Next Article in Journal
Experimental Study of High-Strength Concrete-Steel Plate Composite Shear Walls
Previous Article in Journal
Corrosion of α-Brass in Solutions Containing Chloride Ions and 3-Mercaptoalkyl-5-amino-1H-1,2,4-triazoles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Effect of Nanofluids on Boiling Heat Transfer Performance

School of Energy and Power Engineering, Jiangsu University of Science and Technology, Zhenjiang 212003, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2019, 9(14), 2818; https://doi.org/10.3390/app9142818
Submission received: 31 May 2019 / Revised: 1 July 2019 / Accepted: 4 July 2019 / Published: 15 July 2019
(This article belongs to the Section Nanotechnology and Applied Nanosciences)

Abstract

:
At present, there are many applications of nanofluids whose research results are fruitful. Nanofluids can enhance the critical heat flux, but the effect on boiling heat transfer performance still has disagreement. Base liquids with higher viscosity improve the boiling heat transfer performance of nanofluids. When the base liquid is a multicomponent solution, the relative movement between the different solutions enhances the microscopic movement of the nanoparticles due to the different evaporation order during the boiling process, so that the boiling heat transfer performance is enhanced. Compared with the thermal conductivity of the heated surface, the deposition of the low thermal conductivity nanoparticles reduces the heat dissipation rate of the heated surface and improves the wall superheat. Then the enhancement of the boiling heat transfer coefficient should be attributed to the thermal conductivity improvement of base fluid and the bubble disturbance resulted from the nanoparticle’s microscopic motion.

1. Introduction

In 1756, Leidenfrost [1] dropped droplets on two aluminum plates of different temperatures, and it was found that the water droplets evaporate faster on the low temperature plates. This phenomenon has laid a milestone in the development of boiling heat transfer. In 1931, Jakob used a high-speed camera to study the generation and development of bubbles [2,3]. In 1936, the roughness effect on boiling heat transfer was investigated [4]. However, Kutateladze [5,6] used fluid dynamics to study the boiling critical heat flux in 1950, which promoted the theoretically research process of boiling heat transfer. In the early 1970s, the rapid development of industry further promoted the study of two-phase flow boiling heat transfer.
In 1995, the Argonne National Laboratory proposed to disperse metal or nonmetal particles in water or organic solvents to form a stable suspension solution, which can improve the thermal conductivity of the base liquid. It has been verified by many scholars such as Akhgar [7], Kakavandi [8], Aparna [9], Hamid [10], and so on. The nanofluids stability is affected by the nanoparticles, base fluid, and dispersant. In order to improve the dispersibility of nanoparticles, the dispersant with appropriate concentration and type was added in nanofluids and then oscillated in the oscillator [11,12].
At present, the development of electronic devices in the direction of high integration and high computing speed has brought about the “thermal barrier” and “thermal management” problems that have become a hidden danger for the safe operation of equipment. Therefore, the search for high-efficiency heat conduction material and heat-dissipating methods has become the hot focus in industry [13,14,15,16]. During the boiling heat transfer process, bubbles are formed at the vaporization core point, and continuously absorb heat to disengage, float, and escape the water surface. The bubble detachment cause severe disturbance to the liquid of the heating surface, so that the boiling heat transfer is stronger than the convective heat transfer of a single-phase fluid. Therefore, boiling heat transfer is widely used in heat dissipation and cooling of many industrial fields, such as electronic equipment, because of its small temperature difference and high heat transfer capacity [17,18,19].
Kim [20] et al. carried out the bubble operation of the classic pool boiling curve at different boiling stages from the Figure 1. There is no bubble generation in the natural convection stage. The nuclear boiling phase can be divided into partial development stages and full development stages. There are relatively few bubbles in partial development stages and significant bubble coalescence has not yet occurred. During the full development stage, the bubbles appear to slip on the heated surface and merge into a mushroom-like and then escape from the water surface [21,22]. During the transitional boiling phase and the film boiling phase, the bubbles gradually form a gas film on the heating surface, and the heat transfer is carried out by means of heat conduction and radiation, which is likely to cause the device to burn out.
Nikkhah [23] et al. investigated experimentally the convective boiling heat transfer coefficient of spherical CuO(II) nanoparticles dispersed in water inside the vertical heat exchanger. Results show that by increasing heat and mass fluxes, the heat transfer coefficient considerably increases for both heat transfer regions, while by increasing the nanoparticle weight concentration the heat transfer coefficient increases in convective heat transfer (~35% at the maximum concentration) and deteriorates the heat transfer coefficient (~9% at the maximum concentration) in nucleate boiling region due to the formation of nanoparticle deposition of heating surface.
Sarafraz [24] et al. experimentally performed the flow boiling heat transfer characteristics of MgO/therminol 66 nanofluid at wt.% = 0.1 and wt.% = 0.3 as a potential coolant on a copper-made disc. Results revealed that bubble formation induces a pressure drop within the test section. Heat flux had no influence on the pressure, while an increase in the fluid flow rate caused an increase in the pressure drop. It was also found that the heat transfer coefficient decreased with operating time due to the presence of nanoparticles on the boiling surface resulting in the creation of thermal resistance on the surface. Also, an asymptotic behavior for the fouling thermal resistance over the time was registered. The maximum enhancement for the heat transfer coefficient was 23.7% at wt.% = 0.1. For wt.% = 0.2 and wt.% = 0.3, the maximum enhancement of 16.2% and 13.3%, were achieved, respectively. At the same time, Sarafraz [22] et al. found that by increasing the applied heat flux, the flow boiling heat transfer coefficient increases for both test fluids at both heat transfer regions. In addition, by increasing the flow rate of fluid, the heat transfer coefficient dramatically increases at both regions. Furthermore, higher heat transfer coefficient can be obtained due to interactions of bubbles and local agitations. In addition, Author experimentally measured the nucleate pool boiling heat transfer coefficients of Al2O3–water and Ti2O–water nanofluids on three horizontal tubes with different materials and similar roughness [25]. Results revealed that presence of nanoparticles in the base fluid leads to an increase in pool boiling heat transfer coefficients on stainless steel and brass tubes in contrast to copper tube. Presence of nanoparticles had no effect on the pool boiling heat transfer coefficient for the copper tube. Variations of surface excess temperature for the copper tube were higher in comparison with that of the other tubes tested.
Salari [26] et al. experimentally studied the pool boiling heat transfer characteristics of gamma Fe3O4 aqueous nanofluids on a flat disc heater. The nanoparticle mass concentration was 0.1–0.3% and the heat flux was 0–1546 kW/m2. Results indicated that the pool boiling heat transfer coefficient increases with increasing the mass concentration and applied heat flux. In addition, the rate of bubble formation is significantly intensified at higher heat fluxes, and subsequently larger bubbles detached from the surface due to the intensification of bubble coalescence. In terms of fouling formation, it can be stated that fouling of nanofluids is a strong function of time and rate of deposition is increased over the extended time while the pool boiling heat transfer coefficient was not decreased over the time, as the porous deposited layer on the surface detached from the surface by bubble interactions. In terms of critical heat flux, capillary action of the deposited layer was found to be the main reason responsible for increasing the critical heat flux as liquid is stored inside the porous deposited layer, which enhances the surface toleration against the critical heat flux crisis.
The nanoparticles can increase the thermal conductivity of the base fluid. In addition, the deposition of nanoparticles can reshape the heated surface, thereby affecting the boiling heat transfer performance. So the summary of this article were shown below.
  • Application of nanofluids.
  • Effect of nanoparticle type on boiling heat transfer.
  • Effect of base fluid type on boiling heat transfer.
  • Outline the effect of nanofluids on boiling critical heat flux.
  • Outline the effect of nanofluids on boiling heat transfer coefficient.
This paper summarizes the existing literature and understands the research and value of nanofluids in different application fields, providing readers with a relatively macroscopic understanding. The disagreement and differences in the effects of nanofluids on the boiling heat transfer coefficient [27,28] are explained by the heated surfaces, boiling bubbles, and the microscopic motion of nanofluids. It also provides some ideas for future research for nanofluids.

2. Nanofluid Application in Engineering

Nanofluids have great potential applications in the fields of energy, chemical, automotive, construction, microelectronics, and information, and thus have become a research hotspot in materials, physics, chemistry, heat transfer, and other fields.

2.1. Nanofluid Application in Heat Pipes

Heat pipes operate under a number of physical constraints including the capillary, boiling, sonic and entrainment limits that fundamentally affect their performance [29]. Temperature gradients near the heated end may be high enough to generate significant Marangoni forces that oppose the return flow of liquid from the cold end in heat pipes. In the presence of significant Marangoni forces, dry out is not the initial mechanism limiting performance, but that the physical cause is exactly the opposite behavior: flooding of the hot end with liquid. The observed effect is a consequence of the competition between capillary and Marangoni-induced forces.
Maximum heat transfer ability was estimated as the minimal value from the heat transfer abilities specified by the hydrodynamic, the sonic and the boiling crisis limitations in the HPs. These limitations were calculated as [30] follows.
The hydrodynamic limit,
Q max h y d = 2.19 N · S · F W L e f f
The sonic limit,
Q max s n = F v c · ρ v · r · γ v R μ v T s 2 μ ( γ v + 1 )
The boiling crisis limit,
Q max b = L e v 2 π λ e f f w T s r ρ v ln ( d i s h d v c ) ( 2 β r c r b P c )
Figure 2 depicts the choosing of the optimal wick parameters. It can be seen from Figure 2 that the heat transfer ability at saturation temperatures below −20 °C is extremely low and in the temperature range of −60 °C~0 °C is specified by the sonic limit; at 0 °C~+40 °C, the hydrodynamic limit; and at +40 °C~+60 °C, the boiling crisis limit.
Heat pipe that is a compact and effective heat exchanger includes microchannel heat pipe, mesh wick heat pipe, sintered core heat pipe, oscillating heat pipe and thermosyphon heat pipe, and other heat pipe forms [31]. The researchers conducted experimental studies on different types of nanofluids such as Al2O3 [32,33], CuO [34,35], SiO2 [36,37], Fe2O3 [38,39], and graphene nanosheets [40,41,42] in heat pipes. The results indicated that nanofluids can enhance heat transfer of the heat pipes. Figure 3 shows the working principle of the heat pipe.
Kim [18] et al. studied the heat transfer performance of heat pipes and mesh cores filled with water-based graphene oxide nanofluid (GON) with volume concentrations of 0.01% and 0.03%. The experimental results show that the nanofluid makes the wall temperature lower than the water. In addition, heat pipes containing graphene oxide exhibit lower evaporation heat resistance. The 0.01% GON exhibits better heat transfer due to the different nanoparticle layers deposited on the core structure. The nanoparticle deposition layer changes the capillary force, resulting in an increase in the maximum fluid velocity through the core structure.
Zhou et al. [43] studied the heat transfer characteristics of graphene nanosheets (GNs) nanofluids in an oscillating heat pipe. The volume concentration of graphene was 1.2–16.7% and the filling rate was 45–90%. The results showed that compared with deionized water, GNs nanofluids improve the heat transfer characteristics of the oscillating heat pipe, and the optimum concentration is 2–13.8%, mainly because the addition of GNs reduces the drying phenomenon of the heated surface.
In addition, Li [44], Nazaria [45], Hosseinian [46], and others also studied the effect of nanofluids on the heat transfer of heat pipes, and found that the addition of nanofluids can enhance the heat transfer characteristics of heat pipes. Mainly reason is that the deposition of nanoparticles changes the wettability of the evaporation surface, which enhances the capillary action and reduces the drying of the heated surface.

2.2. Nanofluid Application in Automobiles

The car radiator consists of three parts which are including the inlet chamber, the outlet chamber and the radiator core. The coolant flows in the radiator core and the air passes outside the radiator. The hot coolant cools due to heat dissipation from the air and the cold air warms up by absorbing the heat dissipated by the coolant. Nanofluids can make the car radiators smaller and lighter, and develop in an efficient direction, which reduce the cost of the car [47,48].
Subhedar [47] et al. used a mixed liquid whose volume ratio is 50:50 with ethylene glycol and deionized water (50EG/50DW) Al2O3-based nanofluid as a coolant for automotive radiators to investigate its heat transfer performance. The nanoparticle volume fraction is 0.2–0.8%, the coolant flow is 4–9 L/min, and the inlet temperature is 65–85 °C. Nanofluids can improve the heat transfer performance of heat sinks compared to conventional coolants. When the nanoparticle concentration is 0.2%, the heat transfer increased by 30%. Contreras [48] and others introduced the 50EG/50DW liquid-based graphene and silver nanofluids with the volume concentrations of 0.01%, 0.05%, and 0.1% in the automotive radiator. The coolant flow rate is between 0.08 kg/s and 0.11 kg/s, the coolant inlet temperature is between 55 °C and 85 °C, and the air flow rate is maintained at 2.1 m/s. The experiment found that the silver nanofluid increased the heat transfer rate by 4.4%, while the graphene nanofluid decreased compared with the base fluid. In the same ratio mixture of ethylene glycol and water, Tijani [49] evaluated the heat transfer characteristics of CuO and Al2O3 nanofluids, which include the heat transfer coefficient, thermal conductivity, Nusselt number, and heat transfer rate. The CuO nanofluid exhibited the highest heat transfer performance, whose heat transfer coefficient was recorded as 36,384.41 W/(m2·K), the thermal conductivity was 1.241 W/(m2·K), the Nusselt number was 208.71, and the heat transfer rate was 28.45 W. The Al2O3 nanofluid has a heat transfer coefficient of 31,005.9 W/(m2·K), a thermal conductivity of 1.287 W/(m2·K), a Nusselt number of 173.19, and a heat transfer rate of 28.25 W.
Chaurasla [50] and Ahmed [51] studied the heat transfer performance of automotive radiators using water-based Al2O3 and water-based TiO2 nanofluids, respectively. The fluid poisoning and the increase in the flow rate of the air improve the heat transfer performance. Adding the 0.2% volume of Al2O3 nanoparticles, the heat transfer efficiency of the radiator is increased by up to 44.28%. While adding TiO2 nanoparticles with concentration of 0.1% and 0.3% can improve the efficiency of the car radiator by 47%. The average heat transfer coefficient is directly affected by the increase in Reynolds number and volume concentration of the nanofluid.

2.3. Nanofluid in the Collector

Farhana et al. [52] reviewed the importance of nanofluids for thermal performance for six types solar collectors. In addition, an improved quantification of the performance of solar collectors by different types of nanofluids is compiled. Bellos et al. [53] studied in detail the use of nanofluids in concentrating solar collectors, including heating, cooling, power, and triads. Emphasis is placed on the enhancement of the thermal efficiency of the collector using nanofluids. Figure 4 shows the types of nanoparticles in solar collectors and the range of applications of solar collectors. As can be seen from the figure, there are many types of nanofluids used in solar collectors. The types of solar collectors include Flat plate, evacuated tube, direct absorber, parabolic trough, solar dish, and thermal photovoltaic. Solar collector is a special kind of heat exchanger which convert solar radiation into two types of energy such as thermal energy in solar thermal applications and electrical energy directly in photovoltaic applications. In the case of thermal application, solar devices absorb the solar radiant energy incident on its surface; transform it into heat and transfer the heat to the working fluid flowing through them.
Nanoparticle shape and type have an effect on the heat transfer of direct absorption solar collectors [54]. At room temperature, the thermal conductivity of nanofluids is 0.4% higher than that of pure water. At the surface of the photothermal conversion test of simulated solar irradiation with an intensity of 1 kW/m2, after irradiation for 3000 s, the temperature of all spherical metal nanofluids was increased by 5 K compared with pure water, and the efficiency was improved by 35%. Nonspherical silver and graphene/silver are more efficient and increase by 35%. Water-based single-walled carbon nanotube nanofluids with a volume concentration of 0.2% can increase the efficiency of vacuum tube solar collectors by 66% [55]. Water-based CuO and Al2O3 nanofluids can improve the thermal efficiency of parabolic trough collectors [56]. Verma [57] evaluated the performance of water-based CuO/MWCNT mixture nanofluids and water-based MgO/MWCNT mixture nanofluids for flat plate collectors. The results show that the optimal operating conditions of the plate collector occur at a mass flow rate of 0.025–0.03 kg/s, and the nanoparticle concentration range is 0.75 to 1.0%. Compared with the base fluid, the energy efficiency of the MgO/MWCNT mixture nanofluid is improved, 25.1%, and compared with water-based MgO nanofluid energy efficiency increased by 16.28%. The performance of the MgO/MWCNT mixture nanofluid is closer to that of the water-based MWCNT nanofluid than the CuO/MWCNT mixture nanofluid.
It can be seen that the application of nanofluids in solar collectors can effectively increase the temperature of the base fluid and have a positive impact on the efficiency of the collector.

2.4. Nanofluids in Other Applications

At the same time, the application and research of nanofluids in batteries [58], energy [59], geothermal energy utilization [60], and chemical aspects [61,62] have gradually increased, and some positive research results have been achieved.
It can be seen from the existing literature that nanofluids are applied to the research of enhanced heat transfer in various fields. However, there are many shortcomings and differences on the boiling heat transfer of nanofluids. Summarize the consensus results of nanofluids currently in the field of boiling heat transfer, and it is necessary to sort out some existing differences. The following is mainly to analyze the effects of nanoparticle type, concentration, base type, and structure of heated surface on boiling critical heat flux and boiling heat transfer coefficient. We put forward ourselves views on the divergence of the nanofluid on the boiling heat transfer coefficient and give ideas and insights for future research.

3. Effects of Nanofluid Physical Properties and Heated Surface on Boiling Heat Transfer

The nanofluid will form a sedimentary layer on the heated surface during the boiling process. The deposited layer will change the number of vaporization core points and the surface wettability, resulting in changes in the growth and development and escape of the bubble. Therefore, the boiling heat transfer performance of nanofluids is mainly affected by the size of nanoparticles, base fluids and heating surfaces. In addition, it affected by the microscopic motion of nanoparticles and different types of base fluids, including the Brownian motion of nanoparticles [63], and the Marangoni effect between different base liquids [64]. The applied electric field [65], magnetic field [66,67], photothermal boiling [68], and pressure [69] also have a certain influence on the boiling heat transfer of the nanofluid. The physical properties of nanofluids are influenced by the type of nanoparticles and base fluids. At present, the research parameters of nanoparticles include the type, particle size and concentration of nanoparticles. The base liquid includes water, alcohol, oil, and a refrigerant. The heated surface contains the effects of material, surface shape and surface roughness. The current research results of nanofluids are summarized in Table 1.

3.1. Effect of Nanoparticles on Boiling Heat Transfer

3.1.1. Effect of Nanoparticle Types on Boiling Heat Transfer

The nanoparticles have a large specific surface area and the thermal conductivity is larger than that of the base liquid. However, the thermal conductivity, specific surface area of the different nanoparticles [106,107], the difference in density between the nanoparticles and the base [85], and the affinity for the base liquid are different [107,108]. At present, the study of boiling heat transfer of nanoparticles includes metal oxides Al2O3 [70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86], Fe2O3 [76,81], Fe3O4 [67,92], CuO [71,87,88,89,90,91], ZnO [80,87,100,101], TiO2 [79,89,103,104,105], nonmetal oxide SiO2 [79,81,98,99], carbon nanotubes [71,76,80,95,96,97], graphene [93,94] etc.
Sarafraz et al. [71] studied the flow boiling heat transfer performance of water-based Al2O3, CuO and functionalized multiwalled carbon nanotube (MWCNT) nanofluids in annulus heat transfer. In terms of boiling heat transfer, the functionalized multiwalled carbon nanotube nanofluid has better and lower deposition resistance than Al2O3 and CuO nanofluids. The density of Al2O3 is much lower than that of CuO, so the difference in density between Al2O3 nanoparticles and water is much smaller than the difference in density between CuO nanoparticles and water. The stability of the water-based Al2O3 nanofluid is better, thereby reducing the scaling thermal resistance of the heated surface, and the thermal conductivity is better than that of the water-based CuO nanofluid. Since the deposition of the nanoparticles on the boiling surface reduces the surface roughness and the number of nucleation sites, so that the bubble transport phenomenon is weakened and the heat transferred by the surface transfer is reduced, which make the heat transfer deteriorated. The boiling heat transfer coefficient curves of water-based Al2O3, CuO, and CNT nanofluids are shown in Figure 5. It can be seen from the figure that adding the nanoparticles to deionized water can enhances the boiling heat transfer performance. CNT nanofluids have the best boiling heat transfer performance, followed by CuO nanofluids, and finally Al2O3 nanofluids.
In short, the starting point temperature of the nucleate boiling, the average bubble diameter, the heat transfer coefficient and the heat flux at the nucleate boiling start point depend on the rate of mass flow of the nanofluid.
Shoghl et al. [80] studied the effect of adding ZnO, Al2O3, and CNT nanoparticles with mass concentration of 0.01% and 0.05% to deionized water on boiling heat transfer performance. The results show that ZnO and Al2O3 nanoparticles deteriorated the boiling heat transfer performance and the CNT enhanced boiling heat transfer performance compared with the deionized water. The reason is that the ZnO and Al2O3 nanoparticles reduce the roughness of the heated surface, while CNT improves the roughness of the heated surface, but at low concentrations the surface deteriorates. Figure 6 shows the different kinds of nanometers deposition surface after the experiment.
Neto [76] et al. studied the boiling heat transfer performance of water-based Fe2O3, Al2O3, and CNTs nanofluids on the copper surface at a volume concentration of 0.02% and 0.1%, and the deposition surface of Fe2O3, Al2O3, and CNTs after experiment on the deionized water. The CHF and Zuber correlation curves [109] and Kindlikar correlation curves [110] of different nanoparticle types were compared. The Zuber correlation curve and the Kindlikar correlation curve are as shown in Equations (4) and (5).
q max = K Z h l v ρ v [ σ g ( ρ l ρ v ) ρ v 2 ] 0.25 ( ρ l ρ l ρ v ) 0.5
q max = h l v ρ v 0.5 [ 1 + cos β 16 ] [ 2 π + π 4 ( 1 + cos β ) cos ϕ ] 0.5 [ σ g ( ρ l ρ v ) ] 0.25
where q max is the maximum heat flux, h l v is the latent heat of vaporization, ρ is the density, σ is the surface tension, g is the acceleration of gravity, and the subscripts l and v represent liquid and vapor, respectively; K Z is the empirical constant ( 0.12 K Z 0.16 ). Kandilkar [110] predicted the CHF of pure liquid saturated pool boiling, taking into account fluid dynamics and nonfluid dynamic effects. β is the power receding contact angle and ϕ is the heating surface inclination angle. The experimentally obtained CHF value, Zuber predicted value, Kandlikar predicted value, and the deviation between them are shown in Table 2.
As can be seen from Table 2, the CHF of the coating formed on the heated surface after boiling on the deionized water is higher than the CHF of the nanofluid on the smooth surface, and because the separation of the nanoparticle layer is enhanced, the critical heat flux is increased [79].
Kim [108] and others studied the boiling heat transfer of water-based Al2O3 and RGO (reduced graphene oxide) nanofluids. The volume concentration ratio of alumina nanoparticles to reduced graphene oxide was 1:1 and the solution volume concentration was 0.001%. The experimental results show that the enhancement of the critical heat flux for mixture nanoparticles reaches 473%, which is 9 times that of alumina nanoparticles and 12–47 times the graphene enhancement. The reason for this is because the deposition of alumina nanoparticles on the surface of the reduced graphene oxide greatly enhances the capillary force of the porous deposition surface.
Park [111] and others studied the boiling heat transfer performance of water-based Al2O3, GNs, and GONs nanofluids on the surface of the heating wire at a volume concentration of 0.001%. The results show that the CHF of Al2O3, GNs, and GONs nanofluids increased by 152%, 84%, and 179%, respectively, compared to deionized water. The authors point out that the CHF enhancement cannot be explained simply by surface wettability and capillary action of nanoparticle deposits. Therefore, it is proposed that the ordered porous structure surface structure of nanoparticle deposition changes the critical unstable wavelength [112], resulting in the critical heat flux increases.
Liter [112] et al. proposed a formula for the influence of modulation wavelength on the critical heat flux of a porous surface, as shown in Equation (6).
q p o r p u s = π 8 h f g ( σ ρ g λ m ) 1 / 2
Figure 7a,b shows the contact angle and the wavelength of the porous surface for the critical heat flux of different types of nanofluids, respectively.
The boiling heat transfer of nanofluids is not only affected by nanoparticle type, but also the thermal conductivity of the base liquid [106]. Because of the density [89] and specific surface area [107], the roughness of deposited topography on heated surface is different during the boiling heat transfer process. The wettability and contact angle of the deposited surface have a great influence on the formation, growth, and detachment of boiling bubbles [82,103]. The type of nanoparticles and the size of the nanoparticles have a significant effect on the boiling heat transfer. Compared with the above literature, the thermal conductivity of nanoparticles is higher and the heat transfer performance is better [71,80]. That is, the better the thermal conductivity of the prepared nanofluids, the better the heat transfer performance. However, in view of the long-term static or high temperature conditions, the stability of the nanofluid is affected, so that the thermal conductivity of the nanofluid is reduced while changing the morphology of the heated surface. The morphology of the heated surface changes the number of boiling vaporization core points and bubble formation and detachment, thereby further affecting the boiling heat transfer performance. It can be concluded that the stronger stability and thermal conductivity of the prepared nanofluid, the stronger the heat transfer. However, when the nanofluid cannot guarantee the long-term stability under the test conditions, the deposition of the nanoparticles on the heated surface causes the influence about the change of the physical properties of the heated surface. Firstly, the different nanoparticles type causes the different morphology and thickness of the deposited layers at the same concentrations. Secondly, the different nanoparticle types cause different wettability of the heating surface, thereby affecting the formation and detachment of bubbles, which include the shape and the diameter of the bubbles. Finally, the different kinds of nanoparticles make the different running speed in base liquid. During the boiling process, the difference in the speed caused by the microscopic movement of the nanoparticles causes the frequency of the disturbance and the separation of the bubbles to be different, thus affecting the heat transfer performance.

3.1.2. Effect of Nanoparticle Concentration on Boiling Heat Transfer

The concentration of the nanoparticles largely determines the deposition morphology of the heated surface. Appropriate nanoparticle concentration can increase the number of vaporization core sites and increase the wettability of the heated surface. If the concentration is too high, the surface roughness of the deposited surface will be reduced or the deposited layer will be too thick, resulting in a decrease in the boiling heat transfer coefficient.
Ham [74], Shahmoradi [78], and others studied the boiling critical heat flux curve and boiling heat transfer coefficient of water-based Al2O3 nanofluids with different concentrations. It can be seen from Figure 8a that the addition of Al2O3 nanoparticles makes the boiling curve move to the right and the heat transfer performance is weakened. As the concentration of nanoparticles increases, the heat transfer performance of Al2O3 nanofluids first decreases, and then increases. It can be seen from Figure 8b that the addition of Al2O3 nanoparticles to deionized water, in addition to moving to the left in boiling curve of the Al2O3 nanofluid with a volume concentration of 0.001%, right with the concentration. However, it can be seen from the literature that the addition of Al2O3 nanoparticles increases the boiling critical heat flux of the base liquid. The literature also points out that the increase of the critical heat flux of nanoparticles is due to the enhanced wettability of the porous surface formed by the deposition of Al2O3 nanoparticles on the heated surface, and the increase of the wettability causes the bubble contact angle to decrease. The decrease of the boiling heat transfer coefficient is due to the porous deposition layer formed by the deposition of the nanoparticles, which increases the thermal resistance of the heat conduction, as shown in Figure 9.
Bang [70], Ham [74], Neto [76], and Shahmoradi [78] et al. experimentally studied the boiling heat transfer performance of water-based Al2O3 nanofluids with different concentrations. The literature [76] also studied the boiling heat transfer of water-based CNTs and water-based Fe2O3 nanofluids. Sulaiman [79] et al. studied the boiling heat transfer of water-based SiO2 nanofluids and finally reached the same conclusion. That is, the deposition of the nanoparticles causes an increase in wettability, which in turn increases the critical heat flux. However, the porous deposition layer causes an increase in the thermal resistance of the heat conduction, thereby increasing the wall superheat and reducing the boiling heat transfer coefficient.
Sheikhbahai [92] et al. studied the boiling heat transfer characteristics of ethylene glycol and deionized water mixed with a liquid-based Fe3O4 nanofluid on the Ni–Cr wire surface. The volume concentration was 0.01%, 0.05%, and 0.1%. The volume ratio of ethylene glycol to deionized water is 50:50. The results show that as the concentration increases, the deposition of nanoparticles increases and the critical heat flux increases. At a concentration of 0.1%, the critical heat flux is 100% higher than that of the mixed base liquid. Firstly, the concentration of the nanoparticles increases, resulting in enhanced wettability of the heated surface, as shown in Figure 9. Secondly, the enhanced surface wettability results in an increase in the detachment diameter of the bubble, and the bubble detachment frequency is lowered, as shown in Figure 10.
The study on boiling heat transfer of nanofluids, including Al2O3 nanofluids from Sarafraz [71], Manetti [72], Ahmed [75], Sulaiman [79], Raveshi [84], Diao [85] and others; CuO nanofluids from Shoghl [87], Sarafraz [88], Karimzadehkhouei [89], Umesh [90], Heris [91], and others; as well as GNs [92], Fe3O4 [67], ZnO [101], SiO2 [99], and CNT [71,80,95,97] nanofluids, indicated that nanofluids enhance the boiling heat transfer coefficient. The current explanation for this phenomenon is that nanoparticles increase the number of vaporization core points on the heated surface [67,71,72,84,92] and increase the thermal conductivity of the solution whose stability of the solution enhanced [75,84,86,95]. Decreased surface tension [87,91] leads to a decrease in the diameter of the bubble, which is conducive to detachment of the bubble. While a decrease in the wettability of the heated surface is also advantageous for the detachment of the bubble [101]. The boiling heat transfer coefficient is increased under the same heat flux by reducing the wall superheat. However, Ham [74] pointed out that the boiling heat transfer performance of water-based nanofluids is weaker than that of deionized water. However, because the critical heat flux is faster than the wall superheat, the maximum boiling heat transfer coefficient is enhanced.
It can be seen from the above influence of the nanoparticles concentration on the boiling heat transfer that there have an optimum value of the nanoparticles concentration making the maximum critical heat flux or the boiling heat transfer coefficient. A low concentration of nanoparticles can increase the thermal conductivity of the base liquid, and can increase the number of gasification core points, so that the generation of bubbles increases and the heat transfer is enhanced under the premise of the same bubble overflow frequency. However, high nanoparticle concentration causes a decrease in the operating frequency of the bubbles, which increases liquid viscosity and increases the thickness of the deposited layer. Under the premise of increased roughness of the surface deposited layer, the critical heat flux increases with the increase of the nanoparticles concentration. As the surface roughness begins to decrease as the concentration increases, the critical heat flux begins to decrease. The presence of a certain cavity in the deposited layer formed by the nanoparticles on the heated surface causes the heat to be not dissipated in time, resulting in an increase of wall superheat, so that the critical heat flux increases ratio less or less than the wall superheat, which causes a decrease in the boiling heat transfer coefficient. Therefore, it can be concluded that there is an optimum nanoparticles concentration dispersed in the base liquid, which can increase the number of boiling vaporization core points and the microscopic motion disturbance of the bubbles. At the same time, the formation of the nanoparticle deposition cavity does not cause the wall superheat to rise too fast. Thereby the maximum boiling heat transfer coefficient is improved.

3.2. Effect of Base Fluid Type on Boiling Heat Transfer

When the nanoparticles are dispersed in different base fluids with different physical properties, the boiling heat transfer performance of nanofluids is also affected by the type of base fluid in addition to the influence of nanoparticles. Different base fluids exhibit different thermal conductivity [113,114] and different viscosities [115,116] (as shown in Figure 11), resulting in difference in convective heat transfer performance of the nanoparticles, and the bubble generation and escape rates as well as the surface tension. At present, the base fluids include deionized water, organic solvents, and refrigerants. The thermal conductivity of deionized water is higher than that of organic solvents and refrigerants, but some of the high specific surface area nanoparticles have very poor dispersibility in deionized water, so there are many studies on dispersing in organic solvents or refrigerants [93,117], otherwise the nanoparticles are functionalized. Therefore, the influence of the physical parameters of the base liquid on the heat transfer has also received extensive attention and research.
The viscosity increased with MWCNT concentration in the ethylene glycol and deionized water mixed liquid, but the viscosity decreased with increasing temperature [118]. Xu [119], Sarafraz [83], Raveshi [84], and others studied the physical properties and heat transfer properties of ethylene glycol and deionized water mixed liquid-based Al2O3 nanofluids. The viscosity of the base liquid and the charge on the nanoparticles surface determine the degree of aggregation. Moderate aggregation of nanoparticles can make nanofluids more thermally conductive. In addition, it was found that the ethylene glycol and deionized water mixture has higher thermal conductivity than the single ethylene glycol or deionized water-based Al2O3 nanofluid. As the heat flux increases, the boiling heat transfer coefficient increases remarkably. However, as the concentration increases, the surface roughness and heat transfer coefficient deteriorate significantly. There is an optimum concentration to enhance the boiling heat transfer performance. When the optimum volume concentration of the nanoparticles is 0.75%, the boiling heat transfer coefficient is increased by 64%. Kole [100], He [101], and others studied the boiling heat transfer of ethylene glycol-based ZnO nanofluids, ethylene glycol and water mixed liquid-based ZnO nanofluids respectively. The thermal conductivity increased by 40% when the volume concentration of nanoparticles is 3.75% dispersed in ethylene glycol. As can be seen from Figure 12, the volume fraction of 1.6% ZnO can increase the boiling heat transfer coefficient by 22%, and 2.6% ZnO can increase the critical heat flux by 117%. When ZnO nanoparticles with a mass fraction of less than 7.25% are dispersed in the 80EG/20DW (the volume ratio of ethylene glycol and deionized water is 80:20) mixed liquid, the critical heat flux and boiling heat transfer coefficient are enhanced due to the decrease in surface wettability. Hu [93] then studied the boiling heat transfer performance of the graphene nanosheet in the 60EG/40DW mixed liquid. When the graphene concentration is less than 0.02%, critical heat flux increases with the increase of graphene concentration, because the deposition of graphene on the heated surface causes the wettability to increase. When the mass concentration is greater than 0.02%, the critical heat flux remains basically unchanged with the increase of concentration. In addition, SiO2 [120], SiC [116], Fe3O4 [92], CuO [91,121], and TiO2 [122] were also studied with regard to the heat transfer characteristics of nanoparticles in a mixture of ethylene glycol and deionized water. In the same mixture of ethylene glycol and water, as the concentration of nanoparticles increases, the thermal conductivity of the nanofluid increases. But under the same nanoparticle type and concentration, the thermal conductivity of the nanofluid is weakened with the ethylene glycol ratio. In addition, as the nanoparticles concentration and the proportion of ethylene glycol increases, the viscosity of the nanofluid increases. However, as the temperature increases, the viscosity of the nanofluid decreases. The different types of nanoparticles lead to the different thermal conductivity, but the effect on the boiling heat transfer performance depends largely on the deposition morphology of the nanoparticles on the heated surface. A small amount of deposition of nanoparticles increases the number of vaporization core points, and the boiling heat transfer performance is enhanced. The deposition of the nanoparticles increases the roughness of the heated surface, which increases the wettability, resulting in an increase in critical heat flux. The formation of a closed porous deposit of the nanoparticles on the heated surface causes an increase in the heat transfer resistance and a decrease in the boiling heat transfer coefficient.
In view of the divergence of nanofluids in boiling heat transfer, Nikulin [123] found that the addition of Al2O3 nanoparticles to isopropanol resulted in an increase in boiling heat transfer coefficient at low heat flux and a decrease at high heat flux. The effect of nanofluids on the boiling heat transfer coefficient depends on the concentration of the nanoparticles and the boiling temperature. Trisaksri [124] and Naphon [105] investigated the effect of refrigerant R141b-TiO2 nanofluid on the boiling heat transfer performance. The addition of nanoparticles deteriorated the boiling heat transfer coefficient and deteriorated severely with the increase of nanoparticle concentration. It can be seen that the influence of nanoparticles on the boiling heat transfer performance still differs, but it is certain that the deposition of nanoparticles on the heated surface changes the wettability, resulting in an increase in the critical heat flux.
At present, the effect of nanofluids on boiling heat transfer is mainly aimed at the increase of bubble activation points caused by the deposition of nanoparticles. Therefore, it is necessary to investigate the effect of the nanoparticle deposition surface on the boiling heat transfer performance. Nanoparticle deposition on a smooth surface increases the density of vaporized core points and increases the boiling heat transfer [110]. The deposited layer of nanoparticles on the heated surface will still have partial peeling during the boiling process. As the mass of the nanoparticles in the exfoliated layer increases, the boiling heat transfer coefficient decreases [125].
The difference in the base fluid makes a large difference in the physical properties of the nanofluid. The viscosity of the base fluid is greater, while the viscosity of the nanofluid is greater. The intense disturbance caused by the generation and detachment of bubbles is the main reason for the strong boiling heat transfer in the nucleus. Therefore, the difference in the nanofluid viscosity caused by the difference in the base fluid is crucial for the influence of boiling bubbles. It is easy to see from the above literature that the addition of nanoparticles to a base liquid having a relatively high viscosity greatly enhances the boiling heat transfer performance. Mainly reasons are that, first, the nanoparticles are dispersed in a solution with a high viscosity, and the stability is well guaranteed. Second, the addition of nanoparticles has little effect on the viscosity of the solution, resulting in a small change in bubble detachment and operation. Finally, the stability of the nanofluid is better, the nanoparticle deposition on the heating surface is less, the number of gasification core points on the heating surface is easily increased, and the wettability of the heating surface is enhanced, so that the boiling heat transfer is strengthened.
At present, the boiling heat transfer performance of nanoparticles dispersed in a multicomponent solution has also received great attention. However, the literature has more to attribute the effect on boiling heat transfer performance to the deposition of nanoparticles on the heated surface, resulting in enhanced wettability and increased surface roughness. It can be seen that the analysis is mostly from the macroscopic point of view, but no one has studied the effect of the change of physical properties and interaction of binary solution on boiling heat transfer. The current boiling heat transfer devices basically have a condensing tube located inside the evaporation chamber and disposed above the solution. However, the difference in physical properties of the binary solution makes the boiling evaporation temperature different, and the binary solution in the evaporation chamber will be unevenly distributed during the boiling process. At this time, the interaction between the binary solutions becomes an important factor affecting the boiling heat transfer performance. As a carrier of nanoparticles, the binary solution accelerates the movement of the nanoparticles and causes a substantial change in heat transfer.

3.3. Effect of Heated Surface on Boiling Heat Transfer of Nanofluids

The effect of the roughness of the heated surface on the boiling heat transfer performance is to increase the heat transfer surface area, thereby increasing the influence of the vaporization core on the boiling heat transfer. However, the different heating surfaces and the surface roughness are also different for the heat transfer performance of the nanofluid. In addition, in order to further explore the influence of nanofluids on boiling heat transfer, it is necessary to study the boiling heat transfer of the pure base liquid on the deposition surface formed by the nanofluid boiling heat transfer test.
Compared with deionized water, Al2O3 nanoparticles with a volume concentration of 0.001% can enhance the boiling heat transfer performance of a smooth surface with an average roughness of 25 nm, but the effect on the heat transfer of a rough surface with an average roughness of 420 nm is not obvious [52]. Ham [53] et al. further promoted the boiling heat transfer of water-based Al2O3 nanofluids on different roughness heating surfaces. The volume concentration was 0–0.05% and the heating surface roughness was 177.5 nm and 292.8 nm, respectively. The results show that when the concentration is increased from 0% to 0.05%, the critical heat flux increases by 224.8% and 138.5% on the surfaces of roughness 177.5 nm and 292.8 nm, respectively. The boiling heat transfer performance of Al2O3 nanofluid with a concentration of 0.05% is lower than that of deionized water at Ra = 177.5 nm, but the maximum boiling heat transfer coefficient increases due to the increase of critical heat flux. The weakening of the boiling heat transfer coefficient is mainly caused by the deposition of nanoparticles on the heated surface, which increases the heat transfer resistance, but the wettability of the deposited surface is enhanced and the critical heat flux is increased.
By welding the foam metal on a smooth surface, the boiling heat transfer surface area can be increased. In addition, there is an influence on bubble growth and detachment during boiling process. In order to further enhance the boiling heat transfer of the foam metal surface, Xu et al. [126] investigated the effect of nanoparticles concentration and nanoparticles size on the boiling heat transfer of copper foam. The nanoparticles were Al2O3 and SiC. The foam copper thickness was 7 mm; the pore density was 5 PPI, 60 PPI, 100 PPI; and porosity was 0.9, 0.95, and 0.98. Preliminary studies have found that the addition of nanoparticles can enhance the boiling heat transfer performance of low porosity foamed copper, and the deposition of nanoparticles on the surface of foamed copper can increase the capillary force of copper foam, in addition to increasing the number of vaporization core points. In addition, the boiling heat transfer performance of nanofluids on the gradient hole surface was also studied, because the addition of nanoparticles blocked the voids of high density foam copper (100 PPI), and the bubble escape resistance increased, so that the heat transfer performance decreased [127]. Karthikeyan et al. [128] found that the ethanol-based MWCNT nanofluids increased the boiling heat transfer rate by 59% on the surface of the disk with laser inscribed square column texture.
It can be seen that the boiling heat transfer performance of nanofluids is affected not only by the nanofluids themselves, but also by the heating surface. Besides the roughness of the heated surface, the ratio of the average particle size to the roughness of the heated surface is also related to the boiling heat transfer performance. Using larger ratios, the deposition of nanoparticles can further increase the roughness of the heated surface, and, on the contrary, reduce the roughness. However, the boiling heat transfer performance will be significantly deteriorated by the formation of porous deposits on the surface of nanoparticles with high concentration.
Since the boiling heat transfer strength of the nanofluid is related to the number of boiling vaporization core points. So the boiling heat transfer performance of the nanoparticle deposition surface becomes critical after the boiling experiment in the pure base liquid. Dadjoo et al. [129] examined the boiling heat transfer of water-based SiO2 nanofluids and nanoparticle deposition surfaces in deionized water, and also investigated the effect of heating surface inclination on boiling heat transfer, as shown in Figure 13. It can be seen that when the inclination angle is 0 degrees, the boiling heat transfer curve of the deionized water on the smooth surface, the water-based SiO2 nanofluid on the smooth surface and the deionized water on the deposition surface is sequentially moved to the left. Nanofluids have the highest critical heat flux, followed by the deposition surface, and finally the smooth surface.
It can be seen that the boiling heat transfer of the nanofluid is strong, and besides the physical properties of the nanoparticle itself, it is also affected by the heated surface. The most intuitive effect of the modification of the heated surface is to increase the heat transfer area, thereby increasing the number of vaporization core points. From the indirect point of view, it is because the modification of the heating surface changes the number of boiling bubbles generated and the operating state in addition to increasing the number of gasification core points. Therefore, the nanoparticles have a larger deposition surface, and a higher concentration of the nanofluid solution can be used, so that the thermal conductivity of the base liquid increases greatly and the nanoparticle does not rapidly deteriorate due to the porous deposition layer formed on the heated surface. In addition, the bubbles generated by the irregular surface change the movement form of the bubble due to other forces such as capillary force, and are not singly subjected to upward buoyancy.

4. Effect of Nanofluid on Critical Heat Transfer and Boiling Heat Transfer Coefficient

This paper discusses the boiling heat transfer performance from the four aspects including nanoparticle types, concentrations, base fluids, and heating surfaces. Considering the divergence of nanofluids on boiling heat transfer performance, the differences between the enhancement of critical heat flux and the boiling heat transfer coefficient of nanofluids are considered. The addition of nanoparticles can enhance the critical heat flux, and this conclusion is well supported. The strengthening factor is mainly due to the following two reasons.
(1)
The addition of nanoparticles increases the thermal conductivity of the base fluid.
(2)
The deposition on the heated surface during boiling process reduces the contact angle and changes the wettability, so that the liquid is quickly replenished to the dry spots formed after the bubbles are detached.
When the concentration of the nanoparticles is too large and the heating surface roughness begins to decrease, the critical heat flux no longer increases and begins to deteriorate gradually. In addition, the deposition morphology of different nanoparticles increases the capillary force and enhances the liquid replenishing ability after the bubbles are detached.
From the current research literature, it is concluded that the influence of nanofluids on the boiling heat transfer coefficient is mainly affected by the following factors.
(1)
The influence of the deposition morphology of the nanoparticles on the heated surface.
(2)
The size of the bubble diameter.
(3)
Liquid convection effect on the heated surface.
It is certain that the higher concentration nanoparticles deteriorates the boiling heat transfer coefficient. Since there are many cavities in the deposited layer formed on the heated surface and the contact thermal resistance increases. It can be known from the boiling theory that the amount of bubble detachment from the heated surface determines the boiling heat transfer coefficient in a certain period of time. The number of gasification core points on nano-heated surface plays an important role, so the boiling heat transfer coefficient can be enhanced by grooving [130,131,132] and polishing [129]. In addition, the diameter and detachment state of the bubble are changed by changing the pressure of the boiling environment, and the addition of magnetic fields, electric fields, and gravitational fields enhances the microscopic motion of the nanoparticles. The frequency of the bubble’s detachment determines how quickly the heat on the heated surface is carried away. To some extent, the external field changes the perturbation effect of the microscopic motion of the nanoparticles on the boiling bubbles, thereby enhancing the detachment of the bubbles and enhancing the heat transfer. The low pressure effect is beneficial to the detachment of bubbles [85]. In addition to making the nanofluid more stable, the surfactant can reduce the surface tension, so that the bubble detachment diameter is reduced, the heated surface cannot easily form a film, and the boiling heat transfer performance is enhanced [87,88].
The reasons for the current boiling heat transfer coefficient are presumed to include the following factors.
(1)
The thermal conductivity of nanoparticles compared to heated surfaces.
(2)
Bubble detachment.
(3)
Microscopic motion between multiple solutions.
Copper blocks are used as electronic components with high thermal conductivity, and there are relatively many studies on heating surfaces. However, the thermal conductivity of metal oxides such as Al2O3, CuO, and Fe2O3 is much lower than that of copper, resulting in an increase in thermal resistance at the deposition point of the heated surface and the heat cannot be taken away in time, which increased the wall superheat. However, since the specific surface area of the nanoparticles is large, although the thermal resistance at the deposition point with the heating surface is increased, it is advantageous for the nanoparticles to be in contact with the heating surface and the foaming of the nanoparticles themselves. This greatly increases the number of vaporization core points and also changes the shape of the bubble generation and detachment [91]. During bubble growth, most of the heat is absorbed from the heated surface, and a small portion of the heat is absorbed from the nanoparticles deposited on the heated surface, while the nanoparticles absorb heat at a slower rate than the heated surface to slow the growth of the bubbles. When the thermal conductivity of the nanoparticles is higher than that of the heated surface, the rate of bubble growth is increased because the bubbles absorb heat from the deposited nanoparticles at a higher rate than the heated surface. Therefore, compared to conductivity of the heated surface, the boiling heat transfer of the nanoparticle with the high thermal conductivity becomes critical. However, current high thermal conductivity nanomaterials, including graphene (500–5000 W/(m2·K)), multiwalled carbon nanotubes (6000 W/(m2·K)), and other materials, but they all have a commonality: the density is small, resulting in a specific surface area that is too large. In the boiling process, graphene nanosheets are likely to cover the heated surface in sheet shape, resulting in a cavity between the nanosheets and the covered surface. In addition, because the specific surface base is too large, the boiling speed is relatively small, which restrains the escape frequency of bubbles to a certain extent, thus seriously deteriorating the boiling heat transfer. Therefore, the stability study of graphene nanofluids will not be neglected. At the same concentration, the deposition area of the MWCNT on the heated surface is small, and the boiling heat transfer performance is easily improved [72,80,95,97]. However, it can be seen from the literature [96] that the boiling heat transfer performance of the ethylene glycol and deionized water mixed liquid-based graphene nanosheet nanofluid is improved compared with the pure base liquid because the surface roughness caused by the deposition of the nanoparticles is increased.
In addition, we conclude that the microscopic motion of the multisolution nanofluid base fluid has a significant effect on heat transfer. Firstly, the Marangoni effect [64] and the thermophoresis [133] effect of nanoparticles are enhanced due to the enhancement of the density difference between the base liquids. Secondly, the micromotion of the multisolution during the boiling process causes the disturbance of bubbles on the heated surface to increase. Finally, the micromotion of the nanoparticles makes the flow of the heated surface liquids quicker, which is conducive to the heat being taken away, heat transfer performance is enhanced.
Therefore, it can be seen that when the thermal conductivity and the fluidity of the base liquid are strong, the strength of the boiling heat transfer of the nanofluid is mainly influenced by the thermal conductivity and concentration of the nanoparticle and the shape of the heated surface. When the base liquid is an alcohol with weak fluidity, the viscosity of the base is small, and the influence of the addition of the nanoparticles on the viscosity of the base liquid is large. In addition, the dispersibility and the stability of the nanoparticles in the high viscosity solution are better than lower viscosity base liquid. Therefore, the deposition of the nanoparticles on the heated surface is small, so that the enhancement of the thermal conductivity of the base liquid is stronger than the microscopic movement of the heated surface nanoparticles, and the boiling heat transfer performance is enhanced. When the base liquid is a multicomponent solution, the relative movement between the solutions enhances the microscopic motion of the nanoparticles due to the different evaporation order during the boiling process. Therefore, the disturbance of the bubble on the heating surface is enhanced, and the boiling heat transfer performance is enhanced.

5. The Direction of Nanofluid Research

At present, there are still many scholars who have conducted in-depth research on the boiling heat transfer performance of nanofluids. The innovative ideas include selecting nanoparticles with high thermal conductivity and increasing their properties by surface functionalization or dispersion in a base liquid with a high viscosity to improve boiling heat transfer performance. In terms of analytical methods, the following research methods are included.
(1)
Scanning electron microscopy is used to analyze the sedimentary morphology and atomic force microscopy is used to analyze the roughness of the deposited surface.
(2)
Analysis of the generation and detachment of boiling bubbles by high-speed cameras.
(3)
Analysis of the wettability of the heated surface.
However, through the analysis in this paper, in terms of the thermal conductivity of the heated surface, the deposition of the low thermal conductivity nanoparticles on the heated surface reduces the heat dissipation rate of the heated surface and improves the wall superheat of the heated surface [70,78]. Then, the enhancement of the boiling heat transfer coefficient of the nanofluid should be attributed to the thermal conductivity of the nanoparticle to the base fluid and the influence of the microscopic motion of the nanoparticle on the disturbance of the heated surface bubble. The nanometer particles can be appropriately labeled, and the compositional analyzer can be used to observe the motion path during the boiling process. The influence of the microscopic motion of the nanoparticles on the boiling heat transfer is deeply studied and analyzed.
In addition, because the surface of the nanoparticle is deposited on the heating surface to change the bubble generation state, it is necessary to use a high-speed camera to observe and analyze the bubble generation state around the deposition of a single nanoparticle. Raza [134] first experimentally studied the surface wettability-independent critical heat flux during boiling crisis with foaming solutions. Foaming solution refers to those aqueous surfactant solutions that are very effective in avoiding bubble coalescence and form vapor foam. The slowly rising small bubbles in the foam crowd the heater surface to inhibit rewetting and trigger boiling crisis. Similar to the typical relation between the terminal velocity and the bubble size, we observe a power law exponent of half between the CHF and the bubble size. Such a behavior suggests that the ability of buoyancy to remove bubble swarm away from the heater surface dictates the CHF during boiling with foaming solutions. Resulting premature dry-out not only reduces CHF significantly in comparison to boiling with pure water, but also renders the effect of wettability improvements from micro-/nanotexturing inconsequential to CHF enhancement. Terminal velocity of the rising bubbles is estimated to model the maximum vapor-removal capacity and successfully predict the CHF over a wide range of concentrations. Weakly foaming surfactants, or the strongly foaming surfactants but at lower concentrations, behave similar to water wherein they form large bubbles and wettability improvements via surface modifications exhibit CHF enhancements. The physical insights gained in this study can now be used to devise strategies for CHF enhancement with aqueous surfactant solutions. High-speed visualization of the nucleate boiling process illustrated a fundamental difference in bubble dynamics (Figure 14). It can be seen from the Figure 14, comparison to water and Tween 80, bubbles relatively large in number and small in size were formed with aqueous SDS and Triton X-100 solutions. Boiling videos further suggested that the size and the number of bubbles correlated with the bubble coalescence properties of these solutions. For example, pure water allows the formation of relatively large bubbles, but a few in numbers, due to significant coalescence. Similarly, coalescence also prevails in Tween-80 solution and hence the formation of large bubbles. On the contrary, nucleating bubbles in SDS and Triton X-100 solutions rarely merged/coalesced. Accordingly, high-density small sized bubbles formed with these foaming solutions.
Bottom view images of vapor bubbles at different subcooling and heat fluxes, for aqueous solutions and SDS at critical micelle concentration, during pool boiling on an inverted heater as shown in Figure 15. From the Figure 15, it can be suggested that the effective bubble size was found to decrease with the increase in subcooling due to enhanced condensation. Subsequently, vapor crowding was decreased and the CHF was found to increase with subcooling for ionic liquid and surfactant.
Kumar [135] et al. found that unlike water the aqueous solution of surface active ionic liquid avoids coalescence to form multiple small bubbles with significantly large wet area on the heater surface. Resulting rewetting of the heater surface increases the critical heat flux to ~950 kW/m2, which is an enhancement of 4 : 5 × in comparison to pure water. It can be seen from the above literature that the formation, size and detachment state of boiling bubbles have a significant influence on the boiling heat transfer performance. So the influence of nanofluids on boiling heat transfer can be studied by the following means.
(1)
By disposing the nanoparticles with different thermal conductivity on the heated surface and not escaping due to the severity boiling during the boiling process. The effects on the critical heat flux and boiling heat transfer coefficient in the pure solution are investigated and the bubbles around a single nanoparticle are observed.
(2)
The current boiling heat transfer devices are all boiling in a large container saturated pool boiling, so the influence of the multicomponent solution on the boiling heat transfer can be observed by externally condensing the device, and the relative movement between the multicomponent solutions can be clearly observed. The motion and the disturbance of the bubble were observed and analyzed, and the effects on the critical heat flux and boiling heat transfer coefficient were investigated.
(3)
By investigating the boiling heat transfer performance of mixed nanoparticles in multisolution, the relative motion between base liquids and the influence of micromotion of different nanoparticles due to their density difference and specific surface difference on the critical heat flux and boiling heat transfer coefficient were analyzed.
(4)
To classify the heat transfer between different nanoparticles and different base liquid nanofluids and to correlate the effects of nanofluids on boiling heat transfer to a general purpose test of boiling critical heat flux and boiling heat transfer coefficient. Subsequent further in-depth research laid the foundation.
The excellent physical properties of nanofluids make it gradually applied to many fields of enhanced heat transfer, but the research and conclusions on its boiling heat transfer performance still lack a lot of experimental verification and convincing mechanism explanation. Therefore, it is necessary to examine the microscopic motion of nanoparticles and the operating state of bubbles from deeper and more diverse studies to reveal the mechanism of strengthening or suppressing heat transfer.

6. Conclusions

Nanofluids are widely used in the future, technology, and military industries. Compared with other applications, nanoparticles have relatively more research and application in heat transfer, and the results are better. However, there are still differences in the effects of the boiling heat transfer performance of nanofluids. This paper analyzes and summarizes the current research literature, and obtains the following conclusions and ideas for the future research of nanofluids in the field of boiling heat transfer.
(1)
The boiling heat transfer performance of nanofluids is affected by the type, concentration of nanoparticles, base type, and heating surface. Nanoparticle deposition causes the wettability and the number of gasification core points on the heated surface to change, which has an effect on the boiling heat transfer performance. In addition, the nanoparticle deposition surface changes the generation state of the boiling bubble and the escape frequency per unit time.
(2)
The nanofluid has a strong strengthening effect on the critical heat flux. It is mainly manifested that when the concentration of the nanoparticles is too large and the heating surface roughness begins to decrease, the critical heat flux no longer increases and begins to deteriorate. In addition, the deposition morphology of different nanoparticles increases the capillary force, thereby enhancing the liquid replenishing ability after the bubble is detached. This in turn increases the critical heat flux.
(3)
The influence of nanofluids on the boiling heat transfer coefficient is still different. The main explanation and analysis is attributed to the formation of a porous deposit on the heated surface of the nanoparticles, which increases the heat transfer resistance of the heated surface and reduces the boiling heat transfer coefficient. In addition, the deposition of nanoparticles on the heated surface causes the critical heat flux to increase faster than superheat, resulting in an increase in the maximum boiling heat transfer coefficient. There is also a large increase in the thermal conductivity of the base liquid to the base liquid, so that the boiling heat transfer coefficient is enhanced. The perturbation of the boiling surface by the nanoparticles increases the thinning of the liquid microlayer on the heated surface, the bubble disturbance is enhanced, and the boiling heat transfer coefficient is enhanced.
(4)
In view of the divergence of the nanofluids on the boiling heat transfer coefficient, it is found that the dispersion of nanoparticles in the base liquid with higher viscosity has a positive effect on the boiling heat transfer performance because the nanoparticles have a great effect on the thermal conductivity of the base liquid, and the stability is good. Therefore, the deposition of the nanoparticles on the heated surface is small, and thus the enhancement of the thermal conductivity of the base liquid is stronger than the microscopic movement of nanoparticles on the heated surface and the boiling heat transfer performance is enhanced. When the base liquid is a multicomponent solution, the relative movement between the solutions enhances the microscopic motion of the nanoparticles due to the different evaporation order during the boiling process. Therefore, the disturbance of the bubble on the heating surface is enhanced, and the boiling heat transfer performance is enhanced.
(5)
Compared with the thermal conductivity of the heated surface, the deposition of the low thermal conductivity nanoparticle on the heated surface reduces the heat dissipation rate and improves the wall superheat of the heated surface. Then the enhancement of the boiling heat transfer coefficient of the nanofluid should be attributed to the thermal conductivity of the nanoparticle to the base fluid and the influence of the microscopic motion of the nanoparticle on the disturbance of the heated surface bubble. The movement path in the boiling process can be observed by using a component analyzer by appropriately marking the nanoparticles. In-depth research and analysis on the effects of microscopic motion of nanoparticles on boiling heat transfer. In addition, because the surface of the nanoparticle is deposited on the heating surface to change the bubble generation state, it is necessary to use a high-speed camera to observe and analyze the bubble generation state around the deposition of a single nanoparticle.

Author Contributions

All the authors have contributed equally to the conception and idea of the paper, evaluating and discussing the review results, and writing and revising this manuscript.

Acknowledgments

The work of this paper was funded by the National Natural Science Foundation of China (51176069).

Conflicts of Interest

The authors declare no conflict of interest.

References and Note

  1. Leidenfrost, J.G. De Aquae Communis Nonnullis Qualitatibus Tractatus. (Ovenius, Duisburg, 1756); transl. Wares, C. On the fixation of water in diverse fire. Int. J. Heat Mass Trans. 1966, 9, 1153–1166. [Google Scholar] [CrossRef]
  2. Jakob, M.; Fritz, W. Versuche ueber den verdampfungsvorgang. Forsch. Geb. Ingenieurwes 1931, 2, 437–447. [Google Scholar]
  3. Jakob, M.; Linke, W. Heat Transfer from a Horizontal Plate. Forsch. Geb. Ingenieurwes 1933, 4, 434. [Google Scholar]
  4. Jacob, M. Heat transfer in evaporation and condensation. Mech. Eng. 1936, 58, 643–660. [Google Scholar]
  5. Kutateladze, S.S. AHydrodynamic Theory of Changes in Boiling Process under Free Convection. Isv Akad Nauk, USSR, Otd Tekh Nauk 1951, 4, 529–935. [Google Scholar]
  6. Kutateladze, S.S. Heat Transfer in Condensation and Boiling. Report No. AEC-tr-3770, 1952. [Google Scholar]
  7. Akhgar, A.; Toghraie, D. An experimental study on the stability and thermal conductivity of water-ethylene glycol/TiO2-MWCNTs hybrid nanofluid: Developing a new correlation. Powder Technol. 2018, 338, 806–818. [Google Scholar] [CrossRef]
  8. Kakavandi, A.; Akbari, M. Experimental investigation of thermal conductivity of nanofluids containing of hybrid nanoparticles suspended in binary base fluids and propose a new correlation. Int. J. Heat Mass Transf. 2018, 124, 742–751. [Google Scholar] [CrossRef]
  9. Aparna, Z.; Michael, M.M.; Pabi, S.K.; Ghosh, S. Diversity in thermal conductivity of aqueous Al2O3- and Ag-nanofluids measured by transient hot-wire and laser flash methods. Exp. Therm. Fluid Sci. 2018, 94, 231–245. [Google Scholar] [CrossRef]
  10. Hamid, K.A.; Azmi, W.H.; Nabil, M.F.; Mamat, R.; Sharma, K.V. Experimental investigation of thermal conductivity and dynamic viscosity on nanoparticle mixture ratios of TiO2-SiO2 nanofluids. Int. J. Heat Mass Transf. 2018, 116, 1143–1152. [Google Scholar] [CrossRef]
  11. Ahn, H.S.; Kim, M.H. The boiling phenomenon of alumina nano-fluid near critical heat flux. Int. J. Heat Mass Transf. 2013, 54, 718–728. [Google Scholar] [CrossRef]
  12. Liu, J.; Lu, S.; Chen, Q.; Zhang, J.H.; Yao, S.G. Stability of Nanofluids for Heat Pipe. Mater. Sci. Forum 2014, 789, 6–11. [Google Scholar] [CrossRef]
  13. Ozsoy, A.; Corumlu, V. Thermal performance of a thermosyphon heat pipe evacuated tube solar collector using silver-water nanofluid for commercial applications. Renew. Energy 2018, 122, 26–34. [Google Scholar] [CrossRef]
  14. Ding, Y.L.; Alias, H.; Wen, D.S.; Williams, R.A. Heat transfer of aqueous suspensions of carbon nanotubes (CNT nanofluids). Int. J. Heat Mass Transf. 2006, 49, 240–250. [Google Scholar] [CrossRef]
  15. Subramani, J.; Nagarajan, P.K.; Mahian, O.; Sathyamurthy, R. Efficiency and heat transfer improvements in a parabolic trough solar collector using TiO2 nanofluids under turbulent flow regime. Renew. Energy 2017. [Google Scholar] [CrossRef]
  16. Park, K.J.; Jung, D.; Shim, S.E. Nucleate boiling heat transfer in aqueous solutions with carbon nanotubes up to critical heat fluxes. Int. J. Multiph. Flow 2009, 35, 525–532. [Google Scholar] [CrossRef]
  17. Jaikumar, A.; Gupta, A.; Kandlikar, S.G.; Yang, C.Y.; Su, C.Y. Scale effects of graphene and graphene oxide coatings on pool boiling enhancement mechanisms. Int. J. Heat Mass Transf. 2017, 109, 357–366. [Google Scholar] [CrossRef]
  18. Kim, K.M.; Bang, I.C. Effects of graphene oxide nanofluids on heat pipe performance and capillary limits. Int. J. Therm. Sci. 2016, 100, 346–356. [Google Scholar] [CrossRef]
  19. Mahian, O.; Kianifar, A.; Kalogirou, S.A.; Pop, I.; Wongwises, S. A review of the applications of nanofluids in solar energy. Int. J. Heat Mass Transf. 2013, 57, 582–594. [Google Scholar] [CrossRef]
  20. Kim, D.E.; Yu, D.I.; Jerng, D.W.; Kim, M.H.; Ahn, H.S. Review of boiling heat transfer enhancement on micro/nanostructured surfaces. Exp. Therm. Fluid Sci. 2015, 66, 173–196. [Google Scholar] [CrossRef]
  21. Yao, S.G.; Teng, Z.C.; Huang, X.G.; Song, Y.D.; Zhang, S.L. Study on Boiling Heat Transfer of Ethylene Glycol/Deionized Water Based Al2O3 Nanofluids Under Different Pressures. Nanosci. Nanotechnol. Lett. 2019, 11, 222–228. [Google Scholar] [CrossRef]
  22. Sarafraz, M.M.; Hormozi, F.; Peyghambarzadeh, S.M.; Vaeli, N. Upward Flow Boiling to DI-Water and Cuo Nanofluids Inside the Concentric Annuli. J. Appl. Fluid Mech. 2015, 8, 651–659. [Google Scholar]
  23. Nikkhah, V.; Sarafraz, M.M.; Hormozi, F. Application of Spherical Copper Oxide (II) Water Nano-fluid as a Potential Coolant in a Boiling Annular Heat Exchanger. Chem. Biochem. Eng. Q. 2015, 29, 405–415. [Google Scholar] [CrossRef]
  24. Sarafraz, M.M.; Arya, H.; Saeedi, M.; Ahmadi, D. Flow boiling heat transfer to MgO-therminol 66 heat transfer fluid: Experimental assessment and correlation development. Appl. Therm. Eng. 2018, 138, 552–562. [Google Scholar] [CrossRef]
  25. Sarafraz, M.M.; Peyghambarzadeh, S.M. Nucleate Pool BoilingHeat Transfer to Al2O3-Water and TiO2-Water Nanofluids on Horizontal Smooth Tubes with DissimilarHomogeneous Materials. Chem. Biochem. Eng. Q. 2012, 26, 199–206. [Google Scholar]
  26. Salari, E.; Peyghambarzadeh, S.M.; Sarafraz, M.M.; Hormozi, F.; Nikkhah, V. Thermal behavior of aqueous iron oxide nano-fluid as a coolant on a flat disc heater under the pool boiling condition. Heat Mass Transf. 2017, 53, 265–275. [Google Scholar] [CrossRef]
  27. Ciloglu, D.; Bolukbasi, A. A comprehensive review on pool boiling of nanofluids. Appl. Therm. Eng. 2015, 84, 45–63. [Google Scholar] [CrossRef]
  28. Liang, G.T.; Mudawar, I. Review of pool boiling enhancement with additives and nanofluid. Int. J. Heat Mass Transf. 2018, 124, 423–453. [Google Scholar] [CrossRef]
  29. Kundan, A.; Plawsky, J.L.; Wayner, P.C., Jr.; Chao, D.F.; Sicker, R.J.; Motil, B.J.; Lorik, T.; Chestney, L.; Eustace, J.; Zoldak, J. Thermocapillary Phenomena and Performance Limitations of a Wickless Heat Pipe in Microgravity. Phys. Rev. Lett. 2015, 114, 146105. [Google Scholar] [CrossRef] [PubMed]
  30. Kravets, V.; Alekseik, Y.; Alekseik, O.; Khairnasov, S.; Baturkin, V.; Ho, T.; Celotti, L. Heat pipes with variable thermal conductance property developed for space applications. In Proceedings of the Joint 18th IHPC and 12th IHPS, Jeju, Korea, 12–16 June 2016. [Google Scholar]
  31. Guptaa, N.K.; Tiwaria, A.K.; Ghosh, S.K. Heat transfer mechanisms in heat pipes using—A review. Exp. Therm. Fluid Sci. 2018, 90, 84–100. [Google Scholar] [CrossRef]
  32. Mashaei, P.R.; Shahryari, M.H.; Fazeli, S.M. Hosseinalipour, Numerical simulation of nanofluid application in a horizontal mesh heat pipe with multiple heat sources: A smart fluid for high efficiency thermal system. Appl. Therm. Eng. 2016, 100, 1016–1030. [Google Scholar] [CrossRef]
  33. Sözen, A.; Menlik, T.; Gürü, M.; Boran, K.; Kılıç, F. A comparative investigation on the effect of fly-ash and alumina nanofluids on the thermal performance of two-phase closed thermo-syphon heat pipes. Appl. Therm. Eng. 2016, 96, 330–337. [Google Scholar] [CrossRef]
  34. Kumaresan, G.; Venkatachalapathy, S.; Godson, L.; Wongwises, S. Comparative study on heat transfer characteristics of sintered and mesh wick heat pipes using CuO nanofluids. Int. Commun. Heat Mass Transf. 2014, 57, 208–215. [Google Scholar] [CrossRef]
  35. Venkatachalapathy, S.; Kumaresan, G.; Suresh, S. Performance analysis of cylindrical heat pipe using nanofluids – An experimental study. Int. J. Multiph. Flow 2015, 72, 188–197. [Google Scholar] [CrossRef]
  36. Chen, Y.; Wang, P.; Liu, Z. Application of water-based SiO2 functionalized nanofluid in a loop thermosyphon. Int. J. Heat Mass Transf. 2013, 56, 59–68. [Google Scholar] [CrossRef]
  37. Gunnasegaran, P.; Zulkifly, M.; Zamri, M. Optimization of SiO2 nanoparticle mass concentration and heat input on a loop heat pipe, Case Stud. Therm. Eng. 2015, 6, 238–250. [Google Scholar]
  38. Goshayeshi, H.R.; Izadi, F.; Bashirnezhad, K. Comparison of heat transfer performance on closed pulsating heat pipe for Fe3O4 and Fe2O3 for achieving an empirical correlation. Phys. E Low-Dimens. Syst. Nanostruct. 2017, 89, 43–49. [Google Scholar] [CrossRef]
  39. Goshayeshi, H.R.; Safaei, M.R.; Goodarzi, M.; Dahari, M. Particle size and type effects on heat transfer enhancement of Ferro-nanofluids in a pulsating heat pipe. Powder Technol. 2016, 301, 1218–1226. [Google Scholar] [CrossRef]
  40. Sadeghinezhad, E.; Mehrali, M.; Saidur, R.; Mehrali, M.; Latibari, S.T.; Akhiani, A.R.; Metselaar, H.S.C. A comprehensive review on grapheme nanofluids: Recent research, development and applications. Energy Convers. Manage. 2016, 111, 466–487. [Google Scholar] [CrossRef]
  41. Tharayil, T.; Asirvatham, L.G.; Ravindran, V.; Wongwises, S. Thermal performance of miniature loop heat pipe with graphene–water nanofluid. Int. J. Heat Mass Transf. 2016, 93, 957–968. [Google Scholar] [CrossRef]
  42. Buffone, C.; Coulloux, J.; Alonso, B.; Schlechtendahl, M.; Palermo, V.; Zurutuza, A.; Albertin, T.; Martin, S.; Molina, M.; Chikov, S.; et al. Capillary pressure in graphene oxide nanoporous membranes for enhanced heat transport in Loop Heat Pipes for aeronautics. Exp. Therm. Fluid Sci. 2016, 78, 147–152. [Google Scholar] [CrossRef]
  43. Zhou, Y.; Cui, X.Y.; Weng, J.H.; Shi, S.Y.; Han, H.; Chen, C.M. Experimental investigation of the heat transfer performance of an oscillating heat pipe with graphene nanofluids. Powder Technol. 2018, 332, 371–380. [Google Scholar] [CrossRef]
  44. Li, S.F.; Bao, Y.Y.; Wang, P.Y.; Liu, Z.H. Effect of nano-structure coating on thermal performance of thermosyphon boiling in micro-channels. Int. J. Heat Mass Transf. 2018, 124, 463–474. [Google Scholar] [CrossRef]
  45. Nazaria, M.A.; Ghasempoura, R.; Ahmadib, M.H.; Heydarian, G.; Shafiic, M.B. Experimental investigation of graphene oxide nanofluid on heat transfer enhancement of pulsating heat pipe. Int. Commun. Heat Mass Transf. 2018, 91, 90–94. [Google Scholar] [CrossRef]
  46. Hosseinian, A.; Isfahani, A.H.M.; Shirania, E. Experimental investigation of surface vibration effects on increasing the stability and heat transfer coeffcient of MWCNTs-water nanofluid in a flexible double pipe heat exchanger. Exp. Therm. Fluid Sci. 2018, 90, 275–285. [Google Scholar] [CrossRef]
  47. Subhedar, D.G.; Ramani, B.M.; Gupta, A. Experimental investigation of heat transfer potential of Al2O3/Water-Mono Ethylene Glycol nanofluids as a car radiator coolant. Case Stud. Therm. Eng. 2018, 11, 26–34. [Google Scholar] [CrossRef]
  48. Contreras, E.M.C.; Oliveira, G.A.; Filho, E.P.B. Experimental analysis of the thermohydraulic performance of graphene and silver nanofluids in automotive cooling systems. Int. J. Heat Mass Transf. 2019, 132, 375–387. [Google Scholar] [CrossRef]
  49. Tijani, A.S.; Sudirman, A.S.B. Thermos-physical properties and heat transfer characteristics of water/ anti-freezing and Al2O3/CuO based nanofluid as a coolant for car radiator. Int. J. Heat Mass Transf. 2018, 118, 48–57. [Google Scholar] [CrossRef]
  50. Chaurasia, P.; Kumar, A.; Yadav, A.; Rai, P.K.; Kumar, V.; Prasad, L. Heat transfer augmentation in automobile radiator using Al2O3– water based nanofluid. SN Appl. Sci. 2019, 257. [Google Scholar] [CrossRef]
  51. Ahmed, S.A.; Ozkaymak, M.; Sözen, A.; Menlik, T.; Fahed, A. Improving car radiator performance by using TiO2-water nanofluid. Eng. Sci. Technol. Int. J. 2018, 21, 996–1005. [Google Scholar] [CrossRef]
  52. Farhana, K.; Kadirgama, K.; Rahman, M.M.; Ramasamy, D.; Noor, M.M.; Najafi, G.; Samykano, M.; Mahamude, A.S.F. Improvement in the performance of solar collectors with nanofluids—A state-of-the-art review. Nano-Struct. Nano-Objects 2019, 18, 100276. [Google Scholar] [CrossRef]
  53. Bellos, E.; Said, Z.; Tzivanidis, C. The use of nanofluids in solar concentrating technologies: A comprehensive review. J. Clean. Prod. 2018, 196, 84–99. [Google Scholar] [CrossRef]
  54. Campos, C.; Vasco, D.; Angulo, C.; Burdiles, P.A.; Cardemil, J.; Palza, H. About the relevance of particle shape and graphene oxide on the behavior of direct absorption solar collectors using metal based nanofluids under different radiation intensities. Energy Convers. Manag. 2019, 181, 247–257. [Google Scholar] [CrossRef]
  55. Mahbubul, I.M.; Khan, M.M.A.; Ibrahim, N.I.; Ali, H.M.; Al-Sulaiman, F.A.; Saidur, R. Carbon nanotube nanofluid in enhancing the efficiency of evacuated tube solar collector. Renew. Energy 2018, 121, 36–44. [Google Scholar] [CrossRef]
  56. Allouhia, A.; Amine, M.B.; Saidur, R.; Kousksou, T.; Jamil, A. Energy and exergy analyses of a parabolic trough collector operated with nanofluids for medium and high temperature applications. Energy Convers. Manag. 2018, 155, 201–217. [Google Scholar] [CrossRef] [Green Version]
  57. Verma, S.K.; Tiwari, A.K.; Tiwari, S.; Chauhan, D.S. Performance analysis of hybrid nanofluids in flat plate solar collector as an advanced working fluid. Sol. Energy 2018, 167, 231–241. [Google Scholar] [CrossRef]
  58. Islam, R.; Shabani, B. Prediction of electrical conducivity of TiO2 water and ethylene glycol-based nanofluids for cooling application in low temperature PEM fuel cells. Energy Procedia 2019, 160, 550–557. [Google Scholar] [CrossRef]
  59. Ahmad, L.; Khan, M. Importance of activation energy in development of chemical covalent bonding in flow of Sisko magneto-nanofluids over a porous moving curved surface. Int. J. Hydrog. Energy 2019, 44, 10197–10206. [Google Scholar] [CrossRef]
  60. Sui, D.; Langåker, V.H.; Yu, Z.X. Investigation of thermophysical properties of Nanofluids for application in geothermal energy. Energy Procedia 2017, 105, 5055–5060. [Google Scholar] [CrossRef]
  61. Hayat, T.; Kiyani, M.Z.; Alsaedi, A.; Khan, M.I.; Ahmad, I. Mixed convective three-dimensional flow of Williamson nanofluid subject to chemical reaction. Int. J. Heat Mass Transf. 2018, 127, 422–429. [Google Scholar] [CrossRef]
  62. Khan, M.I.; Qayyum, S.; Hayat, T.; Khan, M.I.; Alsaedi, A. Entropy optimization in flow of Williamson nanofluid in the presence of chemical reaction and Joule heating. Int. J. Heat Mass Transf. 2019, 133, 959–967. [Google Scholar] [CrossRef]
  63. Kathiravan, R.; Kumar, R.; Gupta, A.; Chandra, R. Characterization and Pool Boiling Heat Transfer Studies of Nanofluids. J. Heat Transf. 2009, 131, 1–8. [Google Scholar] [CrossRef]
  64. Yang, Y.M.; Maa, J.R. Pool boiling of dilute surfactant solutions. J. Heat Transf. 1983, 105, 190–192. [Google Scholar] [CrossRef]
  65. Feng, Z.F.; Luo, X.P.; Zhang, J.X.; Xiao, J.; Yuan, W. Effects of electric field on flow boiling heat transfer in a vertical minichannel heat sink. Int. J. Heat Mass Transf. 2018, 124, 726–741. [Google Scholar] [CrossRef]
  66. Özdemir, M.R.; Sadaghiani, A.K.; Motezakker, A.R.; Paraparib, S.S.; Park, H.S.; Acar, H.Y.; Koşar, A. Experimental studies on ferrofluid pool boiling in the presence of external magnetic force. Appl. Therm. Eng. 2018, 139, 598–608. [Google Scholar] [CrossRef]
  67. Abdollahi, A.; Salimpour, M.R.; Etesami, N. Experimental analysis of magnetic field effect on the pool boiling heat transfer of a ferrofluid. Appl. Therm. Eng. 2017, 111, 1101–1110. [Google Scholar] [CrossRef]
  68. Ulset, E.T.; Kosinski, P.; Zabednova, Y.; Zhdaneev, O.V.; Struchalin, P.G.; Balakin, B.V. Photothermal boiling in aqueous nanofluids. Nano Energy 2018, 50, 339–346. [Google Scholar] [CrossRef]
  69. Yao, S.G.; Huang, X.G.; Song, Y.D.; Shen, Y.; Zhang, S.L. Effects of nanoparticle types and size on boiling feat transfer performance under different pressures. AIP Adv. 2018, 8, 025005. [Google Scholar] [CrossRef] [Green Version]
  70. Bang, I.C.; Chang, S.H. Boiling heat transfer performance and phenomena of Al2O3–water nano-fluids from a plain surface in a pool. Int. J. Heat Mass Transf. 2005, 48, 2407–2419. [Google Scholar] [CrossRef]
  71. Sarafraz, M.M.; Hormozi, F. Comparatively experimental study on the boiling thermal performance of metal oxide and multi-walled carbon nanotube nanofluids. Powder Technol. 2016, 287, 412–430. [Google Scholar] [CrossRef]
  72. Manetti, L.L.; Stephen, M.T.; Beck, P.A.; Cardoso, E.M. Evaluation of the heat transfer enhancement during pool boiling using low concentrations of Al2O3-water based nanofluid. Exp. Therm. Fluid Sci. 2017, 87, 191–200. [Google Scholar] [CrossRef]
  73. Pan, B.G.; Wang, W.H.; Chu, D.L.; Mei, L.Q.; Zhang, Q.H. Experimental investigation of hypervapotron heat transfer enhancement with alumina–water nanofluids. Int. J. Heat Mass Transf. 2016, 98, 738–745. [Google Scholar] [CrossRef]
  74. Ham, J.; Kim, H.; Shin, Y.; Cho, H. Experimental investigation of pool boiling characteristics in Al2O3 nanofluid according to surface roughness and concentration. Int. J. Therm. Sci. 2017, 114, 86–97. [Google Scholar] [CrossRef]
  75. Ahmed, O.; Hamed, M.S. Experimental investigation of the effect of particle deposition on pool boiling of nanofluids. Int. J. Heat Mass Transf. 2012, 55, 3423–3436. [Google Scholar] [CrossRef]
  76. Neto, A.R.; Oliveira, J.L.G.; Passos, J.C. Heat transfer coefficient and critical heat flux during nucleate pool boiling of water in the presence of nanoparticles of alumina, maghemite and CNTs. Appl. Therm. Eng. 2017, 111, 1493–1506. [Google Scholar] [CrossRef]
  77. Vafaei, S. Nanofluid pool boiling heat transfer phenomenon. Powder Technol. 2015, 277, 181–192. [Google Scholar] [CrossRef]
  78. Shahmoradi, Z.; Etesami, N.; Esfahany, M.N. Pool boiling characteristics of nanofluid on flat plate based on heater surface analysis. Int. Commun. Heat Mass Transf. 2013, 47, 113–120. [Google Scholar] [CrossRef]
  79. Sulaiman, M.Z.; Matsuo, D.; Enoki, K.; Okawa, T. Systematic measurements of heat transfer characteristics in saturated pool boiling of water-based nanofluids. Int. J. Heat Mass Transf. 2016, 102, 264–276. [Google Scholar] [CrossRef] [Green Version]
  80. Shoghl, S.N.; Bahrami, M.; Jamialahmadi, M. The boiling performance of ZnO, a-Al2O3 and MWCNTs/water nanofluids: An experimental study. Exp. Therm. Fluid Sci. 2017, 80, 27–39. [Google Scholar] [CrossRef]
  81. Minakov, A.V.; Pryazhnikov, M.I.; Guzei, D.V.; Zeer, G.M.; Rudyak, V.Y. The experimental study of nanofluids boiling crisis on cylindrical heaters. Int. J. Therm. Sci. 2017, 116, 214–223. [Google Scholar] [CrossRef]
  82. Ham, J.; Cho, H. Theoretical analysis of pool boiling characteristics of Al2O3 nanofluid according to volume concentration and nanoparticle size. Appl. Therm. Eng. 2016, 108, 158–171. [Google Scholar] [CrossRef]
  83. Sarafraz, M.M.; Hormozi, F. Nucleate pool boiling heat transfer characteristics of dilute Al2O3–ethyleneglycol nanofluids. Int. Commun. Heat Mass Transf. 2014, 58, 96–104. [Google Scholar] [CrossRef]
  84. Raveshi, M.R.; Keshavarz, A.; Mojarrad, M.S.; Amiri, S. Experimental investigation of pool boiling heat transfer enhancement of alumina–water–ethylene glycol nanofluids. Exp. Therm. Fluid Sci. 2013, 44, 805–814. [Google Scholar] [CrossRef]
  85. Diao, Y.H.; Liu, Y.; Wang, R.; Zhao, Y.H.; Guo, L.; Tang, X. Effects of nanofluids and nanocoatings on the thermal performance of an evaporator with rectangular microchannels. Int. J. Heat Mass Transf. 2013, 67, 183–193. [Google Scholar] [CrossRef]
  86. Tang, X.; Zhao, Y.H.; Diao, Y.H. Experimental investigation of the nucleate pool boiling heat transfer characteristics of d-Al2O3-R141b nanofluids on a horizontal plate. Exp. Therm. Fluid Sci. 2014, 52, 88–96. [Google Scholar] [CrossRef]
  87. Shoghl, S.N.; Bahrami, M. Experimental investigation on pool boiling heat transfer of ZnO, and CuO water-based nanofluids and effect of surfactant on heat transfer coefficient. Int. Commun. Heat Mass Transf. 2013, 45, 122–129. [Google Scholar] [CrossRef]
  88. Sarafraz, M.M.; Hormozi, F. Pool boiling heat transfer to dilute copper oxide aqueous nanofluids. Int. J. Therm. Sci. 2015, 90, 224–237. [Google Scholar] [CrossRef]
  89. Karimzadehkhouei, M.; Shojaeian, M.; Sendur, K.; Mengüç, M.P.; Kosar, A. The effect of nanoparticle type and nanoparticle mass fraction on heat transfer enhancement in pool boiling. Int. J. Heat Mass Transf. 2017, 109, 157–166. [Google Scholar] [CrossRef]
  90. Umesh, V.; Raja, B. A study on nucleate boiling heat transfer characteristics of pentane and CuO-pentane nanofluid on smooth and milled surfaces. Exp. Therm. Fluid Sci. 2015, 64, 23–29. [Google Scholar] [CrossRef]
  91. Heris, S.Z. Experimental investigation of pool boiling characteristics of low-concentrated CuO/ ethylene glycol–water nanofluids. Int. Commun. Heat Mass Transf. 2011, 38, 1470–1473. [Google Scholar] [CrossRef]
  92. Sheikhbahai, M.; Esfahany, M.N.; Etesami, N. Experimental investigation of pool boiling of Fe3O4/ethylene glycolewater nanofluid in electric field. Int. J. Therm. Sci. 2012, 62, 149–153. [Google Scholar] [CrossRef]
  93. Hu, Y.W.; Li, H.R.; He, Y.R.; Wang, L.D. Role of nanoparticles on boiling heat transfer performance of ethylene glycol aqueous solution based graphene nanosheets nanofluid. Int. J. Heat Mass Transf. 2016, 96, 565–572. [Google Scholar] [CrossRef]
  94. Zhang, L.; Fan, L.W.; Yu, Z.T.; Cen, K.F. An experimental investigation of transient pool boiling of aqueous nanofluids with graphene oxide nanosheets as characterized by the quenching method. Int. J. Heat Mass Transf. 2014, 73, 410–414. [Google Scholar] [CrossRef]
  95. Xing, M.B.; Yu, J.L.; Wang, R.X. Effects of surface modification on the pool boiling heat transfer of MWNTs/water nanofluids. Int. J. Heat Mass Transf. 2016, 103, 914–919. [Google Scholar] [CrossRef]
  96. Fan, L.W.; Li, J.Q.; Wu, Y.Z.; Zhang, L.; Yu, Z.T. Pool boiling heat transfer during quenching in carbon nanotube (CNT)- based aqueous nanofluids: Effects of length and diameter of the CNTs. Appl. Therm. Eng. 2017, 122, 555–565. [Google Scholar] [CrossRef]
  97. Amiri, A.; Shanbedi, M.; Amiri, H.; Heris, S.Z.; Kazi, S.N.; Chew, B.T.; Eshghi, H. Pool boiling heat transfer of CNT/water nanofluids. Appl. Therm. Eng. 2014, 71, 450–459. [Google Scholar] [CrossRef]
  98. Bolukbasi, A.; Ciloglu, D. Pool boiling heat transfer characteristics of vertical cylinder quenched by SiO2-water nanofluids. Int. J. Therm. Sci. 2011, 50, 1013–1021. [Google Scholar] [CrossRef]
  99. Hu, Y.W.; Li, H.R.; He, Y.R.; Liu, Z.Y.; Zhao, Y.H. Effect of nanoparticle size and concentration on boiling performance of SiO2 nanofluid. Int. J. Heat Mass Transf. 2017, 107, 820–828. [Google Scholar] [CrossRef]
  100. Kole, M.; Dey, T.K. Investigations on the pool boiling heat transfer and critical heat flux of ZnO-ethylene glycol nanofluids. Appl. Therm. Eng. 2012, 37, 112–119. [Google Scholar] [CrossRef]
  101. He, Y.R.; Li, H.R.; Hu, Y.W.; Wang, X.Z.; Zhu, J.Q. Boiling heat transfer characteristics of ethylene glycol and water mixture based ZnO nanofluids in a cylindrical vessel. Int. J. Heat Mass Transf. 2016, 98, 611–615. [Google Scholar] [CrossRef]
  102. Pioro, I. Boiling heat transfer characteristics of thin liquid layers in a horizontally flat two-phase thermosyphon. In Proceedings of the Preprints of the 10th International Heat Pipe Conference, Stuttgart, Germany, 21–25 September 1997; p. 1. [Google Scholar]
  103. Ali, H.M.; Generous, M.M.; Ahmad, F.; Irfan, M. Experimental investigation of nucleate pool boiling heat transfer enhancement of TiO2-water based nanofluids. Appl. Therm. Eng. 2017, 113, 1146–1151. [Google Scholar] [CrossRef]
  104. Suriyawong, A.; Wongwises, S. Nucleate pool boiling heat transfer characteristics of TiO2–water nanofluids at very low concentrations. Exp. Therm. Fluid Sci. 2010, 34, 992–999. [Google Scholar] [CrossRef]
  105. Naphon, P.; Thongjing, C. Pool boiling heat transfer characteristics of refrigerant-nanoparticle mixtures. Int. Commun. Heat Mass Transf. 2014, 52, 84–89. [Google Scholar] [CrossRef]
  106. Chun, S.Y.; Bang, I.C.; Choo, Y.J.; Song, C.H. Heat transfer characteristics of Si and SiC nanofluids during a rapid quenching and nanoparticles deposition effects. Int. J. Heat Mass Transf. 2011, 54, 1217–1223. [Google Scholar] [CrossRef]
  107. Sadeghinezhad, E.; Mehrali, M.; Rosen, M.A.; Akhiani, A.R.; Latibari, S.T.; Mehrali, M.; Metselaar, H.S.C. Experimental investigation of the effect of graphene nanofluids on heat pipe thermal performance. Appl. Therm. Eng. 2016, 100, 775–787. [Google Scholar] [CrossRef] [Green Version]
  108. Kim, J.H.; Kim, J.M.; Jerng, D.W.; Kim, E.Y.; Ahn, H.S. Effect of aluminum oxide and reduced graphene oxide mixtures on critical heat flux enhancement. Int. J. Heat Mass Transf. 2018, 116, 858–870. [Google Scholar] [CrossRef]
  109. Zuber, N. Hydrodynamic Aspects of Boiling Heat Transfer. Ph.D. Thesis, Research Laboratory, Los Angeles and Ramo-Wooldridge Corporation, University of California, Los Angeles, CA, USA, 1959. [Google Scholar]
  110. Kandlikar, S.G. A theoretical model to predict pool boiling CHF incorporating effects of contact angle and orientation. J. Heat Transf. 2001, 123, 1071–1079. [Google Scholar] [CrossRef]
  111. Park, S.D.; Lee, S.W.; Kang, S.; Bang, I.C.; Kim, J.H.; Shin, H.S.; Lee, D.W.; Lee, D.W. Effects of nanofluids containing graphene/graphene-oxide nanosheets on critical heat flux. Appl. Phys. Lett. 2010, 97, 023103. [Google Scholar] [CrossRef] [Green Version]
  112. Liter, S.G.; Kaviany, M. Pool-boiling CHF enhancement by modulated porous-layer coating: Theory and experiment. Int. J. Heat Mass Transf. 2001, 44, 4287–4311. [Google Scholar] [CrossRef]
  113. Nabil, M.F.; Azmia, W.H.; Hamida, K.A.; Mamata, R.; Hagos, F.Y. An experimental study on the thermal conductivity and dynamic viscosity of TiO2-SiO2 nanofluids in water: Ethylene glycol mixture. Int. Commun. Heat Mass Transf. 2017, 86, 181–189. [Google Scholar] [CrossRef]
  114. Chiam, H.W.; Azmi, W.H.; Usri, N.A.; Mamat, R.; Adam, N.M. Thermal conductivity and viscosity of Al2O3 nanofluids for different based ratio of water and ethylene glycol mixture. Exp. Therm. Fluid Sci. 2017, 81, 420–429. [Google Scholar] [CrossRef]
  115. Kole, M.; Dey, T.K. Investigation of thermal conductivity, viscosity, and electrical conductivity of graphene based nanofluids. J. Appl. Phys. 2013, 113, 084307. [Google Scholar] [CrossRef]
  116. Li, X.K.; Zou, C.J. Thermo-physical properties of water and ethylene glycol mixture based SiC nanofluids: An experimental investigation. Int. J. Heat Mass Transf. 2016, 101, 412–417. [Google Scholar] [CrossRef]
  117. Seo, H.; Chu, J.H.; Kwon, S.Y.; Bang, I.C. Pool boiling CHF of reduced graphene oxide, graphene, and SiC-coated surfaces under highly wettable FC-72. Int. J. Heat Mass Transf. 2015, 82, 490–502. [Google Scholar] [CrossRef]
  118. Rahmati, A.R.; Reiszadeh, M. An experimental study on the effects of the use of multi-walled carbon nanotubes in ethylene glycol/water-based fluid with indirect heaters in gas pressure reducing stations. Appl. Therm. Eng. 2018, 134, 107–117. [Google Scholar] [CrossRef]
  119. Xu, J.H.; Bandyopadhyay, K.; Jung, D. Experimental investigation on the correlation between nano-fluid characteristics and thermal properties of Al2O3 nano-particles dispersed in ethylene glycol–water mixture. Int. J. Heat Mass Transf. 2016, 94, 262–268. [Google Scholar] [CrossRef]
  120. Guo, Y.F.; Zhang, T.T.; Zhang, D.R.; Wang, Q. Experimental investigation of thermal and electrical conductivity of silicon oxide nanofluids in ethylene glycol/water mixture. Int. J. Heat Mass Transf. 2018, 117, 280–286. [Google Scholar] [CrossRef]
  121. Ahmadi, M.A.; Ahmadi, M.H.; Alavi, M.F.; Nazemzadegan, M.R.; Ghasempour, R.; Shamshirband, S. Determination of thermal conductivity ratio of CuO/ethylene glycol nanofluid by connectionist approach. J. Taiwan Inst. Chem. Eng. 2018, 91, 383–395. [Google Scholar] [CrossRef]
  122. Sandhya, D.; Reddy, M.C.S.; Rao, V.V. Improving the cooling performance of automobile radiatorwith ethylene glycol water based TiO2 nanofluids. Int. Commun. Heat Mass Transf. 2016, 78, 121–126. [Google Scholar]
  123. Nikulin, A.; Khliyeva, O.; Lukianov, N.; Zhelezny, V.; Semenyuk, Y. Study of pool boiling process for the refrigerant R11, isopropanol and isopropanol/Al2O3 nanofluid. Int. J. Heat Mass Transf. 2018, 118, 746–757. [Google Scholar] [CrossRef]
  124. Trisaksri, V.; Wongwises, S. Nucleate pool boiling heat transfer of TiO2–R141b nanofluids. Int. J. Heat Mass Transf. 2009, 52, 1582–1588. [Google Scholar] [CrossRef]
  125. Watanabe, Y.; Enoki, K.; Okawa, T. Nanoparticle layer detachment and its influence on the heat transfer characteristics in saturated pool boiling of nanofluids. Int. J. Heat Mass Transf. 2018, 125, 171–178. [Google Scholar] [CrossRef]
  126. Xu, Z.G.; Zhao, C.Y. Influences of nanoparticles on pool boiling heat transfer in porous metals. Appl. Therm. Eng. 2014, 65, 34–41. [Google Scholar] [CrossRef]
  127. Xu, Z.G.; Zhao, C.Y. Enhanced boiling heat transfer by gradient porous metals in saturated pure water and surfactant solutions. Appl. Therm. Eng. 2016, 100, 68–77. [Google Scholar] [CrossRef]
  128. Karthikeyan, A.; Coulombe, S.; Kietzig, A.M. Boiling heat transfer enhancement with stable nanofluids and laser textured copper surfaces. Int. J. Heat Mass Transf. 2018, 126, 287–296. [Google Scholar] [CrossRef]
  129. Dadjoo, M.; Etesami, N.; Esfahany, M.N. Influence of orientation and roughness of heater surface on critical heat flux and pool boiling heat transfer coefficient of nanofluid. Appl. Therm. Eng. 2017, 124, 353–361. [Google Scholar] [CrossRef]
  130. Sun, Y.L.; Chen, G.; Zhang, S.W.; Tang, Y.; Zeng, J.; Yuan, W. Pool boiling performance and bubble dynamics on microgrooved surfaces with reentrant cavities. Appl. Therm. Eng. 2017, 125, 432–442. [Google Scholar] [CrossRef]
  131. Gheitaghy, A.M.; Samimi, A.; Saffari, H. Surface structuring with inclined minichannels for pool boiling improvement. Appl. Therm. Eng. 2017, 126, 892–902. [Google Scholar] [CrossRef]
  132. Xu, Z.G.; Qu, Z.G.; Zhao, C.Y.; Tao, W.Q. Experimental correlation for pool boiling heat transfer on metallic foam surface and bubble cluster growth behavior on grooved array foam surface. Int. J. Heat Mass Transf. 2014, 77, 1169–1182. [Google Scholar] [CrossRef]
  133. Oyarzua, E.; Waltherbc, J.H.; Zambrano, H.A. Water thermophoresis in carbon nanotubes: The interplay between thermophoretic and friction forces. Phys. Chem. Chem. Phys. 2018, 20, 3672–3677. [Google Scholar] [CrossRef]
  134. Raza, M.Q.; Kumar, N.; Raj, R. Wettability-independent critical heat flux during boiling crisis in foaming solutions. Int. J. Heat Mass Transf. 2018, 126, 567–579. [Google Scholar] [CrossRef]
  135. Kumar, N.; Raza, M.Q.; Seth, D.; Raj, R. Aqueous ionic liquid solutions for boiling heat transfer enhancement in the absence of buoyancy induced bubble departure. Int. J. Heat Mass Transf. 2018, 122, 354–363. [Google Scholar] [CrossRef]
Figure 1. Boiling curve [20].
Figure 1. Boiling curve [20].
Applsci 09 02818 g001
Figure 2. Choice of optimal wick parameters [30].
Figure 2. Choice of optimal wick parameters [30].
Applsci 09 02818 g002
Figure 3. Heat pipe schematic.
Figure 3. Heat pipe schematic.
Applsci 09 02818 g003
Figure 4. Application of nanofluids in solar collectors.
Figure 4. Application of nanofluids in solar collectors.
Applsci 09 02818 g004
Figure 5. Influence of heat flux on HTC of different nanofluids in forced convection and nucleate boiling regions [71].
Figure 5. Influence of heat flux on HTC of different nanofluids in forced convection and nucleate boiling regions [71].
Applsci 09 02818 g005
Figure 6. Field-emission scanning electron microscopy (FESEM) pictures of applied nanoparticles, (a) ZnO; (b) Al2O3, and SEM of (c) carbon nanotubes (CNTs) [80].
Figure 6. Field-emission scanning electron microscopy (FESEM) pictures of applied nanoparticles, (a) ZnO; (b) Al2O3, and SEM of (c) carbon nanotubes (CNTs) [80].
Applsci 09 02818 g006
Figure 7. CHF phenomena and comparisons between models and experimental data. (a) Effects of a surface wettability (Kandlikar’s model with contact angle). (b) Effects of a geometrically determined critical instability wavelength [111].
Figure 7. CHF phenomena and comparisons between models and experimental data. (a) Effects of a surface wettability (Kandlikar’s model with contact angle). (b) Effects of a geometrically determined critical instability wavelength [111].
Applsci 09 02818 g007
Figure 8. Boiling heat transfer curves and boiling heat transfer coefficient curves of different nanofluids. (a) Al2O3 nanofluid boiling curve [74]. (b) Al2O3 nanofluid boiling curve [78].
Figure 8. Boiling heat transfer curves and boiling heat transfer coefficient curves of different nanofluids. (a) Al2O3 nanofluid boiling curve [74]. (b) Al2O3 nanofluid boiling curve [78].
Applsci 09 02818 g008
Figure 9. Schematic diagram of the deposition of nanoparticles on a heated surface [78].
Figure 9. Schematic diagram of the deposition of nanoparticles on a heated surface [78].
Applsci 09 02818 g009
Figure 10. These two aspects cause a decrease in the boiling heat transfer coefficient.
Figure 10. These two aspects cause a decrease in the boiling heat transfer coefficient.
Applsci 09 02818 g010
Figure 11. Thermal conductivity and viscosity diagram of ethylene glycol water-based GNs nanofluids with temperature [115].
Figure 11. Thermal conductivity and viscosity diagram of ethylene glycol water-based GNs nanofluids with temperature [115].
Applsci 09 02818 g011
Figure 12. Boiling curves for ZnO-EG/DW nanofluids and EG/DW mixtures. (a) Boiling heat transfer curve. (b) Boiling heat transfer coefficient [101].
Figure 12. Boiling curves for ZnO-EG/DW nanofluids and EG/DW mixtures. (a) Boiling heat transfer curve. (b) Boiling heat transfer coefficient [101].
Applsci 09 02818 g012
Figure 13. Boiling curves of deionized water on nanocoated and uncoated surfaces and comparing with 0.0025 vol.% nanofluid at various inclination angles of the heater [129].
Figure 13. Boiling curves of deionized water on nanocoated and uncoated surfaces and comparing with 0.0025 vol.% nanofluid at various inclination angles of the heater [129].
Applsci 09 02818 g013
Figure 14. (a) Snapshot of bubbles during near-saturated boiling with aqueous surfactant solutions at critical micelle concentration and pure water in the early nucleate boiling regime. (b) Schematic representation of the bubble behavior during pool boiling. Relatively large bubbles with higher terminal velocity were formed with water in comparison to surfactant solution [134].
Figure 14. (a) Snapshot of bubbles during near-saturated boiling with aqueous surfactant solutions at critical micelle concentration and pure water in the early nucleate boiling regime. (b) Schematic representation of the bubble behavior during pool boiling. Relatively large bubbles with higher terminal velocity were formed with water in comparison to surfactant solution [134].
Applsci 09 02818 g014
Figure 15. Bottom view images of vapor bubbles at different subcooling and heat fluxes, for aqueous solutions and SDS at critical micelle concentration, during pool boiling on an inverted heater [135].
Figure 15. Bottom view images of vapor bubbles at different subcooling and heat fluxes, for aqueous solutions and SDS at critical micelle concentration, during pool boiling on an inverted heater [135].
Applsci 09 02818 g015
Table 1. Nanofluid boiling heat transfer.
Table 1. Nanofluid boiling heat transfer.
NanoparticlesBase LiquidConcentrationHeated SurfaceCHF/HTCReasons
Al2O3DW0.5, 1, 2, 4 wt.%copper surfaceCHF enhancement
HTC deterioration
CHF enhancement is due to surface characteristics of nanoparticle deposition[70]
0.1, 0.2, 0.3 wt.%stainless steel rodCHF enhancement
HTC enhancement
Nanoparticle deposition causes changes in surface roughness and vaporized core points.[71]
0.0007, 0.007 Vol.%copper blockHTC enhanced 15–75% (0.0007%)HTC is affected by surface roughness.[72]
0.005, 0.05, 0.01, 0.1 wt.%steel plateCHF enhancementThe nanoparticles accelerate the liquid perturbation, increasing the frequency of bubble detachment.[73]
0.001, 0.01, 0.05, 0.1 wt.%copper surfaceCHF enhancement
HTC deterioration
It subject to surface roughness and concentration.[74]
0.01, 0.1, 0.5 vol.%copper surfaceHTC enhancement (0.01%)The effect of the thermal conductivity of the nanofluid is more pronounced than the effect of nanoparticle deposition on the surface.[75]
0.02, 0.1 vol.%copper surfaceCHF enhanced 26–37%
HTC deterioration
CHF enhancement is due to the complete wetting behavior exhibited by the nanoparticles.
HTC deterioration is due to the increase in wettability and thickness of the porous layer structure.
[76]
0.001, 0.01, 0.1 Vol.%copper surface HTC depends not only on cavity size and surface wettability, but also on the heat flux range.[77]
0.001, 0.002, 0.02, 0.05, 0.1 Vol.%flat plate heaterCHF enhancement
HTC deterioration
The deposition of nanofluids on the surface causes changes in wettability and thermal resistance.[78]
0.04, 0.4, 1 kg/m3copper surface CHF enhanced 2.5–3 times
HTC enhancement
Wall superheat is reduced quickly.[79]
0.01, 0.05 wt.%horizontal rod heaterHTC deteriorationNanoparticle deposition causes a reduction in surface roughness.[80]
0.05–1 vol.%cylindrical heaterCHF enhancementNanoparticle deposition causes an increase in surface roughness.[81]
0.025, 0.05, 0.075, 0.1 vol.%Theoretical analysisCHF enhancementThe bubble detachment frequency is lowered, the detachment diameter is increased, and the wettability is enhanced.[82]
EG0.1, 0.2, 0.3 wt.%stainless steel cylinderHTC deteriorationThe surface average roughness is reduced and the number of bubble nucleation is reduced.[83]
EG/DW0.05, 0.1, 0.25, 0.5, 0.75, 1 vol.%copper cylinder HTC enhancement, when concentration is 0.75%, HTC enhance 64%In addition to the increased properties of the base fluid, changes in the state of the heated surface are the main reasons.[84]
R141b0.001, 0.01 vol.%Microchannel surfacesHTC enhancementThe low pressure causes an increase in the heat transfer coefficient.[85]
0.001, 0.01, 0.1 vol.%copper surfaceHTC enhancementThe addition of SDBS resulted in less deposition of nanoparticles.[86]
CuOWater0.1, 0.2, 0.3 wt.%stainless steel rodCHF enhancement
HTC enhancement
Nanoparticle deposition causes changes in surface roughness and vaporization core points.[71]
0.01, 0.02 wt.%surface of rod heaterHTC enhancementSDS significantly reduces surface tension.[87]
0.1, 0.2, 0.3, 0.4 wt.%stainless steel cylinderHTC enhancementSurfactants increase surface wettability and the contact angle of the bubbles with the surface is reduced.[88]
0.001, 0.01, 0.05, 0.075, 0.2 wt.%flat heater plateHTC enhancementThe deposited nanoparticles have a positive effect on heat transfer.[89]
pentane0.005, 0.01 vol.%circular brass surfaces HTC enhancementSurface trenches are the dominant condition for increasing nuclear boiling.[90]
EG/DW0.1, 0.2, 0.3 wt.%cylindrical cartridge heaterHTC enhancement.
When concentration is 0.5%, HTC enhanced 55%
Nanoparticle deposits alter surface roughness and wettability, which changes bubble shape and behavior.[91]
Fe2O3Water0.02, 0.1 vol.% copper surfacecopper surfaceCHF enhance 26–37%
HTC deterioration
CHF enhancement is due to the complete wetting behavior exhibited by the nanoparticles.
HTC deterioration is due to the increase in wettability and thickness of the porous layer structure.
[76]
0.05–1 vol.%cylindrical heaterCHF enhancementNanoparticle deposition causes an increase in surface roughness.[81]
Fe3O4Water0.01, 0.05, 0.075, 0.1, 0.2, 0.4 vol.%cylindrical copper surfaceWhen concentration is 0.1%, the HTC enhanced 43%Surface characteristics have a significant effect on boiling heat transfer.[67]
EG/DW0.01, 0.05, 0.1 vol.%horizontal Ni–Cr wireCHF enhancement
HTC deterioration
The deposited layer increases surface wettability, resulting in significant CHF enhancement.[92]
GNsEG/DW0.005, 0.01, 0.02, 0.05, 0.1 wt.% heater wireCHF enhancement
HTC enhancement
The surface wettability caused by the deposition of nanoparticles is enhanced.[93]
GONsWater0.0001, 0.0002, 0.0005, 0.001 wt.%Copper spheresCHF non-monotonic changeNon-monotonic changes consistent with changes in surface wettability caused by GON[94]
CNTWater0.1, 0.2, 0.3 wt.%stainless steel rodCHF enhancement
HTC enhancement
Nanoparticle deposition causes changes in surface roughness and vaporization core points.[71]
0.02, 0.1 vol.%copper surfaceCHF enhanced 26–37%
HTC deterioration
CHF enhancement is due to the complete wetting behavior exhibited by the nanoparticles.
HTC deterioration is due to the increase in wettability and thickness of the porous layer structure.
[76]
0.01, 0.05 wt.%horizontal rod heaterHTC enhancementThe larger size of the nanoparticles results in an increase in surface roughness.[80]
0.1, 0.3, 0.5, 0.7, 1 wt.%cooper surface HTC enhancementThe covalent modification of the nanofluid can avoid the deposition of nanoparticles and is beneficial to HTC enhancement.[95]
0.5 wt.%stainless steel spheresCHF enhancementLonger and thicker CNTs tend to form microporous layers, increasing surface roughness.[96]
0.01, 0.05, 0.1 wt.%heating element CHF enhancement
HTC enhancement
Using the covalent functionalization of cysteine and silver nanoparticles; the smaller the size, the increased surface area.[97]
SiO2Water0.04, 0.4, 1 kg/m3copper surface CHF enhanced 2.5–3 times
HTC deterioration
The separation of the nanoparticle layer may significantly deteriorate the heat transfer coefficient of boiling.[79]
0.05–1 vol.%cylindrical heaterCHF enhancementNanoparticle deposition causes an increase in surface roughness.[81]
0.001, 0.01, 0.05, 0.1 wt.%vertical cylinderHTC deteriorationThe change in surface properties caused by the deposition of nanoparticles on the surface has a major influence on the quenching process.[98]
EG/DW0.1, 0.25, 0.5, 0.75 vol.%hot-wireHTC enhancement with concentration of 0.25%Boiling heat transfer changes caused by nanoparticle coatings.[99]
ZnOWater0.01, 0.05 wt.%horizontal rod heaterHTC deteriorationNanoparticle deposition causes a reduction in surface roughness.[80]
0.01, 0.02 wt.%surface of rod heaterHTC enhancementSDS significantly reduces surface tension.[87]
EG0.35, 0.62, 0.93, 1.6, 2.6 vol.%cylindrical copper block HTC firstly increase and then decrease with the concentration It is affected by the surface roughness of the nanoparticle deposition.[100]
EG/DW5.25, 7.25, 8.25 wt.%Ni–Cr wireCHF enhancement
HTC enhancement
Heating the coating reduces surface wettability.[101]
TiO2Water0.04, 0.4, 1 kg/m3copper surface CHF enhanced 2.5–3 timesIt depends on the concentration of the nanoparticles.[79]
0.001, 0.01, 0.05, 0.075, 0.2 wt.%flat heater plateHTC deteriorationThe bubbles have a more spherical shape and their size is reduced.[89]
12, 15 wt.%copper surfaceHTC enhanced 124–138%Verified by Pioro correlation [102][103]
0.00005, 0.0001, 0.0005, 0.005, 0.01 wt.%Horizontal circular platesHTC firstly increase and then decrease with concentration, which 0.0001% is best. Subject to surface roughness.[104]
EG0.01, 0.025, 0.05, 0.075 vol.%cylindrical brass surfaceHTC deteriorationThe bubble size decreases with increasing concentration.[105]
R141b0.01, 0.025, 0.05, 0.075 vol.%cylindrical brass surfaceHTC deteriorationThe bubble size decreases with increasing concentration.[105]
Table 2. Predictions of CHF obtained using Zuber [109] and Kandlikar [110] correlations [76].
Table 2. Predictions of CHF obtained using Zuber [109] and Kandlikar [110] correlations [76].
Experimental NumberWorking FluidCHF (kW/m2)Zuber (Kw/m2)Kandlikar (kW/m2)
1Distilled water12001108(−8%)1205(+0.4%)
2Fe2O3–Water15141108(−36%)1507(−0.4%)
3Al2O3–Water15421108(−39%)1507(−2%)
4CNTs–Water (0.02%)15521108(−40%)1507(−3%)
5CNTs–Water (0.1%)14961108(−35%)1507(+0.7%)
6Distilled water/layers of Fe2O315261108(−37%)1507(−1%)
7Distilled water/layers of Al2O316521108(−49%)1507(−9%)
8Distilled water/layers of CNTs16121108(−45%)1507(−7%)

Share and Cite

MDPI and ACS Style

Yao, S.; Teng, Z. Effect of Nanofluids on Boiling Heat Transfer Performance. Appl. Sci. 2019, 9, 2818. https://doi.org/10.3390/app9142818

AMA Style

Yao S, Teng Z. Effect of Nanofluids on Boiling Heat Transfer Performance. Applied Sciences. 2019; 9(14):2818. https://doi.org/10.3390/app9142818

Chicago/Turabian Style

Yao, Shouguang, and Zecheng Teng. 2019. "Effect of Nanofluids on Boiling Heat Transfer Performance" Applied Sciences 9, no. 14: 2818. https://doi.org/10.3390/app9142818

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop