Anda di halaman 1dari 17

International Journal of Thermal Sciences 72 (2013) 1e17

Contents lists available at SciVerse ScienceDirect

International Journal of Thermal Sciences


journal homepage: www.elsevier.com/locate/ijts

Nanouids in thermosyphons and heat pipes: Overview of recent experiments and modelling approaches
Matthias H. Buschmann
Institut fr Luft-und Kltetechnik Dresden, Bertolt-Brecht Allee 22, 01309 Dresden, Germany

a r t i c l e i n f o
Article history: Received 23 November 2012 Received in revised form 24 April 2013 Accepted 24 April 2013 Available online 24 June 2013 Keywords: Nanouids Thermosyphon Heat pipe Thermal performance

a b s t r a c t
Confronted with limited energy and material resources and undesirable manmade climate changes, science is searching for new and innovative strategies to save, transfer and store thermal energy. Currently, one of the most intensively discussed options are the so-called nanouids. Nanouids are suspensions consisting of a liquid baseuid and solid particles of sizes ranging from 10 nm to 200 nm. The higher thermal conductivity of these nanoparticles leads to an increased effective thermal conductivity of the uid which, the general expectation is, should enhance the heat transfer of the device. This overview aims to compile results of the application of nanouids in thermosyphons, heat pipes, and oscillating heat pipes. The general goal is to draw conclusions with respect to the potentials for improvement of the thermal performance of these gadgets. Additionally, possible mechanisms which may generate these improvements are discussed. All together 38 experimental studies and 4 modelling approaches are analyzed. While most investigations recognize nanouids as an advantageous working uid, some others report negative effects. Performance effects which are related to lling ratio, inclination angle, and operation temperature seem to be similar to those for classical working uids. Several authors report a decrease of the thermal resistance or an increase of the efciency with increasing concentration, but also a reversing of this trend if a certain optimal concentration is exceeded. This observation mainly follows with a signicant increase of the evaporator heat transfer coefcient. The condenser heat transfer coefcient seems to be not or only weakly affected. Baseuid, nanoparticle material, size and shape, and the stabilization of the suspension have an inuence on the thermal performance. However, the limited number of experiments does not allow drawing rm conclusions. The main mechanism responsible for the improved thermal performance seems to be a porous layer built from nanoparticles on the evaporator surface. Additional positive effects may follow from the changed thermophysical properties of the working uid. 2013 Elsevier Masson SAS. All rights reserved.

1. Introduction Confronted with limited energy and material resources and undesirable manmade climate changes, science is searching for new and innovative strategies to save, transfer, and store thermal energy. Currently, one of the most intensively discussed options are the so-called nanouids. Rapid advances in manufacturing methods allow the production of nanoparticles of various sizes, shapes, and materials. Nanouids are created by suspending nanoparticles of size 10 nme200 nm in varying baseuids. Fig. 1 compares the normalized numbers of publications in the elds of nanouids, heat transfer, turbulence, and turbulent boundary layer. Normalization is carried out with the 2011 values which are 425 for the keyword nanouids, 8729 for heat transfer, 5736 for turbulence

E-mail address: Matthias.Buschmann@ilkdresden.de. 1290-0729/$ e see front matter 2013 Elsevier Masson SAS. All rights reserved. http://dx.doi.org/10.1016/j.ijthermalsci.2013.04.024

and 896 for turbulent boundary layer. Clearly the exponential increase of publications for nanouids is visible. The motivation for this new step in heat and mass transfer can be found in upward trends in energy density of electronic devices, increasing packing density of heat transfer equipment in general, miniaturization of heat exchangers, and other advanced heat transfer concepts. The general expectation is that the higher thermal conductivity of the nanoparticle materials leads to an increased effective thermal conductivity of the uid which in turn should enhance the heat transfer of the device. The minuteness of the nanoparticles provides hope that these advantages are not counteracted by clogging, sedimentation, and abrasion, issues known for larger particles. Several recent overviews analyzing the current nanouid research (e.g. Sergis and Hardalupas [1], Thomas and Sobhan [2]) support this assumption. The darker facet of nanouids is their extreme complexity, preventing rst-principle solutions and customary physical models from describing uid mechanic and

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

Nomenclature di dnf Lc Le Lhp inner diameter, mm diameter of nanoparticle, nm length of condenser section, mm length of evaporator section, mm geometric length of heat pipe, mm

Greek symbols a angle of inclination against horizontal axis, 4 concentration, vol. %, wt. %, ppm Abbreviations diamond Cd CNT carbon nanotubes

DI EA EG ESS FR gHP HP MWCNT OHP oTS sHP ST TS US FP-OHP

deionized electromagnetic vibration ethylene glycol electrostatic stabilization lling ratio grooved heat pipe heat pipe multiwalled carbon nanotubes oscillating heat pipe open two-phase thermosyphon heat pipe with sintered powder wick surfactant closed two-phase thermosyphon ultrasonic vibration at plate oscillating heat pipe

thermodynamic behaviour (Feja and Buschmann [3]). Therefore, most nanouid investigations are carried out experimentally and so far in rather simple geometries such as straight pipes (Prabhat et al. [4], Buschmann [5]). New challenges occur with the application of nanouids in thermosyphons, heat pipes, and oscillating heat pipes (Liu and Li [6]). Thermosyphons and heat pipes are well understood and widespread applied heat transfer devices. They are highly efcient due to the utilized phase change. Design, operation principles and thermal performance are discussed in detail in several textbooks (e.g. Reay and Kew [7], Faghri [8]). Schematics of the operation principles of these three types of apparatus are shown in Fig. 2. Thermosyphons are devices of passive heat exchange employing natural convection. Fig. 2a shows the main components of thermosyphons, the ways of rising vapour, and down owing condensate lm. Thermosyphons are either open-loop ore closed. In the latter case the working uid returns to the original evaporator via the down running condensate lm. The main parts of a heat pipe (Fig. 2b) are the evaporator were heating takes place, the adiabatic section, and the condenser were cooling takes place. In a standard heat pipe condensate return is ensured via capillary force. Therefore, the inner wall is lined with a wick which has a capillary structure. Usually screen wicks, sintered powder wicks, axial or helical grooves are employed. Alternatively the condensate can be returned by centripetal, electrokinetic, magnetic, osmotic forces or other strategies [7]. In case that a heat pipe is inclined against the horizontal axis its operation is supported by gravity. For large lling

ratios and inclination angles the effect of capillary structure becomes insignicant. In a standard heat pipe the lling ratio is just sufcient to saturate the capillary structure.1 When employing nanouids as working uids, a puzzling manifold of additional parameters e nanoparticle size, shape and material, baseuid characteristics etc. e inuences the thermal performance of these gadgets. The general goal is to lower the thermal resistance, dened as the transported heat divided by the temperature difference between evaporator and condenser. This target seems to be not only reached by the enhanced thermal conductivity of the working uid as it is the case in e.g. laminar pipe ow. It is rather a complex interplay of nanouid thermal properties, nanoparticle interaction with evaporator surface and wick, and changed vapour bubble generation due to porous layers formed from nanoparticles. This overview aims to compile recent experimental results of thermosyphons, heat pipes, and oscillating heat pipes operated with nanouids. The survey is supplemented by several modelling approaches. Possible physical mechanisms acting in gadgets operated with nanouids are discussed. The general goal is to draw conclusions with respect to the potentials for improvement of the thermal performance. Focus is therewith on the differences between reference uid and nanouid. Additionally suggestions for future research are given. However, the application of nanouids and especially in heat pipes and related devices is still associated with severe and unresolved questions. Many issues are open or even not commenced to investigate. Therefore, parts of the following text and the conclusions are formulated with reserve. 2. Organization of paper The study focuses on recent advances in the application of nanouids in closed two-phase and open thermosyphons, heat pipes, and oscillating heat pipes. The main focus of the study is directed towards the effects caused when nanoparticles are added to the working uids of these gadgets. A general rule proposed by Fernholz and Finley [9] in their 1996 review paper on turbulent boundary layers: In the absence of any complete or convenient theoretical approach, our primary function is to describe what we see is applied. In our case this means that the results seen by the different investigators are compiled according to their logical connectedness. Based on that ordering, the four main parts of the survey discuss physical effects which are related to operating

Fig. 1. Normalized number of publications in different elds of uid mechanics and heat transfer. Normalization is carried out with 2011-values. Data taken 2013-04-23 from ISI WEB of KNOWLEDGE.

The author is grateful to an anonymous reviewer for pointing this out to him.

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

(a)
rising vapor

cooling

(b)
wick vapor

heating

(c)
condenser

wick

down flowing condensate film

upward flowing condensate

vapor bubble Q liquid slug

Q selfsustained thermally driven bubble / slug oscillation

working fluid

heating

working fluid

cooling evaporator

Fig. 2. Schematics of operation principles of thermosyphon (a), heat pipe (b) and oscillating heat pipe (c) adapted from Refs. [7] and [56].

parameters of gadgets, effects which are related to nanouid characteristics, the general thermal performance of the gadgets, and modelling approaches. Some of the ndings cannot be assigned to one of these rubrics alone and appear therefore in multiple contexts. Based on the compiled experimental results and physical arguments of technically closely related investigations, statements with respect to the different physical mechanisms acting in gadgets operated with nanouids are discussed. Table 1 compiles all experimental studies discussed. The different gadgets are denoted as done by the original authors. This is also true for some of the so-called heat pipes which are strictly thermosyphons. The serpentines of the oscillating heat pipes are called turns. Their number is identical with the half of the number of straight tubes. In cases where the original authors provide some information with respect to the wick this is mentioned in Table 1. In some occasions gures of the original publications are quoted to draw the readers attention to results which are verbally difcult to describe. 3. Some statistics With this study the results of 38 experimental and 4 modelling approaches are compiled (Table 1). Among the experiments 11 turn their attention to closed two-phase thermosyphons and one to open thermosyphon, 18 to wicked, or grooved heat pipes, and 8 to oscillating heat pipes. 51 nanouids have been tested all together with these experiments. A majority, 41, are water-based nanouids. Other basefluids are the refrigerant R11 (trichlorouoro-methane, CCl3F), ethyleneeglycol mixtures (EG), and acetone ((CH3)2CO). Nanoparticles employed are metals, namely silver (Ag), gold (Au) and copper (Cu), oxides (Al2O3, TiO2, SiO2, CuO, ZnO, Fe2O3) or variations of carbon (diamond, carbon nanotubes). The number of experiments carried out with silver nanoparticles, 13, exceeds by far the other materials. Alumina (Al2O3) follows with 11, CuO with 5, pure copper and titanium (TiO2) with 4, silica (SiO2) and iron (II, III)-oxide (Fe2O3) with 2, and zinc oxide (ZnO) with one experiment. The size of the nanoparticles ranges from 2 nm to about 100 nm. Note that most publications provide only the size of the primary nanoparticles. Due to agglomeration, the actual size within the nanouid might be much larger. Fig. 3 provides an overview of the frequency of the nanoparticle size employed. A maximum exists between 20 nm and 40 nm.

Most authors provide only short comments with respect to the preparation of their nanouids. Four papers give even no information. The overwhelming majority (approx. 91%) employs twostep methods where purchased or at least separately produced nanoparticles are dispersed in the baseuid. Only three experiments (Tsai et al. [10], Hajian et al. [27], Manimaran et al. [28]) were carried out with nanouids produced by one-step methods. Here the nanoparticles were directly created within the baseuid by chemical processes. Most two-step methods employ ultrasonication for dispersing the nanoparticles. The sonication time varies between 1 h and 20 h (Fig. 4). A maximum is found around 4 he6 h. About one third of the experimental groups state that they have not employed any stabilization or surfactant. Particle concentrations are either given in volume or in weight percentage or in parts per million (ppm). Nanoparticle concentrations ranging from 0.003 to 5.3 vol. %, from 0.1 to 0.5 wt. %, and from 1 to 104 ppm are utilized in the experiments (Table 1). Due to the fact that it is rather difcult to value how the baseuid density is changed by adding stabilizers or surfactants (Feja and Buschmann [3]), no attempts have been made to convert these different data. While 16 experiments are conducted only in vertically orientated gadgets, 9 studies investigate only horizontally positioned apparatus. In seven experiments the inclination angle is varied. Only one publication (Riehl and dos Santos [54]) reports inclination angles between 90 (evaporator on top) and 90 (evaporator at bottom) for an oscillating heat pipe. 4. Effects with respect to gadget parameters 4.1. Filling ratio It is known from classical working uids that the lling ratio has a non-negligible inuence on the maximum heat throughput of heat pipes (Reay and Kew [7]). The lling ratio of thermosyphons and gravity supported heat pipes operated in thermosyphon mode is dened as the ratio of working uid volume to internal evaporator volume. Differently for oscillating heat pipes the lling ratio is specied as ratio of working uid volume divided to total internal volume. Experiments by Khandekar et al. [56] employing classical working uids in an OHP showed that the maximum of the heat throughput dependence on the lling ratio is rather at. For water and ethanol the optimal region stretches from 10% to 80% and for

Table 1 Synopses of experimentally investigated thermosyphons, heat pipes and oscillating heat pipes. Author Experimental apparatus Type Tsai et al. [10] Park et al. [11] Wei et al. [12], Kang et al. [13] Naphon et al. [14] Kang et al. [15] Naphon et al. [16] Chen [17] Do et al. [18] Teng et al. [19] Senthilkumar et al. [20,21] B/HP B/HP B/gHP B/HP B/sHP B/HP /HP /HP B/HP B/HP di/LHP/Le/Lc 6a/170/30/50 6/300/50/160 6a/200/e/e 15a/600/200/200 6a/200/e/e 15a/600/200/200 3/200/30/e 3/300/50/150 8/600/e/e 17.6/600/150/150 Specication of nanouid

a
90 90 0 45 0 60 0 0 0, 30, 60, 90 0, 15, 30, 45, 60, 75, 90

FR in % 40 e 66 e 50 e e 20, 40, 60, 80 e

Suspension H2O/Ag H2O/Ag H2O/Ag C2H6O/TiO2 H2O/Ag R11/TiO2 H2O/Ag H2O/Al2O3 H2O/Al2O3 H2O, H2O/C4H10O, H2O/C6H14O, all with Cu H2O/Al2O3 H2O/Ag H2O/Al2O3/Ag H2O/CuO H2O/Al2O3 H2O/Fe3O4 H2O/Ag H2O/CuO H2O/Al2O3 H2O/TiO2 EG/Al2O3 EG/TiO2 EG/ZnO H2O/CNT H2O/CuO H2O/CuO H2O/Al2O3 laponite clay H2O/Al2O3 H2O/Ag

dnf in nm 8e15.6 10, 35 21 10, 35 21 35 30 5 20e30 40

4 103, 104 ppm 1e100 ppm 0.01e0.5 vol. % 1, 10, 100 ppm 0.01e2 vol. % 5, 50, 100 ppm 1, 3 vol. % 0.5, 1, 3 wt. % 100 ppm

Preparation One step Two step US US/no ST US US/no ST US/no ST US/cationic dispersant US M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

Han and Rhi [22]

B/gHP

10/500/167/167

5, 45, 90

27 Ag 89 Al2O3 50 40 10 1.4 e 20 e

0.005, 0.05, 0.1 vol. %b

US 16e20 h

Liu and Zhu [23], Wang et al. [24] Mousa [25] Chiang et al. [26] Hajian et al. [27] Manimaran et al. [28] Putra et al. [29]

B/HP B/HP B/HP B/HP B/HP B/HP

6.8/350/100/150 20 /120/40/40 e/e/e/e 22.36/1000/200/250 20.8/600/150/150 7.44/200/60/60


a

0, 30, 45, 60, 90 90 90 0 0, 30, 60, 90 0

60 25e100 e e 25, 50, 75,100 e

0.5, 0.8, 1, 1.25, 1.5, 2 wt. % 0.25, 0.5, 0.75, 1.2 vol. % 0.003, 0.8 vol. % 50, 200, 600 ppm 1 wt. % 1, 2, 3, 4, 5 vol. %

US 10 h US/no ST e One step/no ST/ US 15 min One step US 1 h

Xue et al. [30] Liu et al. [31] Mehta and Khandekar [32], Khandekar et al. [33] Noie et al. [34] Parametthanuwat et al. [35] Liu et al. [36] Mousa [37] Huminic et al. [38], Huminic and Huminic [39] Lu et al. [40] Parametthanuwat et al. [41,42] Yang and Liu [43] Heris et al. [44] Shanbedi et al. [45] Ma et al. [46] Lin et al. [47]

B/TS B/TS B/TS

20/e/e/e 6/350/100/150 16/620/120/300

90 90 90

20 50 100

B/TS B/TS

20/1000/350/400 7.5, 11.1, 25.4 3 Le

3 di 10 di 20 di

Lc Le

90 90

e 30, 50, 80

150/10,0003 30 8.6e13.5 40e47 25 20 <100

1 vol. % 0.1e2 wt. % 1 wt. %

US 2 h/not ST US 12 h US 2 h

1e3 vol. % 0.5 wt. %

US/no ST US 5 h

B/TS B/TS B/TS

6/350/100/150/ 6.37/300/100/100 13.6/2000/850/850

90 90 30, 45, 60, 90

50 25e100 12

H2O/MWCNT H2O/Al2O3 H2O/Fe2O3

15/5000e15,000c 40 4e 5

1, 1.5, 2, 2.5 wt. % 0.25e1.2 vol. % 2, 5.3 vol. %

US 12 US/no ST Two step

B/oTS B/TS /TS /TS B/TS B/OHP/12 turns B/OHP 5 turns

36/e/1750/300 12.7/750/250/250 8/550/100/350 22/400/200 /150 20a/450/160/200 1.65/222/60/60 2.45/170/22/35


d

30 90 90 e 90 80 e

60 30, 50, 80 e 40 60 50 20e80

H2O/CuO H2O/Ag H2O/SiO2 H2O/Ag H2O/MWCNT H2O/Cd H2O/Ag

50 <100 30 e 10e20 nm/ 5e15 mm3 20e50 20

0.8, 1, 1.2, 1.5 wt. % 0.5 wt. % 0.5, 1, 1.5, 2, 2.5 wt. % 20, 40, 60, 80 ppm 0.2e1.5 wt. % 1 vol. % 100, 450 ppm

US 10 h/no ST US 5 h 0.5e1.5 wt. % oleic acid ST US 12 h ST 3-glycid-oxylproyl e US Two step e

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

US 4 h Mixing 3 d/US 3 h US 4 h/no ST

0.25,0.5, 0.75, 1 wt. %

US 5 h

e 3 104 vol. % 0.1, 1 vol. % 0.1e1.2 wt. % 0.3, 1, 3, 5 vol. % 0.1e0.6 wt. % 0.1e1.2 wt. % 5 wt. %

Mixing

US 1.5 h

Fig. 3. Frequency of the nanoparticle size employed in the experiments compiled in Table 1.

B Indicates cylindrical and rectangular cross section. a Outer diameter. b And combinations of these nanouids. c Diameter and length of primary CNT. d Variable evaporator lengths. e Hydraulic diameter. f Boehmite alumina. g The original authors did not employ a constant lling ratio for all experiments. For details see Ref. [49].

150 50 300 100 450 150 0.76e1.39e/131.4/25.4/90

2/120/50/70 1.65/155/40/64 2/225/50/70

B/OHP 40 turns

B/OHP 6 turns B/OHP 6 turns B/OHP 6 turns

/OHP 8, 10, 16 turns

Wannapakhe et al. [48]

Cheng et al. [49]

Riehl & dos Santos [54], Riehl [55]

Qu et al. [50] Ji et al. [51,52] Qu and Wu [53]

B/OHP 7.5 turns

1.5/300/100/100

R-123 from 35% to 80%. This knowledge motivated several nanouid experiments with varying lling ratios. Lin et al. [47], investigating an OHP with 5 turns, found an optimal lling ratio of 60% for DI-water and for Ag-nanouids (100 ppm and 450 ppm). Similarly, Paramatthanuwat et al. [41,42] showed a maximum of heat transfer at FR 50% for silver nanouid in a circular thermosyphon. But the same was true for the reference uid DI-water. Mousa [25] found that the thermal resistance of a vertical circular heat pipe was minimal for a lling ratio of 40e50% for both DI-water and Al2O3-nanouid. The lling ratio was dened as the percentage of the evaporator lled with working uid. This nding was conrmed for a vertical thermosyphon by Mousa [37]. The optimal lling ratio here was 48%. Manimaran et al. [28] (H2O/CuO in wire meshed HP) showed an increase of thermal efciency when the lling ratio was increased (similar as for water). The maximum was reached for an inclination angle of 30 and FR 75%. Even with an inclination angle of 0 maximal efciency was reached with FR 75%. Investigating an OHP with H2O/Al2O3, Qu et al. [50] found that a lling ratio of 70% gives the largest decrease of the thermal resistance at a power input of 58.8 W. An exception was the experiment by Teng et al. [19]. These authors indicated that at an optimal tilt angle of 60 the thermal efciency was higher the lower the FR was. Maximal values were achieved at FR 20% with H2O/Al2O3. Conversely, DI-water showed a maximum at 60 when FR was 60%. To summarize, an optimal lling ratio seems to exist between 45% and 70% depending on the design of the gadget. Clear trends with respect to different nanouids cannot be derived. In the most

70g

50, 60, 70 50 50 90 90 0

50

50 100 150

0, 90

90, 0, 90

90

50

C3H6O/CD H2O/Ag H2O/CD H2O/Al2O3 EG,H2O/g-AlO(OH)f H2O/SiO2 H2O/Al2O3 H2O/Cu

H2O/Ag

2 3 2 56 50, 80 30 56 29

<100

Fig. 4. Sonication time of the experiments compiled in Table 1.

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

cases the found optimal lling ratios are the same or at least similar as for water. 4.2. Inclination angle The inclination angle of a thermosyphon or heat pipe is of special importance for e.g. solar heating systems and cooling devices with changing spatial position. Similar as for the lling ratio, it is known that the inclination angle affects the thermal performance signicantly (Patil and Yarasu [57], Liu and Li [6]). Wang et al. [24] analyzed a wicked heat pipe and found that the inclination angle has nearly no impact on the temperature distribution of the condenser when H2O/CuO nanouids were employed. Therefore the condenser heat transfer coefcient was only negligibly changed. At an inclination angle of 45 an increase of about 5% compared to the horizontal pipe was found for a concentration of 1 wt. %. The evaporator temperature depended slightly on the inclination angle. Here a minimum was found for an angle of 45 compared to the horizontal heat pipe where the heat transfer coefcient was increased by about 22% (see Ref. [24] Fig. 5b). This maximum was not affected by operation pressure and heat input. However, all of these observations were comparable with the results for pure water. In general it was found that the ratio of the maximum heat uxes of nanouid and water was nearly independent of the inclination angle. The efciency of the wire mesh wicked heat pipe investigated by Senthilkumar et al. [20,21] improved for increasing angles of inclination. However, reaching angles larger than 30 for DI-water and aqueous solution of n-Hexanol and 45 for CuO DI-water and CuO aqueous solution of n-Butanol and n-Hexanol nanouids the efciency started to decrease again. Similarly, Teng et al. [19] found for their heat pipe operated with H2O/Al2O3 nanouid that an increasing tilt angle led to an increase of the thermal efciency. This increase deteriorated above 60 . The same was true for pure water. A continuous decrease of thermal resistance with increasing inclination angle was found by Manimaran et al. [28] for their wire meshed heat pipe using copper oxide nanouid. The trends were similar as for water but at much lower levels. A continuous more or less monotonic increase of the heat ux from evaporator to condenser with inclination angle was found by Wannapakhe et al. [48] for an OHP. This was the case for all investigated concentrations of Ag-nanoparticles and pure water. Exceptions are presented in the studies by Huminic et al. [38] and Huminic and Huminic [39], which investigated a thermosyphon at inclination angles between 30 and 90 with a water based Fe2O3nanouid. While the thermal resistance of the water-lled thermosyphon had a minimum at an angle of 45 , the nanouid-lled thermosyphon showed the lowest values at 90 (see Ref. [39] Fig. 7). The maximal reduction rate of the thermal resistance occurred at an inclination angle of 90 and a concentration of 5.3 vol. %. At an angle of 45 the heat throughput increased by 19% for 2 vol. % and by 22.2% for 5.3 vol. %. The increase was nearly doubled for an angle of 90 (39% at 2 vol. % and 42% at 5.3 vol. %). It should be noted that in these experiments the lling ratio was with 12.6% comparably small and that very small particles (4e5 nm) were employed. To summarize, while some authors nd an optimum inclination angle, others see a continuous change of thermal performance with changing angle. Five from the six studies dealing with the inuence of the inclination angle for nanouid-lled gadgets nd similar behaviour as known for pure water. 4.3. Operation temperature The third gadget parameter discussed is the operation temperature. Despite its known relevance from classical working uids,

only three studies address this parameter directly. For example, Paramatthanuwat et al. [41] found that with increasing working temperature the benets of silver nanouid compared to pure water became slightly more pronounced. Paramatthanuwat et al. [42] conrmed this result for water-based Ag-nanouids with added oleic acid surfactant; however, this was also true for the classical working uid DI-water. A comparable result again for H2O/ Ag nanouid is known from Wannapakhe et al. [48]. Their oscillating heat pipe showed a monotonically increasing maximum heat ux from evaporator to condenser with increasing operation temperature regardless of the nanoparticle concentration. Finally, Wang et al. [24] showed that the ratio of the maximum heat ux of H2O/CuO-nanouid and water depended slightly on the operation temperature and reached a maximum of 1.4 and 1.28 at 40  C and 60  C. Taking all results together it seems clear that water-based nanouids behave similarly to pure water with respect to operation temperature. 5. Effects with respect to nanouid parameters 5.1. Baseuid Certainly the characteristics of the baseuid of a nanouid play a dominant role with respect to its effective thermophysical properties. Nevertheless, only three studies, Senthilkumar et al. [20,21] and Putra et al. [29], address the question if different baseuids change the thermal performance of heat pipes. Senthilkumar et al. [20,21] investigated CuO nanouids based on either DI-water or aqueous solution of n-Butanol or n-Hexanol both at a concentration of 100 mg/l. The thermal efciency of nanouids based on aqueous solutions of n-Butanol or n-Hexanol were in all cases signicantly better than for nanouids based on DI-water. The authors argue that this was caused by the positive surface tension gradient with temperature possessed by n-Butanol and n-Hexanol. However, the experiments carried out with nHexanol by Senthilkumar et al. [21] seem to indicate that this effect was neither enforced nor attenuated by adding nanoparticles. Weak differences of the thermal resistance of DI-water and pure ethylene glycol were found by Putra et al. [29] for a screen mesh wick heat pipe. Adding nanoparticles to these baseuids led to the differences becoming much more pronounced. However, in every case the water based liquids showed a lower thermal resistance. While for H2O/Al2O3 (5 vol. %) the thermal resistance was decreased by about 78% (10 W), only a decrease of 11% (10 W) was achieved for EG/Al2O3 (5 vol. %). A similar trend was observed for TiO2 nanoparticles (5 vol. %, 10 W). Here a reduction of 69% was achieved for the water-based and 2.4% for the ethylene glycol-based nanouid. The effect was slightly more pronounced for higher heating loads. Putra et al. [29] argued that this signicant difference was basically caused by the different enhancements of thermal conductivity of the two basefluids. While by adding 5 vol. % of TiO2 nanoparticles to water a thermal conductivity enhancement of about 14% was achieved, for ethylene glycol only an increase of about 9% was obtained. This in turn should have caused the very different increases of the evaporator heat transfer coefcient (see Ref. [29] Fig. 12). These ndings clearly indicate the inuence of the baseuid on the thermal resistance. However, the number of investigations with successively varied baseuids is still too small to draw rm conclusions. 5.2. Particle size, shape, material, and suspension stabilization This section compiles effects which are related to the characteristics of the nanoparticles directly. Numerous studies (for an

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

overview see Sergis and Hardalupas [1]) indicate that the material of the dispersed nanoparticles as well as their size and shape affect the effective thermophysical properties of nanouids signicantly. For the effective values of density, heat capacity, and thermal expansion coefcient, straightforward models exist which are physically correct. The situation is more complicated for thermal conductivity and dynamic viscosity. Very promising semi-empirical approaches either based on classical models like the Maxwellequation for thermal conductivity (Buongiorno et al. [58], Ehle et al. [59]) or based on physical assumptions for dynamic viscosity (Chevalier et al. [60], Corcione [61], Prasher et al. [62]) have been published in recent years. However, these models may lack generality for several reasons (Feja and Buschmann [3]). Stabilization is considered in this section because it is undertaken to prevent particles from sedimentation or forming agglomerates. Tsai et al. [10] found that enlarging the size of their gold nanoparticles from 8 nm to 24 nm led to a signicant reduction of the thermal resistance of the evaporator section. Kang et al. [13] and Kang et al. [15] investigated Ag/H2O suspensions with particle sizes of 10 nm and 35 nm. The general nding was that while with the smaller particles the thermal resistance could be lowered at a maximum of 50%, with the larger particle a lowering of about 80% was achieved. The numerical model by Do and Jang [63] conrmed this nding. Ji et al. [51] investigated an OHP and found that the thermal resistance was reduced signicantly by adding Al2O3nanoparticles of sizes 50 nm and 80 nm. A slightly lower thermal resistance was achieved for the larger particles. Given unchanged primary surface roughness of the evaporators in all of these experiments,2 the surface particle interaction parameter (ratio of average surface roughness to average particle diameter) would be lowered by a factor of 1.6 [51], 3 [10] and 3.5 [13,15] when employing the larger particles. According to Harish et al. [64], the heat transfer enhancement ratio should then be lowered too. The opposite was found by the experiments of Kang et al. [13] and Kang et al. [15] which may indicate that the nanoparticles have acted in a way which cannot be explained alone by particle surface interaction. Therefore, one may hypothesize that under the inuence of bubble formation and departure from the evaporator surface, nanoparticles might be forced to enhanced agglomeration. This would denitively change the particle size distribution of the working uid. Particle size distributions measured under calm conditions would then be irrelevant for the real situation in the vicinity of the evaporator surface. Ji et al. [52] was the only work found so far which addressed the inuence of the particle shape on the thermal performance of a heat pipe. Employing boehmite alumina (g-AlO (OH)) particle shapes of platelets ( 9 nm), blades ( 60 nm, thickness 10 nm), cylinders ( 10 nm, length 80 nm) and bricks (edge length 40 nm) were investigated. The same nanoparticles were analyzed with respect to their effective thermophysical properties of nanouids by Timofeeva et al. [65]. Ji et al. [52] found that at an operation temperature of 20  C and a volume fraction of 0.3 vol. % for heat loads less than 100 W, heat transfer enhancement of their OHP was best for cylindrical nanoparticles followed by blade, plate, and brick shaped particles, respectively. However, for heat loads larger than 125 W the situation changed and brick shaped particles performed best followed by cylindrical, plate, and blade shaped particles, respectively. The experiments by Putra et al. [29] indicated strong differences between H2O/Al2O3 and H2O/TiO2 nanouids for heat transfer

2 Note that the different gadgets are not compared among themselves. It is rather the change of effective roughness which follows from settling down of nanoparticle having different sizes within one device employing different nanouids which is compared.

coefcient at evaporator (see Ref. [29] Fig. 12). In both cases the concentration was 5 vol. %. A similar but much weaker trend was shown for EG/Al2O3 and EG/TiO2 (5 vol. %). Unfortunately, no information with respect to possible different sizes of Al2O3 and TiO2 nanoparticles were given. As mentioned above, this is of importance because particle size affects, together with primary roughness, the bubble formation process. Paramatthanuwat et al. [42], by adding oleic acid surfactant (0.5 vol. %, 1 vol. %, 1.5 vol.%) to water-based Ag-nanouids, found that at a concentration of 1 vol. % of this dispersant the largest heat ux and the highest effectiveness were achieved. Depending on the operation temperature an increase of the heat ux between 150% and 50% (see Ref. [42] e.g. Fig. 7) was achieved. Rising the surfactant concentration further to 1.5 vol. % damped this increase remarkably. However, in any case the nanouids stabilized with oleic acid surfactant showed a better thermal performance than the pure baseuid (DI-water) or the pure nanouid. Similarly, Hajian et al. [27] found that non-stabilized nanouids have a negative effect on thermal resistance. Qu and Wu [53] investigated an oscillating heat pipe employing water-based nanouids with Al2O3 (30 nm) and SiO2 (56 nm) at concentrations 0.1, 0.3, and 0.6 wt. %. They found that while the silica particles increased the overall thermal resistance, the alumina particles lowered it. The reason seemed to be that the alumina particles deposited on the evaporator increased surface nucleation and thus enhanced heat transfer. The silica particles also formed a precipitate on the evaporator surface but decreased nucleation and contact angle. The straightforward conclusion that the particle material was the relevant parameter should be taken with caution. The TEM images of SiO2 and Al2O3 nanoparticles shown by Qu and Wu [53] indicate signicant differences of effective particle sizes due to clustering (see Ref. [53] Fig. 1). From that observation one rather could conclude that the stabilization and, therewith, the dispersion of the particles is important for the particle layer formed on the evaporator. Yang and Liu [43] investigated two types of water-based silica nanouids. The rst was a conventional nanouid and the second was a specially stabilized nanouid grafting silane (CAS number 2530-83-8) on the particle surface. The later was named functionalized nanouid. Both uids were manufactured employing the same technology (two-step method, 12 h ultrasonic bath). While the conventional nanouid built a porous layer on the evaporator surfaces, this was not the case with the functionalized one. The authors reasoned that the latter enhanced the evaporator heat transfer coefcient due to its changed thermophysical properties, mainly its increased thermal conductivity. The conventional nanouid decreased the evaporating heat transfer coefcient by up to 11%, depending on the operating temperature. Yang and Liu [43] argue that this deterioration was largely caused by the deposition layer on the evaporator surface. The analytical model by Alizad et al. [66] indicated a clear effect of the nanoparticle material on the decrease of the thermal resistance. Keeping baseuid, nanoparticle size, and concentration xed showed that CuO-particles lower thermal resistance the most, followed by Al2O3-particles and TiO2-particles. This is surprising from the point of view of a nanouid as heat carrier because the thermal conductivity of the particle materials order as Al2O3 (35 W/ m-K), CuO (20 W/m-K) and TiO2 (11.7 W/m-K). Therefore, one may assume that the heat transfer process depends additionally on other thermophysical properties such as heat capacity, and density, which is indeed known from classical working uids. Additionally, the model indicated that the inuence of the particle size diminished for particles larger than 20 nm. It is highly likely that nanoparticle size, shape, and material have an inuence on the thermal performance of the different gadgets. The same is true for any stabilization or surfactant employed.

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

However, most of these inuences seem to be different than for heat transfer applications without phase changes where thermophysical properties are in focus. That one third of all experiments was carried out with non-stabilized nanouids points in the same direction. There is some likelihood that stabilization as done for nanouids in other thermodynamic applications is counterproductive for the gadgets investigated here. Especially the interaction between evaporator surface and wick surface on the one side and nanoparticles on the other side has to be taken in consideration. Some studies indicate that the larger the nanoparticles, the stronger the enhancement of the evaporator heat transfer coefcient. Nevertheless, far too few experiments have been carried out to answer this and the much more complex questions with respect to optimal nanoparticle shapes and proper stabilizations. 5.3. Concentration of nanoparticles Nanoparticle concentration is paramount with respect to a nanouids effective thermophysical properties. Therefore, it is not surprising that many studies aim to investigate the dependence of thermal performance on concentration. Numerous groups report a decrease of the thermal resistance or an increase of the thermal efciency with increasing concentration, but also a reversing of this trend if a certain optimal concentration is exceeded. Already the early work by Wei et al. [12] reported a declining dependence of the thermal resistance on particle concentration of H2O/Ag nanouids. At around 100 ppm this dependence seemed to have reached saturation. This nding was basically conrmed by Kang et al. [13] who found a lowering of the evaporator temperature by adding 1, 10 or 50 ppm Ag-nanoparticles to pure water. However, this trend was stopped for a concentration of 100 ppm and no further improvement was observed. Another conrmation came from Lin et al. [47], who showed that a concentration of 100 ppm of Agnanoparticles (20 nm) in DI-water led to a better efciency of their OHP than a concentration of 450 ppm. An OHP with H2O/Ag nanouid was also investigated by Wannapakhe et al. [48], who found that a concentration of 0.5 wt. % delivered the highest heat ux for all operation temperatures, inclination angles, and aspect ratios Le/di. Hajian et al. [27] presented a more detailed study. Investigating a comparably large wicked heat pipe (Le 1000 mm) they found that adding 50 ppm Ag-nanoparticles to pure water led to a maximum lowering of the thermal resistance by about 30%. Higher concentrations (200 ppm and 600 ppm) showed a signicant increase of the thermal resistance. Analyzing the response times of their heat pipe in dependence of the employed working uid, surprisingly no regularity in dependence of the concentration was found. Naphon et al. [14] investigated TiO2-nanouids based on alcohol and Naphon et al. [16] based on the refrigerant R11 (trichlorouoromethane). The rst case showed an increase of heat pipe thermal efciency up to a concentration of 0.1 vol. % (FR 50%). For 0.5 vol. % the effect disappeared completely and the efciency was even lower as for the pure basefluid alcohol. With R11 the highest heat pipe efciency was again found for an optimum concentration of 0.1 vol. %. Nearly no increase of the efciency occurred for a concentration of 0.01 vol. % of TiO2 in R11. At an operation temperature of 100  C and a concentration of 1.2 wt. % (H2O/CuO), an increase of about 15% of the evaporator heat transfer coefcient was found by Lu et al. [40] for an open thermosyphon. Lower and higher concentrations showed less improvement. This result was basically conrmed by Liu and Zhu [23] for a horizontal mesh heat pipe. Besides the signicantly increased heat transfer coefcient at the evaporator with increasing concentration, these authors also reported a higher heat transfer coefcient at the condenser. At the evaporator a distinct maximum

was reached at about 1 vol. %. At the condenser the maximum was atter and ranged from concentrations of about 0.8 vol. % to about 1.25 vol. %. For higher concentrations this trend reversed. At a concentration of 2 vol. % the evaporator heat transfer coefcient amounts to only about two thirds of the maximal value. However, in any case the heat transfer coefcients at evaporator and at condenser were higher for the investigated nanouid (H2O/CuO) than for pure water. These general trends were conrmed by Wang et al. [24] for the same wicked heat pipe but with inclination angles ranging between 0 and 90 . Four independent research groups found for H2O/Al2O3 nanouids no optimal concentration or only a limited dependence of the thermal performance on the concentration. The two concentrations (1, 3 vol. %) investigated by Do et al. [18] clearly showed a decrease of thermal resistance. However, while for the very low heat input of 3 W the decrease signicantly depended on the concentration nearly identical results were found for heat inputs ranging between 5 W and 15 W (see Ref. [18] Fig. 6). No optimal point for thermal resistance of a thermosyphon in dependence of the concentration was found by Mousa [37]. However, thermal resistance was nearly constant from 1 vol. % H2O/Al2O3 nanouid onward. Nanoparticle size was 40 nm. In a more recent study Putra et al. [29] found that the temperature was lowered throughout their heat pipe when the Al2O3 concentration of the working uid was increased from 1 vol. % to 5 vol. % without running through a maximum or reaching saturation. Similarly Noie et al. [34] did not nd an efciency optimum depending on concentration in the range from 1 vol. % to 3 vol. % at a xed input power (see Ref. [34] Fig. 3). The fth group dealing with H2O/Al2O3 nanouids in an oscillating heat pipe (Qu et al. [50]) reported that the maximal decrease of thermal resistance occurred at a concentration of 0.9 wt. %. For higher concentrations the thermal resistance increased again. This optimum was related with a comparable high lling ratio of 70%. To summarize, several results show that for water-based Agnanouids an optimum exists for concentrations of about 100 ppm. Optima were also found for TiO2 and CuO nanoparticles dispersed in different basefluids. Surprisingly, four independent studies dealing with H2O/Al2O3 nanouids in thermosyphons and heat pipes could not nd such an optimum in the range of 1 vol. %e5 vol. %. The latter might be caused by different reasons; among them shape effects as they are reported by Ji et al. [52] or that the optimum simply occurs outside the investigated concentration ranges. However, that an optimum was found for an H2O/Al2O3 nanouid in an OHP [50] indicate that nanouids may act differently in different types of gadgets. 5.4. Latent heat and surface tension of nanouids Khandekar et al. [56] compiled characteristics of thermophysical properties which an optimal working uid should have. Among them are low dynamic viscosity, low latent heat, and low surface tension. These demands should also be valid for heat pipes and thermosyphons employing nanouids. However, it is known from several experimental works (e.g. Chevalier et al. [60]) that the dynamic viscosity of the pure liquid is signicantly increased by the addition of nanoparticles. Therefore it seems to be a priori impossible that the demand for low viscosity can be satised by any nanouid. The two other thermophysical properties, especially the latent heat, are rarely addressed by nanouid studies. The only investigation discovered which deals with the latent heat of nanouids is Worek et al. [67]. These results show that the experimentally determined latent heat of H2O/Al2O3 nanouid (46 nm) and H2O/BiO2 are nearly the same as the one of pure water. Worek et al. [67], in analyzing the same nanouids as for the latent heat, found basically no changes of the surface tension

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

compared with pure water. Similarly, Das et al. [68] indicated that the surface tension of their H2O/Al2O3 nanouid had basically not changed for concentrations up to 4 vol. %. No changes up to about 5 vol. % were also found by Buschmann and Franzke [69] for a conventional H2O/TiO2 nanouid. A steep linear decrease of surface tension with increasing concentration was found by Liu and Zhu [23] for water-based CuO-nanouids. At a concentration of about 2 vol. % the surface tension was lowered by 15% compared to the value of pure water. For higher concentrations the surface tension stayed at this value up to the highest investigated concentration of 4 vol. %. An increase of the surface tension was found by Xue et al. [30]. The analyzed suspension with 1 vol. % CNT showed an increase by about 113.5% compared to pure water. Some experimental results are compiled in Fig. 5. The results for the surface tension are very contradictory and do not allow rm conclusions with respect to thermosyphons and heat pipes despite the fact that more research is urgently needed with respect to these thermophysical properties. Other thermophysical properties relevant for heat pipes and thermosyphons, especially density, heat capacity, and the thermal expansion coefcient, can be calculated in a straightforward manner. Thermal conductivity and viscosity of nanouids are still difcult to model and may have to be measured in each individual case. 5.5. Nanouids with magnetic nanoparticles Ferrouids are suspensions of magnetic particles dispersed in a baseuid. The particles may be Fe3O4, cobalt or any magnetically ordered material. The response of ferrouids to magnetic elds depends on the saturation magnetisation (Popplewell et al. [70]). This ability offers additional options to inuence the working uid and may enhance therewith the thermal performance of thermosyphons and heat pipes. Two publications present results for magnetic nanouids or magnetically enhanced heat pipes. Chiang et al. [26] operated a specic heat pipe with a magnetic nanouid consisting of water and Fe3O4-nanoparticles (0.8 vol. %). The essential difference compared with classical heat pipes were the porous iron nozzles and the magnets at evaporator and condenser (for design details see Fig. 1 in Ref. [26]). Both design components inuenced obviously the formation of slugs and vapour bubbles and therewith the complete ow pattern within the heat pipe. Applying a magnetic eld strength with a magnitude of 200 Oe the thermal resistance of the heat pipe operated with the magnetic nanouid was lowered by about 40%. The vector of the magnetic eld strength denes strength and direction of magnetic eld at any spatial point.

The second investigation employing magnetic effects was published by Heris et al. [44]. This group investigated a closed thermosyphon employing a H2O/Ag-nanouid. A variable magnetic ux density (0.12, 0.35, and 1.20 T) was applied to the evaporator section. It was found that with increasing magnitude of the magnetic ux density the thermal resistance of the thermosyphon was decreased. At 0.35 T (1.20 T) a decrease of about 3.22% (5.37%) compared to pure water was observed for a particle concentration of 20 ppm. The effect became the more pronounced the higher the concentration. However, within the experimentally investigated ranges of concentration and magnetic ux density, no optimal combination with respect to the thermal resistance was found. These are two promising results with respect to magnetic nanouids. However, much more research is needed to establish this very special type of liquid as working uid in heat pipes and thermosyphons. 5.6. Nanouids with carbon nanotubes Due to their large length to diameter ratio, carbon nanotubes (CNTs) do not really belong to nanoparticles. However, an increasing number of investigations pay attention to nanouids with CNTs for their extraordinary mechanical, electrical, and thermophysical properties (Xie and Chen [71]). Carbon nanotubes can be single, double, or multi-walled. One of the most interesting features is their astonishing high thermal conductivity along the tube axis which may reach up to 2000 W/(m-K) for a CNT with a diameter of 9.8 nm (Fujii et al. [72]). However, due to the large length of CNTs such nanouids may behave completely different than nanouids with rigid and more or less spherical nanoparticles. Three research groups analyzed the thermal performance of thermosyphons and heat pipes operated with CNT-nanouids. One of them, Xue et al. [30], reported negative results and difculties with the preparation of suspensions with CNTs keeping their high aspect ratio. Only when the primary nanotubes were chemically cut in pieces, whereby some kept their tube structure and others became fragments, could stability at 1 vol. % be established. Due to the signicantly increased surface tension of this specic nanouid, the evaporation process was modied which led to the formation of larger bubbles, diminished departure frequency, and a tendency of bubble-coalescence. All of that resulted in a decreased performance of the investigated thermosyphon. For example, at a heat ow of 555 W the thermal resistance increased by a factor of 3.3 compared with the water lled thermosyphon. A completely contrary result was reported by Liu et al. [36] investigating multiwalled carbon nanotubes (MWCNT) in a thermosyphon. The thermal resistance of the investigated thermosyphon was clearly lowered when employing MWCNT-nanouids. Additionally, these authors showed that the operation range was signicantly larger for MWCNT suspensions (from 20 W to 465 W) than for DI-water (from 20 W to 200 W). The optimal mass concentration found was about 2 wt. %. Depending on heat ux and operation pressure, the evaporator heat transfer coefcient was increased up to 40%. In a very recent work Shanbedi et al. [45] investigated MWCNTs in a thermosyphon. The nanouid was stabilized by functionalizing the nanotubes with ethylenediamine. Differently to Xue et al. [30] the nanotubes kept their cylindrical structure with a diameter ranging from 10 nm to 20 nm. Thermal resistance of the thermosyphon decreased up to a concentration of 1 wt. % and increased again for higher concentration. In any case, thermal resistance decreased with increasing input power. This effect was mainly caused by a temperature decrease at the evaporator while the temperature at the condenser was only weakly changed. Employing CNT-nanouids in thermosyphons and heat pipes seems to be a promising option. However, one of the main

Fig. 5. Experimentally determined surface tension of water-based nanouids. Broken line denotes pure water and dotted lines show a band of 1% of pure water.

10

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

problems with these special nanouids is their proper stabilization. In addition, the mechanisms responsible for the enhanced thermal performance discussed for ordinary nanouids might be different in the case of CNT-nanouids. 6. Effects with respect to thermal performance of gadgets 6.1. Evaporator It seems to be that one of the major physical phenomena responsible for the decrease of the thermal resistance is a porous layer built by nanoparticles on the evaporator surface. Therefore it is clear that evaporator temperature and evaporator heat transfer coefcient take centre stage of many investigations. Indeed, with one exception (Han and Rhi [22]), all studies indicate a decrease of evaporator temperature depending on concentration and heat input. Investigating a conventional grooved (211 mm wide 217 mm deep) cylindrical heat pipe, Kang et al. [13] found that adding 1 ppm Ag-nanoparticles to pure water lowered the temperature at the evaporator by about 1%. Even stronger effects were reported for water-based Al2O3 (Putra et al. [29], Noie et al. [34]) and CuO (Liu et al. [31], Senthilkumar et al. [20], Wang et al. [24]) nanouids. Putra et al. [29] showed a signicant reduction of the evaporator temperature for H2O/Al2O3 nanouid decreasing nearly linearly with concentration (see Ref. [29] Fig. 7). The analytical model by Alizad et al. [66] conrmed this nding by showing a decrease of the temperature difference between evaporator and condenser up to the maximal investigated concentration of 8 vol. %. For H2O/CuO nanouid Liu et al. [31] found that the evaporator heat transfer coefcient is more than doubled for the optimal concentration of 1 wt. %. The surface temperature of the evaporator was lowered by about 13% when H2O/CuO nanouid (100 mg/l) was employed instead of pure water (Senthilkumar et al. [20]). This temperature was again decreased when 2 ml/l of n-Butanol was added to the nanouid. For classical working uids the evaporator temperature corresponds to the applied heat input in an essentially linear manner.3 This might not be the case for nanouids. The investigation by Lu et al. [40] clearly showed that it was not the heat input alone but, rather the combination of heat input and the characteristics of the employed nanouid that matters. Lu et al. [40] investigated an open thermosyphon employing CuO-nanouid. At low input power the temperature distribution along the evaporator tube was nearly unchanged when nanoparticles were added to the baseuid. However, above a certain threshold the evaporator temperature became lower for the nanouid. For example, at a power input of about 1600 W the temperature decreased from 96  C (DI-water) to 90  C for the nanouid with a concentration of 1.2 wt. %. Evaporator heat transfer coefcient depended also strongly on operation temperature and heat input. For low heat ux (up to about 3200 W/m2) the increase of this parameter compared with the values for pure water was negligible. However, for higher heat input the increase was signicant and reached values up to 20% (see Ref. [40] Fig. 10). Similarly, Wang et al. [24], when employing a H2O/CuO nanouid, reported an increase of the evaporator heat transfer coefcient up to 3e4 times depending on operation temperature and heat input. No or only a weak dependence on the inclination angle was found. To summarize, nearly 90% of all experimental results indicate either a signicant decrease of the evaporator temperature or an

increase of the evaporator heat transfer coefcient. This result is in agreement with the analytical model by Alizad et al. [66]. 6.2. Condenser The question if the condenser heat transfer coefcient is changed when nanouids employed is basically equivalent with the question if nanoparticles are torn out of the working uid and transported by the vapour phase or not. Intuitively one may argue that this is not the case due to the much larger particle size compared with the size of the baseuid molecules. For example, the smallest oxide nanoparticles of all experiments compiled in Table 1 have a size of 2 nme3 nm [4] and 4 nme5 nm [38]. This is about 20e50 times larger than a water molecule which has a size of about 0.1 nm. Many experimental results [31,36,43,50] indicated that the condenser heat transfer was either not or only weakly affected by nanouids. Qu et al. [50] analyzed SEM images of the condenser surface of an oscillating heat pipe and found nearly no changes compared to a reference surface boiled in pure water. Therefore they concluded that the thermal resistance of the condenser should not have been changed. Later, the same group (Qu and Wu [53]) found that for equal concentrations the condenser showed no particle deposit for well dispersed Al2O3 particles, but had some precipitate for less dispersed SiO2 particles of similar primary size. Nanoparticle deposits on the condenser were also found by Han and Rhi [22]. However, the structure of this nanoparticle layer was different than the one found on the evaporator. While on the evaporator surface the depositions were planer and of large size, the deposition shape on the condenser surface looked like spherical structures. In this special experiment the thermal performance of the gadget was decreased with dispersed nanoparticles. An interesting result was reported by Riehl and dos Santos [54]. These authors investigated the temperature uctuations at evaporator, condenser, and in the adiabatic part of an open loop pulsating heat pipe employing H2O/CuO nanouid (5 vol. %). For the nanouid the uctuations in all three components were signicantly increased especially at high heat inputs. It was argued that due to the nanoparticles more bubbles were generated, which led to more vapour in the pipe. Therefore it seemed that the condenser as such did not contribute to the improved thermal performance of this pulsating heat pipe. Besides the majority of experiments which indicated that the condenser was more or less not affected by nanouids, a group from Shanghai Jiaotong University (China) showed the opposite. Liu and Zhu [23] analyzed a horizontal mesh heat pipe employing H2O/ CuO nanouid with 50 nm nanoparticle size. The condenser heat transfer coefcient increased signicantly with increasing concentration but reached a maximum at about 1 vol. % and then fell again. Similarly, Wang et al. [24] found that the temperature distribution at the condenser was slightly larger for the H2O/CuO nanouid than for pure water. The condenser heat transfer coefcient of the horizontal case was increased by about 35% at a concentration of 1 wt. %. This was signicantly lower than the enhancement of the evaporator heat transfer coefcient. It should be noted that the two studies [23,24] indicating an increase of the condenser heat transfer coefcient were carried out with horizontal meshed heat pipes. To conclude, most studies indicate that the condenser is not or only weakly affected by nanouids; however, some experiments see also strong effects. No study analyzed here investigated the vapour phase with respect to the transport of nanoparticles directly. Therefore, the question remains open as to if nanoparticles are torn out from the working uid and transported by the vapour phase during heat pipe/thermosyphon operation. This question can be answered only with adequate experiments.

3 The author thanks one of the reviewers for pointing out that fact and therefore elucidating the differences to nanouids.

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

11

6.3. Thermal resistance of whole gadget Thermal resistance is one of the main characteristics of thermosyphons, heat pipes, and oscillating heat pipes. Therefore, most authors try to show that the thermal resistance of their gadget is lowered when exchanging classical working uids with nanouids. Already in an early work, Tsai et al. [10] found that the thermal resistance of a heat pipe was signicantly lowered when employing H2O/Au nanouid. However, the main decrease occurred at the evaporator (between 33% and 56%) while the condenser section showed only a decrease between 7% and 20%. This result was conrmed by Park et al. [11] for H2O/Ag nanouid, who found that the thermal resistance of their heat pipe was slightly decreased at a concentration of 10,000 ppm. Much smaller concentrations (5, 50, and 100 ppm) were investigated by Chen [17]. Here the results indicated that a lowering of the thermal resistance by about 50% was already achieved with 5 ppm. When adding Al2O3 nanoparticles to pure water, a decrease of the thermal resistance of a wicked heat pipe up to 40% was reported by Do et al. [18], and depending on the heat input up to 78% (10 W), 84% (20 W) and 88% (30 W) by Putra et al. [29]. Huminic et al. [38] employed comparably small Fe2O3-nanoparticles (4 nme5 nm) at high concentration (5.3 vol. %) and found that at an inclination angle of 45 the thermal resistance was decreased by about 10% and at 90 by 42% (see Ref. [38] Fig. 9). Similarly, Manimaran et al. [28] found that, when varying lling ratio and inclination angle, the thermal resistance of their heat pipe was in all cases lower for the H2O/CuO nanouid (20 nm) than for the reference uid water. Senthilkumar et al. [20,21] found for a wire mesh wicked heat pipe that the thermal resistance was lower for H2O/CuO nanouid than for pure water. However, with increasing inclination angle and increasing heat supplied this effect disappeared. At about 60 e70 , depending on the amount of supplied heat, the thermal resistance of both uids was equal. These effects were weaker for n-Butanol based nanouid. Surprisingly no differences between the pure aqueous solution of n-Hexanol and the CuO/aqueous solution of n-Hexanol nanouid were found. A dependence of the thermal resistance on the nanoparticle concentration was reported by Liu and Zhu [23]. Thermal resistance increased with increasing concentration up to 1 vol. %. However, for higher concentrations it began to decrease again. The investigated horizontal mesh heat pipe could be run with DI-water up to a heat ux of 44 kW/m2, but failed to work for higher uxes. Employing nanouid (H2O/CuO) the operation range was enlarged signicantly keeping the lower thermal resistance. A similar enlargement of the operation range was reported by Liu et al. [31] for a vertical miniature thermosyphon employing the same nanouid. Here the maximum heat input at the evaporator was increased from 200 W (DI-water) up to 400 W (H2O/CuO). Some values of thermal resistance reduction are compiled in Fig. 6. This plot shows only data which are given by the original authors as numerical values employing vol. % for the concentration. Results given in wt. % or ppm are omitted because it is not possible to convert these values into vol. %. This is due to the fact that even if the baseuid of a nanouid is known, its density is changed in the most cases by adding stabilizers, surfactants or the like to the nanouid [3]. Fig. 6 does not show signicant trends. The reason for the large scatter is basically the manifold nanouid properties. Therefore this plot provides only an initial notion of how thermal resistance depends on nanouid characteristics. Thermal resistance reached e.g. a nearly constant minimal value above a concentration of 1 vol. % in a circular heat pipe (Mousa [25]) and a thermosyphon (Mousa [37]). The achieved reduction depended on the heat imprinted at the evaporator and varied between 50% and 65% (circular heat pipe,

Fig. 6. Experimentally determined reduction of thermal resistance. Full grey curves Mousa [37] 12%. Broken grey curves Mousa [37] extended above 1.2 vol. %.

40 We60 W) and 50% and 68% (thermosyphon, 10 We60 W). Mousa [37] proposed for the percentage reduction the following empirical relation

RR 0:844 0:3042

(1)

Equation (1) has a relative error of 12% according to its originator. It is in reasonable agreement with results from Senthilkumar et al. [20,21] and Manimaran et al. [28]. However, it should not be applied to concentrations higher than 1.2 vol. %, as Fig. 6 clearly indicates. Similar as for all other parameters discussed so far, negative effects on the thermal resistance caused by nanouids were reported. Mehta and Khandekar [32] and Khandekar et al. [33] investigated water-based nanouids (Al2O3, CuO and Laponite) and found a signicantly increased thermal resistance compared to the reference uid water. In a more recent work Han and Rhi [22] investigated a grooved heat pipe (grooves are 1 mm in depth, 1 mm width on top, and 1.3 mm in bottom) employing water-based Ag- and Al2O3-nanouids. In those cases an increase of the thermal resistance was found. The highest thermal resistance is obtained for a hybrid nanouid consisting of Ag (0.1 vol. %) and Al2O3 (0.005 vol. %). The analytical model by Shafahi et al. [73] showed that the thermal resistance was lower the smaller the particles were or the higher the concentration of those particles was. For concentrations up to 4 vol. % no pronounced optimum was found for water-based Al2O3, CuO, and TiO2 nanouids with particle sizes of 10 nm, 20 nm and 40 nm. To summarize, the majority of the studies indicate a decrease of thermal resistance of the whole gadget. However, the larger part of this decrease occurs at the evaporator. Some negative results blur the picture and indicate an increase of the thermal resistance due to nanouids. Additionally, contradictory results with respect to the dependence of the thermal resistance on the imprinted heat at the evaporator are reported (Putra et al. [29], Senthilkumar et al. [20,21]). Obviously the thermal resistance does not only depend on nanouid characteristics alone but also on heat input, inclination angle and lling ratio. Therefore it is difcult to value the inuences of the different nanouids in a straightforward manner. Detailed studies varying the different parameters successively are urgently needed. 6.4. Specics of oscillating heat pipes In 1990 Akachi [75] patented the rst oscillating heat pipe. An oscillating heat pipe consists typically of a serpentine tube of capillary diameter which is evacuated and partly lled with

12

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

working uid. Due to surface tension slugs of liquid intersperse with vapour bubbles [7,8]. Heating an OHP at the evaporator leads to a bubble growth which in turn moves the liquid to the cooler part of the OHP. Heat is transported via the sensible heat of the liquid phase and via the latent heat of the vapour phase. The lling ratio of OHP is differently as for thermosyphons and heat pipes dened as the ratio of the volume of the working uid to the total inner volume of the device. Advantages of OHPs are, among others, the enforcement of heat transfer additionally to the phase change by the oscillatory motion and that liquid and vapour ow do not interfere [46,51]. General information with respect to OHPs employing classical working uids can be found e.g. in Karimi and Culham [76] and Khandekar et al. [77]. Ma et al. [46] were probably the rst who investigated nanouids in an oscillating heat pipe. They found that by adding 1 vol. % diamond nanoparticles to DI-water, the thermal resistance was reduced by more than 40%. However, with increasing heat input this decrease degenerated and reached a value of about 14% at 336 W, the maximal heat input investigated. An increase of the thermal resistance was also observed for a heat input below about 30 W (see Ref. [46] Fig. 3). Cheng et al. [49] investigated a at plate oscillating heat pipe (for design details see Ref. [49] Fig. 1) employing diamond and gold nanouids. When adding 0.1 vol. % diamond nanoparticles to acetone, no differences between the two working uids were found except the power limit of the FP-OHP, which was extended dramatically from about 190 W to 360 W. Only a slight decrease of thermal resistance compared with pure water was found for the 0.0003 vol. % H2O/Au nanouid. When charged with diamond water nanouid of 1.0 vol. %, an increase of the thermal resistance of the FP-OHP was observed. According to the authors of this study, one of the reasons for this behaviour might have been that the comparably high concentration was not stable, settled, and led to clogging. An OHP was also analyzed by Ji et al. [51]. Here Al2O3-particles sizes of 50 nm, 80 nm, 2.2 mm, 20 mm were employed. It was found that particles in general lower the startup temperature. The startup temperature was understood as the temperature when the oscillatory motion in the OHP began. Especially the nanoparticles signicantly reduced this characteristic temperature (pure water: 54.2  C, nanouid 50 nm: 40.6  C). In addition, thermal resistance was reduced signicantly by adding nanoparticles. However, it seems that the size of the nanoparticles (50 nm or 80 nm) did not have a signicant inuence on the startup temperature and the thermal resistance. Riehl and dos Santos [54] investigated an open loop oscillating heat pipe in vertical, horizontal, and top down positions (evaporator above condenser). In all three orientations the evaporation temperature of the nanouid lled OHP was lower than the reference case with DI-water. It was found that the plug/slug formation was more dynamic meaning more liquid was believed to be pumped and kept the evaporator temperature lower than for the working uid DI-water. Additional the authors found that the temperature uctuations at evaporator and condenser were much more pronounced when employing the nanouid. 6.5. Long term tests Despite its importance for practical application of nanouids, their long term stability is still a rarely analyzed subject. Only a few experiments have been noted so far. Among them are the pipe ow experiments by Walsh et al. [78]. These authors found that a fresh H2O/Al2O3 (10 5 nm) nanouid showed a HagenePoiseuille prole in laminar pipe ow indicating Newtonian behaviour. However, repeating the experiments two weeks later, that changed to a clear non-Newtonian behaviour. Obviously the nanouid

underwent an ageing. According to the authors, the most probable explanation for that nding was the formation of aggregates which changed the nanouid from Newtonian to non-Newtonian. However, an interaction with the material of the test rig might also have played a role, as the nanouid might not have improved with an anticorrosive. Liu et al. [31] repeated their thermosyphon experiments also about two weeks after the rst runs. The general nding was that the results had basically not changed. The authors argued that the suspension was refreshed due to the movement of the working uid caused by the boiling process. However, keeping in mind that a practical thermosyphon or heat pipe, for example for cooling an electronic device, would have to work several months to years, these test periods are far too short to draw conclusions from. The buildup of the porous layer is a time dependent process. Therefore ageing and changing over time are natural. White [79] investigated the enhancement of boiling surfaces using nanouid particle deposition. He showed that nanouid boiling performance was increased by about 25% initially before the deposited nanoparticle layer was completely formed. After this process had nished the nanouid boiling coefcient declined again due to suppression of bubble nucleation and transport. These results show clearly that the time dependent character of the porous layer has to be considered in more detail. 7. Modelling and prediction Models can basically be split into two groups. The rst group comprises engineering approaches which are more or less based on semi-empirical correlations. Hereto belong also approaches which order experimental data in a way to create such correlations. Models solving mass, momentum, and energy conservation equations, either analytically or numerically, in combination with evaporation and condensation models build the second group. 7.1. Semi-empirical models An example for the ordering of experimental results to obtain semi-empirical correlations was given by Paramatthanuwat et al. [35]. These authors provided a correlation for the Kutateladze number (ratio of heat ux to critical heat ux) in dependence of ten similarity numbers, among them Prandtl, Grashof, Archimedes, Weber and Froude number. The correlation was in reasonable agreement with their data. Leong et al. [80] developed a model for the prediction of an air to air thermosyphon heat recovery exchanger. The model was based on semi-empirical correlations for the heat transfer coefcients at evaporator and condenser. It took the thermal resistances of the airside, the thermosyphon wall, the working uid at evaporator, and at the condenser into consideration. Thermal conductivity and viscosity of nanouids were considered as temperature and concentration dependent. It was assumed that nanoparticles were transported by vapour and the condensate had the same characteristic as the working uid. As discussed above, it is rather open as to if nanoparticles are transported by the vapour phase. For this reason, the assumption by Leong et al. [80] that the condensate has the same quality as the working uid (mainly effective thermophysical properties caused by nanoparticles) is rather surprising. One of the main results of the predictions was that the change of nanouid thermophysical properties was not of great importance for heat transfer in the analyzed thermosyphon. 7.2. Models based on conservation equations Do and Jang [63] provide two models to study the inuence of effective thermophysical properties and the inuence of the porous

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

13

layer built on the evaporator separately. Both models were based on continuity and momentum equations for vapour and liquid phase. Additionally, the energy equation and the YoungeLaplace equation (capillary pressure difference between liquid and vapour) were solved numerically. The simulations indicated an optimum nanoparticle concentration. The authors argued that the optimum was caused by two competing effects. One was the extension of the evaporator surface due to nanoparticle deposition and the other was the reduced evaporation heat transfer rate due to a higher pressure drop in the porous layer built by the nanoparticles. Moreover, it was shown that the thermal resistance of the heat pipe decreased with increasing nanoparticle size, which was in agreement with the experimental results of Kang et al. [15]. An analytical model for the prediction of the thermal performance of at-shape heat pipes operated by nanouids was provided by Shafahi et al. [74]. A similar model was applied to cylindrical heat pipes [73]. The model was based on descriptions of pressure, velocity, and temperature distributions inside the heat pipe. Pressure distribution was derived by integrating momentum equation and temperature distribution following from energy balances at evaporator and condenser. The velocity distribution in the wicks followed from Darcys law and the mass balance. The YueChoi model for the thermal conductivity of nanouids (Yu and Choi [81]) was modied to account for the porous layer. One of the main ndings of this model was that there exists an optimum for wick thickness and nanoparticle concentration. The rst nding was unique in the sense that none of the experiments address the wick characteristic as an independent parameter. An extension of the Shafahi-Biancoe VafaieManca model [73,74] for the startup characteristics of atshaped heat pipes was proposed by Alizad et al. [66]. Modelling and simulation are a priori condemned to failure if the physical mechanisms causing the experimentally found improvement of the thermal performance of thermosyphons and heat pipes operated with nanouids are not understood. The statement that any model fails if the physics behind an observation are either not known or only poorly understood is a general one. An extraordinary example is the debate of the last decade with respect to wall-bounded turbulence on at plates and in channel and pipe ows (for an overview see Buschmann and Gad-el-Hak [82]). Therefore, all models for thermosyphons and heat pipes mentioned above bear some preliminary assumptions. This does not negate the fact that the majority of the physical, mathematical and numerical components of these models are correct and represent the state of the art. In all cases, the results depend strongly on the assumptions made with respect to thermophysical properties [3]. Sensitivity studies carried out for forced [83] and natural convection [84] for different ceramic/water nanouids conrm this. However, the models available so far capture main features of nanouids in thermosyphons and heat pipes found experimentally. 8. Possible mechanisms In essence, nanouids are two-phase liquids. Their behaviour is in many cases not comparable with classical single-phase liquids like water or acetone. Relevant physical processes occur on micro and macro length scales and on different time scales. In the case of phase change, additionally signicant interactions between free surfaces of apparatus and nanouids have to be considered. The huge number of parameters characterizing nanouids e particle material, size, shape; baseuid characteristics etc. e makes the drawing of rm conclusions even more complicated. Thus, nding explanations for the aforementioned enhancements or deteriorations of heat transfer performance of thermosyphons and heat pipes is extraordinary challenging. Therefore, the explanations given below may appear partly vague and seem not to address the underlying physical

phenomena completely. However, as already stated, there are many open issues with respect to nanouids and the understanding of the underlying physics is partly poor. This may also explain the phenomenon that several authors having found positive results but do not try to explain the physical mechanism behind them. Rather, there are usually the research teams which found deterioration look for clarication. Here mainly discussions which are given by different authors analyzing their experimental data are reected. Only minor valuing comments are made. The explanation attempts for improvements discussed by several authors can be categorized according to the following assumed underlying mechanisms:  nanouids improve thermal performance due to enhanced thermal conductivity  nanouids enlarge operation range due to much higher critical heat ux  bubble departure frequency at evaporator is increased due to bubble bombardment by nanoparticles  nanoparticles coat the wick, which enlarges heat transfer area or capillary forces or both of them  the often found concentration optimum follows from counter acting forces  a porous nanoparticle layer forms on the evaporator, which changes wettability Several authors ascribe improvements to more than one of these reasons. It is only logical to seek explanations for positive effects in a similar way as done for other heat transfer applications employing nanouids, such as pipe ow (Prabhat et al. [4]). For example, Kang et al. [13], Noie et al. [34], and Mousa [25] argued that nanouids can atten temperature gradients due to the increased effective thermal conductivity and therefore the global thermal resistance of thermosyphons and heat pipes should be reduced. However, due to the usually low concentration of nanoparticles utilized and the therefore weak changes of thermophysical properties, large alterations, e.g. of the thermal resistance, cannot be explained that way. You et al. [85] showed that, by immersing a polished copper plate as heater surface in a H2O/Al2O3 nanouid (0e 0.04 g/l), the critical heat ux was increased up to 200% compared with pure water. This effect may explain the increased operation range found by Liu et al. [31], Liu et al. [36], and Liu and Zhu [23]. Likewise, Wei et al. [12] argued that critical heat ux and convective heat transfer coefcient of their nanouids were higher than for the pure baseuid. Therefore it was expected that the thermal performance of a heat pipe with nanouid as working uid was better than employing pure water. Already Tsai et al. [10] argued that a large component of the thermal resistance of their heat pipe was due to the formation of vapour bubbles at the surface of the evaporator. It was then expected that the particles suspended in nanouids bombard the bubbles during formation and, therefore, the nucleation size of the bubbles became much smaller, which in turn should have reduced the global thermal resistance of the heat pipe. Huminic et al. [38] argued similarly that the reduction of the thermal resistance of their thermosyphon was caused by a bombardment of the vapour bubbles during bubble formation. Likewise, Shanbedi et al. [45] used that argument to explain the enhancement of thermal performance by MWCNT in their thermosyphon. Do et al. [18] inspected the surface of the wick after employing H2O/Al2O3 nanouids of 1 vol. % and 3 vol. % utilizing an optical microscope. In both cases the wick was covered with a thin porous coating consisting obviously of nanoparticles. They argued that this coating provided additional heat transfer area for evaporation. A

14

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

coating on the screen mesh surface consisting of nanoparticle material was also found by Putra et al. [29]. However, differently to Do et al. [18], this research team concluded that the coating provided improved capillary effects. Wannapakhe et al. [48] argued that the found optimum at a concentration of 0.5 wt. % was caused by the interplay of thermophysical properties, mainly dynamic viscosity and thermal conductivity. Higher particle concentrations increased viscosity signicantly and therefore obstructed bubble generation and liquid movement in their oscillating heat pipe. Qu et al. [50] employed the theory by Mikic and Rohsenow [86] to explain the found optimum. According to this approach the thermal resistance due to nucleate boiling at the evaporator can be written as

Re

2Nb D2 b

1 q f pknf rnf cp;nf

(2)

Due to the low concentration the product knf rnf cp,nf is comparably less changed compared to the baseuid. It was rather the increase of the active nucleation site density Nb, the bubble release diameter db, and the bubble release frequency f, that decreased the thermal resistance at the evaporator. Qu et al. [50], utilizing twodimensional Atomic Force Microscopy (2d-AFM) images, conrmed that Al2O3-nanoparticles had deposited on the evaporator and modied its surface completely. A nanometre-roughness was formed on top of the original micrometre-roughness. This new surface quality inuenced bubble release frequency and bubble diameter signicantly. The level of particle agglomeration seemed to be increased by concentration, which in turn would enlarge structures of the nanometre-roughness. From that observation it appeared to be clear to the authors that an optimal concentration with respect to the decrease of the thermal resistance of the evaporator should exist. Liu and Zhu [23] investigated a cylindrical wicked heat pipe. While a very thin porous layer consisting mainly of cooper and oxygen, the two components of the CuO-nanoparticles, was found on the inner wall of the evaporator, almost no deposits were found on the mesh surface. An extremely thin porous layer on the inner side of the evaporator after employing nanouid was also found by Lu et al. [40] in their open thermosyphon. Liu and Zhu [23] concluded from their experiments that the improvement mainly followed from the changed thermophysical properties, including decreased surface tension and solideliquid contact angle. Additional reasons for the enhanced heat transfer may have occurred from the increased effective thermal conductivity and the random Brownian motion of the nanoparticles. It should be noted that in this special experiment the thermal performance of the heat pipe was dominated by the convective heat transfer in the liquid lm. Supporting explanations for the experimental ndings came from the two models by Do and Jang [63]. For the test cases with H2O/Al2O3-nanouid, both models indicated a reduction of the evaporator temperature. However, this effect was signicantly weaker without the porous layer (Model I) than with (Model II). The lowering found by Model I was mainly explained by the increased thermal conductivity, which allowed for an improvement of the evaporating heat transfer in the liquid lm. One of the major ndings of these simulations was that the porous layer on the n tops provides an additional evaporation surface which led to an almost doubling of heat transfer. However, the performance optimum was due to competing effects which followed from increasing concentration. As the particle concentration increased, the pressure drop of the growing porous layer became larger and consequently thermal resistance was enforced. Minimal thermal resistance was achieved at nanoparticle concentrations of 0.3 vol. % (40 W) and 0.8 vol. % (80 W).

As already discussed in the foregoing sections, not all experiments found an improvement of the thermal performance of thermosyphons and heat pipes. Han and Rhi [22] indicated a general deterioration of thermal performance of their grooved heat pipe. They analyzed the uctuation (amplitude between low and high peak temperature) at evaporator and at condenser. These uctuations were more intense at the condenser than at the evaporator. The authors argued that this was due to the abrupt release of the heat transported by the nanoparticle in the condenser. The whole effect might have been intensied by particle agglomeration. Experiments with pure water were carried out by Han and Rhi [22] after nanouid experiments. The results indicated that the performance of the heat pipe was the same as with nanouids. It seemed to be that the porous layer established during nanouid experiments remained and acted further in following experiments. The shown SEM photos from different positions of the inner surface of the heat pipe clearly indicated deposits of nanoparticles everywhere. However, the quality of the deposits varied along the pipe. While in the evaporator the shape of the deposition was planer in the condenser, the deposition looked more like spherical particles. Increased thermal uctuations were also observed by Riehl and dos Santos [54] and Riehl [55] in an oscillating heat pipe. Riehl and dos Santos [54] calculated the critical bubble diameter for water 5.45 104 m and for nanouid (H2O/ Cu, 5 wt. %) 5.33 104 m at 20  C. Differently to Han and Rhi [22], the authors argued that the smaller critical bubble diameter directly contributed to the increase of vapour bubble formation and thus the pulsations were more intensive with nanouid. Additionally, it has to be kept in mind that Han and Rhi [22] employed a grooved heat pipe and that therefore the physical reason for the increased thermal uctuations might be different in the two cases. Mehta and Khandekar [32] and Khandekar et al. [33] drew basically two conclusions from the found deterioration of the thermal performance. First, the adverse effect of nanoparticles on pool boiling was stronger than any effect which may have followed from an increased thermal conductivity and, second, the interplay between nanoparticles and heater surface affected pool boiling negatively. Mehta and Khandekar [32] argued that most models describing nucleate boiling heat ux show a power law dependence on nucleation site density na and superheating (Tw Tsat).
y q00 nx a Tw Tsat

(3)

The nucleation site density is affected by physical adhesion of nanoparticles and therewith changing cavity mouth radius and/or cavity mouth angle and by changing the wetting characteristics. A simple model showed that the necessary superheat indeed can increase when the cavity mouth radius is decreased due to nanoparticle deposition. In Khandekar et al. [33] the investigation was extended with respect to the initial surface roughness (Ra 0.24 mm). The surface interaction parameter (ratio of surface roughness to particle size), according to Harish et al. [64], for the CuO particles (8.6 nme13.5 nm) would then be between 17.8 and 27.9 and for the Al2O3 particles (40 nme47 nm) between 5.1 and 6.0. In both cases this would indicate an enhancement of heat transfer (see Ref. [64] Fig. 11). However, as discussed by Khandekar et al. [33], the particles may have undergone agglomeration under the boiling process. This in turn would have lowered the surface interaction parameter signicantly and therewith the expected heat transfer enhancement. 9. Summary Despite their seemingly simple design, thermosyphons, heat pipes, and oscillating heat pipes are complex heat transfer devices.

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17

15

Numerous parameters affect their thermal performance already when classical one-phase liquids are employed as working uids. The number of relevant parameters becomes even larger when nanouids are brought into action. Additional issues occur due to the diminutiveness of the particles and their interaction with the surfaces of these gadgets. Not all phenomena, such as the effective thermophysical properties, especially viscosity, surface tension, and thermal conductivity and a signicantly increased critical heat ux, are still explainable using the trusted continuum approaches. The buildup of porous layers forced by boiling processes reaches far beyond uid mechanic and thermodynamic knowledge. This complexity explains why the experimental and modelling work carried out so far is partly contradictory and does not answer all physical and technical questions relevant for the proper design of thermosyphons, heat pipes, and oscillating heat pipes operated with nanouids. However, based on the examination of 38 experimental studies and 4 modelling approaches, the following conclusions are drawn.  The majority of the studies indicate that lling ratio, inclination angle, and operation temperature affect thermosyphons and heat pipes operated with nanouids in the same or at least in a similar way as classical uids do. However, a few experiments also indicate specic trends for nanouids.  Only three, but very clear, experiments indicate the inuence of the baseuid on the thermal resistance.  It is highly likely that nanoparticle size, shape, and material and chemical stabilization have an inuence on the thermal performance. However, these inuences seem to be different than for heat transfer applications without phase changes. Especially the interaction between evaporator surface and wick surface on one side and nanoparticles on the other side has to be take into consideration.  Optima of the thermal performance of thermosyphons and heat pipes in dependence on the concentration were found for Ag, TiO2 and CuO nanoparticles dispersed in different basefluids. Surprisingly, no optimum was found in thermosyphons and heat pipes by several independent research groups dealing with H2O/Al2O3 nanouids.  Magnetic nanouids and nanouids with carbon nanotubes are promising options because they allow the use of additional effects like the application of additional external forces or the extraordinary high thermal conductivity along the main axis of the CNT. However, they are still far from practical use because the number of studies is far too less to create reliable knowledge with respect to the behaviour of these specic liquids in thermosyphons and heat pipes.  Nearly 90% of all experimental results indicate either a significant decrease of the evaporator temperature or an increase of the evaporator heat transfer coefcient.  The majority of the studies indicate a decrease of thermal resistance of the whole gadget. The larger part of this decrease occurs at the evaporator. However, because thermal resistance does not only depend on nanouid characteristics alone, but also on heat input, inclination angle and lling ratio it is difcult to value the inuences of the different nanouids in a straightforward manner.  Most studies indicate that the condenser is not or only weakly affected by nanouids. However, some experiments see also strong effects. The question if nanoparticles are torn out from the working uid and transported by the vapour phase is still open.  Several experimental studies show that working uids with nanoparticles allow an operation at higher input powers without dryout.

 Despite its strong importance for practical applications, long term stability of nanouids and long term operation behaviour of gadgets operated with nanouids is not investigated in detail so far.  First modelling approaches available cover main features of nanouids in thermosyphons and heat pipes found experimentally. However, as long as some physical mechanisms are not understood more detailed these models will lack generality. The buildup of porous nanoparticle layers on the evaporator surface or on the wick surface seems to be one of the key features related with nanouids employed in thermosyphons and heat pipes. Interestingly enough, different research teams tried to explain the found improvements, as well as the found deteriorations, utilizing these nanoparticle layers. The general effects related to such layers which signicantly improve surface wettability, as shown by a reduction of the static contact angle, were investigated by Kim et al. [87]. The reduction of the static contact angle in turn is due to the changed surface energy and surface morphology caused by the nanoparticle layer. However, these experiments were not carried out in the more complex thermosyphons and heat pipes. Here different effects may act simultaneously or even interfere. There are several open questions with respect to the porous layer formed from nanoparticles. Among them is the question if these layers could be built without nanouids and if such articial layers would be sustainable without renewal by nanoparticles from a nanouid employed as working liquid. Solomon et al. [88] coated a wick with nanoparticles by simply immersing the wick into a nanouid and drying it afterwards. Employing pure water as working uid they found a decrease of the thermal resistance up to 19%. Here much more detailed experiments are needed. To summarize nanouids are a new and promising option as working uids in thermosyphons, heat pipes, and oscillating heat pipes. The majority of the experimental works carried out so far indicates a lowering of the thermal resistance and therewith a signicant improvement of the thermal performance. Several general questions with respect to gadget parameters seem to be answered. However, many questions with respect to nanoparticle characteristics, optimization, and acting physical mechanisms are still open. Despite promising modelling approaches existing, the limited knowledge with respect to the dependencies of the thermal performance on nanouid characteristics prevents the development of reliable design tools for practical application. In any case, more detailed experiments which address specic questions of nanouids employed as working uids in thermosyphons, heat pipes, and oscillating heat pipes are needed. Acknowledgement This work was supported by BMWi under the grants MF 110030 and MF 120019. References
[1] A. Sergis, Y. Hardalupas, Anomalous heat transfer modes of nanouids: a review based on statistical analysis, Nanoscale Res. Lett. 6 (2011) 391. http:// dx.doi.org/10.1186/1556-276X-6-391. [2] S. Thomas, C.B.P. Sobhan, A review of experimental investigations on thermal phenomena in nanouids, Nanoscale Res. Lett. 6 (2011) 377. http://dx.doi.org/ 10.1186/1556-276X-6-377. [3] S. Feja, M.H. Buschmann, Nanouids e potentials illusions, Adv. Comput. Methods Exp. Heat Transf. XII (2012) 195e205. [4] N. Prabhat, J. Buongiorno, L.W. Hu, Convective heat transfer enhancement in nanouids: real anomaly or analysis artifacts?, in: Proceedings ASME/JSME 2011 8th AJTEC, USA, 2011.

16

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17 [35] T. Paramatthanuwat, S. Rittidech, A. Pattiya, A correlation to predict heattransfer rates of a two-phase closed thermosyphon (TPCT) using silver nanouid at normal operating conditions, Int. J. Heat Mass Transf. 53 (2010) 4960e4965. http://dx.doi.org/10.1016/j.ijheatmasstransfer.2010.05.046. [36] Z.H. Liu, X.F. Yang, G.S. Wang, G.L. Guo, Inuence of carbon nanotube suspension on the thermal performance of a miniature thermosyphon, Int. J. Heat Mass Transf. 53 (2010) 1914e1920. http://dx.doi.org/10.1016/j.ijheatmasstransfer.2009.12.065. [37] M.G. Mousa, Thermal performance of thermosyphon charged by nanouid for cooling electronic component, Int. J. Sci. Eng. Res. 2 (2011b) 1e9. [38] G. Huminic, A. Huminic, I. Morjan, F. Dumitrache, Experimental study of the thermal performance of thermosyphon heat pipe using iron oxide nanoparticles, Int. J. Heat Mass Transf. 54 (2011) 656e661. http://dx.doi.org/ 10.1016/j.ijheatmasstransfer.2010.09.005. [39] G. Huminic, A. Huminic, Heat transfer characteristics of a two-phase closed thermosyphons using nanouids, Expt. Therm. Fluid Sci. 35 (2011) 550e557. http://dx.doi.org/10.1016/j.expthermusci.2010.12.009. [40] L. Lu, Z.H. Liu, H.S. Xiao, Thermal performance of an open thermosyphon using nanouids for high-temperature evacuated tubular solar collectors Part 1: indoor experiment, Solar Energy 85 (2011) 379e387. http://dx.doi.org/ 10.1016/j.solener.2010.11.008. [41] T. Paramatthanuwat, S. Boothaisong, S. Rittidech, K. Booddachan, Heat transfer characteristics of a two-phase closed thermosyphon using deionized water mixed with silver nano, Heat Mass Transf. 46 (2009) 281e285. http:// dx.doi.org/10.1007/s00231-009-0565-y. [42] T. Paramatthanuwat, S. Rittidech, A. Pattiya, Y. Ding, S. Witharana, Application of silver nanouid containing oleic acid surfactant in a thermosyphon economizer, Nanoscale Res. Lett. 6 (2011) 315. http://dx.doi.org/10.1186/1556276X-6-315. [43] X.F. Yang, Z.H. Liu, Application of functionalized nanouid in thermosyphon, Nanoscale Res. Lett. 6 (2011) 494. http://dx.doi.org/10.1186/1556276X-6-494. [44] S.Z. Heris, H. Salehi, S.H. Noie, The effect of magnetic eld nanouid on thermal performance of two-phase closed thermosyphon (TPCT), Int. J. Phys. Sci. 7 (2012) 534e543. http://dx.doi.org/10.5897/IJPS11.1019. [45] M. Shanbedi, S.Z. Heris, M. Baniadam, A. Amiri, M. Maghrebi, Investigation of heat-transfer characterization of EDA-MWCNT/DI-Water nanouid in a twophase closed thermosyphon, Ind. Eng. Chem. Res. 51 (2012) 1423e1428. http://dx.doi.org/10.1021/ie202110g. [46] H.B. Ma, C. Wilson, B. Borgmeyer, K. Park, Q. Yu, S.U.S. Choi, M. Tirumala, Effect of nanouid on the heat transport capability in an oscillating heat pipe, Appl. Phys. Lett. 88 (2006) 143116. http://dx.doi.org/10.1063/1.2192971. [47] Y.H. Lin, S.W. Kang, H.L. Chen, Effect of silver nano-uid on pulsating heat pipe thermal performance, Appl. Therm. Eng. 28 (2008) 1312e1317. http:// dx.doi.org/10.1016/j.applthermaleng.2007.10.019. [48] S. Wannapakhe, S. Rittidech, B. Bubphachot, O. Watanabe, Heat transfer rate of a closed-loop oscillating heat pipe with check valves using silver nanouid as working uid, J. Mech. Sci. Technol. 23 (2009) 1576e1582. http://dx.doi.org/ 10.1007/s12206-009-0424-2. [49] P. Cheng, S. Thompson, J. Boswell, H.B. Ma, An investigation of at-plate oscillating heat pipes, J. Electron. Packag. 132 (2010) 041009. http:// dx.doi.org/10.1115/1.4002726. [50] J. Qu, H.Y. Wu, P. Cheng, Thermal performance of an oscillating heat pipe with Al2O3ewater nanouids, Int. Comm. Heat Mass Transf. 37 (2010) 111e115. http://dx.doi.org/10.1016/j.icheatmasstransfer.2009.10.001. [51] Y. Ji, H. Ma, F. Su, G. Wang, Particle size effect on heat transfer performance in an oscillating heat pipe, Exp. Therm. Fluid Sci. 35 (2011) 724e727. http:// dx.doi.org/10.1016/j.expthermusci.2011.01.007. [52] Y. Ji, C. Wilson, H. Chen, H. Ma, Particle shape effect on heat transfer performance in an oscillating heat pipe, Nanoscale Res. Lett. 6 (2011) 296. http:// dx.doi.org/10.1186/1556-276X-6-296. [53] J. Qu, H. Wu, Thermal performance comparison of oscillating heat pipes with SiO2/water and Al2O3/water nanouids, Int. J. Therm. Sci. 50 (2011) 1954e 1962. http://dx.doi.org/10.1016/j.ijthermalsci.2011.04.004. [54] R.R. Riehl, N. dos Santos, Water-copper nanouid application in an open loop pulsating heat pipe, Appl. Therm. Eng. 42 (2012) 6e10. http://dx.doi.org/ 10.1016/j.applthermaleng.2011.01.017. [55] R.R. Riehl, private communication, 2012. [56] S. Khandekar, N. Dollinger, M. Groll, Understanding operational regimes of closed loop pulsating heat pipes: an experimental study, Appl. Therm. Eng. 23 (2003) 707e719. http://dx.doi.org/10.1016/S1359-4311(02)00237-5. [57] A.D. Patil, R.B. Yarasu, Factors affecting the thermal performance of two phase closed thermosyphon: a review, Int. J. Emerging Technol. Adv. Eng. 2 (2012) 202e206. [58] J. Buongiorno, D.C. Venerus, N. Prabhat, T. McKrell, J. Townsend, R. Christianson, et al., A benchmark study on the thermal conductivity of nanouids, J. Appl. Phys. 106 (2009) 094312. http://dx.doi.org/10.1063/ 1.3245330. [59] A. Ehle, S. Feja, M.H. Buschmann, Temperature dependency of ceramic nanouids shows classical behavior, J. Thermophys. Heat Transf. 25 (2011) 378e385. http://dx.doi.org/10.2514/1.T3634. [60] J. Chevalier, O. Tillement, F. Ayela, Structure rheology of SiO2 nanoparticle suspensions under very high shear rates, Phys. Rev. E 80 (2009) 051403. http://dx.doi.org/10.1103/PhysRevE.80.051403. [61] M. Corcione, Empirical correlating equations for predicting the effective thermal conductivity dynamic viscosity of nanouids, Energy Conversion

[5] M.H. Buschmann, Thermal conductivity heat transfer of ceramic nanouids, Int. J. Therm. Sci. 62 (2012) 19e28. http://dx.doi.org/10.1016/j.ijthermalsci.2011.09.019. [6] Z.H. Liu, Y.Y. Li, A new frontier of nanouid research e application of nanouids in heat pipes, Int. J. Heat Mass Transf. 55 (2012) 6786e6797. http:// dx.doi.org/10.1016/-j.ijheatmasstransfer.2012.06.086. [7] D.A. Reay, P.A. Kew, Heat Pipes, fth ed., Elsevier, 2006. [8] A. Faghri, Heat Pipe Science and Technology, Mechanical Engineering, Taylor & Francis Group, 1995. [9] H.H. Fernholz, P.J. Finley, The incompressible zero-pressure-gradient turbulent boundary layer: an assessment of the data, Prog. Aerospace Sci. 32 (1996) 6245e6311. [10] C.Y. Tsai, H.T. Chien, P.P. Ding, B. Chan, T.Y. Luh, P.H. Chen, Effect of structural character of gold nanoparticles in nanouid on heat pipe thermal performance, Mater. Lett. 58 (2004) 1461e1465. http://dx.doi.org/10.1016/ j.matlet.2003.10.009. [11] K.H. Park, W.H. Lee, K.W.K.W. Lee, I.H. Baek, S.H. Rhi, D.R. Shin, Study on the operating characteristics in small size heat pipe using nanouids, in: Proceedings 3rd IASME/WSEAS Int. Conf. on Heat Transfer, Thermal Eng. Environment (2005), pp. 106e109. [12] W.C. Wei, S.H. Tsai, S.Y. Yang, S.W. Kang, Effect of nanouid concentration on heat pipe thermal performance, IASME Trans. (2005) 1432e1439. [13] S.W. Kang, W.C. Wei, S.H. Tsai, S.Y. Yang, Experimental investigation of silver nano-uid on heat pipe thermal performance, Appl. Therm. Eng. 26 (2006) 2377e2382. http://dx.doi.org/10.1016/j.applthermaleng.2006.02.020. [14] P. Naphon, P. Assadamongkol, T. Borirak, Experimental investigation of titanium nanouids on the heat pipe thermal efciency, Int. Commun. Heat Mass Transf. 35 (2008) 1316e1319. http://dx.doi.org/10.1016/j.icheatmasstransfer.2008.07.010. [15] S.W. Kang, W.C. Wei, S.H. Tsai, C.C. Huang, Experimental investigation of nanouids on sintered heat pipe thermal performance, Appl. Therm. Eng. 29 (2009) 973e979. http://dx.doi.org/10.1016/j.applthermaleng.2008.05.010. [16] P. Naphon, D. Thongkum, P. Assadamongkol, Heat pipe efciency enhancement with refrigerant-nanoparticles mixtures, Energy Conversion Manage. 50 (2009) 772e776. http://dx.doi.org/10.1016/j.enconman.2008.09.045. [17] Y.T. Chen, Experimental study of silver nanouid on at heat pipe thermal performance, J. Marine Sci. Technol. 18 (2010) 731e734. [18] K.H. Do, H.J. Ha, S.P. Jang, Thermal resistance of screen mesh wick heat pipes using the water-based Al2O3 nanouids, Int. J. Heat Mass Transf. 53 (2010) 5888e5894. http://dx.doi.org/10.1016/j.ijheatmasstransfer.2010.07.050. [19] T.P. Teng, H.G. Hsua, H.E. Mob, C.C. Chenc, Thermal efciency of heat pipe with alumina nanouid, J. Alloys Compd. 504S (2010) S380eS384. http:// dx.doi.org/10.1016/j.jallcom.2010.02.046. [20] R. Senthilkumar, S. Vaidyanathan, B. Sivaraman, Performance analysis of heat pipe using copper nanouid with aqueous solution of n-Butanol, Int. J. Mech. Mater. Eng. 4 (2010) 251e256. [21] R. Senthilkumar, S. Vaidyanathan, B. Sivaraman, Experimental analysis of cylindrical heat pipe using copper nanouid with an aqueous solution of nHexanol, Front. Heat Pipes 2 (2011) 033004. http://dx.doi.org/10.5098/ fhp.v2.3.3004. [22] W.S. Han, S.H. Rhi, Thermal characteristics of grooved heat pipe with hybrid nanouids, Therm. Sci. 15 (2011) 195e206. http://dx.doi.org/10.2298/ TSCI100209056H. [23] Z.H. Liu, Q.Z. Zhu, Application of aqueous nanouids in a horizontal mesh heat pipe, Energy Conversion Manage. 52 (2011) 292e300. http://dx.doi.org/ 10.1016/j.enconman.2010.07.001. [24] P.Y. Wang, X.J. Chen, Z.H. Liu, Y.P. Liu, Application of nanouid in an inclined mesh wicked heat pipes, Thermoch. Acta 539 (2012) 100e108. http:// dx.doi.org/10.1016/j.tca.2012.04.011. [25] M.G. Mousa, Effect of nanouid concentration on the performance of circular heat pipe, Int. J. Sci. Eng. Res. 2 (2011) 1e8. [26] Y.C. Chiang, J.J. Chieh, C.C. Ho, The magnetic-nanouids heat pipe with superior thermal properties through magnetic enhancement, Nanoscale Res. Lett. 7 (2012) 322. http://dx.doi.org/10.1186/1556-276X-7-322. [27] R. Hajian, M. Layeghi, K.A. Sani, Experimental study of nanouid effects on the thermal performance with response time of heat pipe, Energy Conversion Manage. 5 (2012) 63e68. http://dx.doi.org/10.1016/j.enconman.2011.11.010. [28] R. Manimaran, K.N. Palaniradja, N. Alagumurthi, Effect of lling ratio on thermal characteristics of wire-mesh heat pipe using copper oxide nanouids, Front. Heat Pipes 3 (2012) 023004. http://dx.doi.org/10.5098/fhp.v3.2.3004. [29] N. Putra, W.N. Septiadi, H. Rahman, R. Irwansyah, Thermal performance of screen mesh wick heat pipes with nanouids, Exp. Therm. Fluid Sci. 40 (2012) 10e17. http://dx.doi.org/10.1016/j.expthermusci.2012.01.007. [30] H. Xue, J. Fan, Y. Hu, R. Hong, K. Cen, The interface effect of carbon nanotube suspension on the thermal performance of a two-phase closed thermosyphon, J. Appl. Phys. 100 (2006) 104909. http://dx.doi.org/10.1063/1.2357705. [31] Z.H. Liu, J.G. Xiong, R. Bao, Boiling heat transfer characteristics of nanouids in a at heat pipe evaporator with micro-grooved heating surface, Int. J. Multiphase Flow 33 (2007) 1284e1295. http://dx.doi.org/10.1016/j.ijmultiphaseow.2007.06.009. [32] B. Mehta, S. Khandekar, Two-phase closed thermosyphon with nanouids, in: 14th International Heat Pipe Conference (2007), pp. 22e27. [33] S. Khandekar, Y. Joshi, B. Mehta, Thermal performance of closed two-phase thermosyphon using nanouids, Int. J. Therm. Sci. 47 (2008) 659e667. http://dx.doi.org/10.1016/j.ijthermalsci.2007.06.005. [34] S.H. Noie, S.Z. Heris, M. Kahani, S.M. Nowee, Heat transfer enhancement using Al2O3/water nanouid in a two-phase closed thermosyphon, Int. J. Heat Fluid Flow 30 (2009) 700e709. http://dx.doi.org/10.1016/j.ijheatuidow.2009.03.001.

M.H. Buschmann / International Journal of Thermal Sciences 72 (2013) 1e17 Manage. 52 (2011) 789e793. http://dx.doi.org/10.1016/j.enconman.2010. 06.072. R. Prasher, D. Song, J. Wang, Measurements of nanouid viscosity and its implications for thermal application, Appl. Phys. Lett. 89 (2006) 133108. http://dx.doi.org/10.1063/1.2356113. K.H. Do, S.P. Jang, Effect of nanouids on the thermal performance of a at micro heat pipe with a rectangular grooved wick, Int. J. Heat Mass Transf. 53 (2010) 2183e2192. http://dx.doi.org/10.1016/j.ijheatmasstransfer.2010.07.050. G. Harish, V. Emlin, V. Sajith, Effect of surface particle interactions during pool boiling of nanouids, Int. J. Therm. Sci. 50 (2011) 2318e2327. http:// dx.doi.org/10.1016/j.ijthermalsci.-2011.06.019. E.V. Timofeeva, J.L. Routbort, D. Singh, Particle shape effects on thermophysical properties of alumina nanouids, J. Appl. Phys. 106 (2009) 014304. http:// dx.doi.org/10.1063/1.3155999. K. Alizad, K. Vafai, M. Shafahi, Thermal performance operational attributes of the startup characteristics of at-shaped heat pipes using nanouids, Int. J. Heat Mass Transf. 55 (2012) 140e155. http://dx.doi.org/10.1016/ j.ijheatmasstransfer.2011.08.050. W.M. Worek, M.N. Golubovic, H.D.M. Hettiarachchi, Nanouids critical heat ux, an experimental analytical study, in: 10th UK National Heat Transfer Conference (2007). S.K. Das, N. Putra, W. Roetzel, Pool boiling characteristics of nano-uids, Int. J. Heat Mass Transf. 46 (2003) 851e862. M.H. Buschmann, U. Franzke, Improvement of thermosyphon performance by employing nanouid, in: 4th IRR Conference, Paper no. TP-014. The Netherland (Delft), June 16e20 (2013). J. Popplewell, A. Al-Qenaie, S.W. Charles, R. Moskowitz, K. Raj, Thermal conductivity measurements on ferrouids, Colloid Polym. Sci. 260 (1982) 333e338. H. Xie, L. Chen, Review on the preparation thermal performances of carbon nanotube contained nanouids, J. Chem. Eng. Data 56 (2011) 1030e1041. http://dx.doi.org/10.1021/je101026j. M. Fujii, X. Zhang, H. Xie, H. Ago, K. Takahashi, T. Ikuta, H. Abe, T. Suimizu, Measuring the thermal conductivity of a single carbon nanotube, Phys. Rev. Lett. 95 (2005) 065502. http://dx.doi.org/10.1103/PhysRevLett.95.065502. M. Shafahi, V. Bianco, K. Vafai, O. Manca, An investigation of the thermal performance of cylindrical heat pipes using nanouids, Int. J. Heat Mass Transf. 53 (2010) 376e383. http://dx.doi.org/10.1016/j.ijheatmasstransfer.2009.09.019. M. Shafahi, V. Bianco, K. Vafai, O. Manca, Thermal performance of at-shaped heat pipes using nanouids, Int. J. Heat Mass Transf. 53 (2010) 1438e1445. http://dx.doi.org/10.1016/j.ijheatmasstransfer.2009.12.007.

17

[62]

[63]

[64]

[65]

[66]

[67]

[68] [69]

[70]

[71]

[72]

[73]

[74]

[75] H. Akachi, U.S. Patent No. 4, 921, 041, May 1990. [76] G. Karimi, J.R. Culham, Review assessment of pulsating heat pipe mechanism for high heat ux electronic cooling, in: Int. Society Conference on Thermal Phenomena (2004). [77] S. Khandekar, P.K. Panigrahi, F. Lefvre, J. Bonjour, Local hydrodynamics of ow in a pulsating heat pipe: a review, Front. Heat Pipes 1 (2010) 023003. http://dx.doi.org/10.5098/fhp.v1.2.3003. [78] P.A. Walsh, V.M. Egan, E.J. Walsh, Novel micro-PIV study enables a greater understanding of nanoparticle suspension ows: nanouids, Microuid Nanouid 8 (2009) 837e842. http://dx.doi.org/10.1007/s10404-009-0553-z. [79] S.B. White, Enhancement of Boiling Surfaces Using Nanouid Particle Deposition. PhD-Thesis, The University of Michigan, 2010. [80] K.Y. Leong, R. Saidur, T.M.I. Mahlia, Y.H. Yau, Performance investigation of nanouids as working uid in a thermosyphon air preheater, Int. Commun. Heat Mass Transf. 39 (2012) 523e529. http://dx.doi.org/10.1016/ j.icheatmasstransfer.2012.01.014. [81] W. Yu, S.U.S. Choi, The role of interfacial layers in the enhanced thermal conductivity of nanouids: a renovated Maxwell model, J. Nanoparticle Res. 5 (2003) 167e171. [82] M.H. Buschmann, M. Gad-el-Hak, Recent developments in scaling of wallbounded ows, Prog. Aerospace Sci. 42 (2007) 419e467. http://dx.doi.org/ 10.1016/j.paerosci.2007.01.001. [83] Y. He, Y. Mena, Y. Zhao, H. Lu, Y. Ding, Numerical investigation into the convective heat transfer of TiO2 nanouids owing through a straight tube under the laminar ow conditions, Appl. Therm. Eng. 29 (2009) 1965e1972. http://dx.doi.org/10.1016/j.applthermaleng.2008.09.020. [84] E. Abu-Nada, Effects of variable viscosity thermal conductivity of CuO-water nanouid on heat transfer enhancement in natural convection: mathematical model simulation, J. Heat Transf. 132 (2010) 052401. http://dx.doi.org/ 10.1115/1.4000440. [85] S.M. You, J.H. Kim, K.H. Kim, Effect of nanoparticles on critical heat ux of water in pool boiling heat transfer, Appl. Phys. Lett. 83 (2003) 3374e3376. http://dx.doi.org/10.1063/1.1619206. [86] B.B. Mikic, W.M. Rohsenow, A new correlation of pool-boiling data including the effect of heating surface characteristics, J. Heat Transf. 12 (1969) 245e250. [87] S.J. Kim, I.C. Bang, J. Buongiorno, L.W. Hu, Effects of nanoparticle deposition on surface wettability inuencing boiling heat transfer in nanouids, Appl. Phys. Lett. 89 (2006) 153107. http://dx.doi.org/10.1063/1.2360892. [88] B. Solomon, K. Ramachandran, B.C. Pillai, Thermal performance of a heat pipe with nanoparticles coated wick, Appl. Therm. Eng. 36 (2012) 106e112. http:// dx.doi.org/10.1016/j.applthermaleng.2011.12.004.

Anda mungkin juga menyukai