Author's personal copy - University of Michigan

12
Author's personal copy Pool boiling experiments in reduced graphene oxide colloids. Part I – Boiling characteristics Ho Seon Ahn a , Ji Min Kim b , Massoud Kaviany c,d , Moo Hwan Kim c,a Division of Mechanical System Engineering, Incheon National University, Incheon, Republic of Korea b Department of Mechanical Engineering, POSTECH, Pohang, Republic of Korea c Division of Advanced Nuclear Engineering, POSTECH, Pohang, Republic of Korea d Department of Mechanical Engineering, University of Michigan, Ann Arbor, USA article info Article history: Received 5 February 2013 Received in revised form 31 October 2013 Accepted 9 January 2014 Available online 27 January 2014 Keywords: Pool boiling Critical heat flux Boiling heat transfer Graphene abstract Nucleate pool boiling characteristics were investigated in a reduced graphene oxide (RGO) colloid with a water base. Critical heat flux (CHF) pool boiling experiments were performed to analyze both the boiling heat transfer and the CHF on a silicon dioxide plate in colloids with concentrations of 0.0001, 0.0005, and 0.0010 wt.%. The RGO colloid was prepared by chemical reduction of hydrazine from graphene oxide, and diluted for the experiments. In the boiling experiments, boiling heat transfer and CHF enhancement were shown to be related to the concentration. Special attention was focused on two phenomena: the enhanced CHF and improved boiling heat transfer in the RGO colloid. Possible mechanisms for these phe- nomena were examined from various viewpoints, including the RGO coating layer characteristics, wetta- bility, water absorption by the RGO porous medium, and history of the RGO coating layer development. Ó 2014 Elsevier Ltd. All rights reserved. 1. Introduction Phase-change heat transfer is highly efficient, due to the large latent heat of the working fluids. Representative applications of boiling heat transfer (liquid–vapor phase change) include those found in nuclear power plant energy transfer [1], electronic chip cooling [2], and various chemical processes [3]. During boiling heat transfer, the two critical issues are performance (boiling heat transfer coefficient) and critical heat flux (CHF), which is the limit of boiling heat transfer in a thermal system such as a nuclear power plant. Thus, thermal systems of this type require enhanced boiling heat transfer and increased CHF. Over the past century, numerous techniques for providing high boiling heat transfer and CHF have been developed. Most recently, the effects of nanofluids on boiling heat transfer have been extensively studied; it has been shown that nanoparticles on a heated surface cause a CHF enhancement while boiling a nanofluid by changing the wettability characteristics of the surface [4]. Das et al. [5] reported degraded boiling heat transfer in a nanofluid, because the nanoparticles filled the cavity of the heater surface. Bang and Chang [6] found that the CHF during alumina nanofluid pool boiling increased, but the boiling heat transfer decreased. They observed a roughness change on the heater surface, and hypothesized that the reason for the variation in CHF performance was due to the nanoparticle surface coating on the heater. Kim et al. [7] performed pool boiling exper- iments with pure water on nanoparticle-deposited surfaces that were produced while pool boiling water–TiO 2 and water–Al 2 O 3 nanofluids. The CHF of the pure water boiling on the nanoparti- cle-deposited surface was greater than the nanofluid CHF in all cases. This clearly indicates that the nanoparticles deposited on the heating surface were the main cause of CHF enhancement. Kim et al. [8] also conducted pool boiling experiments with nano- fluids, and described the CHF enhancement in terms of surface wettability and absorption of the nanoparticle deposition layer. They suggested a simple model, consisting of an absorption layer of nanoparticles, to describe how the nanoparticles changed the surface wettability. Using results obtained with a thin-wire heater, Kim and Kim [9] postulated that the wettability and capillarity of the heating surface could influence the CHF enhancement. They ex- plained that the CHF increases the liquid supply on the heating sur- face, due to the capillarity of the absorption layer of nanoparticles. Recent general conclusions about the effects of nanofluids on boiling heat transfer and CHF are: The CHF increases dramatically during nanofluid boiling because of the nanoparticle coating on the heater during nucle- ate boiling [6–13]. The boiling heat transfer decreases during nanofluid boiling because the nanoparticle coating reduces the active nucleation sites [5,6]. 0017-9310/$ - see front matter Ó 2014 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.01.022 Corresponding author. Address: Korea Institute of Nuclear Safety, Daejeon, South Korea. E-mail address: [email protected] (M.H. Kim). International Journal of Heat and Mass Transfer 74 (2014) 501–512 Contents lists available at ScienceDirect International Journal of Heat and Mass Transfer journal homepage: www.elsevier.com/locate/ijhmt

Transcript of Author's personal copy - University of Michigan

Author's personal copy

Pool boiling experiments in reduced graphene oxidecolloids. Part I – Boiling characteristics

Ho Seon Ahn a, Ji Min Kim b, Massoud Kaviany c,d, Moo Hwan Kim c,⇑a Division of Mechanical System Engineering, Incheon National University, Incheon, Republic of Koreab Department of Mechanical Engineering, POSTECH, Pohang, Republic of Koreac Division of Advanced Nuclear Engineering, POSTECH, Pohang, Republic of Koread Department of Mechanical Engineering, University of Michigan, Ann Arbor, USA

a r t i c l e i n f o

Article history:Received 5 February 2013Received in revised form 31 October 2013Accepted 9 January 2014Available online 27 January 2014

Keywords:Pool boilingCritical heat fluxBoiling heat transferGraphene

a b s t r a c t

Nucleate pool boiling characteristics were investigated in a reduced graphene oxide (RGO) colloid with awater base. Critical heat flux (CHF) pool boiling experiments were performed to analyze both the boilingheat transfer and the CHF on a silicon dioxide plate in colloids with concentrations of 0.0001, 0.0005, and0.0010 wt.%. The RGO colloid was prepared by chemical reduction of hydrazine from graphene oxide, anddiluted for the experiments. In the boiling experiments, boiling heat transfer and CHF enhancement wereshown to be related to the concentration. Special attention was focused on two phenomena: theenhanced CHF and improved boiling heat transfer in the RGO colloid. Possible mechanisms for these phe-nomena were examined from various viewpoints, including the RGO coating layer characteristics, wetta-bility, water absorption by the RGO porous medium, and history of the RGO coating layer development.

� 2014 Elsevier Ltd. All rights reserved.

1. Introduction

Phase-change heat transfer is highly efficient, due to the largelatent heat of the working fluids. Representative applications ofboiling heat transfer (liquid–vapor phase change) include thosefound in nuclear power plant energy transfer [1], electronic chipcooling [2], and various chemical processes [3]. During boiling heattransfer, the two critical issues are performance (boiling heattransfer coefficient) and critical heat flux (CHF), which is the limitof boiling heat transfer in a thermal system such as a nuclearpower plant. Thus, thermal systems of this type require enhancedboiling heat transfer and increased CHF. Over the past century,numerous techniques for providing high boiling heat transfer andCHF have been developed. Most recently, the effects of nanofluidson boiling heat transfer have been extensively studied; it has beenshown that nanoparticles on a heated surface cause a CHFenhancement while boiling a nanofluid by changing the wettabilitycharacteristics of the surface [4]. Das et al. [5] reported degradedboiling heat transfer in a nanofluid, because the nanoparticles filledthe cavity of the heater surface. Bang and Chang [6] found that theCHF during alumina nanofluid pool boiling increased, but theboiling heat transfer decreased. They observed a roughness changeon the heater surface, and hypothesized that the reason for the

variation in CHF performance was due to the nanoparticle surfacecoating on the heater. Kim et al. [7] performed pool boiling exper-iments with pure water on nanoparticle-deposited surfaces thatwere produced while pool boiling water–TiO2 and water–Al2O3

nanofluids. The CHF of the pure water boiling on the nanoparti-cle-deposited surface was greater than the nanofluid CHF in allcases. This clearly indicates that the nanoparticles deposited onthe heating surface were the main cause of CHF enhancement.Kim et al. [8] also conducted pool boiling experiments with nano-fluids, and described the CHF enhancement in terms of surfacewettability and absorption of the nanoparticle deposition layer.They suggested a simple model, consisting of an absorption layerof nanoparticles, to describe how the nanoparticles changed thesurface wettability. Using results obtained with a thin-wire heater,Kim and Kim [9] postulated that the wettability and capillarity ofthe heating surface could influence the CHF enhancement. They ex-plained that the CHF increases the liquid supply on the heating sur-face, due to the capillarity of the absorption layer of nanoparticles.

Recent general conclusions about the effects of nanofluids onboiling heat transfer and CHF are:

� The CHF increases dramatically during nanofluid boilingbecause of the nanoparticle coating on the heater during nucle-ate boiling [6–13].� The boiling heat transfer decreases during nanofluid boiling

because the nanoparticle coating reduces the active nucleationsites [5,6].

0017-9310/$ - see front matter � 2014 Elsevier Ltd. All rights reserved.http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.01.022

⇑ Corresponding author. Address: Korea Institute of Nuclear Safety, Daejeon,South Korea.

E-mail address: [email protected] (M.H. Kim).

International Journal of Heat and Mass Transfer 74 (2014) 501–512

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer

journal homepage: www.elsevier .com/locate / i jhmt

Author's personal copy

The boiling heat transfer coefficient and the CHF are bothimportant factors in a thermal system. Thus, the advantages anddisadvantages of nanofluid boiling arise concurrently from thenanoparticle coating on the heater surface. Recently, a number ofresearchers have investigated ways of increasing both factors innanofluids. Kwark et al. [14] reported that the boiling heat transfercoefficient and CHF could be increased simultaneously by control-ling the thickness of the nanoparticle coating. They controlled theheat flux and the coating time as the coating parameters. However,there appears to be a need for further studies on increasing boththe CHF and the boiling heat transfer.

Graphene has recently attracted considerable attention due toits high thermal conductivity (900 to 5000 W K�1 m�1) [15], hightransparency [16], ultra-high electric conductivity [17], and novelmechanical properties [18]. However, little information about thethermal characteristics or applications of graphene is available.Since graphene normally exists in the form of ultra-thin and micro-scale flakes, a great deal of effort has been expended to synthesizelarge-area films that retain its superb properties. In particular,soluble graphene in water solvent has been synthesized via theoxidation of graphene. Through a chemical process with hydrazine,graphene oxide (GO) can be changed to a reduced graphene oxide(RGO), which can exhibit properties similar to graphene. Jang et al.[19] estimated the thermal conductivity of RGO film to be 110 to1100 W K�1 m�1 at room temperature. They found that themeasured thermal conductivity of an RGO film on a siliconsubstrate increased as the film grew thicker (from 110 to1100 W K�1 m�1 when its thickness increased from 1 to 100 nmat room temperature). Even though the thermal conductivity of agraphene film is much lower than that of graphene flakes, it is still

higher than that of other substrates, and hence graphene remainsattractive for thermal applications. We chose RGO colloid with adistilled water base as a nanofluid, in anticipation of enhanced heattransfer performance.

The boiling and CHF experiments with the RGO colloids wereconducted in an orderly manner, to examine the characteristics ofRGO colloid boiling and the surface coating of RGO flakes. The RGOcoating layers were characterized by varying the concentration ofthe colloids (0.0001, 0.0005 and 0.0010 wt.%). The effects of the coat-ing layers on the boiling heat transfer and CHF were then revealedthrough surface examination of the RGO coating layers, such as scan-ning electron microscopy (SEM) observation, through considerationof the effects of the coating thermal conductivity, wettability, andwater absorption, and through parametric experiments on the coat-ing layer history in relation to the heat flux.

2. Experiments

2.1. Experimental pool boiling facility

Fig. 1 shows a diagram of the experimental pool boiling facility,which was designed to carry out pool boiling experiments for boil-ing heat transfer and CHF under atmospheric pressure via the elec-trical Joule heating method. It was composed of a test sample and amain pool chamber. The test sample consisted of a silicon dioxideheater (SiO2 surface) and a polyetheretherketone (PEEK) test sam-ple frame. PEEK is a thermoplastic that has high thermal resistance,and is compatible with an aqueous/chemical environment. The testsamples were waterproofed with adhesive sealant (Permatex clearroom-temperature vulcanizing (RTV) silicone) and fixed with

Fig. 1. Schematic diagram of the pool boiling experimental facility.

502 H.S. Ahn et al. / International Journal of Heat and Mass Transfer 74 (2014) 501–512

Author's personal copy

binary epoxy (Duralco 4461). The main pool chamber was a rectan-gular assembled bath made of 10-mm aluminum plate with a4-liter capacity. The test sample was placed on the bottom of thepool chamber. A reflux condenser was installed at the top of themain pool chamber to prevent evaporation of the water.

2.2. Test heater and data acquisition

Fig. 2 shows images of the test heater used in the present work,consisting of double-polished silicon dioxide (SiO2) backed with aplatinum thin-film heater. The test heater was composed ofrectangular silicon wafer plate. The silicon substrate measured25 � 20 mm, and had a SiO2 layer to eliminate native oxidation ef-fects on both the top and bottom surfaces. The heating elementwas a platinum thin film with a thickness of 1200 Å, layered ontothe bottom of the substrate using an electron-beam (E-beam) evap-orator. A titanium thin film with a thickness of 120 Å was used for theadhesion layer between the silicon wafer and the platinum film. Thecomplete platinum film heater was H-shaped (15 � 10 mm), andwas the main heating area. Wire electrodes were attached by leadsoldering, so that the resistance of the electrodes was less than 1%of the total sample resistance. The Joule heating method uses electri-cal power to heat samples directly via conductor resistance. For thepurpose of calibrating the wall temperature, it is important to knowthe exact voltage and current during actual operation, so that theresistance of each test sample can be correlated as a function oftemperature (calibration chart), as shown in Fig. 3. A direct current(DC) power supply (120 V/18 A) was used to provide power to theJoule heating system. The voltage applied to the test sample wasmeasured directly, using a power measuring line and data acquisi-tion system with a 1-s period. The reference resistance unit (Vishay,RH02510R00FC02) was immersed in a constant-temperature bath at10 �C, and used to calculate the exact current passing through thetest sample. The heat flux (q00) was estimated from the voltage ap-plied to the test heater (Vheater), the reference resistor voltage load(Icircuit = Vref/Rref , where Vref and Rref denote the voltage load on thereference resistance unit and the constant reference resistance,respectively), and the heating area (Aheater) as follows:

q00 ¼ VheaterIcircuit

Aheaterð1Þ

The wall temperature was estimated from the calibration chart foreach sample, using the measured resistance of the test heater(Rheater = Vheater/Icircuit).

2.3. Experimental procedure and uncertainties

Prior to the boiling experiments, the working fluid was heated inthe pool for 2 h via a submerged cartridge heater for degassingpurposes. During the experiments, the heat flux was increased in100-kW/m2 steps after 2 min for steady-state data gathering. At heatfluxes less than 100 kW/m2, a 10-kW/m2 step was used to observethe onset of nucleate boiling (ONB). When the CHF occurred duringwater boiling, the wall temperature increased rapidly, and theelectric power was immediately shut down to prevent heater failure.

The analysis of the experimental uncertainty was based on thework of Holman [20]. Using our experimental data, the general for-mula for heat flux calculation is

q00 ¼ Vheater � Vref

Rref � Aheaterð2Þ

The functional form of Eq. (2) is

q00 ¼ q00ðVheater ;Vref ;Rref ;AheaterÞ ð3Þ

where Aheater is assumed to be constant. The uncertainty analysisexpression becomes

U2q00 ¼

@q00

@Vheater

� �2

U2Vheater

þ @q00

@Vref

� �2

U2Vrefþ @q00

@Rref

� �2

U2Rref

ð4Þ

Fig. 2. Test heater (SiO2 substrate and platinum film heater).

Fig. 3. Calibration charts of the test heaters for the wall temperaturemeasurements.

H.S. Ahn et al. / International Journal of Heat and Mass Transfer 74 (2014) 501–512 503

Author's personal copy

where Uq00 , UVheater, UVref

, and URrefrepresent the uncertainties in q00,

Vheater, Vref, and Rref, respectively. Eq. (4) can be simplified to

Uq00

q00¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiU2

Vheater

V2heater

þU2

Vref

V2ref

þU2

Rref

R2ref

vuut ð5Þ

The heater surface temperature (wall temperature) is assumedto be equal to the heater temperature, which is calculated asfollowing.

Twall ¼1aðRheaterÞ ¼

1a

Vheater � Rref

Vref

� �ð6Þ

The general expression for the uncertainty of the surface tempera-ture is

Uwall

DTwall¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiU2

Vheater

V2heater

þU2

Vref

V2ref

þU2

Rref

R2ref

vuut ð7Þ

Accounting for all instrument errors, the maximum uncertain-ties of the heat flux and wall temperature were estimated to be lessthan 1.8% and 3.0%, respectively, over the expected CHF range. Inaddition, we tried to consider the effect of heat spreading (lateralconduction due to the difference of size between the actual heatingarea (10 � 15 mm) and the surface area (25 � 20 mm). To includethe effect of heat spreading and heat loss from backside PT heater,the numerical simulation was conducted as COMSOL 4.0 and Mat-lab 7.0 in our previous study. [21] Based on measured bottom heatflux and temperature information, the reduction procedure is pro-gressed with several heat transfer coefficients till the calculatedvalues match the experimental magnitude. When one used heattransfer coefficient for numerical simulation satisfy experimentallyobtained target condition’s magnitude, we determined on that heattransfer coefficient as our result for the evaluation of heat lossamount. Finally, the heat loss was 6.7–21% regarding to the appliedheat flux, however, the comparison with boiling performance andcritical heat flux of water and RGO colloid boiling could bereasonable.

2.4. Preparation of the RGO colloid

The procedure for the synthesis of the RGO flakes was asfollows. A mixture of 55 mg of GO [22] in 55 mL of water wasultra-sonicated in a high-intensity ultrasonic processor (AutotuneSeries, 750 W) for 30 min. The solution was then centrifuged(MF500 centrifuge, Hanil) at 3000 rpm for 30 min. Next, 50 mL ofthe GO solution was diluted with 50 mL of distilled water contain-ing 50 lL of hydrazine solution (35 wt.%, Aldrich) and 250 lL ofammonia solution (�30%). After a few minutes of vigorous mixing,the solution was maintained at 95 �C for 3 h [23]. The solubility ofRGO is approximately 0.3 mg/mL. The amount of RGO in the solu-tion was measured by filtering a known amount of the solution andweighing the RGO and filter paper. A photograph of the RGO flakes

suspended in deionized water is shown in Fig. 4(a). A transmissionelectron microscopy (TEM) image of typical RGO flakes synthesizedby the above method is shown in Fig. 4(b). The inset image of

Fig. 4. Characteristics of the RGO colloid. (a) Optical image of the RGO colloid, (b) TEM image of an RGO flake (inset: SEAD image), (c) AFM image of an RGO flake.

Fig. 5. Boiling curve of RGO colloid and water.

Fig. 6. CHF enhancement vs. RGO colloid concentration.

Table 1Thermal conductivities of RGO colloids.

Water 0.0001 wt.% 0.0005 wt.% 0.001 wt.%

Thermal conductivity(W/m K)

0.5786 0.585445 0.59491 0.605808

504 H.S. Ahn et al. / International Journal of Heat and Mass Transfer 74 (2014) 501–512

Author's personal copy

Fig. 4(b) shows the selected-area electron diffraction (SAED)performed over the area indicated by the box in the TEM image.The region was oriented along the [001] zone axis. The 24 spots inthe first ring, corresponding to reflections from the (1100) plane, re-vealed hexagonal symmetry in the [0001] diffraction pattern [24].These 24 bright spots represent four highly crystalline sheets ofgraphene that were overlapped. The thickness and size of the sus-pended RGO flakes in water, characterized via atomic force micros-copy (AFM) measurements, were 0.675 nm and 0.5 to 1.0 lm,respectively. This was in agreement with RGO samples preparedby other processes [23,25], but exceeded that of a single graphenelayer (0.334 nm). The apparent increase in the height of the graph-ene on mica was caused by residual epoxy and carbonyl groups thathad not been fully reduced. Water-based RGO colloids wereprepared at concentrations of 0.0001, 0.0005, and 0.0010 wt.%.

3. Experimental results and discussion

Prior to the experiments, the RGO colloids were boiled for 2 h toremove the dissolved gas. A heat flux was applied to the test heaterin 100-kW/m2 steps, with a waiting time of 2 min to achieve asteady state. The boiling phenomena were characterized by the ap-plied heat flux and the wall temperature. We verified the reliabilityand repeatability of the boiling heat transfer and CHF during waterboiling on a clean test heater (bare heater) in a previous study [21].The heat flux and wall temperature at the ONB for every bare hea-ter were 50 to 100 kW/m2 and 10 to 15 �C, respectively. Each pointwas calculated by time-averaging at each step. The CHF was de-fined as the maximum heat flux given by the boiling curve.

3.1. Boiling heat transfer characteristics in the RGO colloids

RGO colloids were prepared at concentrations of 0.0001, 0.0005,and 0.0010 wt.%. Fig. 5 shows the boiling curves for each of theRGO colloids, and indicates that the ONB point was slightly earlierwhen the concentration of the RGO colloid was higher. Overall, theboiling heat transfer performance of the RGO colloid was higherthan that of water for all concentrations. The boiling performancesof the 0.0001, 0.0005, and 0.0010 wt.% RGO colloids improved withdecreasing concentration. The CHF is shown in Fig. 6 in relation tothe concentration of the RGO colloid. As the figure indicates, theCHF increased as the concentration of the RGO colloid decreased.

For the 0.0001 wt.% RGO colloid, the CHF increased by nearly200%. These results raise three questions. Why did ONB occurslightly earlier during RGO colloid boiling than during water boil-ing? Why did the CHF increase? Why did the boiling heat transferperformance increase with decreasing concentration?

First, we will explain why the ONB of the 0.0010 wt.% RGO col-loid was the fastest. According to classical ONB theory [26,27],bubble nucleation can occur when the liquid temperature nearthe heater exceeds the equilibrium temperature DTsat = 2rTsatvg/hfgr⁄ of the nuclear vapor, where r, Tsat, vg, hfg, and r⁄ are the surfacetension of the liquid, saturated temperature, specific volume of thevapor, latent heat, and critical radius, respectively. Han and Griffith[27] proposed the following formula for the ONB temperature:

Fig. 7. Optical images of the test heaters after the CHF experiments with RGO colloids.

Fig. 8. SEM images of the BGL, SFG, and TGL.

Fig. 9. Raman spectroscopy of the BGL, SFG, and TGL.

H.S. Ahn et al. / International Journal of Heat and Mass Transfer 74 (2014) 501–512 505

Author's personal copy

ðTW � TsatÞONB ¼8rq00Tsatvg

hfgkf

� �1=2

ð8Þ

where TW, q00, and kf are the wall temperature, heat flux, and thermalconductivity of the liquid, respectively. We measured the thermalconductivities and heat capacities of the 0.0001, 0.0005, and0.0010 wt.% RGO colloids using the hot-wire method; the resultsare listed in Table 1. The early ONB of the RGO colloids can beexplained by their improved thermal conductivity. For the

quantitative result of ONB, the thermal conductivities of RGO col-loids at 373–378 K were needed, because of the superheated liquidnear the heated surface for onset of nucleate boiling. The thermalconductivity of RGO colloid at only room temperature wasmeasured, because the amount of test colloid was very small, thusthe temperature of RGO colloid could not be maintained. However,the thermal conductivities of the RGO colloids were lower than ourexpectations, due to the superb thermal conductivity of a singlegraphene flake.

Fig. 11. Contact angle measurement on (a) BGL, (b) SFG, (c) TGL, (d) interface between BGL and SiO2 substrate, and (e) interface between BGL and SFG.

Fig. 10. Detailed SEM images of the BGL, SFG, and TGL.

506 H.S. Ahn et al. / International Journal of Heat and Mass Transfer 74 (2014) 501–512

Author's personal copy

Second, we will explain why the CHF increased during RGO col-loid boiling. From the answer to the first question, we can concludethat the thermal conductivity of RGO colloid increases withincreasing concentration. In a number of CHF models, such as thoseof Zuber [28], Haramura and Katto [29] and Kandlikar [30], thethermal conductivity of the working fluid is not carefully consid-ered. Their models incorporate only the physical meaning and rela-tionships between properties of the working fluid and the CHFphenomena, so that the small increase in the thermal conductivi-ties of the RGO colloids cannot be examined as the main parameterof CHF enhancement. We will carefully examine the CHF enhance-ment during RGO boiling in Section 3.3.

Third, we will explain why the boiling heat transfer performanceof the RGO colloids during nucleate boiling improved with decreas-ing concentration. As the working fluids, the RGO colloids had adistilled water base. Even though the thermal conductivities of theRGO colloids were slightly improved, the improvement was not suf-ficient to explain the boiling heat transfer phenomena that occurredduring nucleate boiling. We considered the additional effect of thesurface characteristics on the boiling heat transfer, including possi-ble nucleation sites, since a number of previous studies indicate thatnucleate boiling of nanofluids creates surface modifications viananoparticle deposition [5–14,31]. We will examine the boiling heattransfer performance more thoroughly in Section 3.4.

3.2. Surface investigation following the CHF tests

Fig. 7 shows the test heaters after the CHF experiments with theRGO colloids. RGO flakes were deposited on the heater surfaces.According to Kwark et al. [14], the thickness of the nanoparticlelayer deposited during nucleate boiling is controlled by the appliedheat flux and the coating time. As Fig. 6 indicates, the CHF valuewas highest for the 0.0001 wt.% RGO colloid, and lowest for the0.0010 wt.% RGO colloid. Having the highest CHF implies a high ap-plied heat flux and long coating time. However, the coating of the0.0001 wt.% RGO colloid was the thinnest, contrary to [14]. Thismeans that the RGO coating needs to be carefully investigated.Fig. 7 shows that the gray levels of the RGO coatings depositedon the test heaters spanned a range from light gray to dark grayand black. As the concentration of the RGO colloid increased, thegray level of the RGO coating became darker. Accordingly, weinvestigated three points, (a)–(c), of Fig. 7 via SEM observation.Fig. 8 shows the SEM images at each point. In accordance withour previous study of self-assembled foam-like graphene struc-tures via nucleate boiling [32], the light gray, dark gray, and blacklayers of the RGO coating were referred to as the base graphene

layer (BGL), the self-assembled foam-like graphene structure(SFG), and the thickly aggregated graphene layer (TGL).

Before investigating the characteristics of the RGO coating, weclarified the material characteristics of these three layers via Ra-man spectroscopy, as shown in Fig. 9, since double carbon–carbonbonds lead to high Raman intensities at certain wavelengths. Thethree highly ordered RGO coating layers exhibited similar peakpositions in the G and D bands, at 1346 and 1598 cm�1, respec-tively. Thus, we may infer that the characteristics of the RGO flakeswere nearly same, regardless of the layer.

For further insight into the effects of the RGO coating layers onthe CHF enhancement, we examined representative SEM images ofthe test heaters, shown in Fig. 10, after the CHF experiments withthe 0.0001, 0.0005, and 0.0010 wt.% RGO colloids. Based on theselayers, we studied the CHF enhancement via examination of thesurface wettability changes [8]. In previous studies of CHFenhancement in nanofluids, the improved surface wettability dueto nanoparticle deposition had a strong effect on the CHF. Accord-ingly, we measured the contact angle of a 1-lL water droplet onrepresentative RGO coating layers, including the BGL + baby-SFG,SFG, and TGL. Unfortunately, we were unable to isolate a sectionof the BGL layer large enough to accommodate a 1-lL water drop-let, since the RGO coating layers were formed by nucleate boiling.

In Fig. 11, the contact angles show the different characteristicsof the RGO coating layers. On the BGL + baby-SFG layer, the contactangle revealed a hydrophilic nature, similar to that of the bare SiO2

substrate (62�). However, on the SFG and TGL layers, the contactangle revealed a highly hydrophobic nature. The TGL appeared tobe more hydrophobic than the SFG. It is possible that the rough-ness of the RGO coating layers affected the hydrophobic wettingaccording to Wenzel’s relationship [33]:

cos ha ¼ rðcos hiÞ ð9Þ

where ha, r, and hi are the apparent contact angle, roughness ratio(defined as the ratio of the true area of the solid surface to its pro-jected area), and intrinsic contact angle, respectively. The improvedroughness merely amplifies the intrinsic wettability of the originalsubstrate material, since r is always greater than 1. From Eq. (9), it isclear that ha will be greater than hi if the surface is originally hydro-phobic. In other words, the roughness of the TGL layer is greaterthan that of the SFG layer, and hence the contact angle of the TGLlayer should be greater than that of the SFG layer. In addition, con-tact angle hysteresis can be observed at the interfaces between thelayers in Fig. 11(d) and (e), which show the contact angles at theinterfaces between BGL/SiO2 substrate and the BGL/SFG, respec-tively. This raises a question. Why did the BGL and BGL + baby-

Fig. 12. TEM images of (a) BGL and (b) junction of SFG.

H.S. Ahn et al. / International Journal of Heat and Mass Transfer 74 (2014) 501–512 507

Author's personal copy

SFG layers exhibit hydrophilic wettability, even though the otherlayers were hydrophobic?

We reconsidered the basic characteristics of the RGO flakes to ex-plain this dual wettability of the RGO coating layers. The RGO flakesconsisted of a few layers of graphene combined with the carboxylgroup for increased water solubility, since the original graphenewas hydrophobic [34]. Liang et al. [35] reported that few multilay-ered graphene layers (fMG) exhibited a contact angle of nearly 90�,with hydrophobic and hydrophilic characteristics according to therespective vertical, horizontal, and mixed directions of the fMG. Theyalso considered the effect of the carboxyl group (hydrophilic func-tionalization on fMG) on the wetting phenomena, which could beresponsible for the reduced contact angle, due to water absorption

in the numerous microchannels of the fMG layer. Accordingly, weexamined the alignment of the RGO flakes in the hydrophilic BGLand hydrophobic SFG using TEM, as shown in Fig. 12. Well alignedRGO flakes were observed on the BGL, in good agreement with theprevious report of Liang et al. [35]. The RGO flakes at the SFG junctionwere also well aligned. However, this was contrary to expectation, inview of the results for the BGL. Since the SFG consisted of severalpores with numerous junctions, and was also comprised of a few lay-ers of RGO flakes, water droplets could not meet along the directionof these well-aligned junctions. Thus, the SFG exhibited a hydropho-bic nature. However, in both the BGL and SFG, there was a functionalpart of the well-aligned RGO flakes and the carboxyl group forattracting water.

Fig. 13. Water absorption phenomenon on the SFG.

Fig. 14. The schematic of boiling on porous media [39].

508 H.S. Ahn et al. / International Journal of Heat and Mass Transfer 74 (2014) 501–512

Author's personal copy

Finally, we investigated the water absorption phenomena usingenvironmental scanning electron microscopy (E-SEM). At 0.6 barand room temperature, water droplets were formed on the SFG. Asthe size of the water droplets increased, the droplets were immedi-ately absorbed into the SFG, as shown in Fig. 13. In nature, it is wellknown that a hydrophobic material cannot absorb water, because itrepels water. In addition, hydrophobic micro/nanoscale structuresare absolutely unable to absorb water because of the air packing be-tween the structures. Nevertheless, the hydrophobic SFG was able toabsorb water, as shown in Fig. 13. This could prove to be reliable evi-dence in favor of the dual wettability of the SFG, due to the presenceof the carboxyl group. For further consideration, the SFG and BGLcould be fully saturated with water during nucleate boiling.

3.3. Effects of the RGO layers on CHF enhancement

We now attempt to answer the following unanswered questionfrom Section 3.1: Why did the CHF of the RGO colloid increase? InSection 3.2, we investigated the characteristics of the three RGOcoating layers (BGL, SFG, and TGL). First, we considered the rela-tionship between the concentration of the RGO colloid, thedominant RGO layer, and the CHF enhancement. After the CHFtests of the 0.0001, 0.0005 and 0.0010 wt.% RGO colloids, the CHFenhancement and dominant RGO coating layers differed. For the0.0001 wt.% RGO colloid, the CHF enhancement had the maximumratio, and the dominant RGO layer was the BGL and the baby-SFG.For the 0.0005 wt.% RGO colloid, the dominant layer was mostly

the SFG, and the TGL was only partially deposited. For the0.0010 wt.% RGO colloid, the CHF enhancement had the minimumratio, and the dominant layer was entirely the TGL. Thus, the dom-inant RGO coating layer was different for each RGO colloid concen-tration, and seemed to affect the CHF enhancement. Generallyspeaking, an RGO coating layer with low contact angle (hydro-philic), such as the BGL, is favorable for CHF enhancement. TheCHF enhancement during pool boiling can be estimated withrespect to the wettability via Kandlikar’s model [30]:

q00CHF ¼hfgq1=2g

1þcosb16

� �2pþp

4ð1þcosbÞcos/

� �1=2

rgðql�qgÞh i1=4

ð10Þ

where qg, ql, b, and / are the gas density, liquid density, angle ofinclination, and contact angle, respectively. The improved wettabilityof the BGL could explain some degree of the CHF enhancement. How-ever, it seems inadequate to fully explain a CHF enhancement of200%. Moreover, it is also insufficient to fully explain the CHFenhancement while boiling 0.0005 and 0.0010 wt.% RGO colloids,since the SFG and TGL exhibited highly hydrophobic characteristics.Therefore, we reconsidered the water absorption phenomenon onthe SFG, as shown in Fig. 13. We inferred that the SFG played the roleof a porous medium, which could be saturated by the water. As thewetted area increases, the water (liquid) has a greater chance ofencountering liquid between the pores of the porous medium. Therole of water absorption by the BGL and SFG could adequatelyexplain the CHF enhancement in terms of the supplementary liquid

Fig. 15. Structure of the deposited layer of RGO flakes. (a) Dominant TGL after the CHF test in 0.0010 wt.% RGO colloid, (b) dominant SFG after the CHF test in 0.0005 wt.% RGOcolloid and (c) simple schematic diagram of the stacked layers of RGO flakes.

H.S. Ahn et al. / International Journal of Heat and Mass Transfer 74 (2014) 501–512 509

Author's personal copy

obtained from the capillary flow [36–38]. According to Kaviany [39],the vapor thickness (dg) and the two-phase length (dgl) in the porousmedium are the main parameters that determine the CHF for boilingon a liquid-saturated porous medium, as shown in Fig. 14. He provedthat dgl tends to infinity when the viscous and gravitational forces ex-actly balance. The capillary pressure and saturation gradients in theporous medium then become zero, and finally the CHF occurs. Thismeans that the balance between the drag and the capillary forcesin a liquid-saturated porous medium could affect the CHF enhance-ment. According to Liter and Kaviany [38], the CHF could increasethe viscous-drag liquid choking limit by five times. Taken together,the following factors would provide a reasonable explanation forthe CHF enhancement on the BGL, SFG, and TGL layers during theRGO colloid boiling experiments with the various concentrations:the improved surface wettability, the dynamics of the contact angleduring water absorption, and the liquid-saturated porous mediumresulting from water absorption by the BGL and SFG.

Finally, there are some additional points of interest regardingthe BGL, where the RGO flakes appeared to be well aligned, as

shown in Fig. 12. Here, the alignment of the RGO (graphene)flakes in the layer (thin film) is important for determiningcharacteristics such as the electric and thermal conductivities(the thermal conductivities of in-plane and out-of-plane graph-ene flakes are significantly different – 5000 and 5 W/m K,respectively) [15]. Even though the edges of the RGO flakes in-duce thermal/electric resistance [40], there are a number ofpublished reports on the improved thermal conductivity (111to 1000 W/m K) of a well-aligned RGO film, such as the BGL[19,41,42]. Thus, we considered the improved thermal conduc-tivity of the BGL on the heater surface, which affected the CHFenhancement. According to Arik and Bar [43] and Bang et al.[44], improved thermal activity (S ¼ d

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiqhchkh

p, where S, d, qh,

ch, and kh denote the thermal activity, thickness of the layer,density of the layer, heat capacity of the layer, and thermalconductivity of the layer, respectively) on a heater surface couldincrease the CHF by delaying the hot/dry spot growth. However,since SFG is a porous material, the effective thermal conductiv-ity of the SFG could not be assumed to have a high value, even

Fig. 16. Development of the RGO flake layer with applied heat flux.

510 H.S. Ahn et al. / International Journal of Heat and Mass Transfer 74 (2014) 501–512

Author's personal copy

though the SFG junction exhibited well-aligned RGO flakes, asshown in Fig. 12.

3.4. Relationship between RGO layer development and boiling heattransfer

Next, we attempt to answer the following unanswered questionfrom Section 3.1: Why did the boiling heat transfer performance ofthe RGO colloids during nucleate boiling improve with decreasingconcentration? From the ONB, the boiling heat transfer performancechanges smoothly and gradually with increasing heat flux. Our boil-ing curves had no inflection points, but exhibited a continuous ten-dency. Accordingly, we investigated the history of the RGO coatinglayer development in relation to the applied heat flux. First, we ob-tained side view SEM images of the RGO coating layers of the testheater after colloid boiling, as shown in Fig. 15. After 0.0010 wt.%RGO colloid boiling (Fig. 15(a)), all three RGO layers appeared,stacked in the following order: BGL, SFG, and TGL. On the other hand,after 0.0005 wt.% RGO colloid boiling (Fig. 15(b)), only the SFG andBGL appeared, and the TGL was absent. According to the SEM obser-vations, the BGL, SFG, and TGL were stacked in order. This means thateach layer affected the boiling heat transfer in common with the CHFenhancement. Thus, we investigated the relationship between theheat flux and RGO coating layer development by observing theRGO flake coatings at heat fluxes of 50, 100, 400, 800, and1200 kW/m2. The representative RGO colloid concentration was0.0005 wt.%. Fig. 16 shows the RGO coating layer development inrelation to the applied heat flux. At a heat flux of 800 kW/m2, SFG for-mation was observed. As the heat flux increased, the SFG had moreuniform pores and smaller pore sizes, which could lead to increasedboiling heat transfer resulting from more activated nucleation sitesand hydrophobic properties [45]. In the 0.0001 wt.% RGO colloid,the formation of the RGO coating layer could be slower than in the0.0005 wt.% RGO colloid, whereas in the 0.0010 wt.% RGO colloid, itcould be faster. The TGL layer, which was observed in 0.0010 wt.%RGO colloid boiling, seemed to be an aggregation of RGO flakes.The SFG would provide better enhancement of the boiling heattransfer and CHF, since the TGL could contribute to the thermal resis-tance. In addition, because of the early ONB resulting from the en-hanced thermal conductivity of the working fluid, the enhancedboiling heat transfer over the entire boiling curve could also be ex-plained by the surface deposition of RGO flakes.

4. Conclusions

We studied the pool boiling heat transfer characteristics of RGOcolloids. Through CHF pool boiling experiments, the effect of col-loid concentration on the boiling process was estimated quantita-tively. The results showed that the CHF and boiling heat transfer inthe RGO colloids were both increased in comparison to plain water,and that as the concentration increased, (1) the ONB occurred at anearlier stage, (2) the rate of CHF enhancement increased, and (3)the coefficient of boiling heat transfer decreased. Based on theexperimental results, we focused on the effect of the concentrationon the working fluid itself and the RGO flake layer characteristics,and postulated possible mechanisms for the CHF and boiling heattransfer enhancement.

� As the colloid concentration increased, the thermal conductivityof the fluid also increased, as shown in Table 1. The increasedthermal conductivity of the working fluid could explain theearly occurrence of ONB at relatively low wall temperatures.� Nucleate boiling of the RGO colloid formed unique structures on

the heater surface, namely the RGO flake layers, BGL, SFG, andTGL. Even though these layers consisted of the same material

(RGO flakes, as demonstrated by the Raman spectroscopy anal-ysis), contact angle measurements revealed significant differ-ences between the dominant RGO layers remaining on theheater surface: The BGL was hydrophilic, while the SFG andTGL were hydrophobic. TEM analysis identified wall-alignedRGO flakes in both the BGL and SFG, but the hydrophilic andhydrophobic characteristics were determined by how the watercontacted the layers. On the other hand, water absorption phe-nomena were observed in the BGL and SFG layers. This dualwettability of SFG could be explained by carboxyl groupactivation.� The CHF enhancement was examined in terms of the water

absorption phenomena in the porous medium of the RGO flakelayers. The concentration of the colloid directly affected thedominant RGO layer on the surface, which determined the char-acteristics of the porous medium. Another possible mechanismwas the improved thermal activity of the well-aligned RGOflake layers due to delayed hot/dry spot growth.� The decreased boiling heat transfer with increasing concentra-

tion was explained by the RGO flake layer formation mecha-nism. By controlling the heat flux, we were able to observehow the RGO flakes formed the layers, and found that the SFGcould contribute to better boiling heat transfer and CHF, whilethe TGL could contribute to thermal resistance.

Acknowledgments

This work was supported by the National Research Foundationof Korea (NRF) Grant funded by the Korea Government (MEST)(2012R1A1A2002900).

References

[1] M.G. Kang, Experimental investigation of tube length effect on nucleate poolboiling heat transfer, Ann. Nucl. Energy 25 (4–5) (1998) 295–304.

[2] I. Mudawar, Assessment of high-heat-flux thermal management schemes, IEEETrans. Compon. Packag. Technol. 24 (2) (2001) 122–141.

[3] J.R. Thome, Enhanced Boiling Heat Transfer, Hemisphere, New York, 1990.[4] H.D. Kim, J.B. Kim, M.H. Kim, Effect of nanoparticles on CHF enhancement in

pool boiling of nano-fluids, Int. J. Heat Mass Transfer 49 (25–26) (2006) 5070–5074.

[5] S.K. Das, N. Putra, W. Roetzel, Pool boiling characteristics of nano-fluids, Int. J.Heat Mass Transfer 46 (2003) 851–862.

[6] I.C. Bang, S.H. Chang, Boiling heat transfer performance and phenomena ofAl2O3–water nano-fluids from a plain surface in a pool, Int. J. Heat MassTransfer 48 (2005) 2407–2419.

[7] H.D. Kim, J.B. Kim, M.H. Kim, Experimental studies on CHF characteristics ofnano-fluids at pool boiling, Int. J. Multiphase Flow 33 (2007) 691–706.

[8] S.J. Kim, I.C. Bang, J. Buongiorno, L.W. Hu, Effects of nanoparticle deposition onsurface wettability influencing boiling heat transfer in nanofluids, Appl. Phys.Lett. 89 (15) (2006) 153107.

[9] H.D. Kim, M.H. Kim, Effect of nanoparticle deposition on capillary wicking thatinfluences the critical heat flux in nanofluids, Appl. Phys. Lett. 91 (1) (2007)014104.

[10] M.N. Golubovic, H.D. Madhawa Hettiarachchi, W.M. Worek, W.J. Minkowycz,Nanofluids and critical heat flux, experimental and analytical study, Appl.Therm. Eng. 29 (2009) 1281–1288.

[11] Z. Liu, L. Liao, Sorption and agglutination phenomenon of nanofluids on a planeheating surface during pool boiling, Int. J. Heat Mass Transfer 51 (2008) 2593–2601.

[12] J.S. Coursey, J. Kim, Nanofluid boiling: the effect of surface wettability, Int. J.Heat Fluid Flow 29 (2008) 1577–1585.

[13] Y.H. Jeong, W.J. Chang, S.H. Chang, Wettability of heated surfaces under poolboiling using surfactant solutions and nano-fluids, Int. J. Heat Mass Transfer 51(2008) 3025–3031.

[14] S.M. Kwark, R. Kumar, G. Moreno, J. Yoo, S.M. You, Pool boiling characteristicsof low concentration nanofluids, Int. J. Heat Mass Transfer 53 (5–6) (2010)972–981.

[15] A.A. Balandin, S. Ghosh, W. Bao, I. Calizo, D. Teweldebrhan, F. Miao, C.N. Lau,Superior thermal conductivity of single-layer graphene, Nano Lett. 8 (3) (2008)902–907.

[16] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, M.I. Katsnelson, I.V.Grigorieva, S.V. Dubonos, A.A. Firsov, Two-dimensional gas of massless Diracfermions in graphene, Nature 438 (2005) 197–200.

H.S. Ahn et al. / International Journal of Heat and Mass Transfer 74 (2014) 501–512 511

Author's personal copy

[17] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, Y. Zhang, S.V. Dubonos, I.V.Grigorieva, A.A. Firsov, Electric field effect in atomically thin carbon films,Science 306 (2004) 666–669.

[18] C. Lee, X.D. Wei, J.W. Kysar, J. Hone, Measurement of the elastic properties andintrinsic strength of monolayer graphene, Science 321 (2008) 385–388.

[19] W. Jang, Z. Chen, W. Bao, C.N. Lau, C. Dames, Thickness-dependent thermalconductivity of encased graphene and ultrathin graphite, Nano Lett. 10 (10)(2010) 3909–3913.

[20] J.P. Holman, Experimental Methods for Engineers, seventh ed., McGraw-Hill,New York, 2001 (Chapter 3).

[21] H.J. Jo, H.S. Ahn, S.H. Kang, M.H. Kim, A study of nucleate boiling heat transferon hydrophilic, hydrophobic and heterogeneous wetting surfaces, Int. J. HeatMass Transfer 54 (2011) 5643–5652.

[22] W.S. Hummers, R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc.80 (6) (1958) 1339.

[23] D. Li, M.B. Müller, S. Gilje, R.B. Kaner, Processable aqueous dispersions ofgraphene nanosheets, Nat. Nanotechnol. 3 (2008) 101–105.

[24] M.J. McAllister, J.L. Li, D.H. Adamson, H.C. Schniepp, A.A. Abdala, J. Liu, M.Herrera-Alonso, D.L. Milius, R. Car, R.K. Prud’homme, I.A. Aksay, Single sheetfunctionalized graphene by oxidation and thermal expansion of graphite,Chem. Mater. 19 (18) (2007) 4396–4404.

[25] I.K. Moon, J. Lee, S.R. Ruoff, H. Lee, Reduced graphene oxide by chemicalgraphitization, Nat. Commun. 1 (2010) 73.

[26] Y.Y. Hsu, On the size range of active nucleation cavities on a heating surface,ASME J. Heat Transfer 84 (1962) 207.

[27] C.Y. Han, P. Griffith, The mechanism of heat transfer in nucleate pool boiling –Part II, Int. J. Heat Mass Transfer 8 (1965) 887.

[28] N. Zuber, Hydrodynamic aspects of boiling heat transfer (Ph.D. thesis),University of California, Los Angeles, AECU-4439, 1959

[29] Y. Haramura, Y. Katto, A new hydrodynamic model of critical heat flux,applicable widely to both pool and forced convective boiling on submergedbodies in saturated liquids, Int. J. Heat Mass Transfer 26 (1983) 389–399.

[30] S.G. Kandlikar, A theoretical model to predict pool boiling CHF incorporatingeffects of contact angle and orientation, ASME J. Heat Transfer 123 (2001)1071–1079.

[31] B. Truong, L.W. Hu, J. Buongiorno, T. McKrell, Modification of sandblasted plateheaters using nanofluids to enhance pool boiling critical heat flux, Int. J. HeatMass Transfer 53 (2010) 85–94.

[32] H.S. Ahn, J.-W. Jang, M. Seol, J.M. Kim, D.-J. Yun, C. Park, H. Kim, D.H. Youn, J.Y.Kim, G. Park, D.I. Yu, S.C. Park, J.M. Kim, K. Yong, M.H. Kim, J.S. Lee, Self-Assembled foam-like graphene networks formed through nucleate boiling, Sci.Rep. 3 (2013) 1396.

[33] R.N. Wenzel, Resistance of solid surfaces to wetting by water, Ind. Eng. Chem.28 (1936) 988–994.

[34] J. Rafiee, M.A. Rafee, Z.-Z. Yu, N. Korakar, Superhydrophobic to superhydrophilicwetting control in graphene films, Adv. Mater. 22 (2010) 2151–2154.

[35] Q. Liang, X. Yao, W. Wang, Y. Liu, C.P.A. Wong, Three-dimensional verticallyaligned functionalized multilayer graphene architecture: an approach forgraphene-based thermal interfacial materials, ACS Nano 5 (2011) 2392–2401.

[36] H.S. Ahn, H.J. Jo, S.H. Kang, M.H. Kim, Effect of liquid spreading due to nano/microstructures on the critical heat flux during pool boiling, Appl. Phys. Lett.98 (2011) 071908.

[37] H.S. Ahn, C. Lee, J. Kim, M.H. Kim, The effect of capillary wicking action ofmicro/nano structures on pool boiling critical heat flux, Int. J. Heat MassTransfer 55 (2012) 89–92.

[38] S.G. Liter, M. Kaviany, Pool-boiling CHF enhancement by modulated porous-layer coating: theory and experiment, Int. J. Heat Mass Transfer 44 (2001)4287–4311.

[39] M. Kaviany, Principles of Heat Transfer in Porous Media, Springer-Verlag, NewYork, 1991. pp. 582–585.

[40] S. Shin, M. Kaviany, Interflake thermal conductance of edge-passivatedgraphene, Phys. Rev. B 84 (2011) 235433.

[41] S. Chen, A.L. Moore, W. Cai, J.W. Suk, J. An, C. Mishra, C. Amos, C.W. Magnuson, J.Kang, L. Shi, R.S. Ruoff, Transport in suspended monolayer graphene of variablesizes in vacuum and gaseous environments, ACS Nano 5 (2011) 321–328.

[42] J.H. Seol, I. Jo, A.L. Moore, L. Lindsay, Z.H. Aitken, M.T. Pettes, X. Li, Zhen Yao, R.Huang, D. Broido, N. Mingo, R.S. Ruoff, L. Shi, Two-dimensional phonontransport in supported graphene, Science 328 (2010) 213–216.

[43] M. Arik, A. Bar-Cohen, Effusivity-based correlation of surface property effectsin pool boiling CHF of dielectric liquids, Int. J. Heat Mass Transfer 46 (20)(2003) 3755–3764.

[44] I.C. Bang, J. Buongirno, L.W. Hu, H. Wang, Measurement of key pool boilingparameters in nanofluids for nuclear applications, J. Power Energy Syst. 2 (1)(2008) 340–351.

[45] Y. Takata, S. Hidaka, T. Uraguchi, Boiling feature on a super water-repellentsurface, Heat Transfer Eng. 27 (8) (2006) 25.

512 H.S. Ahn et al. / International Journal of Heat and Mass Transfer 74 (2014) 501–512