Next Article in Journal
Experimental and Theoretical Research on Bending Behavior of Photovoltaic Panels with a Special Boundary Condition
Next Article in Special Issue
Experimental Study on the Evaporation and Condensation Heat Transfer Characteristics of a Vapor Chamber
Previous Article in Journal
Deep Learning Based on Multi-Decomposition for Short-Term Load Forecasting
Previous Article in Special Issue
Magnetohydrodynamics Flow Past a Moving Vertical Thin Needle in a Nanofluid with Stability Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Impacts of Heat-Conducting Solid Wall and Heat-Generating Element on Free Convection of Al2O3/H2O Nanofluid in a Cavity with Open Border

by
Mikhail A. Sheremet
1,*,
Hakan F. Oztop
2,
Dmitriy V. Gvozdyakov
3 and
Mohamed E. Ali
4
1
Laboratory on Convective Heat and Mass Transfer, Tomsk State University, 634050 Tomsk, Russia
2
Department of Mechanical Engineering, Technology Faculty, Firat University, 23119 Elazig, Turkey
3
The Butakov Research Center, Tomsk Polytechnic University, 634050 Tomsk, Russia
4
Department of Mechanical Engineering, College of Engineering, King Saud University, 11421 Riyadh, Saudi Arabia
*
Author to whom correspondence should be addressed.
Energies 2018, 11(12), 3434; https://doi.org/10.3390/en11123434
Submission received: 7 November 2018 / Revised: 2 December 2018 / Accepted: 5 December 2018 / Published: 7 December 2018
(This article belongs to the Special Issue Heat Transfer Enhancement)

Abstract

:
Development of modern electronic devices demands a creation of effective cooling systems in the form of active or passive nature. More optimal technique for an origination of such cooling arrangement is a mathematical simulation taking into account the major physical processes which define the considered phenomena. Thermogravitational convection in a partially open alumina-water nanoliquid region under the impacts of constant heat generation element and heat-conducting solid wall is analyzed numerically. A solid heat-conducting wall is a left vertical wall cooled from outside, while a local solid element is placed on the base and kept at constant volumetric heat generation. The right border is supposed to be partially open in order to cool the local heater. The considered domain of interest is an electronic cabinet, while the heat-generating element is an electronic chip. Partial differential equations of mathematical physics formulated in non-primitive variables are worked out by the second order finite difference method. Influences of the Rayleigh number, heat-transfer capacity ratio, location of the local heater and nanoparticles volume fraction on liquid circulation and thermal transmission are investigated. It was ascertained that an inclusion of nanosized alumina particles to the base liquid can lead to the average heater temperature decreasing, that depends on the heater location and internal volumetric heat generation. Therefore, an inclusion of nanoparticles inside the host liquid can essentially intensify the heat removal from the heater that is the major challenge in different engineering applications. Moreover, an effect of nanosized alumina particles is more essential in the case of low intensive convective flow and when the heater is placed near the cooling wall.

1. Introduction

Convective thermal transmission in partially open cavities with solid heat-conducting walls and heat-generating elements is very important in different engineering supplements, e.g., cooling of electronics, heat-transfer devices, chemical apparatus and solar collectors. It should be noted that nowadays one of the major challenges is a creation of effective cooling system for reducing the working temperature for heat-generating elements inside different electronic cabinets. Optimal approach for solution to this problem is an employment of computer power of modern computational systems combined with numerical methods of hydrodynamics and heat transfer. Different interesting and useful theoretical and experimental data have been announced during last decades.
Thus, Dwesh K. Singh and S.N. Singh [1] examined numerically 2D conjugate free convection-radiation heat transfer in an open air domain under the influence of uniform volumetric heat generating element on the vertical border. They used Hottel’s Crossed-string technique for calculation of view factors during the surface radiation analysis. Taking into account the obtained results, correlations were derived for maximum dimensionless heater temperature for various positions. Mikhailenko et al. [2] conducted a computational analysis of convective heat transfer and thermal radiation inside a rotating cavity with a heat-generating element. They showed that domain rotating velocity and surface emissivity of surrounding walls and heated source are very effective control parameters for creation of optimal cooling system. Chen et al. [3] evaluated the convection-conduction thermal transmission in an open partially porous domain. They chose the lattice Boltzmann approach to solve equations. Naylor et al. [4] studied numerically thermogravitational convection from a window glazing with an insect shield simulated like a porous layer and found that the shield has negligible influence on convective thermal transmission for high values of Ra. Astanina et al. [5] analyzed numerically thermogravitational convection in a partially porous cavity with a heat-generating source. The authors demonstrated that porous layer thickness and variable viscosity parameter of the working medium allows optimizing the passive cooling arrangement owing to an enhancement of cooling impact from the cold vertical borders. Rahman et al. [6] examined the influence of Ohmic heating and magnetic field on mixed convection heat and mass transfer in a horizontal channel using finite element technique.Their results demonstrated that the aforesaid fields affect the flow structure, thermal and mass transmnission. Saha et al. [7] worked on MHD mixed convection under the influence of internal heat generation/absorption. They revealed that the average Nusselt number is a decreasing function of the heat generation and an increasing function of heat absorption. Gangawane [8] made a numerical study on MHD thermograviational convection in an open cavity. The author found that the domain of interest with the inclined magnetic field at the angle π/4 illustrates the highest thermal transmission restriction than other analyzed conditions. Dong and Li [9] studied the conjugate free convection in a composite closed space by vorticity–stream function technique. They found that used material, geometry and Ra are efffective parameters for an optimization of thermal transmission. Bilgen [10] performed a work on free convection in closed spaces having thin obstacle on heated border. He observed that length and location of this obstacle are very good control elements for an intensification of liquid circulation and thermal transmission.
Nowadays, for an intensification of heat transfer within different engineering systmes the nanosized metal or metal oxide particles are used [11,12,13]. Different experimental studies showed that an inclusion of nanoparticles of low concentration can improve the heat transfer performance inside the system [14,15]. Thus, Asirvatham et al. [14] performed experiments for steady convection thermal transmission of de-ionized water combined with copper oxide nanoparticles of low concentration (0.003% by volume) in a copper tube. They showed 8% enhancement of the convective heat transfer coefficient of the nanofluid even with a low volume concentration of nanosized copper oxide particles. Raei et al. [15] experimentally studied the circulation and thermal transmission parameters of three kinds of the nanofluids within a fin-and-tube heat exchanger. It was revealed that an inclusion of alumina nanosized particles into the water enhances the overall heat transfer parameter and friction factor up to 20% and 5%, respectively. Taking into account such heat transfer enhancement, these approach can be used for an intesification of cooling process inside the electronic cabinets. Several theoretical and experimental works were published. Thus, Mehrez et al. [16] solved numerically a problem on mixed convection combined with entropy production with an open region filled with nanoliquid under the bottom heating impact. They found that the thermal transmission combined with entropy production are enhanced with the growth of Re, Ri and nanoparticles concentration, and changed with region AR and nanoparticle types. Sheremet et al. [17] analyzed mixed convective heat and mass transport inside a porous nanoliquid domain with open border under the impacts of thermophoresis and Brownian diffusion. They revealed that the average Nusselt parameter is a growing mapping of Ra and Re and a reduced mapping of Le. Sheremet et al. [18] investigated nanoliquid free convection within a semi-open tall wavy cavity under the effects of horizontal temperature difference and Buongiorno’s nanofluid model. They ascertained that a growth of the shape coefficient (<1.0) results in the thermal transmission intensification, while a rise of this parameter in a range of (>1.0) results in non-monotonic behavior for the average dimensionless heat transfer coefficient owing to significant heating of the wave trough. Bondareva et al. [19] examined MHD nanofluid free convective thermal transmission in a tilted wavy tall porous domain with a local heater using streamlines and heatlines. Optimal parameters for the thermal transmission intensification were found. Miroshnichenko et al. [20] studied computationally natural convective heat transfer of Al2O3/H2O nanoliquid in an inclined open chamber with a heat-generating source. They illustrated that the major cooling of the heater occurs for central element position with chamber inclination angle of π/3. Bondarenko et al. [21] performed a numerical work on free convection of alumina/water nanoliquid in an enclosure with heat-generating solid source. It was ascertained that within hermetic electronic cabinet having a heat-generating source under cold vertical isothermal borders the thermal removal from the heater can be enhanced by including nanosized aluminum oxide particles of small volume fraction (1%) and by the position of the heated element near the cold vertical border.
The abovementioned brief review shows that the problem of heat transfer enhancement within electronic cabinets is very important and usage of nanoparticles with additional steps can help to solve such challenge. Therefore, the aim of this work is to calculate free convection thermal transmission in a partially open alumina-water nanoliquid area under the effect of vertical solid mural and local heat-generating source. The present study is devoted to development of passive cooling arrangement for heat-generating elements using the nanoliquids under the impacts of heat-conducting walls and local opening ports. The present study is a continuation of passive cooling systems analysis that was presented in the case of open inclined cavity [20] and closed vertical chamber [21]. Novelty of the present investigation is an examination of free convection under the combined impacts of open border, cooled solid wall of finite thickness and heat-generating element.

2. Basic Equations

The considered region with frame of reference and boundary conditions is demonstrated in Figure 1. It is a vertical cavity having isothermally cooling left vertical wall, while this solid wall has finite thickness l and finite thermal conductivity kw. Horizontal walls of neglecting thickness are adiabatic. Right vertical wall is adiabatic also but it has an open port of size d where ambient cold nanofluid can penetrate into the cavity and it also can exit from the cavity. Solid heater is placed at the base of the cavity and it has a permanent volumetric heat flux Q. Length from the left border to this heated element is s. The nanoliquid is a solution of water with solid spherical alumina nanosized particles [20,21,22].
The partial differential equations of mathematical physics for 2D free convection nanoliquid circulation and thermal transmission within the considered domain (see Figure 1) employing the conservation laws are written:
  • for the nanoliquid
    u ¯ x ¯ + v ¯ y ¯ = 0
    ρ n f ( u ¯ t + u ¯ u ¯ x ¯ + v ¯ u ¯ y ¯ ) = p x ¯ + μ n f ( 2 u ¯ x ¯ 2 + 2 u ¯ y ¯ 2 )
    ρ n f ( v ¯ t + u ¯ v ¯ x ¯ + v ¯ v ¯ y ¯ ) = p y ¯ + μ n f ( 2 v ¯ x ¯ 2 + 2 v ¯ y ¯ 2 ) + ( ρ β ) n f g ( T T c )
    T t + u ¯ T x ¯ + v ¯ T y ¯ = k n f ( ρ c ) n f ( 2 T x ¯ 2 + 2 T y ¯ 2 )
  • for the heater
    ( ρ c ) h s T t = k h s ( 2 T x ¯ 2 + 2 T y ¯ 2 ) + Q
  • for the solid wall
    ( ρ c ) w T t = k w ( 2 T x ¯ 2 + 2 T y ¯ 2 )
The considered nanoliquid viscosity and heat-transfer capacity are the mappings of nanosized particles concentration as obtained by Ho et al. [22].
Including the stream function ( u ¯ = ψ ¯ y ¯ , v ¯ = ψ ¯ x ¯ ) , vorticity ( ω ¯ = v ¯ x ¯ u ¯ y ¯ ) , and the non-dimensional parameters:
x = x ¯ / L ,   y = y ¯ / L ,   τ = t g β Δ T / L ,   θ = ( T T c ) / Δ T , u = u ¯ / g β Δ T L ,   v = v ¯ / g β Δ T L ,   ψ = ψ ¯ / g β Δ T L 3 ,   ω = ω ¯ L / g β Δ T
the control equations of mathematical physics become:
  • for the nanofluid
    2 ψ x 2 + 2 ψ y 2 = ω
    ω τ + u ω x + v ω y = H 1 ( ϕ ) P r R a ( 2 ω x 2 + 2 ω y 2 ) + H 2 ( ϕ ) θ x
    θ τ + u θ x + v θ y = H 3 ( ϕ ) R a P r ( 2 θ x 2 + 2 θ y 2 )
  • for the heater
    θ τ = α h s / α f R a P r ( 2 θ x 2 + 2 θ y 2 + 1 )
  • for the solid wall
    θ τ = α w / α f R a P r ( 2 θ x 2 + 2 θ y 2 )
The considered additional conditions are
τ = 0 : ψ = ω = θ = 0   at   0 x 1   and 0 y A = H / L τ > 0 : θ = 0   at   x = 0   and 0 y A ψ = 0 ,   ω = 2 ψ y 2 ,   θ y = 0   at   y = 0 ,   A   and 0 < x < 1 ψ = 0 ,   ω = 2 ψ x 2 ,   θ x = 0   at   x = 1   and 0 < y ζ = h / L ψ = 0 ,   ω = 2 ψ x 2 ,   θ x = 0   at   x = 1   and h / L + d / L = ζ + η y < A
τ > 0 : ψ x = 0 , ω x = 0 , { if   ψ y 0   then   θ x = 0 if   ψ y < 0   then   θ = 0   at   x = 1   and ζ < y < ζ + η ψ = 0 ,   ω = 2 ψ n 2 ,   { θ h s = θ n f θ h s n = H 4 ( ϕ ) θ n f n   at   heat   source   surface ψ = 0 ,   ω = 2 ψ x 2 ,   { θ w = θ n f θ w x = H 5 ( ϕ ) θ n f x   at   x = ξ = l / L   and 0 y A
Here additional parameters are
R a = ρ g β Δ T L 3 α μ , P r = μ ρ α , Δ T = Q L 2 k h s , H 1 ( ϕ ) = 1 + 4.93 ϕ + 222.4 ϕ 2 1 ϕ + ϕ ρ p / ρ f , H 2 ( ϕ ) = 1 ϕ + ϕ ( ρ β ) p / ( ρ β ) f 1 ϕ + ϕ ρ p / ρ f , H 3 ( ϕ ) = 1 + 2.944 ϕ + 19.672 ϕ 2 1 ϕ + ϕ ( ρ c ) p / ( ρ c ) f , H 4 ( ϕ ) = 1 + 2.944 ϕ + 19.672 ϕ 2 k h s / k f , H 5 ( ϕ ) = 1 + 2.944 ϕ + 19.672 ϕ 2 k w / k f
The heat transfer rate can be described by the average Nusselt number at the heater surface:
N u ¯ = 1 γ 0 γ ( k n f k f θ n ) d γ

3. Numerical Technique

The governing Equations (7)–(11) with additional conditions (12) and (13) were resolved by the finite difference technique using the uniform grid. The detailed verification of the developed computational code was conduced earlier in refs. [17,18,19,20,21].
In the case of nanoliquid free convection inside the differentially heated square chamber, the obtained data were compared with experimental results of Ho et al. [22] and numerical results of Saghir et al. [23]. Table 1 shows a very good agreement between the considered data.
In the case of free convection inside the partially open chamber, the developed computational code was validated successfully using the numerical data of Mohamad et al. [24] and Kefayati [25] (see Table 2).
The solution technique was tested using various uniform grids. Figure 2 demonstrates the impact of the considered meshes on N u ¯ , average heater temperature and nanofluid flow intensity for Ra = 105, Pr = 6.82, Kr = kw/kf = 10, φ = 0.02, δ = s/L = 0.3, h/L = 0.2, l/L = 0.1 and d/L = 0.5.
The presented data in Figure 2 shows that differences between the considered uniform grids of 200 × 200 nodes and 300 × 300 nodes are non-essential. Therefore, taking into account the calculation accuracy and the computational time, the mesh of 200 × 200 nodes was chosen for the base investigation.

4. Results

This paragraph deals with computational study of the nanoliquid circulation and heat transfer within the considered domain under the influence of the following characteristics: Rayleigh number (103Ra ≤ 105), Prandtl number (Pr = 6.82), thermal conductivity ratio (1 ≤ Kr = kw/kf ≤ 20), heater location (0.1 ≤ δ = s/L ≤ 0.5) and nanoparticles volume fraction (0 ≤ φ ≤ 0.04) for h/L = 0.2, l/L = 0.1 and d/L = 0.5. The analysis pays attention in the description of Ra, Kr, heater location and nanoparticles concentration impacts. Isolines of stream function and temperature, N u ¯ , nanoliquid circulation rate and mean heater temperature for various control characteristics are pictured in Figure 3, Figure 4, Figure 5, Figure 6, Figure 7, Figure 8, Figure 9, Figure 10, Figure 11 and Figure 12. It is worth noting that such investigation can help to understand the nanoliquid behavior and heat transfer performance within the considered partially open cavity with the heat-conducting solid wall and heat-generating silicon element.
Figure 3 presents isolines of stream function and temperature within the considered region for Kr = 5, δ = 0.2 and various Ra. For the present study an increase in Ra is related to a raise of the internal heater generation Q.
In the case of low and moderate Ra (103 and 104) one convective cell is formed near the internal surface of the heat-conducting solid wall and another one characterizes an intrusion of cold nanoliquid from the open port inside the region. An appearance of the convective cell near the solid wall is explained by the formation of temperature difference between the cold solid border and hot heat-generating element. Therefore, this circulation reflects a counter-clockwise motion of nanofluid. Moreover, the penetrative flow from the open border begins to interact with the abovementioned vortex and the line of such interaction corresponds to the thermal plume appeared above the heat source. A growth of the internal heat generation leads to a deformation of the vortex near the solid wall due to more intensive penetrative flow. For Ra = 105 (Figure 3c) this convective cell is divided into two recirculations located in top and lower parts of the solid wall. Isotherms reflect a distribution of temperature within the region and solid wall. Heating occurs from the element located on the bottom wall, while cooling occurs from outside near the left solid border and from the open part of the right boundary. For Ra = 103 (Figure 3a) the dominated thermal transmission mode is a thermal conduction and as a result the isotherms are parallel to the heated and cooled elements. Moreover, more essential cooling occurs in the upper part of the left vertical wall due to less intensive heating from the heater. It is interesting to note that the heat-generating element is a solid block and heat conduction Equation (10) is solved within this element, but isotherms are not presented inside the heater. The main reason for such behavior is the high value of silicon thermal conductivity as the heater material in comparison with the thermal conductivity of working medium. As a result, this element is heated quickly.
An increment in the temperature difference (Ra = 104 in Figure 3b) leads to more essential cooling from the cold parts and heating from the element with intensification of convective thermal transmission in the central part of the domain. It is worth noting here that thermal plume is more essential with pronounced plume’s shape. Due to the prevalence of thermal conduction in a spacing between the solid border and heated element, heating of the solid wall occurs in the bottom part from the heater. While for high value of Ra (105 in Figure 3c) heating of the solid wall from the heater occurs in the bottom part but over the wall from the tilted thermal plume due to the convective thermal transmission intensification. A growth of Ra results also in more essential cooling from the open port. It is worth noting, that distortion of thermal plume for Ra = 105 occurs owing to the significant impact of penetrative nanoliquid on convective flow over the heater.
An inclusion of alumina nanosized particles to the clear water results in the small changes for the thermal conduction dominated mode (Ra = 103 in Figure 3a), while for convective heat transfer, cooling of the cavity occurs essentially owing to high magnitude of the effective heat-transfer capacity (see ref. [22]).
Impacts of Ra and φ on the mean parameters are shown in Figure 4. The presented parameters have been normalized using the temperature difference that depends on the volumetric heat generation. Therefore, for comparison between different Rayleigh numbers it is necessary to normalize using the general scale. An increase in the temperature difference results in an essential rise of N u ¯ owing to more intensive circulation of the working liquid. While a raise of φ results in a diminution of N u ¯ . The nanoliquid circulation rate | ψ | max R a P r increases with Ra, due to more intensive circulation for high ΔT. An increment of nanoparticles volume fraction reduces | ψ | max also due to a rise of the nanofluid viscosity. A growth of ΔT reflects a rise of θavg using the normalization with general scale. An enhancement of φ allows decreasing the average temperature for heat conduction regime (Ra = 103), while for the transition heat transfer (Ra = 104) and heat convection (Ra = 105) θavg increases with the nanoparticles concentration.
Figure 5 indicates the impacts of Kr between solid border material and host fluid on isolines of stream function and temperature for Ra = 105, δ = 0.2. Regardless of the thermal conductivity ratio values, two recirculations are appeared close to the solid boundary in the upper and bottom parts, while major circulation appears from the open port. An enhancement of the solid border heat-transfer capacity results in an intensification of recirculations near the solid wall due to more intensive cooling of this wall, where cold temperature is transported intensively from the external surface till internal one and temperature gradient at this internal surface increases. The latter characterizes an intensification of convective flow near this wall that can deform the nanofluid penetration flow from the open border. Essential cooling of the heat-conducting solid wall can be found in Figure 5 with distributions of isotherms. An increment of isothermal lines density close to the solid-fluid interface (x = 0.1) illustrates an origin of strong temperature gradient at this wall. More significant temperature differences are found along the right and top surfaces of the local heater owing to a reciprocal action between hot and cold temperature waves. Addition of nanoparticles reflects a formation of more stable circulation close to the left solid border, where for Kr ≥ 5 one can find alone convective cell, while for φ = 0 we have two different circulations in upper and bottom parts. As a result, it is possible to conclude that a growth of thermal conductivity ratio related with an increment of solid wall thermal conductivity leads to an intensification of convective recirculation near the upper part of the internal surface of this heat-conducting wall owing to high cooling rate from the border x = 0.
Figure 6, Figure 7 and Figure 8 present the effects of Kr, φ and Ra on the integral parameters such as N u ¯ , | ψ | max and θavg for δ = 0.2. As can be found from this figure, impacts of Ra and nanoparticles concentration were described in Figure 4. Within the considered Figure 6, Figure 7 and Figure 8 the particular efforts were devoted to the influence of thermal conductivity ratio and mutual effects of all considered parameters. In the case of Ra = 103 (Figure 6) a raise of Kr results in a diminution of all considered parameters, while all these parameters decrease with φ also. Significant diminution for N u ¯ , | ψ | max and θavg is between Kr = 1 and Kr = 5, while for Kr ∈ (5, 20) changes are not so essential. As it has been mentioned above for Figure 4, an increase in φ can decrease θavg for Ra = 103.
For Ra = 104 (Figure 7) and Ra = 105 (Figure 8) the behavior of the considered parameters with Kr is similar to the case of Ra = 103 (Figure 6), but the main differences are the values of these characteristics. For these Ra, an increment of φ characterizes an enhancement of the heater average temperature that does not allow to improve the cooling arrangement. Moreover, for high Ra a transition between Kr = 5 and Kr = 20 becomes insignificant.
An influence of heater location on isolines of stream function and temperature for Ra = 105, Kr = 5 is demonstrated in Figure 9. A displacement of the heater from the left wall to the right one results in a combination of two recirculations placed in top and bottom part of the domain close to the solid wall and this combined vortex becomes more stable and strong when heater locates near the right vertical border. At the same time, such displacement characterizes a deformation of the penetrative flow from the open port. It is interesting to note that initially for δ = 0.1 (Figure 9a) the power of the upper circulation is high in comparison with the bottom one, while a rise of δ leads to a strengthening of the bottom vortex and an attenuation of the upper one.
For δ = 0.3 (Figure 9c), after a combination of these circulations, we have a convective circulation with two cores and the following increase in δ reflects a growth of the intensity and size of the bottom convective core. Such variations of hydrodynamics are related with modification of temperature field. A presence of penetrative cold flow from the open port characterizes an origin of tilted thermal plume above the heat source. The heating effect of the solid wall from the heater becomes insignificant with an increment of distance between the heat source and solid border. Such feature characterizes a strengthening of the convective circulation appeared near the solid wall. An inclusion of alumina nanoparticles characterizes an origin of combined convective cell near the solid boundary for less value of the length between the heat source and solid border (δ = 0.2) and for δ = 0.5 sizes of this vortex become more powerful in comparison with the case of clear base fluid. Heat conduction effect can be found in the heat-generating element for δ = 0.1 and δ = 0.3 due to more essential interaction between the cooling temperature waves from the solid wall and open port. It is interesting to note a displacement of the heating zone inside the solid wall from the bottom part (for δ = 0.1) along the vertical coordinate owing to the impact of the tilted thermal plume.
Figure 10, Figure 11 and Figure 12 illustrate the effects of heater location, thermal conductivity ratio, nanoparticles concentration and Rayleigh number on the average Nusselt number, nanoliquid circulation rate and heater average temperature.
In the case of heat conduction dominated regime (Ra = 103 in Figure 10), a rise of the length between the solid border and heat source results in an increment of N u ¯ and this increment is significant for low values of δ, while for δ ∈ (0.3, 0.5) the rate of N u ¯ is low in comparison with the range of δ ∈ (0.1, 0.3). Nanofluid flow rate and heater average temperature have non-monotonic behavior with δ. In the case of Kr = 1, a growth of δ ∈ (0.2, 0.5) reflects a reduction of | ψ | max , while for Kr = 5 and Kr = 20 | ψ | max increases with δ. The average heater temperature enhances with δ for δ ∈ (0.1, 0.4) and for δ > 0.4 it decreases. An enhancement of the thermal conductivity ratio results in a diminution of all considered parameters. The impact of nanoparticles concentration is non-linear for N u ¯ , namely, N u ¯ reduces with φ for δ ∈ (0.1, 0.33) and it increases for other values of δ. At the same time, | ψ | max and θavg are diminished with φ. It is interesting to note the negative values of the average Nusselt number for Kr = 20. It means that the temperature gradient along the heater surface is negative due to change of the heat flux direction.
A non-linear influence of δ on all considered parameters can be found for Ra = 104 (see Figure 11), namely, for δ ∈ (0.1, 0.2) N u ¯ and | ψ | max are increased, while for δ ∈ (0.2, 0.5) these parameters are reduced. θavg increases with δ for Kr = 5 and Kr = 20, while for Kr = 1 it has a non-linear behavior. An effect of Kr on the considered parameters is the same like for Ra = 103 (see Figure 10). N u ¯ and | ψ | max are decreased with the nanoparticles volume fraction, while the average heater temperature is changed non-linear with φ.
Behavior of N u ¯ with all considered parameters for Ra = 105 (see Figure 12) is similar to the case of Ra = 104 (see Figure 11), while the nanoliquid circulation rate and mean heater temperature have the specific behavior.

5. Conclusions

Free convection of alumina–water nanoliquid in an open domain with a solid border of finite thickness and heat-generating element has been investigated. Partial differential equations of mathematical physics written using the non-primitive variables have been solved numerically by the finite difference technique. Impacts of Ra, Kr, heater location and nanoparticles volume fraction on liquid circulation and thermal transmission have been investigated. It has been ascertained that an increment of Ra illustrates an enhancement of all considered average characteristics, while a rise of Kr decreases all these parameters. More important parameter for analysis of passive cooling system is an average heater temperature. It is possible to reduce θavg near the heat-conducting solid wall. A growth of Kr from 1 till 20 for φ = 0, δ = 0.1 and Ra = 103 allows to decrease the average heater temperature at 30%, while for φ = 0, δ = 0.1 and Ra = 104 average heater temperature decreases at 20% and for φ = 0, δ = 0.1 and Ra = 105 θavg reduces at 12%. An addition of nanosized particles results in a diminution of the average heater temperature near the cooled solid border for all values of Ra and for all heater locations at Ra = 103. In the case of δ = 0.1, Kr = 1 and Ra = 103 a growth of nanoparticles volume fraction from 0.0 till 0.04 results in a decrease in the average heater temperature at 6%, for Kr = 5 the diminution is 10% and for Kr = 20 the average heat temperature reduction is 11%. Therefore, alumina-water nanoparticles and cooling solid wall (thermal conductivity ratio) can be effective characteristics for the passive cooling system of the heat-generating solid element (θavg can be reduced at 30%).

Author Contributions

M.A.S. and H.F.O. conceived the main concept. M.A.S., H.F.O., and D.V.G. contributed to the investigation and data analysis. M.A.S. and H.F.O. wrote the manuscript. All authors contributed in writing the final manuscript.

Funding

This work of the third author was performed as a government task of the Ministry of Education and Science of the Russian Federation (Project Number 13.7644.2017/BCH). The second and last authors extend their appreciation to the International Scientific Partnership Program ISPP at King Saud University for funding this research work through ISPP#131.

Acknowledgments

Authors thank the respected reviewers for their constructive comments which clearly enhanced the quality of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
Aaspect ratio (–)
cspecific heat (J kg−1 K−1)
ddimensional open port size (m)
ggravitational acceleration (m s−2)
Hdimensional height of the cavity (m)
hdimensional distance between the bottom wall and open port (m)
H1(φ), H2(φ), H3(φ), H4(φ), H5(φ)special functions (–)
kthermal conductivity (W m−1 K−1)
Ldimensional length of the cavity (m)
ldimensional thickness of the left solid wall (m)
N u ¯ average Nusselt number (–)
pdimensional pressure (Pa)
PrPrandtl number (–)
Qinternal volumetric heat flux (W m−3)
RaRayleigh number (–)
sdimensional distance between the left solid wall and local heater (m)
Tdimensional temperature (K)
tdimensional time (s)
Tcambient nanofluid temperature (K)
u, vdimensionless velocity components (–)
u ¯ ,   v ¯ dimensional velocity components (m s−1)
x, ydimensionless Cartesian coordinates (–)
x ¯ ,   y ¯ dimensional Cartesian coordinates (m)
Greek symbols
αthermal diffusivity (W m−2 K−1)
βthermal expansion coefficient (K−1)
ΔTreference temperature difference (K)
δdimensionless distance between the left wall and the heater (–)
ζdimensionless distance between the bottom wall and open port (–)
ηdimensionless open port size (–)
θdimensionless temperature (–)
μdynamic viscosity (kg m−1 s−1)
ξdimensionless thickness of the left solid wall (–)
ρdensity (kg m−3)
ρcheat capacitance (J m−3 K−1)
ρβbuoyancy coefficient (kg m−3 K−1)
τdimensionless time (–)
γcoordinate along the heater surface (m)
φnanoparticles volume fraction (–)
ψ ¯ dimensional stream function (m2 s−1)
ψdimensionless stream function (–)
ω ¯ dimensional vorticity (s−1)
ωdimensionless vorticity (–)
Subscripts
avgaverage
ccold
ffluid
hsheat source
maxmaximum value
nfnanofluid
p(nano) particle
wsolid wall

References

  1. Singh, D.K.; Singh, S.N. Conjugate free convection with surface radiation in open top cavity. Int. J. Heat Mass Transf. 2015, 89, 444–453. [Google Scholar] [CrossRef]
  2. Mikhailenko, S.A.; Sheremet, M.A.; Mohamad, A.A. Convective-radiative heat transfer in a rotating square cavity with a local heat-generating source. Int. J. Mech. Sci. 2018, 142–143, 530–540. [Google Scholar] [CrossRef]
  3. Chen, S.; Gong, W.; Yan, Y. Conjugate natural convection heat transfer in an open-ended square cavity partially filled with porous media. Int. J. Heat Mass Transf. 2018, 124, 368–380. [Google Scholar] [CrossRef]
  4. Naylor, D.; Foroushani, S.S.M.; Zalcman, D. Free convection heat transfer from a window glazing with an insect screen. Energy Build. 2017, 138, 206–214. [Google Scholar] [CrossRef]
  5. Astanina, M.S.; Sheremet, M.A.; Umavathi, J.C. Transient natural convection with temperature-dependent viscosity in a square partially porous cavity having a heat-generating source. Numer. Heat Transf. A 2018, 73, 849–862. [Google Scholar] [CrossRef]
  6. Rahman, M.M.; Saidur, R.; Rahim, N.A. Conjugated effect of joule heating and magneto-hydrodynamic on double-diffusive mixed convection in a horizontal channel with an open cavity. Int. J. Heat Mass Transf. 2011, 54, 3201–3213. [Google Scholar] [CrossRef]
  7. Saha, L.K.; Uddin, K.M.S.; Taher, M.A. Effect of internal heat generation or absorption on MHD mixed convection flow in a lid driven cavity. Am. J. Appl. Math. 2015, 3, 20–29. [Google Scholar] [CrossRef]
  8. Gangawane, K.M. Effect of angle of applied magnetic field on natural convection in an open ended cavity with partially active walls. Chem. Eng. Res. Des. 2017, 127, 22–34. [Google Scholar] [CrossRef]
  9. Dong, S.-F.; Li, Y.-T. Conjugate of natural convection and conduction in a complicated enclosure. Int. J. Heat Mass Transf. 2004, 47, 2233–2239. [Google Scholar] [CrossRef]
  10. Bilgen, E. Natural convection in cavities with a thin fin on the hot wall. Int. J. Heat Mass Transf. 2005, 48, 3493–3505. [Google Scholar] [CrossRef]
  11. Kristiawan, B.; Santoso, B.; Wijayanta, A.T.; Aziz, M.; Miyazaki, T. Heat transfer enhancement of TiO2/water nanofluid at laminar and turbulent flows: A numerical approach for evaluating the effect of nanoparticle loadings. Energies 2018, 11, 1584. [Google Scholar] [CrossRef]
  12. Vryzas, Z.; Kelessidis, V.C. Nano-based drilling fluids: A review. Energies 2017, 10, 540. [Google Scholar] [CrossRef]
  13. Patil, M.S.; Seo, J.-H.; Kang, S.-J.; Lee, M.-Y. Review on synthesis, thermo-physical property, and heat transfer mechanism of nanofluids. Energies 2016, 9, 840. [Google Scholar] [CrossRef]
  14. Asirvatham, L.G.; Vishal, N.; Gangatharan, S.K.; Lal, D.M. Experimental study on forced convective heat transfer with low volume fraction of CuO/water nanofluid. Energies 2009, 2, 97–119. [Google Scholar] [CrossRef]
  15. Raei, B.; Peyghambarzadeh, S.M.; Asl, R.S. Experimental investigation on heat transfer and flow resistance of drag-reducing alumina nanofluid in a fin-and-tube heat exchanger. Appl. Therm. Eng. 2018, 144, 926–936. [Google Scholar] [CrossRef]
  16. Mehrez, Z.; Bouterra, M.; el Cafsi, A.; Belghith, A. Heat transfer and entropy generation analysis of nanofluids flow in an open cavity. Comput. Fluids 2013, 88, 363–373. [Google Scholar] [CrossRef]
  17. Sheremet, M.A.; Pop, I.; Ishak, A. Double-diffusive mixed convection in a porous open cavity filled with a nanofluid using Buongiorno’s model. Transp. Porous Media 2015, 109, 131–145. [Google Scholar] [CrossRef]
  18. Sheremet, M.A.; Pop, I.; Shenoy, A. Unsteady free convection in a porous open wavy cavity filled with a nanofluid using Buongiorno’s mathematical model. Int. Commun. Heat Mass Transf. 2015, 67, 66–72. [Google Scholar] [CrossRef]
  19. Bondareva, N.S.; Sheremet, M.A.; Oztop, H.F.; Abu-Hamdeh, N. Heatline visualization of MHD natural convection in an inclined wavy open porous cavity filled with a nanofluid with a local heater. Int. J. Heat Mass Transf. 2016, 99, 872–881. [Google Scholar] [CrossRef]
  20. Miroshnichenko, I.V.; Sheremet, M.A.; Oztop, H.F.; Abu-Hamdeh, N. Natural convection of Al2O3/H2O nanofluid in an open inclined cavity with a heat-generating element. Int. J. Heat Mass Transf. 2018, 126, 184–191. [Google Scholar] [CrossRef]
  21. Bondarenko, D.S.; Sheremet, M.A.; Oztop, H.F.; Ali, M.E. Natural convection of Al2O3/H2O nanofluid in a cavity with a heat-generating element. Heatline visualization. Int. J. Heat Mass Transf. 2019, 130, 564–574. [Google Scholar] [CrossRef]
  22. Ho, C.J.; Liu, W.K.; Chang, Y.S.; Lin, C.C. Natural convection heat transfer of alumina-water nanofluid in vertical square enclosures: An experimental study. Int. J. Therm. Sci. 2010, 49, 1345–1353. [Google Scholar] [CrossRef]
  23. Saghir, M.Z.; Ahadi, A.; Mohamad, A.; Srinivasan, S. Water aluminum oxide nanofluid benchmark model. Int. J. Therm. Sci. 2016, 109, 148–158. [Google Scholar] [CrossRef]
  24. Mohamad, A.A.; El-Ganaoui, M.; Bennacer, R. Lattice Boltzmann simulation of natural convection in an open ended enclosure. Int. J. Therm. Sci. 2009, 48, 1870–1875. [Google Scholar] [CrossRef]
  25. Kefayati, G.H.R. Effect of a magnetic field on natural convection in an open cavity subjugated to water/alumina nanofluid using Latttice Boltzmann Method. Int. Commun. Heat Mass Transf. 2013, 40, 67–77. [Google Scholar] [CrossRef]
Figure 1. Analyzed region and reference frame.
Figure 1. Analyzed region and reference frame.
Energies 11 03434 g001
Figure 2. Behavior of the average Nusselt number (a), nanoliquid flow rate (b) and average heater temperature (c) with the non-dimensional time and grid characteristics.
Figure 2. Behavior of the average Nusselt number (a), nanoliquid flow rate (b) and average heater temperature (c) with the non-dimensional time and grid characteristics.
Energies 11 03434 g002
Figure 3. Streamlines ψ and isotherms θ for Kr = 5, δ = 0.2: Ra = 103 (a), Ra = 104 (b), Ra = 105 (c).
Figure 3. Streamlines ψ and isotherms θ for Kr = 5, δ = 0.2: Ra = 103 (a), Ra = 104 (b), Ra = 105 (c).
Energies 11 03434 g003
Figure 4. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on φ and Ra for Kr = 5, δ = 0.2.
Figure 4. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on φ and Ra for Kr = 5, δ = 0.2.
Energies 11 03434 g004
Figure 5. Streamlines ψ and isotherms θ for Ra = 105, δ = 0.2: Kr = 1 (a), Kr = 5 (b), Kr = 20 (c).
Figure 5. Streamlines ψ and isotherms θ for Ra = 105, δ = 0.2: Kr = 1 (a), Kr = 5 (b), Kr = 20 (c).
Energies 11 03434 g005
Figure 6. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on φ and Kr for Ra = 103, δ = 0.2.
Figure 6. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on φ and Kr for Ra = 103, δ = 0.2.
Energies 11 03434 g006
Figure 7. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on φ and Kr for Ra = 104, δ = 0.2.
Figure 7. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on φ and Kr for Ra = 104, δ = 0.2.
Energies 11 03434 g007
Figure 8. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on φ and Kr for Ra = 105, δ = 0.2.
Figure 8. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on φ and Kr for Ra = 105, δ = 0.2.
Energies 11 03434 g008
Figure 9. Streamlines ψ and isotherms θ for Ra = 105, Kr = 5: δ = 0.1 (a), δ = 0.2 (b), δ = 0.3 (c), δ = 0.4 (d), δ = 0.5 (e).
Figure 9. Streamlines ψ and isotherms θ for Ra = 105, Kr = 5: δ = 0.1 (a), δ = 0.2 (b), δ = 0.3 (c), δ = 0.4 (d), δ = 0.5 (e).
Energies 11 03434 g009
Figure 10. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on the heater location, thermal conductivity ratio and nanoparticles concentration for Ra = 103.
Figure 10. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on the heater location, thermal conductivity ratio and nanoparticles concentration for Ra = 103.
Energies 11 03434 g010
Figure 11. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on the heater location, Kr and nanoparticles concentration for Ra = 104.
Figure 11. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on the heater location, Kr and nanoparticles concentration for Ra = 104.
Energies 11 03434 g011
Figure 12. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on the heater location, Kr and φ for Ra = 105.
Figure 12. Dependences of the average Nusselt number (a), nanoliquid circulation rate (b) and mean heater temperature (c) on the heater location, Kr and φ for Ra = 105.
Energies 11 03434 g012
Table 1. Mean Nu at heated border compared with results of other authors.
Table 1. Mean Nu at heated border compared with results of other authors.
Dimensionless ParametersφObtained DataResults of Ho et al. [22]Results of Saghir et al. [23]
Ra = 8.663 × 107,
Pr = 7.002
0.0131.604332.203730.657
0.0231.253831.090530.503
0.0330.82929.076930.205
Table 2. Mean Nu at heated border compared with results of other authors.
Table 2. Mean Nu at heated border compared with results of other authors.
RaObtained DataResults of Mohamad et al. [24]Results of Kefayati [25]
1043.3213.3773.319
1057.3257.3237.391
10514.32214.38014.404

Share and Cite

MDPI and ACS Style

Sheremet, M.A.; Oztop, H.F.; Gvozdyakov, D.V.; Ali, M.E. Impacts of Heat-Conducting Solid Wall and Heat-Generating Element on Free Convection of Al2O3/H2O Nanofluid in a Cavity with Open Border. Energies 2018, 11, 3434. https://doi.org/10.3390/en11123434

AMA Style

Sheremet MA, Oztop HF, Gvozdyakov DV, Ali ME. Impacts of Heat-Conducting Solid Wall and Heat-Generating Element on Free Convection of Al2O3/H2O Nanofluid in a Cavity with Open Border. Energies. 2018; 11(12):3434. https://doi.org/10.3390/en11123434

Chicago/Turabian Style

Sheremet, Mikhail A., Hakan F. Oztop, Dmitriy V. Gvozdyakov, and Mohamed E. Ali. 2018. "Impacts of Heat-Conducting Solid Wall and Heat-Generating Element on Free Convection of Al2O3/H2O Nanofluid in a Cavity with Open Border" Energies 11, no. 12: 3434. https://doi.org/10.3390/en11123434

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