1.8)

    (a)

    0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.41.0

    1.1

    1.2

    1.3

    1.4

    1.5

    1.6

    1.7

    Rela

    tive

    visc

    osity

    μr

    Volume fraction φ (v%)

    Experiment (NF2) Infinitely dilute (Eqs. (14) & (15)) Dilute (Eqs.(15) & (18)) Semi-dilute (Eq. (19)) Concentrated (Eq. (20))

    (b)

    (c)

    0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.41.0

    1.1

    1.2

    1.3

    1.4

    1.5

    1.6

    1.7

    R2 = 0.9920

    aa/a = 1.48, M = 1.95

    Rela

    tive

    visc

    osity

    μr

    Volume fraction φ (v%)

    Experiment (NF2)

    Infinitely dilute (Eqs. (14) & (15)) Dilute (Eqs. (15) & (18)) Semi-dilute (Eq. (19)) Concentrated (Eq. (20))

    Fig. 3. The relative viscosity of nanofluids at various nanoparticle volume fractions.

    L. Yu et al. / Experimental Thermal and Fluid Science 37 (2012) 72–83 77

    where the shape factor is n = 3/w, and w is the sphericity defined asthe ratio of the surface area of a sphere (with the same volume asthe given particle) to the surface area of the particle. For sphericalparticles, w = 1. Eq. (26) reduces to the classical Maxwell model

    kr ¼1þ 2b/1� b/ ð27Þ

    where b = (kp � kf)/(kp + 2kf). To account for the interfacial thermalresistance, the Maxwell–Garnett model [33] predicts

    kr ¼ð1þ 2aÞ þ 2/ð1� aÞð1þ 2aÞ � /ð1� aÞ ð28Þ

    where a = 2Rb kf/D and Rb is the interfacial resistance. The actual va-lue of Rb is difficult to measure directly, but molecular dynamicssimulations suggest it is typically on the order of 10�8–10�9 m2 K/W [15].

    For dilute suspensions containing randomly dispersed spheroi-dal particles, Nan et al. [33] developed a model for the relativethermal conductivity

    kr ¼ 1þ/ð2b11 þ b33Þ

    3� /ð2b11L11 þ b33L33Þð29Þ

    where

    bii ¼ ðkii � kf Þ=½kf þ Liiðkii � kf Þ� ð30Þ

    The depolarization factors for prolate spheroids, Lii (i = 1, 2 and 3),are

  • 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.41.00

    1.05

    1.10

    1.15

    1.20

    R2 = 0.9582

    Rb = 1.24 x 10-8 m2K/W

    k/k f

    Volume fraction φ (v%)

    Experiment (NF2)

    Nan et al. model (Eq. (29)) Nan et al. model (Eqs. (21) & (29)) (aa = 1.48 and M = 1.95)

    Fig. 5. The relative thermal conductivity of NF2 as a function of nanoparticlevolume fraction.

    78 L. Yu et al. / Experimental Thermal and Fluid Science 37 (2012) 72–83

    L11 ¼r2

    2ðr2 � 1Þ �r

    2ðr2 � 1Þ3=2cosh�1r and L33 ¼ 1� 2L11 ð31Þ

    and the equivalent thermal conductivities along the two spheroidalaxes are

    k11 ¼kp

    1þ 2kpRb=dand k33 ¼

    kp1þ 2kpRb=l

    ð32Þ

    Fig. 4 illustrates the comparison of the measured relative ther-mal conductivity of NF1 with the predictions from the Hamilton–Crosser model and the Maxwell–Garnett model. In their originalforms, both models significantly underpredict the experimentaldata, as shown in Fig. 4a. If the particle aggregation effect is consid-ered, the effective volume fraction used in the viscosity calculationof NF1 (Eq. (21) with aa/a = 4.12 and M = 1.8) can be applied in theestimate of the effective thermal conductivity. The new results areplotted in Fig. 4b. The Hamilton–Crosser model is seen to overesti-mate the thermal conductivity. The accuracy of the Maxwell–Gar-nett model is affected by the choice of Rb, which is used here as afitting parameter. By using regression analysis, it is found that forRb = 6.5 10�8 m2 K/W, the Maxwell–Garnett model offers thebest prediction of the experimental data with R2 = 0.9962. For sim-plicity, the experimental data are correlated by a linear function ofthe nanoparticle volume fraction

    k=kf ¼ 1þ 7:6661/ for NF1 ð33Þ

    Fig. 5 shows the comparison of the measured relative thermalconductivity of NF2 with predictions from the Nan et al. model[34]. Clearly, the original model fails to predict the experimentaldata satisfactorily. However, when the effective volume fraction(Eq. (21) with aa/a = 1.48 and M = 1.95) and Rb = 1.24 10�8 m2 K/W are used, the modified model offers a much betterprediction of the experimental data (R2 = 0.9582). In light of thecomplexity of the Nan et al. model, a simple correlation was pro-posed to predict the effective thermal conductivity of nanofluidscontaining nanorods

    k=kf ¼ 1þ 9:4539/ for NF2 ð34Þ

    In the above discussions, the nanofluids are regarded as a ma-trix-based composite material whose effective thermal conductiv-ity depends on the constituent materials, the volume fraction andthe size/shape of the filler nanoparticles. On the other hand, thethermal conductivity of nanofluids is also critically affected by

    the distribution and the relative orientation of nanoparticles withrespect to the temperature field [40]. As depicted in Fig. 6, thereare four possible particle configurations in the particle–fluid sus-pension, namely, the parallel, series, Hashin–Shtrikman (or H–S)and EMT configurations. In each representative cell, the heat fluxis applied from the bottom boundary to the top. When the particlesare continuously configured in parallel with or perpendicular tothe direction of the temperature gradient, the effective thermalconductivity can be described by the parallel or the series model[35,36]

    Parallel model k ¼ ð1� /Þkf þ /kp ð35Þ

    Series model k ¼ 1ð1� /Þ=kf þ /=kpð36Þ

    The parallel and series models represent the upper and lowerbounds over all possible structures of a heterogeneous material.In the H–S configuration, the discrete particles are uniformly dis-tributed without direct contact with each other. The correspondingeffective thermal conductivity is given by [37]

    ke ¼ kp þ1� /1

    km�kp þ/

    3kp

    ð37Þ

    In real particle suspensions, the particle distribution is neither con-tinuous nor uniform, but rather in the form of the EMT structure,i.e., a random distribution. For such kind of particle dispersion,the effective thermal conductivity is predicted by the EMT model[38,39]

    k ¼ ð3/� 1Þkp þ ½3ð1� /Þ � 1�kf�þ

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi½ð3/� 1Þkp þ ½3ð1� /Þ � 1�kf �2 þ 8kf kp

    q �=4 ð38Þ

    For these four particle configurations of the Al2O3-PAO nanofluids(kp/kf = 348), the theoretical predictions of the relative thermalconductivity are plotted in Fig. 7 over the full range of volume frac-tion (0 6 / 6 1). At any given particle concentration, the highestthermal conductivity is obtained in the parallel structure whereheat is conducted through the parallel pathways formed by thealigned particles. In contrast, the lowest possible thermal conduc-tivity occurs in the series structure.

    In almost all heat transfer experiments of nanofluids (includingthe present one), the effective thermal conductivity was measuredunder static conditions, where the nanoparticles were assumed todisperse either randomly or uniformly in the base fluid. However,the thermal conductivity data of NF2 (shown as the inset in Fig. 7)lie between the predictions of the parallel model and the EMT/H–Smodel, suggesting that the actual distribution of the nanorods isneither completely orderly nor completely random. Furthermore,when the nanofluid flows through a circular channel, a velocitygradient exists along the radial direction which gives rise to a shearfield. The shear stress will align the non-spherical particles withthe flow direction, as schematically depicted in Fig. 8 [20]. Afterthe flow becomes hydrodynamically fully developed, a particle–fluid structure similar to the series configuration in Fig. 7 will beformed. Consequently, it is reasonable to expect that the actualthermal conductivity of nanofluids containing rod-like particlesin the convective flow would be less than that is measured understatic conditions. The reduced effective thermal conductivity willadversely affect the convective heat transfer performance of thenanofluids.

    5.3. Pressure drop

    Control experiments were first performed with the base fluid(PAO) to obtain baseline information for single-phase thermal

  • series TMElellarap H-S

    q’’

    particle

    series TMElellarap H-S

    q’’q’’

    particle

    Fig. 6. Four possible particle configurations in particle suspensions.

    φ(v%)

    k/k f

    0 20 40 60 80 1000

    50

    100

    150

    200

    250

    300

    350kparallelkEMTkHashin-Shtrikmankseries

    series

    parallel

    EMT

    H-S

    q’’q’’

    particlemediumparticlemedium

    0 0.5 1 1.51

    1.05

    1.1

    1.15

    1.2

    1.25

    1.3

    1.35

    1.4Experiment (NF2)kparallelkEMTkHashin-Shtrikmankseries

    Volume fraction φ (v%)

    Fig. 7. Effective thermal conductivity for various particle configurations.

    L. Yu et al. / Experimental Thermal and Fluid Science 37 (2012) 72–83 79

    transport of nanofluids. For laminar flow, the Hagen–Poiseuilleequation [15] can be used to predict the friction factor for hydrody-namically fully developed condition

    f � Re ¼ 64 ð39Þ

    and the Shah equation [40] is used to account for the developinglength effect

    fappRe 43:44ffiffiffi

    1p þ 16þ 0:31251� 3:44=

    ffiffiffi1p

    1þ 2:12 10�4=12

    !ð40Þ

    where f = (x/D)/Re. Fig. 9 depicts the measured friction factor of thePAO fluid as a variation of the Reynolds number. The experiments inthis work were conducted over the same flow rate range(50�200 mLPM) as for water in the previous work [13]; however,

    u δ

    δ

    Hydrodynamic entrance reg

    u δ

    δ

    Hydrodynamic entrance reg

    Fig. 8. Shear-induced alignment of rod-like nanopar

    the maximum Re achieved was much lower (Re 6 460) due to thehigh viscosity of PAO. Consequently, the hydrodynamic entrancelength (L+ 0.056 Re�D) is less than 9.2% of the total channel lengtheven at the highest Reynolds number. Therefore, the entranceregion effect is insignificant in all the experiments conducted.Fig. 9 shows the experimental data agree quite well with the predic-tions from Eqs. (39) and (40), except for the low Re region where themeasurement uncertainty is relatively high (the maximum uncer-tainty is 13.8%.).

    The pressure drop experiments were conducted for NF1 andNF2 with volume concentrations of 0.65 vol% and 1.3 vol%, respec-tively. Pressure drop across the length of the minichannel is firstpresented in Fig. 10 as a function of the flow rate. It shows nanofl-uids (both NF1 and NF2) incur higher pressure drop than the basefluid at the same flow rate, and the difference enlarges withincreasing particle volume fraction. Further, the pressure drop ofNF2 is always greater than that of NF1 at the same volume fraction,an observation consistent with the viscosity measurements inFig. 3. Fig. 11 shows the measured friction factor of the nanofluidsas a variation of the Reynolds number. At medium to high Re, the fmeasurements of NF1 match the Shah prediction fairly well. Incontrast, the data for NF2 show appreciable deviation in the samerange of Re, i.e., dropping below the theoretical prediction. Thismay be attributed to the strong alignment of the nanorods underthe shear stress causing the effective viscosity of nanofluids to de-crease in a manner similar to shear thinning.

    5.4. Convection heat transfer

    The local Nusselt number results of PAO are shown in Fig. 12 asa function of x� (=x/(D Re Pr)). Due to the large Prandtl number ofPAO (61.6 < Pr < 87.1), the thermally developing flow behaviorcan be clearly identified: Nux approaches infinity at x� = 0 and rap-idly decays as x� decreases, and Nux does not reach the asymptot-ical (fully developed) value, i.e., Nu = 4.36, over the range of x�

    studied in this work. The experimental data can be reasonablycorrelated by the Shah–London equation [41].

    ion

    r

    Fully developed region

    q’’

    ion

    r

    Fully developed region

    q’’

    ticles in the convective flow through a channel.

  • Flow rate Q (mLPM)

    Pres

    sure

    drop

    ΔP(k

    Pa)

    50 100 150 20060

    80

    100

    120

    140

    160

    180

    200

    220

    240

    260

    PAO0.65 v% NF11.3 v% NF10.65 v% NF21.3 v% NF2

    Fig. 10. Pressure drop across the minichannel versus flow rate.

    Reynolds number

    Fric

    tion

    fact

    or f

    100 200 300 400 500

    0.2

    0.4

    0.6

    0.8

    10.65 v% NF11.3 v% NF10.65 v% NF21.3 v% NF2Hagen-Poiseulle equationShah equation

    Fig. 11. Friction factor of nanofluids as a function of Reynolds number.

    x*

    Loca

    lNu

    ssel

    tnu

    mbe

    rN

    u x

    0 0.01 0.02 0.030

    2

    4

    6

    8

    10

    12

    14

    16

    18

    TC1TC2TC3TC4TC5Shah-London correlation

    Nu = 4.36

    Fig. 12. Local Nusselt number versus x� (heat flux q00 = 6.5 kW/m2).

    Reynolds number

    Fric

    tion

    fact

    or f

    100 200 300 400 500

    0.2

    0.4

    0.6

    0.8

    1

    PAOShah equationHagen-Poiseuille equation

    Fig. 9. Friction factor of PAO versus Reynolds number.

    80 L. Yu et al. / Experimental Thermal and Fluid Science 37 (2012) 72–83

    Nux¼1:302=ðx�Þ1=3 � 1 for x� 6 0:000051:302=ðx�Þ1=3 � 0:5 for 0:0005 6 x� 6 0:00154:364þ8:68ð1000x�Þ�0:506e�41x� for x� P 0:0015

    8><>:

    ð41Þ

    Convective heat transfer experiments were conducted for NF1and NF2 with the volume concentrations of 0.65 and 1.3 vol%,respectively. The local heat transfer coefficients measured at fiveaxial locations (TC1 through TC5) are presented in Fig. 13 forRe = 350 and 490. It is found that the convective heat transfer ofnanofluids is enhanced as compared to the base fluid, and theincrement increases proportionally to Re and /. Further, the per-centile enhancement of local heat transfer coefficient is higher nearthe channel entrance than at the downstream locations. For in-stance, in Fig. 13a, the enhancement in hx for 1.3% NF2 is 28.7%near the entrance, but it decreases to less than 21% near the chan-nel exit. This trend strengthens with increasing Re and /.

    More interestingly, it can be observed that hx for 0.65 vol% NF2always exceeds hx for NF1 at both concentrations (0.65 and1.3 vol%). This finding is unexpected because the measurementdata in Figs. 5 and 6 suggest that the thermal conductivity for0.65 vol% NF2 (k/kf = 1.076) is only slightly higher than that for0.65 vol% NF1 (k/kf = 1.053), but is lower than that for 1.3 vol%NF1 (k/kf = 1.098). When the shear-induced particle alignment isconsidered, an even larger discrepancy emerges. Since the seriesstructure can form more easily in nanofluids containing rod-likeparticles (as shown in Fig. 8), which represents the least effectivepathway for thermal energy transport, the convective heat transferof NF2 would be lower than that of NF1 at the same particlevolume concentration. However, this conjecture is in direct contra-diction to the results shown in Fig. 13.

    The paradox may be explained qualitatively by examining theperiodic orientational motions of rod-like particles in a shear flow[42–45]. Single rod-like particle tends to align its long axis with theflow direction. Unless perfectly aligned, the shear velocities will be

  • (a)

    (b)

    x/D

    h x(W

    /m2 .K

    )

    0 50 100 150 2000

    500

    1000

    1500

    2000

    PAO0.65 v% NF11.3 v% NF10.65 v% NF21.3 v% NF2

    Re = 350

    x/D

    h x(W

    /m2 .K

    )

    0 50 100 150 2000

    500

    1000

    1500

    2000

    PAO0.65 v% NF11.3 v% NF10.65 v% NF21.3 v% NF2

    Re = 490

    Fig. 13. Local convective heat transfer of nanofluids (heat flux q00 = 6.5 kW/m2).

    (a)

    (b)

    x*

    Loca

    lNus

    selt

    num

    berN

    u x

    0 0.01 0.02 0.030

    2

    4

    6

    8

    10

    12

    14

    16

    18

    TC1TC2TC3TC4TC5Shah-London correlationNu = 4.36

    x*

    Loca

    lNus

    selt

    num

    berN

    u x

    0 0.01 0.02 0.030

    2

    4

    6

    8

    10

    12

    14

    16

    18

    TC1TC2TC3TC4TC5Shah-London correlationNu = 4.36

    Fig. 14. The local Nusselt number for NF1 as a function of x� (heat flux q00 =6.5 kW/m2) for: (a) 0.65 vol%, and (b) 1.3 vol%.

    L. Yu et al. / Experimental Thermal and Fluid Science 37 (2012) 72–83 81

    slightly different at the two ends of the rod, causing the rod torotate periodically about its center to trace out the so-called Jefferyorbits [46]. In colloidal suspensions of sufficiently high concentra-tion, the particles interact with their neighbors and are forced toperform the orientational motion collectively, which can be de-scribed by the ‘‘director’’, i.e., the average orientation of the rods.At low shear rates, the director performs a ‘‘tumbling’’ or ‘‘kaya-king’’ motion in which the director rotates slowly when it is nearlyaligned with the flow direction, and rotates rapidly when its longaxis makes a non-zero angle with the shear plane. At intermediateshear rates, the director will oscillate up and down in the shearplane in a symmetrical way about the flow direction, transitioninginto the ‘‘wagging’’ motion. At further increased shear rate, thewagging motion is suppressed until the director is arrested makinga stationary orientation at a small angle with respect to the flowdirection. Clearly, the periodic orientational motions of the rod-likeparticles, either tumbling, kayaking or wagging, will create distur-bances to the local flow field, acting in the role of turbulent eddiesto promote the convective heat transfer. Furthermore, when thenanorods rotate in the thermal boundary layer, their two endsexperience periodically higher temperature in the near-wall regionand lower temperature in the near-bulk region. Since heat can be

    conducted effectively from the hot end to the cold end of the highlyconductive nanorods, they may act as nanoscale heat pumps totransfer heat into the bulk of the fluid. Indeed, the results inFig. 13 suggest that the heat transfer enhancement due to the peri-odic orientational motion of the nanorods in NF2 not only offsetsthe adverse effect of reduced effective thermal conductivity dueto the shear-induced particle alignment, but it also surpasses theheat transfer augmentation in NF1 due to higher particle volumefraction.

    Finally, the local Nusselt number measurements are plotted inFigs. 14 and 15 as a function of x⁄ for NF1 and NF2, respectively.Fig. 14 shows that the experimental data of NF1 closely matchthe predictions from Shah–London’s correlation at both volumefractions of 0.65 and 1.3 vol%. In Fig. 15, the measured Nux initiallyfollows the Shah–London prediction, but drops rapidly with a slopemuch steeper than the theoretical curve as x� increases. Unlike PAOand NF1, the Nux results for NF2 at different axial locations are dis-persive and no common trend line can be found. Rather than a newphysical phenomenon, this should be interpreted as an artifactwhich may be attributed to the effects of shear-induced motionof rod-like particles on the effective thermal conductivity. In

  • (a)

    (b)

    x*

    Loca

    lNus

    selt

    num

    berN

    u x

    0 0.01 0.02 0.030

    2

    4

    6

    8

    10

    12

    14

    16

    18

    TC1TC2TC3TC4TC5Shah-London correlationNu = 4.36

    x*

    Loca

    lNus

    selt

    num

    berN

    u x

    0 0.01 0.02 0.030

    2

    4

    6

    8

    10

    12

    14

    16

    18

    TC1TC2TC3TC4TC5Shah-London correlationNu = 4.36

    Fig. 15. The local Nusselt number for NF2 as a function of x� (heat flux q00 =6.5 kW/m2) for: (a) 0.65 vol%, and (b) 1.3 vol%.

    82 L. Yu et al. / Experimental Thermal and Fluid Science 37 (2012) 72–83

    calculating Nux (=hxD/k), hx is taken from the data of convectiveheat transfer experiments using Eq. (5), but k is from conductivitymeasurements obtained under static conditions. As already dis-cussed, the actual thermal conductivity of NF2 depends stronglyon the shear field as the result of the shear-induced alignmentand orientational motion of the nanorods. Hence it is the use ofstatic measurement of k that leads to the artifact shown inFig. 15. Consequently, certain caution must be exercised whenthe convective heat transfer performance is evaluated on the basisof conventional dimensionless numbers, such as Nu, for nanofluidscontaining non-spherical particles.

    6. Conclusions

    An experimental study was conducted to investigate the ther-mophysical properties and convective heat transfer characteristicsof Al2O3-PAO nanofluids containing both spherical and rod-likenanoparticles. The effective viscosity and thermal conductivity ofthe nanofluids were measured, and the results were comparedwith predictions from several existing theories in the literature.It was found that, in addition to the particle volume fraction, other

    parameters including the aspect ratio, dispersion state and aggre-gation of nanoparticles as well as the shear field have a significantimpact on the effective properties, especially for nanofluids con-taining non-spherical particles. The pressure drop and convectionheat transfer coefficient were also measured for the nanofluids inthe laminar flow regime. Although established theoretical correla-tions provide satisfactory prediction of the friction factor and Nus-selt number for nanofluids containing spherical nanoparticles, theyfail for nanofluids containing rod-like nanoparticles. The resultsindicate that in a convective flow, the shear-induced alignmentand orientational motion of the particles must be considered inorder to correctly interpret the experimental data of nanofluidscontaining non-spherical nanoparticles. Findings from this studyimply that, if external means (such as electromagnetic force orshear field) are applied to manipulate the states of dispersionand orientation of the nanoparticles in the base medium, thermaltransport of the nanofluids can be strategically controlled to yieldoptimal performance.

    Acknowledgments

    The authors are grateful to Prof. Suresh Garimella at PurdueUniversity and Dr. Lois Gschwender at Air Force Research Labora-tory for their assistance with initiating this work. They acknowl-edge the financial supports from the University of Houston, theCooling Technologies Research Center (a National Science Founda-tion Industry/University Cooperative Research Center) at PurdueUniversity and the National Science Foundation (Grant No.0927340).

    References

    [1] S.V. Garimella, D. Liu, Electromechanical actuation of nanofluids, in: S.C.Ramesh, R.V. Ramanujan (Eds.), Nanoparticles: Synthesis, Characterization andApplications, American Scientific Publishers, Stevenson Ranch, CA, 2010, p. 30.

    [2] Y. Yang, Z.G. Zhang, E.A. Grulke, W.B. Anderson, G.F. Wu, Heat transferproperties of nanoparticle-in-fluid dispersions (nanofluids) in larninar flow,Int. J. Heat Mass Transfer 48 (2005) 1107–1116.

    [3] C. Nan, Interface effect on thermal conductivity of carbon nanotubecomposites, Appl. Phys. Lett. 85 (2004) 3549.

    [4] A.S. Cherkasova, J.W. Shan, Particle aspect-ratio and agglomeration-stateeffects on the effective thermal conductivity of aqueous suspensions ofmultiwalled carbon nanotubes, J. Heat Transfer 132 (2010) 082402–082411.

    [5] H. Chen, Y. Ding, A. Lapkin, Rheological behaviour of nanofluids containingtube/rod-like nanoparticles, Powder Technol. 194 (2009) 132–141.

    [6] H. Chen, Y. Ding, A. Lapkin, X. Fan, Rheological behaviour of ethylene glycol-titanate nanotube nanofluids, J. Nanopart. Res. 11 (2009) 1513–1520.

    [7] Y. Ding, H. Alias, D. Wen, R.A. Williams, Heat transfer of aqueous suspensionsof carbon nanotubes (CNT nanofluids), Int. J. Heat Mass Transfer 49 (2006)240–250.

    [8] H. Chen, W. Yang, Y. He, Y. Ding, L. Zhang, C. Tan, A.A. Lapkin, D.V. Bavykin,Heat transfer and flow behaviour of aqueous suspensions of titanatenanotubes (nanofluids), Powder Technol. 183 (2008) 63–72.

    [9] I.C. Nelson, D. Banerjee, R. Ponnappan, Flow loop experiments usingpolyalphaolefin nanofluids, J. Thermophys Heat Transfer 23 (2009) 752–761.

    [10] Y. Yang, E.A. Grulke, Z.G. Zhang, G.F. Wu, Thermal and rheological properties ofcarbon nanotube-in-oil dispersions, J. Appl. Phys. 99 (2006).

    [11] S.-Q. Zhou, R. Ni, D. Funfschilling, Effects of shear rate and temperature onviscosity of alumina polyalphaolefins nanofluids, J. Appl. Phys. 107 (2010)054316–054317.

    [12] S. Shaikh, K. Lafdi, R. Ponnappan, Thermal conductivity improvement in carbonnanoparticle doped PAO oil: an experimental study, J. Appl. Phys. 101 (2007).

    [13] D. Liu, L.Y. Yu, Single-phase thermal transport of nanofluids in a minichannel, J.Heat Transfer-Trans ASME 133 (2011).

    [14] D. Liu, S.V. Garimella, Investigation of liquid flow in microchannels, J.Thermophys. Heat Transfer 18 (2004) 65–72.

    [15] R.D. Blevins, Applied Fluid Dynamics Handbook, Krieger, Malabar, FL, 1992.[16] F.P. Incropera, D.P. DeWitt, Fundamentals of Heat and Mass Transfer, John

    Wiley & Sons, NY, 1996.[17] J.R. Taylor, An Introduction to Error Analysis, University Science Books,

    Sausalito, CA, 1997.[18] A. Einstein, A new determination of molecular dimensions, Ann. Phys. 4 (1906)

    17.[19] G.K. Batchelor, The effect of Brownian motion on the bulk stress in a

    suspension of spherical particles, J. Fluid Mech. 83 (1977) 97–117.

  • L. Yu et al. / Experimental Thermal and Fluid Science 37 (2012) 72–83 83

    [20] S.M.P.a.S.S. Helvaci, Solid and Liquid Two Phase Flow, Elsevier, Amsterdam,Netherlands, 2008.

    [21] A.M. Wierenga, A.P. Philipse, Low-shear viscosity of isotropic dispersions of(Brownian) rods and fibres; a review of theory and experiments, ColloidsSurfaces A – Physicochem. Eng. Aspects 137 (1998) 355–372.

    [22] L. Onsager, Viscosity and particle shape in colloid solutions, Phys. Rev. 40(1932).

    [23] w. Kuhn, K. H, Die abhiingigkeit der viskositfit vom stromungsgeffille beihochverdiinnten suspensionen und losungen, Helv. Chim. Acta 28 (1945)30.

    [24] E.J. Hinch, L.G. Leal, Effect of Brownian motion on phelogogical properties of asuspension of non-spherical particles, J. Fluid Mech. 52 (1972). 683–&.

    [25] D.H. Berry, W.B. Russel, The rheology of dilute suspensions of slender rods inweak flows, J. Fluid Mech. 180 (1987) 475–494.

    [26] I.M. Krieger, Rheology of monodisperse lattices, Adv. Colloid Interface Sci. 3(1972) 25.

    [27] A.P. Philipse, The random contact equation and its implications for (colloidal)rods in packings, suspensions, and anisotropic powders (vol 12, pg 1127,1996), Langmuir 12 (1996) 5971.

    [28] A.P. Philipse, A. Verberkmoes, Statistical geometry of caging effects in randomthin-rod structures, Physica A – Stat. Mech. Appl. 235 (1997) 186–193.

    [29] H.S. Chen, Y.L. Ding, A. Lapkin, X.L. Fan, Rheological behaviour of ethyleneglycol-titanate nanotube nanofluids, J. Nanopart. Res. 11 (2009)1513–1520.

    [30] H.S. Chen, Y.L. Ding, A. Lapkin, Rheological behaviour of nanofluids containingtube rod-like nanoparticles, Powder Technol. 194 (2009) 132–141.

    [31] A.J.G.a.W.C. Tang, Methodology for comparison of hydraulic and thermalperformance of alternative heat transfer fluids in complex systems (1994).

    [32] O.K.C.R.L. Hamilton, IEC Fundamentals, 2, 1962.

    [33] C.W. Nan, R. Birringer, D.R. Clarke, H. Gleiter, Effective thermal conductivity ofparticulate composites with interfacial thermal resistance, J. Appl. Phys. 81(1997) 6692–6699.

    [34] C.W. Nan, Z. Shi, Y. Lin, A simple model for thermal conductivity of carbonnanotube-based composites, Chem. Phys. Lett. 375 (2003) 666–669.

    [35] G.C.J. Bart, Thermal conduction in nonhomogeneous and phase change media,Delft university of Techology, Netherlands, 1994.

    [36] J.K. Carson, S.J. Lovatt, D.J. Tanner, A.C. Cleland, Thermal conductivity boundsfor isotropic, porous materials, Int. J. Heat Mass Transfer 48 (2005) 2150–2158.

    [37] Z. Hashin, S. Shtrikman, A variational approach to the theory of the effectivemagnetic permeability of multiphase materials, J. Appl. Phys. 33 (1962) 3125–3131.

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

    [39] S. Kirkpatrick, Percolatio and conduction, Rev. Mod. Phys (1973) 574–588.[40] F.M. White, Viscous Fluid Flow, McGraw-Hill, NY, 1991.[41] R.K. Shah, A.L. London, Laminar flow forced convection in ducts, Adv. Heat

    Transfer (Suppl. 1) (1978).[42] Y.G. Tao, W.K. den Otter, W.J. Briels, Kayaking and wagging of rods in shear

    flow, Phys. Rev. Lett. 95 (2005).[43] Y.G. Tao, W.K. den Otter, W.J. Briels, Periodic orientational motions of rigid

    liquid-crystalline polymers in shear flow, J. Chem. Phys. 124 (2006).[44] Y.G. Tao, W.K. den Otter, W.J. Briels, Kayaking and wagging of liquid crystals

    under shear: comparing director and mesogen motions, EPL 86 (2009).[45] Y.G. Tao, W.K. den Otter, J.T. Padding, J.K.G. Dhont, W.J. Briels, Brownian

    dynamics simulations of the self- and collective rotational diffusioncoefficients of rigid long thin rods, J. Chem. Phys. 122 (2005).

    [46] G.B. Jeffery, The motion of ellipsoidal particles immersed in a viscous fluid,Proc. Roy. Soc. Lond. A 102 (1922) 18.

    Laminar convective heat transfer of alumina-polyalphaolefin nanofluids containing spherical and non-spherical nanoparticles1 Introduction2 Experiments and methods2.1 Preparation of nanofluids2.2 Viscosity and thermal conductivity measurements2.3 Convective heat transfer experiments

    3 Data reduction4 Measurement uncertainties5 Results and discussion5.1 Viscosity5.2 Thermal conductivity5.3 Pressure drop5.4 Convection heat transfer

    6 ConclusionsAcknowledgmentsReferences