You are on page 1of 11

Chemical Engineering Science 64 (2009) 3290 -- 3300

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: w w w . e l s e v i e r . c o m / l o c a t e / c e s

Investigating the efficacy of nanofluids as coolants in plate heat exchangers (PHE)


M.N. Pantzali, A.A. Mouza, S.V. Paras
Laboratory of Chemical Process and Plant Design, Department of Chemical Engineering, Aristotle University of Thessaloniki, University Box 455, GR 54124 Thessaloniki, Greece

A R T I C L E

I N F O

A B S T R A C T

Article history: Received 14 January 2009 Received in revised form 24 March 2009 Accepted 1 April 2009 Available online 9 April 2009 Keywords: Heat transfer Nanofluid Plate heat exchanger Turbulent flow Laminar flow Thermophysical properties

The efficacy of nanofluids as coolants is investigated in the present study. For the nanofluids tested, systematic measurements confirmed that the thermophysical properties of the base fluid are considerably affected by the nanoparticle addition. A typical nanofluid, namely a 4% CuO suspension in water, is selected next and its performance in a commercial herringbone-type PHE is experimentally studied. The new experimental data confirmed that besides the physical properties, the type of flow inside the heat exchanging equipment also affects the efficacy of a nanofluid as coolant. The fluid viscosity seems also to be a crucial factor for the heat exchanger performance. It is concluded that in industrial heat exchangers, where large volumes of nanofluids are necessary and turbulent flow is usually developed, the substitution of conventional fluids by nanofluids seems inauspicious. 2009 Elsevier Ltd. All rights reserved.

1. Introduction The new technological developments as well as the industrial process intensification have made the need for more efficient heat exchanging systems a contemporary demand. Therefore, the scientific interest is focused both on improving the equipment design and on enhancing the thermal capability of the working fluids. The progress in equipment design has led to the development of compact plate heat exchangers (PHE) with modulated surface. The high surface density of such devices combined with a type of flow that involves successive flow separation and reattachment inside the narrow PHE passages augment heat transfer rate. At the same time the complexity induced by the modulations significantly increases the friction losses, making the design of this type of equipment an optimization problem that must compromise between heat transfer enhancement and pumping power demands (Kanaris et al., 2006). The need for working fluids with improved performance has increased the scientific interest in nanofluids, i.e., colloidal suspensions of nanometer-sized solid particles. Das et al. (2006), Trisaksri and Wongwises (2007) and Wang and Mujumdar (2007) have recently summarized the work done in this area. The available literature indicates that the thermal conductivity of the nanofluid is higher than that of the base fluid and it depends strongly on the size, shape and volume fraction of the nanoparticles as well as on the type of the nanoparticles and of the base fluid (Trisaksri and Wongwises, 2007). The published rheological studies of nanofluids are limited and the

Corresponding author. Tel.: +30 2310 996174. E-mail address: paras@cheng.auth.gr (S.V. Paras). 0009-2509/$ - see front matter 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.ces.2009.04.004

Newtonian (or not) behaviour of the various nanofluids has not been completely clarified (Wang and Mujumdar, 2007). For example, Das et al. (2003a) report the Newtonian behaviour of an Al2 O3 water nanofluid with up to 4% particle volume concentration whereas Pak and Cho (1998), who consider similar nanofluids, detected a nonNewtonian behaviour. Nguyen et al. (2007a) reported that the viscosity of Al2 O3 and CuOwater nanofluids is significantly affected by the particle concentration and the temperature. They also questioned the applicability of the aforementioned nanofluids in heat exchanging equipment, since the hysteresis phenomenon on viscosity, observed after subsequent heating and cooling of the fluid, strongly affects the entire flow field and consequently the heat transfer behaviour. The heat capacity of nanofluids is usually estimated using theoretical equations, since to the authors' best knowledge the relevant published experimental studies are limited. It must be pointed out that the increase of thermal conductivity of nanofluids is necessary but not sufficient condition for achieving high performance in heat exchanging equipment, and hence further investigation is required in this area (Das et al., 2006). The majority of the available experimental studies on the application of nanofluids at single-phase forced convective heat transfer concern laminar flow mostly in externally heated circular tubes or microchannels. Among the various types of nanoparticles and concentrations used, metal oxides with particle volume concentrations up to 4% are the most common, probably due to their lower price. The reported findings and conclusions vary considerably. For example, Ding et al. (2006) studied the performance of a nanofluid containing carbon nanotubes and reported that the heat transfer coefficient can be 3.5 times higher than the respective water value, while Zeinali Heris et al. (2007), who used Al2 O3 water nanofluids, found an enhancement of less than 40%. Wen and Ding (2004) noted that the

M.N. Pantzali et al. / Chemical Engineering Science 64 (2009) 3290 -- 3300 Table 1 Available experimental studies on single-phase heat transfer using nanofluids. Reference Wen and Ding (2004) Ding et al. (2006) Yang et al. (2005) Jung et al. (2009) Hwang et al. (2009) Zeinali Heris et al. (2006) Zeinali Heris et al. (2007) Lee and Mudawar (2007) Chein and Chuang (2007) Rea et al. (2009) Nguyen et al. (2007b) Pantzali et al. (2009) He et al. (2007) Pak and Cho (1998) Williams et al. (2008) Xuan and Li (2003)
a

3291

Nanofluid

(%)

Nanofluid conduit 4.5 mm i.d. horizontal tube 4.5 mm i.d. horizontal tube 4.6 mm i.d. horizontal tube Rectangular channels5050, 50100, 100100 m 1.8 mm i.d. horizontal tube 6 mm i.d. horizontal tube 6 mm i.d. horizontal tube MCHSa Single-phase MCHS 4.5 mm i.d. vertical tube Miniature PHE (elec. coolingliquid system) Miniature PHE (elec. cooling-liquid system) 4 mm i.d. vertical tube 10.7 mm i.d. horizontal tube 9.4 mm i.d. horizontal tube 10 mm i.d. horizontal tube

Type of flow Laminar Laminar Laminar Laminar Laminar Laminar Laminar Laminar Laminar Laminar Mildly turbulent Mildly turbulent Lamin. & Turb. Turbulent Turbulent Turbulent

Result +47% in h +350% in h +22% in h +32% in h +8% constant enhancement in h +30% in Nu Better performance with Al2 O3 < +40% in h h enhancement The enhancement reduces as V increases No abnormal enhancement Empirical correlations valid for nanofluids +42% in h

Al2 O3 0.6, 1, 1.6 CNT0.5 Graphite-oil Al2 O3 0.6, 1.2, 1.8 Al2 O3 < 0.3 Al2 O3 0.2, 1, 2, 3 CuO0.2, 1, 2, 3 Al2 O3 0.2, 0.5, 1, 1.5, 2, 2.5 Al2 O3 1, 2 CuO0.204, 0.256, 0.294, 0.4 Al2 O3 0.9, 1.8, 3.6 ZrO2 0.2, 0.5, 0.9 Al2 O3 1, 3.1, 6.8 CuO0.04 TiO2 0.24, 0.6, 1.1 Al2 O3 1.34, 2.78 TiO2 0.99, 2.04, 3.16 Al2 O3 0.9, 1.8, 3.6 ZrO2 0.2, 0.5, 0.9 Cu0.3, 0.5, 0.8, 1, 1.2, 1.5, 2

< +20% increase in heat transfer rate


+12% in h (laminar) > +40% in h (turbulent) +75% for a given Re 12% for a given V No abnormal enhancement Empirical correlations valid for nanofluids > +39% in Nu

Micro-channel heat sink.

heat transfer coefficient enhancement is much higher than the increase in the thermal conductivity, which is not in agreement with the observations by Yang et al. (2005). The heat transfer enhancement is reported either to increase with flow rate (e.g. Jung et al., 2009), to remain constant (e.g. Hwang et al., 2009) or even to reduce with flow rate (e.g. Chein and Chuang, 2007). Xuan and Li (2003) studied turbulent flow in a horizontal tube using a copper nanofluid and reported about 40% enhancement, while Williams et al. (2008) observed no abnormal enhancement with oxide nanofluids in turbulent flow and reported that if accurate thermophysical properties are employed the existing correlations are capable to predict the convective heat transfer. Nguyen et al. (2007b), who experimentally studied heat transfer in a miniature PHE, reported an enhancement of the heat transfer rates. Pantzali et al. (2009), who investigated the effect of the use of a CuO nanofluid in a similar apparatus both experimentally and numerically, also confirmed the aforementioned trend. Mansour et al. (2007) based on empirical correlations for heat transfer in a uniformly heated tube have demonstrated that the advantages of utilizing nanofluids may vary significantly, depending strongly not only on their thermophysical properties but also on the geometrical characteristics of the heat exchanging equipment and the operating conditions. Bergman (2009) has shown that the performance of a nanofluid is geometry- and flow rate-dependent and that a particular nanofluid, although efficient for one application, can be proved inadequate for another one. The relevant experimental studies are summarized in Table 1. It is clear that most of the experimental works have been conducted in very simple geometrical configurations (e.g. tubes), whereas limited experiments in commercial heat exchangers are available. Moreover, when laminar flow is encountered, the use of a nanofluid is always accompanied by heat transfer augmentation. The results concerning turbulent flow are very limited and seem rather ambiguous. In the present study, the performance of a typical nanofluid used as a coolant in a typical commercial herringbone-type PHE (where the flow complexity is also intense) is experimentally investigated and compared to that of the base fluid (i.e., water). Prior to the experiments, the systematic measurement of all the thermophysical properties involved in heat exchanging processes for various nanofluids will lead to the selection of a typical nanofluid, appropriate for the

heat transfer study in this commercial PHE. Finally and in an effort to assess the efficacy of nanofluids in heat exchanging equipment the results of the present study will be discussed in conjunction with results of other available relevant studies. 2. Preparation of nanofluids Various nanofluids are used in the present study and are presented in Table 2. In all cases water was used as the base fluid. More specifically: Three nanofluids were prepared using commercially available nanoparticles, i.e., Al2 O3 , CuO nanoparticles and carbon nanotubes (CNT), purchased from Nanostructured and Amorphous Materials Inc. (NanoAmor). Two commercially available suspensions containing Al2 O3 and TiO2 nanoparticles were obtained from Sigma Aldrich . A 50% w/w commercial CuO nanofluid was purchased from Alfa Aesar , which was diluted by adding distilled water (Carlo Erba Reagenti SpA, Water Plus for HPLC) under mechanical stirring, to obtain the desirable concentrations. A main concern during the study was the proper homogenization of the suspension. The uniform dispersion and stabilization of the suspensions was accomplished using techniques proposed in the literature (e.g. Wang and Mujumdar, 2007), that is surfactant addition (anionic, cationic or non-ionic), ultrasonic vibration and control of the suspension pH. Among the nanofluids prepared, the CuO suspension gave the most promising results concerning thermal conductivity; however, its stability was poor, even though several surfactants were tested and intense ultrasonic vibration was applied. In an effort to improve its stability, the zeta potential , of the CuO suspension with cetyl trimethylammonium bromide (CTAB) was recorded (Zetasizer nano, ZS90) versus the corresponding pH value and the results are presented in Fig. 1. It must be noted that the zeta potential measurements must be conducted in very dilute suspensions. It is also known that in acidic environment (i.e., pH < 5) the CuO nanoparticles are diluted. As it is shown in Fig. 1, the zeta potential attains

3292 Table 2 Nanofluids studied in the present work. Index Particle size, (APS, by manufacturer) 11 nm 3050 nm i.d.: 35 nm o.d.: 815 nm length: 1050 m < 20 nm Not known 30 nm 30 nm 30 nm

M.N. Pantzali et al. / Chemical Engineering Science 64 (2009) 3290 -- 3300

(% v/v)

Stabilizing agent

Preparation method

Al2 O3 -4 CuO-3 CNT

4.0 3.0 0.5

CTAB CTAB

Dispersion of commercial nanoparticles Dispersion of commercial nanoparticles Dispersion of commercial nanoparticles

Al2 O3 -3 TiO2 CuO-8 CuO-4 CuO-2

2.9 2.4 8.0 4.0 2.0

not known not known not known

Commercial nanofluid by Sigma Aldrich Commercial nanofluid by Sigma Aldrich Dilution of commercial nanofluid by Alfa Aesar Dilution of commercial nanofluid by Alfa Aesar Dilution of commercial nanofluid by Alfa Aesar

40

30

20

(mV)

10

-10

introduced, which, in the case of nanofluids, is better than 2%, is available elsewhere (Assael et al., 2004). The density is calculated by weighing a known volume of the nanofluid. The estimated measuring accuracy is about 5%. The specific heat capacity is measured using differential scanning calorimetry (Setaram C80D). The temperature of both the sample and a reference substance is increased with the same rate and the specific heat of the sample is calculated by measuring the difference in heat required to raise the temperature. The uncertainty of the measurements is better than 3%. The rheological behaviour was studied using a rheometer with coaxial cylinders (Haake RheoStress RS600) and the accuracy of the viscosity measurements is calculated to be about 5%. The surface tension is measured with accuracy better than 2% using the pendant drop method (KSV CAM 200).

-20

-30 5 6 7 8 9 pH 10 11 12 13

The results of the thermophysical property measurements are presented in Table 3. Concerning the thermal conductivity, the general trend is the enhancement of the respective base fluid value. In Table 3 the thermal conductivity values, kn , t , predicted by the Hamilton-Crosser model, Eq. (1), are also included for comparison; kn,t = kp + (n 1)kw (n 1) (kw kp ) kp k w kp + (n 1) + kw kw (1)

Fig. 1. Zeta potential for CuO suspension in water (containing CTAB).

strongly positive and negative values for pH < 5.5 and pH > 11, respectively, indicating that for these values the electrical charge of the particles is strong enough to prevent aggregation and settling. Still, CuO nanofluids with higher particle concentration that would lead to a significant thermal conductivity enhancement and at the same time satisfying stability could not be prepared, independent of the pH values. Consequently, commercial nanofluids that present satisfying stability were considered. In this case aggregation and settling needs at least three weeks at rest to be observed, while ultrasonic vibration can significantly improve the dispersion of the nanoparticles. Among the commercial nanofluids tested, the CuO nanofluid is the most stable, even after it is diluted to prepare suspensions with lower particle volume concentrations. 3. Measurement of thermophysical properties In the present work the thermophysical properties of the nanofluids are systematically measured: The thermal conductivity is measured using the transient hot-wire technique as developed in the Thermophysical Properties Laboratory of the Aristotle University of Thessaloniki. A detailed description of the method, along with an estimation of the uncertainty

where kp and kw are the thermal conductivities of the particles and the particle volume concentration and n the water, respectively, shape factor. It is assumed that all the particles are spherical, i.e., n = 3, except for CNTs, whose shape is considered cylindrical, i.e., n = 6 (Lee et al., 2006). The thermal conductivities of the Al2 O3 , CuO, TiO2 nanoparticles and carbon nanotubes are taken equal to 40, 32.6 (Khandekar et al., 2008), 8.4 (Pak and Cho, 1998) and 3000 W/m K (Nan et al., 2004), respectively. It is worth mentioning that the values measured are generally in good agreement with those predicted by the HamiltonCrosser model (less than 5% deviation). The Al2 O3 and CuO nanofluids measurements confirmed that the thermal conductivity depends strongly on the type of nanoparticles and on the particle volume concentration. The enhancement recorded for these nanofluids is slightly lower than the corresponding values reported in the literature (e.g. Das et al., 2003b; Kim et al., 2007; Lee et al., 1999). The increase in thermal conductivity measured for the TiO2 nanofluid is in agreement with the value reported by Kim et al. (2007). The CuO nanofluids exhibit the highest increase. In the case of the CuO-3 nanofluid, however, the rapid aggregation and settling of the nanoparticles result in a significant change in thermal conductivity with time. For example, three successive measurements of the CuO-3 nanofluid conducted with a time interval of 30 min gave 22%, 11% and 6% enhancement of the thermal conductivity compared to water, showing that the measured value depends on the uniform dispersion and stability of the nanofluid.

M.N. Pantzali et al. / Chemical Engineering Science 64 (2009) 3290 -- 3300 Table 3 Measured thermophysical properties of water and the nanofluids (25 C). Index Thermal conductivity kn (W/m K) 0.607 0.648 0.643 0.628 0.643 0.643 0.720 0.669 0.631 (0.679)a (0.660) (0.620) (0.661) (0.645) (0.756) (0.679) (0.642) Density 3 n (kg/m ) 998 (1076) (1076) (1430) (1214) (1106) Heat capacity cP , n (J/kg K) 4180 (3930) (4031) (2730) (3321) (3694) Viscosity n (mPa s) 1.0 2.9 2.0 5.6 2.0 1.3

3293

Surface tension n (mN/m) 72 35 36 68 66 38 51 51

Base fluid: water Al2 O3 -4 CuO-3 CNT Al2 O3 -3 TiO2 CuO-8 CuO-4 CuO-2
a

1050 1030 1510 1250 1130

3910 3960 2700 3280 3720

(1.1) (1.1) (1.2) (1.1) (1.1)

Values in parentheses denote the respective theoretical prediction.

The thermal conductivity of the CNT nanofluid was found to be slightly higher than that of water, but significantly lower than the one previously measured in the same Laboratory (Assael et al., 2004). It must be pointed out that even during the preparation of the nanofluid the CNT particles used in the aforementioned cases exhibited different behaviour; i.e., the sample examined in this study could not be easily dispersed in the base fluid, while the surfactants used for stabilization by Assael et al. (2004) were proved ineffective in the present case. Presumably, this difference in behaviour can be attributed to the origin of the CNT sample (i.e., supplier and production method) as well as their characteristics (e.g. diameter, length to diameter ratio) and it seems to significantly affect the nanofluid performance (e.g. Das et al., 2006). If the poor stability of the Al2 O3 4, CuO-3 and CNT nanofluids in conjunction with their insufficient thermal conductivity enhancement is considered, these nanofluids will not be further examined in the present study. The density of nanofluids, n , increases compared to water and it is generally in good agreement (as also presented in Table 3) with the values predicted, n , t , by Eq. (2) for a solidliquid suspension:
n,t

+ (1 )

(2)

where p and w are the densities of the nanoparticles and water, respectively. The difference between experimental data and predicted values is less than 5%. The heat capacity measurements show that the specific heat capacity of the base fluid, cP , n , generally decreases when nanoparticles are added. The heat capacity depends almost linearly on the volume fraction of the particles in the suspension and is quite accurately predicted by Eq. (3) (Xuan and Roetzel, 2000). Eq. (4), which is also frequently used for the prediction of nanofluids heat capacity (Pak and Cho, 1998), is found to overpredict its value, as it is also recently reported by Zhou and Ni (2008); therefore its use in calculations would not yield reliable results. cP,n,t =
p cP ,p

+ (1 )
n

w cP ,w

(3) (4)

It has been proved that the addition of nanoparticles in the base fluid results in both a decrease of the heat capacity (which dictates the total amount of heat that can be absorbed by the stream) and an increase of the thermal conductivity (which in turn is related to the heat transfer rate). Thus, the nanofluid use seems advantageous compared to the base fluid in heat exchangers, where the heat transfer rate must be high, that is they have small heat transfer area and high heat flux flow rate, as it has been also speculated by Williams (2006). The rheological experiments were conducted by changing the shear rate applied on the sample under constant temperature. The rheological behaviour and the viscosity are generally affected by the particle volume concentration and the kind of nanoparticles used. The shear stress recorded versus the respective shear rate exerted on the samples is presented in Fig. 2, while the viscosity values are given in Table 3. Both the TiO2 and the Al2 O3 -3 nanofluid exhibit Bingham behaviour that is they begin to flow only after the shear stress exceeds a certain value. Additionally, the TiO2 nanofluid shows a hysteresis phenomenon (Fig. 2a). The CuO nanofluids retain Newtonian behaviour regardless of the particle volume concentration, at least for the concentrations tested (Fig. 2b). The measured viscosity of the nanofluid was generally higher than that of the base fluid and was found to depend both on the type of particles and on their concentration. The results for the CuO nanofluids are in agreement with the measurements of Nguyen et al. (2007a). In the case of Al2 O3 -3 nanofluid the value obtained in the present work is in agreement with the viscosity measured by Pak and Cho (1998) for a 3% particle volume concentration whereas it is significantly higher than that reported by Nguyen et al. (2007a). The correlation proposed by Williams et al. (2008) provides a relatively good prediction. The TiO2 nanofluid exhibits higher viscosity than that measured by Pak and Cho (1998) for a 3% particle volume concentration. These deviations could be attributed to the differences in the production procedure, since the samples are purchased from different suppliers. The values predicted by the commonly used Einstein's equation are included in Table 3 for comparison:
n,t

cP,n,t = cP,p + (1 )cP,w

= (1 + 2.5 )

(5)

where cP , p and cP , w are the heat capacities of the nanoparticles and water, respectively. For the application of the aforementioned models the heat capacities of the Al2 O3 , TiO2 and CuO nanoparticles have been taken equal to 1000 (Wang et al., 2001), 692 (Pak and Cho, 1998) and 530 J/kg K (Leitner et al., 2000), respectively. The values calculated by Eq. (3) are presented in Table 3 along with the results of heat capacity measurements for comparison. The measured value for the Al2 O3 -3 nanofluid is in accordance with the results published by Zhou and Ni (2008). It is worth mentioning at this point that in the cases studied, the product of density by heat capacity is practically same as that of water, as also reported by Williams (2006).

where w is the dynamic viscosity of water. This equation, however, only considers the liquid particle interactions and hence is valid only up to a volume fraction of about 0.01 (Williams, 2006). Beyond this value, as in the present case, it highly underestimates the suspension viscosity. Modifications of the Einstein equation, e.g. Batchelor (1977) and Brinkman (1952), also provide considerably lower values compared to the experimental results, a fact that is also reported by Nguyen et al. (2007a). The surface tension , results are also shown in Table 3. The measured surface tension values of some of the nanofluids that were prepared in our Laboratory are also included. It is observed that the

3294

M.N. Pantzali et al. / Chemical Engineering Science 64 (2009) 3290 -- 3300

1.5 Al2O3-3 1.1 TiO2

(Pa)

0.8

(v/v) CuO suspension presents only a marginal increase in thermal conductivity (i.e., about 4%), while the viscosity of the 8% v/v suspension is too high (almost 6 times higher than water). Thus, the 4% particle volume concentration suspension is selected as a typical nanofluid for the heat transfer study. A typical TEM image of the specific nanofluid is presented in Fig. 3. The nominal diameter of the particles is 30 nm, as stated by the provider, which is also verified by the measurements based on the TEM images. The minimum and maximum particle diameters were found to be 13 and 150 nm, respectively. 4. Experimental setup

0.4

0.0 0 100 200 300 (s-1) 400 500 600

1.2

0.9 CuO-8 (Pa)

CuO-4

0.6 CuO-2

0.3

0.0 0 100 200 300 (s-1) 400 500 600

Fig. 2. Shear stress versus shear rate for: (a) Al2 O3 , TiO2 and (b) CuO nanofluids.

surface tension of Al2 O3 -3 and TiO2 nanofluids, which are prepared with no surfactant addition (as claimed by Sigma-Aldrich ), is not considerably lower than that of water. When a surfactant is used for the stabilization of the suspension, e.g. CuO-3 and CNT nanofluids, as expected, a significant reduction is observed, which should be attributed to the surfactant addition. The CuO nanofluids, prepared using the suspension purchased from Alfa-Aesar , also exhibit a noticeable reduction, which could indicate that a surfactant is present, although no relevant information is provided by the manufacturer. A Fourier transform infrared spectroscopy (FTIR) analysis of the sample gives also evidence of the presence of aliphatic compounds, confirming the aforementioned hypothesis. Taking into consideration the thermophysical property measurements, the rheological behaviour and the stability of all nanofluids examined, a CuO nanofluid was chosen as the most appropriate to be applied in the PHE. According to Williams et al. (2008), the advantages of using nanofluids depend on the relative increase of thermal conductivity and viscosity. The 2% particle volume fraction

The experiments are conducted in a welded commercial PHE (Alfa Laval). The PHE comprises 16 stainless steel corrugated plates that create two isolated fluid paths for the hot and cold fluid flow respectively, forming eight flow channels per stream. The plates have chevron-type corrugations with a height of 2 mm and a wavelength (in a direction normal to the crests) 8.6 mm. The corrugations form a herringbone pattern (Fig. 4a) with an angle of 50 relative to the direction of the flow. Successive plates are arranged with the corrugation pattern pointing in opposite directions. The geometrical characteristics of the plate are presented in Table 4. The PHE is thermally insulated to eliminate heat losses. A sketch of the experimental setup is presented in Fig. 4b. Water is used as the hot fluid and its temperature is controlled at approximately 50 C by means of a heater, while its flow rate is adjusted using a high-accuracy valve and is measured by a float-type flow meter. Either water or the nanofluid is used as cooling liquid in the PHE with the intention of comparing their performance. The cooling liquid is stored in a 5-litre container and is recirculated by a centrifugal pump. Its inlet temperature is controlled at approximately 30 C using another similar PHE with tap water as the working fluid. The flow rate of the cooling liquid is adjusted using a high-accuracy valve and, when water is used as cooling liquid, it is measured using an electro-optical flow meter (McMillan, S-111). As the aforementioned flow meter is suitable only for transparent liquids, in the case of the black-coloured CuO nanofluid the flow rate is measured using an online weighing system connected to a PC. The uncertainty of the liquid flow measurements is estimated to be better than 4%. Temperature data are acquired using high accuracy T-type thermocouples (Omega ), located at the inlet and outlet of the cooling liquid and hot-water streams, respectively. The thermocouples, whose accuracy is 0.2 K, are connected through an A/D converter (Advantech , PCI-1710HG) to a PC for temperature monitoring and recording. The pressure drop between two taps located at the entrance and the exit of the cooling liquid conduit was measured using a differential pressure transducer (Validyne , DP103), whose accuracy is 0.25%. During each experiment the flow rates of the two streams were set and after steady state was achieved, the temperatures were recorded. Then, the cooling liquid flow rate was increased and new data were obtained. The aforementioned procedure was also repeated for various hot water flow rates. 5. Results 5.1. Heat transfer The heat removed from the hot water, Qh , and the heat absorbed by the cooling liquid, Qc , are calculated by Eqs. (6) and (7) using the temperature and mass flow rate data recorded during the experiments. It is confirmed that they are practically equal, i.e., within the accuracy of the measuring technique, which is found to be

M.N. Pantzali et al. / Chemical Engineering Science 64 (2009) 3290 -- 3300

3295

Fig. 3. A typical TEM image of the CuO nanofluid.

cooling water

Tci

Tco

Tho

Thi

digital flowmeter

drain hot water drain cooling liquid Weight scale

Fig. 4. Schematic representation of: (a) a herringbone-type plate (Shah and Sekulic, 2003) and (b) the heat exchanging set-up.

3296 Table 4 Plate geometrical characteristics. Plate length Plate width Mean spacing between plates, 2bc Plate sheet thickness, x Port-to-port length Plate width inside gasket, w Heat transfer area per plate Mean flow cross section per channel

M.N. Pantzali et al. / Chemical Engineering Science 64 (2009) 3290 -- 3300

1000 PP, water


0.20 m 0.075 m 0.004 m 0.0005 m 0.173 m 0.07 m 0.017 m2 1.5104 m2

100 PP, CuO

800

Vw VCuO

80

PP, c, W

400

40

4000 200 20

3000

0 1.6

2.0

2.4 Q, kW

2.8

0 3.2

Q (W)

2000

Fig. 6. Necessary pumping power for the cooling liquid and respective volumetric flow rate vs. heat flow rate.

1000 open symbols: water filled symbols: nanofluid 0 0 100 200 Rec 300

Reh 240 275 335 400

4000

3000 Q (W)

and cold (c) stream, respectively. It must be pointed out that the way the experimental data are presented and interpreted might lead to completely different conclusions. In some works (e.g. Chein and Chuang, 2007) the comparison between the nanofluid and the base fluid results are made on a velocity basis whereas in other works (e.g. Pak and Cho, 1998) on a Reynolds number basis. Furthermore, for the nanofluid Reynolds number calculation the properties used are either measured, or estimated using available models and correlations (which are not always accurate, as discussed previously) and in some cases even the base fluid properties are used, thus making the situation highly ambiguous. The heat flow rate Q, measured in the present study is plotted in terms of both Reynolds numbers, Re, (Fig. 5a) and volumetric flow rate, Vc , (Fig. 5b). The physical properties used for the Reynolds number calculation, Eq. (8), are those measured during this study; Re = us Dh (8)

2000 Vh, ml/s 40 open symbols: water filled symbols: nanofluid 1000 0 25 50 Vc (ml/s)
Fig. 5. Heat flow rate vs.: (a) Reynolds number and (b) cooling liquid volumetric flow rate for various hot water flow rates.

47 56 75 100

less than 15%. Qh = Q = mh cp,h (Thi Tho ) Qc = Q = mc cp,c (Tco Tci ) (6) (7)

In the above equations m is the mass flow rate, cP the heat capacity and Ti and To the inlet and outlet temperatures for the hot (h)

where Dh is the conduit hydraulic diameter and us is the superficial velocity inside the conduit, defined as Dh = 2bc and us = V/wbc N, respectively, and N is the number of passes per stream (Vlasogiannis et al., 2002). For a given Reynolds number (Fig. 5a) an increase of heat transfer rates for the nanofluid is clearly indicated, but in this case the nanofluid flow rate is higher than that of water, due to its higher kinematic viscosity. However, the conclusions drawn from Fig. 5b are different. It is observed that for a given flow rate the performance of the PHE is not improved, when the nanofluid is applied as the cooling liquid instead of water, while the heat removed from the hot water remains practically the same. It must be noted that, when the water is replaced by the nanofluid, the aim is not to retain a constant Reynolds number, but to handle the same heat duty using less fluid and with the same or less pumping power. Therefore, a comparison in terms of the necessary pumping power for both cooling liquids, as well as the respective volumetric flow rate, versus the heat transfer rate is considered essential (Fig. 6). It is observed that for a given heat duty the required volumetric flow rates for both the water and the nanofluid are practically equal, while the necessary pumping power in the case of the nanofluid is up to two times higher than the corresponding value for water due to the higher kinematic viscosity of the fluid.

Vc, ml/s

600

60

M.N. Pantzali et al. / Chemical Engineering Science 64 (2009) 3290 -- 3300

3297

CFX

Velocity 1.87

1.40

0.94

0.47

0.00 [m s-1] x

z x

Fig. 7. Typical flow field inside the PHE, as calculated by CFD.

As already mentioned, PHEs consist of modulated plates aiming to the enhancement of turbulent structures and consequently of heat transfer rates. According to Shah and Wanniarachchi (1991) in this kind of equipment turbulent structures exist, even for very low Reynolds numbers. To gain insight of the flow inside the PHE, a CFD simulation is also performed. More details about the simulation can be found elsewhere (Kanaris et al., 2006). In Fig. 7 a typical flow pattern predicted by CFD simulations for the present PHE is shown (the flow-lines represent the tangent of the velocity vector at each point). From this figure the flow complexity can be clearly seen, proving that the flow in this PHE is not laminar, but it is highly disturbed by the plate corrugations. As already pointed out, most of the available studies are focused mainly on laminar flow (Table 1), where a heat transfer coefficient enhancement of varying magnitude is reported. The references concerning turbulent flow, however, are limited and the results seem contradictory at first sight. For example, Pak and Cho (1998) report an increase of heat transfer rates when the results are compared on the basis of same Reynolds numbers. Yet, if the data are plotted in terms of volumetric flow rates, which are considered, as already reported, a more appropriate way of comparing this kind of data, a 12% decrease is observed. Williams (2006) found that the nanofluids tested provide no enhancement for turbulent convective heat transfer at a given pumping power. On the other hand, He et al. (2007) and Xuan and Li (2003) report a significant enhancement of the heat transfer rate when titanium oxide and copper nanofluids, respectively, are used. Williams (2006) has claimed that the effectiveness of nanofluids as cooling liquids depends mainly on the viscosity and not the thermal conductivity. Thus an efficient nanofluid should have both considerably enhanced thermal conductivity and a low viscosity. Furthermore, Lee and Mudawar (2007) based on empirical correlations for heat transfer in a uniformly heated tube, report that for laminar flow the heat transfer coefficient is proportional only to the fluid thermal conductivity, while for turbulent flow it depends also on the heat capacity and viscosity. Therefore, the effect of thermal conductivity in turbulent flow becomes less pronounced, since, in the presence of nanoparticles, heat capacity and viscosity are also considerably influenced. So, the aforementioned contradiction could be raised by taking into consideration the above comments concerning the ther-

mophysical properties of the nanofluids used. A closer inspection of the relevant studies has shown that the nanofluids used by Pak and Cho (1998) and Williams (2006), as well as by the present authors, exhibit a maximum increase in thermal conductivity about 22%, 18% and 10%, respectively, while in all cases a twofold or higher increase in viscosity is recorded. On the other hand, although the nanofluids employed by Xuan and Li (2003) and He et al. (2007) have exhibited a comparable increase in thermal conductivity, their viscosity was only slightly affected (about 10% increase with respect to the base fluid). From the above discussion it can be concluded that the results reported for turbulent flow are not eventually inconsistent. As already reported, in laminar flows, where convection is not the dominant mechanism, the addition of nanoparticles in a conventional fluid (e.g. water) enhances heat transfer rates. However, when heat transfer is mainly controlled by the development of turbulence structures and the existence of flow separation and reattachment, the performance of a nanofluid depends considerably on all the fluid thermophysical properties. This can be possibly attributed to the fact that a significant increase of the nanofluid viscosity suppresses turbulence, and consequently convection, while the increased thermal conductivity cannot always compensate for this suppression. Thus the performance of a nanofluid depends considerably both on the fluid thermophysical properties and on the flow structure inside the heat exchanging channel. This is in accordance with the results of an accompanying paper (Pantzali et al., 2009), where the performance of a CuO nanofluid in a miniature PHE in the absence of intense turbulence structures is investigated. It is also possible, as stated by Bergman (2009), that the heat capacity of the nanofluids might also play a significant role, but unfortunately relevant measurements were not available in the aforementioned studies, in order to verify this statement. Williams et al. (2008) have found that the existing empirical correlations are reliable for the prediction of the convective heat transfer coefficient when nanofluids are involved. To examine the validity of the aforementioned statement, an empirical correlation for the specific PHE is developed in this study, based merely on experiments conducted with water. It is common to use a correlation of the form Nu = a Reb Prc for the prediction of heat transfer coefficient in heat

3298

M.N. Pantzali et al. / Chemical Engineering Science 64 (2009) 3290 -- 3300

exchangers, where the Nusselt and Prandtl numbers are defined as follows: Nu = hDh k (9) (10)

40 water nanofluid 30 Nuc = 0.247 Rec0.66Prc0.4

cP Pr = k

The exponent of the Prandtl number c = 0.4 seems to be the most appropriate for flows in corrugated passages according to Shah and Wanniarachchi (1991). Following the procedure suggested by Vlasogiannis et al. (2002), experiments with equal mass flow rates of water flowing in both streams are carried out. Since both the geometry and the flow are identical, the ratio of the cold over the hot heat transfer coefficient, hc /hh , is only a function of the physical properties. Neglecting the variations in water density and heat capacity, the aforementioned ratio becomes: hc = hh kc kh
0.6 c h 0.4b

Nuc

20

10

=a

(11)

dashed lines represent 12% deviation 0 0 30 60 Rec0.66Prc0.4


Fig. 8. Comparison of the cold stream experimental data with the proposed correlation.

where the physical properties are calculated at the average of the entrance and exit temperature of each stream. By a trial-and-error procedure the value of b that best fits the above experimental data is calculated and the correlation obtained is: Nu = 0.247Re0.66 Pr0.4 (12)

90

120

This correlation is subsequently used for the calculation of the hot stream heat transfer coefficient, hh , which is water for all measurements. On the other hand the cold stream heat transfer coefficient, hc , was independently calculated as follows: The total heat transfer coefficient U, is calculated by Eqs. (13) and (14) based on the experimental data; U= Q A F LMTD (Tho Tci ) (Thi Tco ) (T Tci ) ln ho (Thi Tco ) (13) (14)

10.0 water nanofluid 7.5

P (kPa)

LMTD =

5.0

where A is the total heat transfer area of the PHE, A = 0.26 m2 , and F is the temperature correction factor, which in the case of the countercurrent flow can be taken equal to 1 (Gut et al., 2004). Given the thermal conductivity of the plate, kSS , and its thickness, x, (the heat transfer area of both streams being the same) the cold stream heat transfer coefficient, hc , can be calculated by Eq. (15): 1 x 1 1 + + = U hc kSS hh (15)

2.5

0.0 0 25 50 Vc (ml/s) 75 100

The experimental data for both the water and the nanofluid are plotted in Fig. 8 in terms of the respective Nusselt, Prandtl and Reynolds numbers, where it becomes obvious that they also agree with Eq. (12). This means that the correlation is capable of predicting the heat transfer also in the case of the nanofluid. More specifically, if the measured values of the nanofluid physical properties are replaced in the aforementioned correlation, it is predicted that for a given volumetric flow rate the heat transfer coefficient of the nanofluid will be almost equal (95%) to the respective value for water. This means that the predicted heat transfer rates are practically equal in both cases, which is in agreement with the experimental results. 5.2. Pressure drop An important parameter in the application of nanofluids in heat exchanging equipment is the pressure drop developed during the flow through the PHE. In Fig. 9 the total pressure drop P, measured

Fig. 9. Pressure drop of the cooling liquid inside the PHE vs. the respective volumetric flow rate.

inside the PHE, is plotted versus the cooling liquid volumetric flow rate for both the water and the nanofluid. As already reported, the measured viscosity of the suspension (Table 3) exhibits a twofold increase compared to water. The results show, as expected, that this leads to a significant increase in the measured pressure drop and consequently in the necessary pumping power when the nanofluid is applied. From the experimental data it is calculated that the pumping power could be increased about 40% compared to water for a given flow rate. The friction factor values f, for both the water and the 0.135 nanofluid are found to follow a correlation of the form fc = a Re c (Fig. 10), as also proposed by Kanaris et al. (2009) for a similar geometry. The constant a (a = 14.5) is found to be higher than the value proposed by the aforementioned author (a 9.5). This is attributed to the fact that in the present study a welded PHE is used, unlike

M.N. Pantzali et al. / Chemical Engineering Science 64 (2009) 3290 -- 3300

3299

20 w 15 nanofluid

fc =14.5 Rec-0.135 10 fc

industrial heat exchangers, where large volumes of nanofluids are involved and turbulent flow is usually developed, the substitution of conventional fluids by nanofluids seems inauspicious. However, in micro-scale equipment with increased thermal duties, where also volume is a matter, and especially in laminar flow, the use of a nanofluid instead of a conventional fluid seems advantageous. In each case the nanofluid properties should be defined carefully in order to evaluate its efficacy in a specific heat exchanger. Notation A bc cP f k LMTD m N Nu P Pr PP Q Re T U us V w x Greek letters shear rate, s1 zeta potential, V dynamic viscosity, kg/ms density, kg/m3 surface tension, kg/s2 shear stress, Pa particle volume concentration, %v/v Subscripts c h i n o p SS t w cold stream hot stream inlet nanofluid outlet nanoparticles stainless steel theoretical water total heat transfer area, m2 mean spacing between plates, m heat capacity, J/kg K friction factor, dimensionless thermal conductivity, W/m K log mean temperature difference, K mass flow rate, kg/s number of channels per stream, dimensionless Nusselt number, dimensionless pressure drop, Pa Prandtl number, dimensionless pumping power, W heat flow rate, W Reynolds number, dimensionless temperature, K total heat transfer coefficient, W/m2 K superficial velocity, m/s volumetric flow rate, m3 /s plate width inside gasket, m plate sheet thickness, m

0 0 100 200 Rec 300 400

Fig. 10. Friction factor of the cooling liquid vs. Reynolds number.

the PHE used by Kanaris et al. (2009), and, as expected, higher pressure drop is produced. Furthermore, the pressure drop calculated by Kanaris et al. (2009) refers only to the flow from port to port of a plate, while the pressure drop measured in the present study includes the entrance and exit sections of the PHE. It has to be pointed out as well that, according to Kanaris et al. (2009), the exact shape of the corrugation significantly influences the friction factor and even very small manufacturing variations can lead to considerable differences of the pressure drop. 6. Concluding remarks In the present study the efficacy of nanofluids as coolants has been investigated. Prior to this the thermophysical properties of several nanofluids have been systematically measured confirming the general trends reported in the literature, when nanoparticles are added in the base fluid, that is: increase of thermal conductivity, increase of density, decrease of heat capacity increase of viscosity and possibly non-Newtonian behaviour.

The new experimental data concerning the use of nanofluids in a commercial heat exchanger confirmed that, besides the physical properties, the type of flow (laminar or turbulent) inside the heat exchanging equipment plays an important role in the effectiveness of a nanofluid. When the heat exchanging equipment operates under conditions that promote turbulence, the use of nanofluids is beneficial if and only if the increase in their thermal conductivity is accompanied by a marginal increase in viscosity, which seems very difficult to be achieved. On the other hand, if the heat exchanger operates under laminar conditions the use of nanofluids seems advantageous, the only disadvantages so far being their high price and the potential instability of the suspension. The empirical correlations characterizing the heat transfer processes seem to be reliable for the convective heat transfer coefficient prediction in the case of nanofluids under the condition that accurate values of physical properties are provided. In conclusion, in

Acknowledgements Financial support by the General Secretariat for Research and Technology and the European Union (PENED 2003) is greatly acknowledged. Prof. M.J. Assael, Dr J. Tihon and Mr. K. Antoniadis are acknowledged for their contribution to this work. The authors would also like to thank Mr. A. Lekkas, Mr. M. Bridakis, Mr. T. Tsilipiras and Mr. F. Lambropoulos for the technical support. References
Assael, M.J., Chen, C.-F., Metaxa, I., Wakeham, W.A., 2004. Thermal conductivity of suspensions of carbon nanotubes in water. Int. J. Thermophys. 25, 971985.

3300

M.N. Pantzali et al. / Chemical Engineering Science 64 (2009) 3290 -- 3300

Batchelor, G.K., 1977. The effect of Brownian motion on the bulk stress in a suspension of spherical particles. J. Fluid Mech. 83 (1), 97117. Bergman, T.L., 2009. Effect of reduced specific heats of nanofluids on single phase, laminar internal forced convection. Int. J. Heat Mass Transfer 52 (56), 12401244. Brinkman, H.C., 1952. The viscosity of concentrated suspensions and solutions. J. Chem. Phys. 20 (4), 571581. Chein, R., Chuang, J., 2007. Experimental microchannel heat sink performance studies using nanofluids. Int. J. Thermal Sci. 46 (1), 5766. Das, S., Choi, S., Patel, H., 2006. Heat transfer in nanofluidsa review. Heat Transfer Eng. 27 (10), 319. Das, S.K., Putra, N., Roetzel, W., 2003a. Pool boiling characteristics of nano-fluids. Int. J. Heat Mass Transfer 46, 851862. Das, S.K., Putra, N., Thiesen, P., Roetzel, W., 2003b. Temperature dependence of thermal conductivity enhancement for nanofluids. J. Heat Transfer 125 (4), 567574. Ding, Y., Alias, H., Wen, D., Williams, R.A., 2006. Heat transfer of aqueous suspensions of carbon nanotubes (CNT nanofluids). Int. J. Heat Mass Transfer 49 (12), 240. Gut, J.A.W., Fernandes, R., Pinto, J.M., Tadini, C.C., 2004. Thermal model validation of plate heat exchangers with generalized configurations. Chem. Eng. Sci. 59 (21), 45914600. He, Y., Jin, Y., Chen, H., Ding, Y., Cang, D., Lu, H., 2007. Heat transfer and flow behaviour of aqueous suspensions of TiO2 nanoparticles (nanofluids) flowing upward through a vertical pipe. Int. J. Heat Mass Transfer 50 (1112), 22722281. Hwang, K.S., Jang, S.P., Choi, S.U.S., 2009. Flow and convective heat transfer characteristics of water-based Al2 O3 nanofluids in fully developed laminar flow regime. Int. J. Heat Mass Transfer 52 (12), 193199. Jung, J.-Y., Oh, H.-S., Kwak, H.-Y., 2009. Forced convective heat transfer of nanofluids in microchannels. Int. J. Heat Mass Transfer 52 (12), 466472. Kanaris, A.G., Mouza, A.A., Paras, S.V., 2009. Optimal design of a plate heat exchanger with undulated surfaces. Int. J. Thermal Sci. 48 (6), 11841195. Kanaris, A.G., Mouza, A.A., Paras, S.V., 2006. Flow and heat transfer prediction in a corrugated plate heat exchanger using a CFD code. Chem. Eng. Technol. 29 (8), 923930. Khandekar, S., Joshi, Y.M., Mehta, B., 2008. Thermal performance of closed two-phase thermosyphon using nanofluids. Int. J. Thermal Sci. 47 (6), 659667. Kim, S.H., Choi, S.R., Kim, D., 2007. Thermal conductivity of metal-oxide nanofluids: particle size dependence and effect of laser irradiation. J. Heat Transfer 129 (3), 298307. Lee, J., Mudawar, I., 2007. Assessment of the effectiveness of nanofluids for singlephase and two-phase heat transfer in micro-channels. Int. J. Heat Mass Transfer 50 (34), 452463. Lee, S., Choi, S.U.S., Li, S., Eastman, J.A., 1999. Measuring thermal conductivity of fluids containing oxide nanoparticles. Trans. ASME J. Heat Transfer 121 (2), 280289. Leitner, J., Sedmidubsky, D., Dousova, B., Strejc, A., Nevriva, M., 2000. Heat capacity of CuO in the temperature range of 298.151300 K. Thermochim. Acta 348 (12), 4951. Mansour, R.B., Galanis, N., Nguyen, C.T., 2007. Effect of uncertainties in physical properties on forced convection heat transfer with nanofluids. Appl. Therm. Eng. 27 (1), 240249. Nan, C.-W., Liu, G., Lin, Y., Li, M., 2004. Interface effect on thermal conductivity of carbon nanotube composites. Appl. Phys. Lett. 85 (16), 35493551.

Nguyen, C.T., Desgranges, F., Roy, G., Galanis, N., Mare, T., Boucher, S., Angue Mintsa, H., 2007a. Temperature and particle-size dependent viscosity data for water-based nanofluidshysteresis phenomenon. Int. J. Heat Fluid Flow 28 (6), 14921506. Nguyen, C.T., Roy, G., Gauthier, C., Galanis, N., 2007b. Heat transfer enhancement using Al2 O3 water nanofluid for an electronic liquid cooling system. Appl. Therm. Eng. 27 (89), 15011506. Pak, B.C., Cho, Y.I., 1998. Hydrodynamic and heat transfer study of dispersed fluids with submicron metallic oxide particles. Exp. Heat Transfer 11 (2), 151170. Pantzali, M.N., Kanaris, A.G., Antoniadis, K.D., Mouza, A.A., Paras, S.V., 2009. Effect of nanofluids on the performance of a miniature plate heat exchanger with modulated surface. Int. J. Heat Fluid Flow, in press, doi:10.1016/j.ijheat fluidflow.2009.02.2005. Rea, U., McKrell, T., Hu, L.-w., Buongiorno, J., 2009. Laminar convective heat transfer and viscous pressure loss of aluminawater and zirconiawater nanofluids. Int. J. Heat Mass Transfer 52 (78), 20422048. Shah, R.K., Sekulic, D.P., 2003. Fundamentals of Heat Exchanger Design. Wiley, New Jersey. Shah, R.K., Wanniarachchi, A.S., 1991. Plate heat exchanger design theory. In: Buchlin, J.-M. (Ed.), Plate Heat Exchanger Design Theory. von Karman Institute Lecture Series 199104. Trisaksri, V., Wongwises, S., 2007. Critical review of heat transfer characteristics of nanofluids. Renew. Sust. Energy Rev. 11 (3), 512523. Vlasogiannis, P., Karagiannis, G., Argyropoulos, P., Bontozoglou, V., 2002. Airwater two-phase flow and heat transfer in a plate heat exchanger. Int. J. Multiphase Flow 28, 757772. Wang, L., Tan, Z., Meng, S., Liang, D., Li, G., 2001. Enhancement of molar heat capacity of nanostructured Al2 O3 . J. Nanoparticle Res. 3, 483487. Wang, X.-Q., Mujumdar, A.S., 2007. Heat transfer characteristics of nanofluids: a review. Int. J. Thermal Sci. 46 (1), 119. Wen, D., Ding, Y., 2004. Experimental investigation into convective heat transfer of nanofluids at the entrance region under laminar flow conditions. Int. J. Heat Mass Transfer 47, 51815188. Williams, W., 2006. Experimental and theoretical investigation of transport phenomena in nanoparticle colloids (nanofluids), Ph.D. Thesis, Massachusetts Institute of Technology, Boston. Williams, W., Buongiorno, J., Hu, L.-W., 2008. Experimental Investigation of turbulent convective heat transfer and pressure loss of alumina/water and zirconia/water nanoparticle colloids (nanofluids) in horizontal tubes. J. Heat Transfer 130 (4), 042412042417. Xuan, Y., Li, Q., 2003. Investigation on convective heat transfer and flow features of nanofluids. J. Heat Transfer 125, 151155. Xuan, Y., Roetzel, W., 2000. Conceptions for heat transfer correlation of nanofluids. Int. J. Heat Mass Transfer 43, 37013707. Yang, Y., Zhang, Z.G., Grulke, E.A., Anderson, W.B., Wu, G., 2005. Heat transfer properties of nanoparticle-in-fluid dispersions (nanofluids) in laminar flow. Int. J. Heat Mass Transfer 48 (6), 1107. Zeinali Heris, S., Etemad, S.G., Nasr Esfahany, M., 2006. Experimental investigation of oxide nanofluids laminar flow convective heat transfer. Int. Commun. Heat Mass Transfer 33 (4), 529535. Zeinali Heris, S., Nasr Esfahany, M., Etemad, S.G., 2007. Experimental investigation of convective heat transfer of Al2 O3 /water nanofluid in circular tube. Int. J. Heat Fluid Flow 28 (2), 203210. Zhou, S.Q., Ni, R., 2008. Measurement of the specific heat capacity of water-based Al2 O3 nanofluid. Appl. Phys. Lett. 92 (9),

You might also like