Next Article in Journal
Joint Security and Energy-Efficient Cooperative Architecture for 5G Underlaying Cellular Networks
Next Article in Special Issue
Finite Difference Simulation of Nonlinear Convection in Magnetohydrodynamic Flow in the Presence of Viscous and Joule Dissipation over an Oscillating Plate
Previous Article in Journal
An Efficient Method for Split Quaternion Matrix Equation XAf(X)B = C
Previous Article in Special Issue
Analytical Solution for the MHD Flow of Non-Newtonian Fluids between Two Coaxial Cylinders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magneto-Nanofluid Flow via Mixed Convection Inside E-Shaped Square Chamber

1
Department of Mathematics, College of Science and Humanities in Al-Kharj, Prince Sattam bin Abdulaziz University, Al-Kharj 11942, Saudi Arabia
2
Department of Basic Engineering Science, Faculty of Engineering, Menoufia University, Shebin El-Kom 32511, Egypt
3
Department of Mathematics, Faculty of Science, Aswan University, Aswan 81528, Egypt
4
Department of Mathematics, Faculty of Science, Assuit University, Assuit 71515, Egypt
5
Basic and Applied Sciences Department, College of Engineering and Technology, Arab Academy for Science & Technology and Maritime Transport (AASTMT), Aswan Branch, Aswan P.O. Box 11, Egypt
*
Author to whom correspondence should be addressed.
Symmetry 2022, 14(6), 1159; https://doi.org/10.3390/sym14061159
Submission received: 13 May 2022 / Revised: 28 May 2022 / Accepted: 1 June 2022 / Published: 4 June 2022
(This article belongs to the Special Issue Symmetry in CFD: Convection, Diffusion and Dynamics)

Abstract

:
Nanofluids play a crucial role in the augmentation of heat transfer in several energy systems. They exhibit better thermal conductivity and physical strength compared to normal fluids. Here, we conduct an evaluative investigation of the magnetized flow of water–copper nanofluid and its heat transport inside a symmetrical E-shaped square chamber via mixed convective impact with a heated corner. The chamber was constructed symmetrically with an inclined magnetic field strength, and the upper surface of the chamber was isolated and set to move at a fixed velocity. The heated corner was set at a fixed hot temperature in both the left and lower directions. The right side was maintained at a fixed cold temperature, while the remaining portions of the left and lower parts were isolated. The investigation was implemented computationally, solving each of the energy and Navier–Stokes models via the application of a symmetrical finite volume method. The following topics have been addressed in this study: the consequences of the magnetic field, the volumetric fraction of nanoparticles, the heat generation–absorption parameters, and the effects of heat-source length and Richardson number on the fluid comportment and heat transport. The outputs of this symmetric study enabled us to arrive at the following derivation: the magnetic field reduces the fluid circulation inside the E-shaped square chamber. The augmentation of the Richardson number leads to an increase in the heat transfer. Moreover, the decrease in heat generation coefficient lowers the nanofluid temperature and weakens the flow fields.

1. Introduction

Mixed convection heat transport is the most utilized phenomenon in several industrial and engineering applications, including nuclear power plants, heat exchangers, solar collectors, and electronic cooling systems [1,2,3,4,5]. Thus, it is a topic of interest for those seeking to understand how it can be managed and controlled to serve its desired purpose. Gangawane [6] investigated mixed convection flow in a square lid-driven enclosure that includes a triangular source of heat. The upper wall of the enclosure moved horizontally whereas the other walls remained fixed. Nayak et al. [7] studied combined convection in a lid-driven cavity. The outputs explained that flow, alongside heat and mass transport, are highly affected by the length and position of the heating and cooling sources. These outputs, which regulate the flow and heat transport, were emphasized by Razera et al. [8], who studied a semi-elliptical heat source in a square lid-driven cavity. Gangawane et al. [9] investigated the effect of the position of a triangular source block on mixed convection in a symmetrical square lid-driven chamber. They concluded that the value of the Nuselt number decreases with an increase in the Richardson number (Ri); furthermore, maximum heat transfer can be achieved when the source is positioned at the center of the cavity. Manchanda and Gangawane [10] studied mixed convection in a dual lid-driven square cavity and a triangular source of heat. The upper and bottom surfaces moved various different directions, whereas the vertical walls remained fixed. They found that the impact of Ri is unnoticeable at power law indexes equaling or less than unity due to having great shear stress, and therefore less viscosity impact. Zhou et al. [11] analyzed mixed convection flow in a double lid-driven enclosure filled with nanofluid. The data suggested that the best configuration for improving heat transport was the fourth of the four cases they examined. Gangawane and Oztop [12] explored compound convective flow in a lid-driven semicircular cavity with an isolated block. They concluded that the best heat-transport improvement resulted from the application of a triangular block shape.
Lately, investigators have been working with several novel types of fluids to achieve more efficient results. The host fluid plays a significant role in a nanofluid mix. In the past, researchers only worked with simple base fluids, but in the last few decades, a preference for nanofluids has arisen. Pioneer analysis of the thermophysical properties of nanofluids has been provided by Choi and Eastman [13]. They noticed that the heat transfer and thermal conductivity of nanofluids have improved compared to basic single-phase base-fluid solvents. Nanofluids are liquids comprised of nanosized particles. Generally, metals, carbides, and oxides are used to form nanofluids in a highly volatile liquid, for example, ethylene glycol, water, oil, paraffin, etc. Nanoparticles are made from ingredients that are completely saturated. To obtain effective thermal conductivity, present-day researchers assemble nanoparticles with unique constructions, as variations in size, basic medium, and particle selection influence effective conductivity. Nanoparticles are useful in a wide range of applications such as car engines, generators, cooling systems and more. Their flexibility allows for their application in advanced industries such as electronic cooling, automotive cooling, heating systems, surfactants, lubricants, metal welding, medical production, solar heating, cooling equipment, integrated engines, cooling engines, cooling integrated circuits (ICs), heat transfer to nuclear reactors, etc. Later, various investigators studied these fluids, as Shahi et al. [14] analyzed combined convection in a ventilated square cavity by producing a nanofluid via explosion. They concluded that, as the concentration of solids increases, so does the heat transfer rate. These outputs have developed via [15,16], who examined combined convection in a cavity with an inclined lid, filled with nanofluid. Ismael et al. [17] explored mixed convection in a lid-driven cavity filled with nanofluid and heated from the bottom corner. They found that when the mixed convection parameter is weak, the effect of the nanoparticles on heat improvement is not remarkable, though this impact becomes pronounced with large Richardson number values. Alsabery et al. [18] analyzed mixed convection in a lid-driven cavity filled with nanofluid, containing a square obstacle. They examined various positions for the obstacle. They concluded that adding nanoparticles improves the heat transport as long as the Re is high. However, the effect of adding more nanoparticles to a cavity with low Re can adversely affect heat transfer. Cho et al. [19] and pal et al. [20] studied the impact of a lid-driven cavity with a wavy wall on nanofluids. Karbasifar et al. [21] studied mixed convection in a square, lid-driven, inclined cavity with a hot elliptical centric cylinder. They deduced that Nusselt number trends are inversely affected by cavity angle, volume fraction, and fluid velocity. Ali et al. [22] used a finite volume scheme to investigate the mixed convection flow of hybrid nanofluid in a double lid-driven cavity. Elshehabey and Ahmed [23] analyzed magneto-mixed convection in a lid-driven cavity that was heated from the sides and filled with nanofluid. Mehmood et al. [24] examined magneto-combined convection in a lid-driven cavity filled with nanofluid. Hussain et al. [25] considered the magnetized nanofluid flow via mixed convection in a dual lid-driven cavity. Mondal and Mahapatra [26] studied the magneto-nanofluid flow via a compound convection in a trapezoidal cavity. Many studies by numerous researchers are devoted to nanofluid flow via mixed convection in enclosures [27,28,29,30,31,32,33,34].
In the view of applied study, especially in the field of fluid dynamics, magnetic fields play an essential role. Scientists from past decades conducted a great deal of research on the properties of magnetic fields. Aspects such as heat transfer, fluid characteristics, and other factors influencing fluid transfer have been studied. However, most of the studies had only a numerical and theoretical basis, meaning not much experimental work has been done regarding magnetic exchange and heat transfer. In the present day, we need experimental results, which are advantageous for industrial application. Cooling devices such as air conditioners and refrigerators are essential for industries in which heat transfer is involved. The use of advanced, modern cooling technology is expected in these situations. In most heat-transfer appliances, upgrades require an increase in surface area, which, in turn, require an increase in overall size and volume. Therefore, overcoming these problems requires more efficient cooling methods. The interactions of Nanofluids with magnetic fields offer possible new solutions regarding the transmission of nanofluid heat. There are two ways to approach the problem. Well-regulated magnetic fields can be used effectively for efficient conversion and thermal management of multi-physical transport. The fixed magnet field changes the speed of the fluid. The multiple physical conditions involved, such as the porous surface and the various properties of nanofluid, make the transport process extremely difficult. For this reason, the magnetic field-based control strategy is widely accepted in many modern systems, and its use is rapidly increasing in various fields of engineering and industry, see [35,36,37,38,39,40,41,42,43].
In light of the gaps in the above investigative literature, we realized that no research was done to examine the magnetized convection flow and heat transport characteristics of water–Cu-based nanofluid inside an E-shaped square chamber. Hence, the current study investigates the magnetized mixed convective of (Cu–water) nanofluid flow inside an E-shaped square cavity with heated corners. Impacts of several related parameters, such as volumetric fraction of nanoparticles, Hartmann number, heat generation/absorption parameters, length of heat source, and Reynolds number are reported. However, nano-colloidal dispersion provides a highly encapsulated method to promote heat transfer, suggesting that the study of a magnetic field’s influence on the thermophysical characteristics of such a fluid would be advantageous. We believe that our efforts will prove valuable to forthcoming nanoscience technologies, from both the scientific and applicative angles.

2. Modeling

An examination of magneto-nanofluid mixed convective flow inside an E-shaped square chamber with length H is performed with the heat corners indicated in Figure 1. The following parameters are established for the exploration:
Steady viscous mixed convection flow.
The chamber filled with copper-water nanofluid is considered to be steady Newtonian, Laminar, and incompressible.
The upper surface of the chamber is adiabatic and moves with a fixed velocity U0.
The heated corner is submitted to a fixed hot temperature Th in both left and lower directions with length sb1 and b2.
The right side is maintained at a fixed cold temperature, TC, while the remaining portions of the left and lower parts are considered adiabatic.
The height and width of the portions on the right side are measured as l1 and l2, respectively.
The gravity force is directed downward and internal heat generation at a constant rate of Q0 is included.
An application of magnetic strength B0 is utilized on the left side of the cavity, with angle Φ along the positive horizontal direction.
It is assumed that the fluid flow within the enclosure is in thermal equilibrium and the working fluid is incompressible nanofluid.
There will be no jump of temperature between base fluid and nanosized particles, and both the viscous dissipation term and Joule heating term due to magnetic field has been waived.
The magnetic field created by induction is small enough to be negligible compared to the magnetic field which is applied.
The thermophysical characteristics of the nanoparticles and the base liquid are addressed in Table 1.
By keeping the above-mentioned presumptions in mind and following Refs. [17,18], the flow and energy equations are addressed below:
u x + v y = 0
u u x + v u y = 1 ρ n f p x + ν n f ( 2 u x 2 + 2 u y 2 ) + σ n B 0 2 ρ n f ( v sin Φ cos Φ u sin 2 Φ )
u v x + v v y = 1 ρ n f p y + ν n f ( 2 v x 2 + 2 v y 2 ) + σ n f B 0 2 ρ n f ( u sin Φ cos Φ v cos 2 Φ ) + ( ρ β ) n f ρ n f g ( T T c )
u T x + v T y = α n f ( 2 T x 2 + 2 T y 2 ) + Q 0 ( ρ c p ) n f ( T T c )
The boundary conditions are:
x = 0 , T = T h , 0 y b 2 , T x = 0 , otherwise u = 0 , v = 0
y = 0 , T = T h , 0 x b 1 , T y = 0 , otherwise u = 0 , v = 0
x = H , T = T c a t 0 y d , d + l 1 y H ( d + l 1 ) a n d 2 d + 2 l 1 y H T x = 0 , otherwise u = 0 , v = 0
y = H , T y = 0 v = 0 , u = U 0
where:
ρ n f , ( ρ c p ) n f , ( ρ β ) n f , μ n f , α n f are defined as (see [44,45]);
ρ n f = ( 1 ϕ ) ρ f + ϕ ρ p
( ρ c p ) n f = ( 1 ϕ ) ( ρ c p ) f + ϕ ( ρ c p ) p
( ρ β ) n f = ( 1 ϕ ) ( ρ β ) f + ϕ ( ρ β ) p
α n f = k n f ( ρ c p ) n f
k n f = k f ( k p + 2 k f ) 2 ϕ ( k f k p ) ( k p + 2 k f ) + ϕ ( k f k p )
μ n f = μ f ( 1 ϕ ) 2.5
σ n f = σ f ( 1 + ( 3 ( γ 1 ) ϕ ) ( ( γ + 2 ) ( γ 1 ) ϕ ) ) ,   γ = σ p σ f
Dimensionless parameters:
L 1 = l 1 H , X = x H , L 2 = l 2 H , Y = y H , B 1 = b 1 H , D = d H , B 2 = b 2 H , V = v U 0 , U = u U 0 , θ = ( T T c ) Δ T , R i = G r Re 2 , Δ T = ( T h T c ) , P = p ρ n f U 0 2 , Q = H 2 k f Q 0
in Equations (1)–(4) produces the following non-dimensional equations:
U X + V Y = 0
U U X + V U Y = P X + 1 Re . ( ν n f ν f ) ( 2 U X 2 + 2 U Y 2 ) + ( ρ f ρ n f ) ( σ n f σ f ) . H a 2 Re ( V sin Φ cos Φ U sin 2 Φ )
U V X + V V Y = P Y + 1 Re . ( ν n f ν f ) ( 2 V X 2 + 2 V Y 2 ) + R i . ( ρ β ) n f ρ n f . β f . θ + ( ρ f ρ n f ) ( σ n f σ f ) H a 2 Re ( U sin Φ cos Φ V cos 2 Φ )
U θ X + V θ Y = ( 1 Pr . Re ) α n f α f ( 2 θ X 2 + 2 θ Y 2 ) + 1 Re . Pr ( ρ c p ) f ( ρ c p ) n f . Q . θ
The non-dimensional boundary conditions (14)–(17) are:
X = 0 , θ = 1 , 0 Y B 2 , θ X = 0 , otherwise U = V = 0
Y = 0 , θ = 1 , 0 X B 1 , θ Y = 0 , otherwise U = 0 , V = 0
X = 1 , θ = 0 at 0 Y D , D + L 1 Y 1 ( D + L 1 ) a n d 2 D + 2 L 1 Y 1 , θ X = 0 , otherwise U = 0 , V = 0
Y = 1 , U = U 0 , V = 0 , θ Y = 0
The local Nusselt number is:
N u s = k n f k f ( θ Y ) Y = 0 , N u s = k n f k f ( θ X ) X = 0
The average Nusselt number is:
N u m = N u X = 0 + N u Y = 0 2
where: N u Y = 0 = 1 B 2 0 B 2 N u s d Y , N u X = 0 = 1 B 1 0 B 1 N u s d X .

3. Numerical Solution and Validation

In this study, the finite volume method with the SIMPLE algorithm [46] has been used to solve the continuity, momentum, and energy equations. The first step in this method is that Equations (15)–(17) are rewritten in the following general form:
( u ϕ ) x + ( v ϕ ) x = x ( Γ ϕ x ) + y ( Γ ϕ y ) + S ϕ
Integrating the previous system over the control volume gives:
F e ϕ e F w ϕ w + F n ϕ n F s ϕ s = D e ( ϕ E ϕ P ) D w ( ϕ P ϕ W ) + D n ( ϕ N ϕ P ) D s ( ϕ P ϕ S ) + S ϕ   V o l
where e ,   w ,   n ,   s refer to the east, west, north, and south locations, respectively. In addition, the discretized form of the continuity equation is expressed as:
F e F w + F n F s = 0
where:
F e = A e u e ,   F w = A w u w ,   F n = A n v n ,   F s = A s v s
D e = Γ e A e δ x P E ,   D w = Γ w A w δ x P W ,   D n = Γ n A n δ y P N ,   D s = Γ s A s δ y P S
The upwind sachem is used to approximate the cell face values of the convective term as:
ϕ e = { ϕ P           f o r       F e > 0 ϕ E           f o r     F e 0
Also, it should be noted that the central difference method is used for the diffusive term. The general form of the previous system is given as:
a P ϕ P = a E ϕ E + a W ϕ W + a N ϕ N + a S ϕ S + S p
where:
a E = D e + max ( F e , 0 )
a W = D w + max ( F w , 0 )
a N = D n + max ( F n , 0 )
a S = D s + max ( F s , 0 )
a P = a E + a W + a N + a S + F e F w + F n F s
S p = S ϕ v o l
The under-relaxed form is:
ϕ P = α ϕ a P ( a E ϕ E + a W ϕ W + a N ϕ N + a S ϕ S + S p   ) + ( 1 α ϕ ) ϕ P 0
The alternating direct implicit (ADI) procedure is used to solve the resultant algebraic equations (Equation (25)), while the SIMPLE algorithm is used to correct for the velocities and pressure. The following criteria of convergence was implemented for unknown dependent variables:
i , j | χ i , j new χ i , j old | 10 6
Here, in order to choose the most suitable grid, a grid test is performed and presented in Table 1. It is observed that a grid size of 101 × 101 is most suitable for all calculations.
The obtained outcomes are independent of the grid-size number. The grid independency results are given at Ha = 10, Ri = 1.0, Q = 1.0, Φ = 45, ϕ = 0.05, L1 = L2 = 0.5, B1 = B2 = 0.5, and presented in Table 2.
Also, Figure 2 exhibits the comparison of the outcomes of this investigation with those of Khanafer and Chamkha [47] and Iwatsu et al. [48]. As observed in this figure, the outcomes are in agreement with these data.

4. Discussions

Numerical simulations are carried out for nanofluid flow via mixed convective (Cu/H2O) inside an E-shaped square chamber with a heated corner. The effects of Richardson number (0.1 ≤ Ri ≤ 100), Hartmann number (0 ≤ Ha ≤ 100), volumetric fraction of nanoparticles (0 ≤ φ ≤ 0.1), heat generation/absorption parameter (−2 ≤ Q ≤ 2), and length of the heat source (B1, B2) on isotherms and streamlines are analyzed. The outputs are obtained for the following fixed parametric values: L1 = L2 = 0.5, Φ = π/4, Gr = 104.

4.1. Effect of Richardson Number Ri

For fixed values of the volumetric fraction of nanoparticles (φ = 0.05) and the (Ha = 10), the isotherms and streamlines at several values of Richardson number (0.1 ≤ Ri ≤ 100) are presented in Figure 3a,b.
The examination of the framework of streamlines exhibited in this figure displays that the circulations inside the chamber are greatly affected by the Richardson number. For low Ri = 0.1, 1, 10 the streamlines are described by the existence of the circulation cell that exists inside the upper surface of the chamber. At the same time, it is clear that the structure of isotherms is equivalent to the symmetrical isothermal surfaces and orthogonal to the adiabatic surface, thus leading to a reduction in temperature gradient. In instances where the Richardson number is weak, the energy-transport impact is governed by the conduction. Moreover, the Richardson number (Ri) is related to Grashof Number per square Reynolds number (Re), and in this section, the Re = 100 and the Richardson number changes via the changing of the Grashof number. However, at Ri = 100, the influence of the Richardson number on the streamlines and isotherms is more remarkable than at Ri = 0.1, where the variation between the compared streamlines and isotherms improves remarkably. By boosting the Richardson number, the impact of the convective heat transport becomes more considerable, and diverse boundary layers form alongside the corner heater parts (B1 and B2) of the chamber.
Figure 4a,b explains the local Nusselt number on the heated parts (B1 and B2) at varying Richardson number values. It is detected from these figures that the local Nusselt number is boosted with each increment in the Richardson number. This is because as Ri grows, the buoyancy force improves, which impacts their combined convective flow. The most extreme impact of Ri on the local Nusselt number is found in various cases where X = 0.5 and Y = 0.4 at the heated corner side walls (B1 and B2). However, to the contrary, alongside the vertical heated wall (B2), the maximum of the Nusselt number is decreased; Ri value increases from 6.14 to 25.3. It should be noted that the increment of heat transfers outweighs the increment of dynamic viscosity, which causes a decrease in the pressure and force related to Ri = 3.52.

4.2. Effect of the HARTMANN Number Ha

Figure 5a,b demonstrates the influence of the Hartmann number on the isotherms and streamlines with φ = 0.05 and Ri = 1.0 and Q = 1.0. Examining Equation (15), the indication of Ha is reverse to the Reynolds number Re in exporter manifestation. Hence, there is a reverse effectiveness of Ri and Ha on mixed convective flow regime.
The magnetic field reduces the fluid circulation inside the E-shaped square chamber. It is revealed that the temperature gradient on the heated part declines as the Hartmann number boosts. Moreover, the obtained data explain that the strength of the nanofluid temperature in the chamber diminishes as a result of the implementation of the magnetic domain. However, from Ha = 0 to Ha = 25, the increment of Ha causes the diminution in the intensity of the cells and flow, and to some extent, the flow regime rests uncharged. This tendency is also visible for the isothermal lines. That is, the form of these lines does not alert observably up to Ha = 25, while it converts at Ha = 100. The Lorentz force and velocity boundary drive to a different behavior than the Hartmann number.
Figure 6a,b exhibits the local Nusselt number with several Hartmann numbers. The behavior of this figure depicts that the augmentation in the magnetic force effect causes a decrease of the Nusselt number. This can be explained by the existence of magnetic forces in the term of volume forces.

4.3. Effect of Nanoparticle Volume Fraction φ

Figure 7a,b manifests the behavior of volume fraction φ (φ = 0.01, 0.05, 0.08 and 0.1) on the frames of streamlines and temperature contours. It is clear that the circulations inside the chamber and the temperature contours are practically unaffected by the addition of Cu nanoparticles to regular fluid. So the flow direction is not modified.
At the same time, the Nusselt coefficient alongside the heat source grows by boosting the φ, as demonstrated in Figure 8a,b. This is due to the fact that enhancing the thermal conductivity of Cu/H2O yields an improvement in heat transport. However, it is apparent that the augmentation of the dynamic viscosity due to the existence of the nanoparticles encourages additional resistance, due to shear forces versus the nanofluid motion and decline in fluid velocity. Moreover, any reduction in the fluid velocity would directly minimize the heat transport by diminishing the advection mechanism. The reduction in the advective heat transport mechanism causes a decline in the aggregate heat transport in the chamber. However, the heat transfer rate increases by raising the volume fraction of nanoparticles. Moreover, it is important to point out that the presence of nanoparticles leads to a heightened convective heat transfer coefficient.

4.4. Effect of Heat Generation Coefficient Q

Figure 9a,b manifests the effect of the heat generation/absorption coefficient on both the flow and temperature contours for Ha = 10, φ = 0.05, B1 = 0.5, B2 = 0.5, Q = 1.0 and Ri = 1.0. It can be seen from these plots that the decrease in the heat generation coefficient Q lowers the nanofluid temperature, and that the flow fields for various Q values are almost identical, which results in the presence of nanofluid with reduced heat transport capabilities.
Moreover, in Figure 10a,b, the local Nusselt number with altered Q values are charted. It is visualized here that the Nusselt number decreases with Q; however, heat transport declines with an increase in Q. This is because positive heat generation boosts the temperature scale inside the chamber, hence the temperature gradient around the flux source falls. In contrast, the temperature gradients around the cold parts improve at the same time; that is to say, the Nusselt number at the cold parts slightly enlarges with Q. Alternately, a heat sink (negative Q) enhances the Nusselt number along the bottom and left parts.

4.5. Effect of Length of the Heat Source B1 and B2

Figure 11a,b show the streamlines and isotherms for four different heat source length values (B1, B2): case 1 (B1 = 0.2, B2 = 1.0), case 2 (B1 = 0.4, B2 = 0.8), case 3 (B1 = 0.6, B2 = 0.6) and case 4 (B1 = 1.0, B2 = 0.5). It is clearly shown that the circulations inside the chamber are practically unaffected by the length of the heat source. Thus, the direction of the flow is not modified. A comparison of the temperature contours obtained for the four heat source lengths mentioned above shows that the most intense fluid temperature is found in case two, and the least significant is observed in case four. For cases one, two, and three, the strength of the nanofluid temperature in the chamber near the vertical left wall is more important, and various boundary layers form along the active walls of the cavity. Conversely, an opposing effect occurs in the intensities of the temperature at the bottom side. For case four, when the length of the heat source (B1) increases, the intensities of the temperature of nanofluid increase near the bottom side.

5. Conclusions

In this investigation, a symmetric simulation is designed to describe the magnetized flow of a water-copper nanofluid and its heat transport in inside E-shaped square chamber via mixed convective impact with a heated corner. The outcomes of various embedded parameters on various physiological quantities are scrutinized. A few of the challenging significances are as follows:
-
The heat transfer rate increases by raising the volume fraction of nanoparticles. It is important to point out that the presence of nanoparticles leads to height convective heat transfer coefficient.
-
The Nusselt number increases as the thermal conductivity and the size of nanoparticles increase.
-
The increment in heat transfers outweighs the increment in dynamic viscosity, which causes a decrease in pressure and force.
-
The magnetic field reduces the fluid circulation inside the E-shaped square chamber.
-
The augmentation of the Richardson number leads to an increase in the heat transfer.
-
The circulations inside the chamber are practically unaffected by the length of heat source.
-
The circulations and temperature contours are practically unaffected by the addition of nanoparticles to regular fluid.
-
The decrease in heat generation coefficient declines along with the nanofluid temperature and the flow fields.

Author Contributions

All authors have equal work. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the Deputyship for Research &Innovation, Ministry of Education in Saudi Arabia for funding this research work through the project number (IF-PSAU-2021/01/17862).

Data Availability Statement

Data available upon Request.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

b1, b2Heat source lengths (m)
B1, B2Dimensionless Heat source lengths
B 0 Magnetic field strength, T
C p Specific heat at constant pr, Messure, J . k g . K 1
Dcold length
GAcceleration due to gravity, m s−2
GrGrashofnumber, g β f H 3 Δ T / υ 2 f
HLength of cavity
HaHartmann number, B 0 L σ f / ρ f ν f
k Thermal conductivity, Wm−1K−1
l1Length of cold temperutre (m)
l2width of cold temperutre (m)
L1Dimensionless Length of cold temperutre
L2Dimensionlesswidth of cold temperutre
NusLocal Nusselt number
N u m Average Nusselt number of heat source
p Fluid pressure, Pa
P Dimensionless   pressure ,   pH / ρ n f α f 2
Pr Prandtl   number ,   υ f / α f
Re Reynolds   number ,   U 0 H / υ f
RiRichardson number G r / Re 2
TTemperature, K
TcCold wall temperature, K
ThHeated wall temperature, K
u, vVelocity components in x, y directions, ms−1
U , V Dimensionless velocity components, u/U0, v/U0
x , y Cartesian coordinates
X, YDimensionlesscoordinates, x/H, y/H
Greek symbols
α Thermal   diffusivity ,   m 2 . s 1 ,   k / ρ c p
β Thermal expansion coefficient, K−1
ϕ Solid volume fraction
σ Effective   electrical   conductivity ,   μ S / c m
θ Dimensionless temperature,
μ Dynamic viscosity, N.S.m−2
ν Kinematic   viscosity ,   m 2 . s 1
ρ Density,
ΦInclination angle
Subscripts
c Cold
f Pure fluid
h Hot
m Average
n f Nanofluid
PNanoparticle

References

  1. Boulemtafes-Boukadoum, A.; Abid, C.; Benzaoui, A. 3D Numerical study of the effect of aspect ratio on mixed convection air flow in upward solar air heater. Int. J. Heat Fluid Flow 2020, 84, 108570. [Google Scholar] [CrossRef]
  2. Hadavand, M.; Yousefzadeh, S.; Akbari, O.A.; Pourfattah, F.; Nguyen, H.M.; Asadi, A. A numerical investigation on the effects of mixed convection of Ag-water nanofluid inside a simcircular lid-driven cavity on the temperature of an electronic silicon chip. Appl. Therm. Eng. 2019, 162, 114298. [Google Scholar] [CrossRef]
  3. Shin, D.-H.; Kim, C.S.; Park, G.-C.; Cho, H.K. Experimental analysis on mixed convection in reactor cavity cooling system of HTGR for hydrogen production. Int. J. Hydrogen Energy 2017, 42, 22046–22053. [Google Scholar] [CrossRef]
  4. Tassone, A.; Caruso, G.; Giannetti, F.; del Nevo, A. MHD mixed convection flow in the WCLL: Heat transfer analysis and cooling system optimization. Fusion Eng. Des. 2019, 146, 809–813. [Google Scholar] [CrossRef] [Green Version]
  5. Xiao, C.; Liao, Q.; Fu, Q.; Huang, Y.; Xia, A.; Chen, H.; Zhu, X. Numerical investigation of laminar mixed convection of microalgae slurry flowing in a solar collector. Appl. Therm. Eng. 2020, 175, 115366. [Google Scholar] [CrossRef]
  6. Gangawane, K.M. Computational analysis of mixed convection heat transfer characteristics in lid-driven cavity containing triangular block with constant heat flux: Effect of Prandtl and Grashof numbers. Int. J. Heat Mass Transf. 2017, 105, 34–57. [Google Scholar] [CrossRef]
  7. Nayak, A.K.; Haque, A.; Banerjee, A. Thermosolutal mixed convection of a shear thinning fluid due to partially active mixed zones within a lid-driven cavity. Int. J. Heat Mass Transf. 2017, 106, 686–707. [Google Scholar] [CrossRef]
  8. Razera, A.L.; da Fonseca, R.J.C.; Isoldi, L.A.; Santos, E.D.d.; Rocha, L.A.O.; Biserni, C. Constructal design of a semi-elliptical fin inserted in a lid-driven square cavity with mixed convection. Int. J. Heat Mass Transf. 2018, 126, 81–94. [Google Scholar] [CrossRef]
  9. Gangawane, K.M.; Oztop, H.F.; Abu-Hamdeh, N. Mixed convection characteristic in a lid driven cavity containing heated triangular block: Effect of location and size of block. Int. J. Heat Mass Transf. 2018, 124, 860–875. [Google Scholar] [CrossRef]
  10. Manchanda, M.; Gangawane, K.M. Mixed convection in a two-sided lid-driven cavity containing heated triangular block for non-Newtonian power-law fluids. Int. J. Mech. Sci. 2018, 144, 235–248. [Google Scholar] [CrossRef]
  11. Zhou, W.; Yan, Y.; Liu, X.; Chen, H.; Liu, B. Lattice Boltzmann simulation of mixedconvection of nanofluid with different heat sources in a double lid-driven cavity. Int. Commun. Heat Mass Transf. 2018, 97, 39–46. [Google Scholar] [CrossRef]
  12. Gangawane, K.M.; Oztop, H.F. Mixed convection in the semi-circular lid-driven cavity with heated curved wall subjugated to constant heat flux for non-Newtonian power-law fluids. Int. Commun. Heat Mass Transf. 2020, 114, 104563. [Google Scholar] [CrossRef]
  13. Choi, S.U.S.; Eastman, J.A. Enhancing thermal conductivity of fluids withnanoparticles. In Proceedings of the 1995 International Mechanical Engineering Congress and Exhibition, San Francisco, CA, USA, 12–17 November 1995. [Google Scholar]
  14. Shahi, M.; Mahmoudi, A.H.; Talebi, F. Numerical study of mixed convective cooling in a square cavity ventilated and partially heated from the below utilizing nanofluid. Int. Heat Mass Transf. 2010, 37, 201–213. [Google Scholar] [CrossRef]
  15. Hussain, S.; Ahmad, S.; Mehmood, K.; Sagheer, M. Effects of inclination angle on mixed convective nanofluid flow in a double lid-driven cavity with discrete heat sources. Int. Heat Mass Transf. 2017, 106, 847–860. [Google Scholar] [CrossRef]
  16. Esfe, M.H.; Akbari, M.; Karimipour, A.; Afrand, M.; Mahian, O.; Wongwises, S. Mixed-convection flow and heat transfer in an inclined cavity equipped to a hot obstacle using nanofluids considering temperature-dependent properties. Int. J. Heat Mass Transf. 2015, 85, 656–666. [Google Scholar] [CrossRef]
  17. Ismael, M.A.; Abu-Nada, E.; Chamkha, A.J. Mixed convection in a square cavity filled with CuO-water nanofluid heated by corner heater. Int. J. Mech. Sci. 2017, 133, 42–50. [Google Scholar] [CrossRef]
  18. Alsabery, A.I.; Ismael, M.A.; Chamkha, A.J.; Hashim, I. Mixed convection of Al2O3-water nanofluid in a double lid-driven square cavity with a solid inner insert using Buongiorno’s twophasemodel. Int. J. Heat Mass Transf. 2018, 119, 939–961. [Google Scholar] [CrossRef]
  19. Cho, C.-C. Heat transfer and entropy generation of mixed convection flow in Cu-waternanofluid-filled lid-driven cavity with wavy surface. Int. J. Heat Mass Transf. 2018, 119, 163–174. [Google Scholar] [CrossRef]
  20. Pal, S.K.; Bhattacharyya, S.; Pop, I. Effect of solid-to-fluid conductivity ratio on mixedconvection and entropy generation of a nanofluid in a lid-driven enclosure with a thick wavywall. Int. J. Heat Mass Transf. 2018, 127, 885–900. [Google Scholar] [CrossRef]
  21. Karbasifar, B.; Akbari, M.; Toghraie, D. Mixed convection of Water-Aluminum oxidenanofluid in an inclined lid-driven cavity containing a hot elliptical centric cylinder. Int. J. Heat Mass Transf. 2018, 116, 1237–1249. [Google Scholar] [CrossRef]
  22. Ali, I.; Alsabery, A.I.; Bakar, N.; Roslan, R. Mixed Convection in a Double Lid-Driven Cavity Filled with Hybrid Nanofluid by Using Finite Volume Method. Symmetry 2020, 12, 1977. [Google Scholar] [CrossRef]
  23. Elshehabey, H.M.; Ahmed, S.E. MHD mixed convection in a lid-driven cavity filled by ananofluid with sinusoidal temperature distribution on the both vertical walls using Buongiorno’s nanofluid model. Int. J. Heat Mass Transf. 2015, 88, 181–202. [Google Scholar] [CrossRef]
  24. Mehmood, K.; Hussain, S.; Sagheer, M. Mixed convection in alumina-water nanofluidfilled lid-driven square cavity with an isothermally heated square blockage inside withmagnetic field effect: Introduction. Int. J. Heat Mass Transf. 2017, 109, 397–409. [Google Scholar] [CrossRef]
  25. Hussain, S.; Öztop, H.F.; Mehmood, K.; Abu-Hamdeh, N. Effects of inclined magnetic fieldon mixed convection in a nanofluid filled double lid-driven cavity with volumetric heatgeneration or absorption using finite element method. Chin. J. Phys. 2018, 56, 484–501. [Google Scholar] [CrossRef]
  26. Mondal, P.; Mahapatra, T.R.; Parveen, R. Entropy generation in nanofluid flow due todouble diffusive MHD mixed convection. Heliyon 2021, 7, e06143. [Google Scholar] [CrossRef] [PubMed]
  27. Rasool, G.; Wakif, A. Numerical spectral examination of EMHD mixed convective flow of second-grade nanofluid towards a vertical Riga plate using an advanced version of the revised Buongiorno’s nanofluid model. J. Therm. Anal. Calorim. 2021, 143, 2379–2393. [Google Scholar] [CrossRef]
  28. Punith Gowda, R.J.; Naveen Kumar, R.; Jyothi, A.M.; Prasannakumara BCSarris, I.E. Impact of binary chemical reaction and activation energy on heat and mass transfer of marangoni driven boundary layer flow of a non-Newtonian nanofluid. Processes 2021, 9, 702. [Google Scholar] [CrossRef]
  29. Shah, N.A.; Wakif, A.; El-Zahar, E.R.; Ahmad, S.; Yook, S.-J. Numerical simulation of a thermally enhanced EMHD flow of a heterogeneous micropolar mixture comprising (60%)-ethylene glycol (EG),(40%)-water (W), and copper oxide nanomaterials (CuO). Case Stud. Therm. Eng. 2022, 35, 102046. [Google Scholar] [CrossRef]
  30. Benos, T.L.; Karvelas, E.G.; Sarris, I.E. Crucial effect of aggregations in CNT-water nanofluid magnetohydrodynamic natural convection. Therm. Sci. Eng. Prog. 2019, 11, 263–271. [Google Scholar] [CrossRef]
  31. Song, Y.Q.; Khan, M.I.; Qayyum, S.; Gowda, R.P.; Kumar, R.N.; Prasannakumara, B.C.; Elmasry, Y.; Chu, Y.-M. Physical impact of thermo-diffusion and diffusion-thermo on marangoni convective flow of hybrid nanofluid (MnZiFe2O4–NiZnFe2O4–H2O) with nonlinear heat source/sink and radiative heat flux. Mod. Phys. Lett. B 2021, 35, 2141006. [Google Scholar] [CrossRef]
  32. Gowda, R.P.; Al-Mubaddel, F.S.; Kumar, R.N.; Prasannakumara, B.; Issakhov, A.; Rahimi-Gorji, M.; Al-Turki, Y.A. Computational modelling of nanofluid flow over a curved stretching sheet using Koo–Kleinstreuer and Li (KKL) correlation and modified Fourier heat flux model. Chaos Solitons Fractals 2021, 145, 110774. [Google Scholar] [CrossRef]
  33. Xiong, P.-Y.; Hamid, A.; Chu, Y.-M.; Khan, M.I.; Gowda, R.J.P.; Kumar, R.N.; Prasannakumara, B.C.; Qayyum, S. Prasannakumara, and Sumaira Qayyum. Dynamics of multiple solutions of Darcy–Forchheimer saturated flow of Cross nanofluid by a vertical thin needle point. Eur. Phys. J. Plus 2021, 136, 315. [Google Scholar] [CrossRef]
  34. Li, Y.-X.; Khan, M.I.; Gowda, R.J.P.; Ali, A.; Farooq, S.; Chu, Y.-M.; Khan, S.U. Dynamics of aluminum oxide and copper hybrid nanofluid in nonlinear mixed Marangoni convective flow with entropy generation: Applications to renewable energy. Chin. J. Phys. 2021, 73, 275–287. [Google Scholar] [CrossRef]
  35. El-Zahar, E.; Rashad, A.M.; Seddek, L.F. Impacts of Viscous Dissipation and Brownian Motion on Jeffrey Nanofluid Flow over an Unsteady Stretching Surface with Thermophoresis. Symmetry 2020, 12, 1450. [Google Scholar] [CrossRef]
  36. Ali, B.; Shafiq, A.; Manan, A.; Wakif, A.; Hussain, S. Bioconvection: Significance of mixed convection and MHD on dynamics of Cassonnanofluid in the stagnation point of rotating sphere via finite element simulation. Math. Comput. Simul. 2022, 194, 254–268. [Google Scholar] [CrossRef]
  37. Habib, D.; Salamat, N.; Hussain, S.; Ali, B.; Abdal, S. Significance of Stephen blowing and Lorentz force on dynamics of Prandtlnanofluid via Keller box approach. Int. Commun. Heat Mass Transf. 2021, 128, 105599. [Google Scholar] [CrossRef]
  38. Asjad, M.I.; Sarwar, N.; Ali, B.; Hussain, S.; Sitthiwirattham, T.; Reunsumrit, J. Impact of Bioconvection and Chemical Reaction on MHD Nanofluid Flow Due to Exponential Stretching Sheet. Symmetry 2021, 13, 2334. [Google Scholar] [CrossRef]
  39. Wang, F.; Asjad, M.I.; Ur Rehman, S.; Ali, B.; Hussain, S.; Gia, T.N.; Muhammad, T. MHD Williamson Nanofluid Flow over a Slender Elastic Sheet of Irregular Thickness in the Presence of Bioconvection. Nanomaterials 2021, 11, 2297. [Google Scholar] [CrossRef]
  40. Haider, S.M.A.; Ali, B.; Wang, Q.; Zhao, C. Stefan Blowing Impacts on Unsteady MHD Flow of Nanofluid over a Stretching Sheet with Electric Field, Thermal Radiation and Activation Energy. Coatings 2021, 11, 1048. [Google Scholar] [CrossRef]
  41. Benos, L.; Sarris, I.E. Analytical study of the magnetohydrodynamic natural convection of a nanofluid filled horizontal shallow cavity with internal heat generation. Int. J. Heat Mass Transf. 2019, 130, 862–873. [Google Scholar] [CrossRef]
  42. Sarris, I.E.; Zikos, G.K.; Grecos, A.P.; Vlachos, N.S. On the limits of validity of the low magnetic Reynolds number approximation in MHD natural-convection heat transfer. Numer. Heat Transf. Part B Fundam. 2006, 50, 157–180. [Google Scholar] [CrossRef]
  43. Raja, M.A.Z.; Shoaib, M.; Tabassum, R.; Khan, M.I.; Gowda, R.J.P.; Prasannakumara, B.C.; Malik, M.Y.; Xia, W.-F. Intelligent computing for the dynamics of entropy optimized nanofluidic system under impacts of MHD along thick surface. Int. J. Mod. Phys. B 2021, 35, 2150269. [Google Scholar] [CrossRef]
  44. Aminossadati, S.M.; Ghasemi, B. Natural convection cooling of a localized heat source at thebottom of a nanofluid-filled enclosure. Eur. J. Mech. B/Fluids 2009, 28, 630–640. [Google Scholar] [CrossRef]
  45. Brinkman, H.C. The viscosity of concentrated suspensions and solution. J. Chem. Phys. 1952, 20, 571–581. [Google Scholar] [CrossRef]
  46. Patankar, S.V. Numerical Heat Transfer and Fluid Flow; Hemisphere: New York, NY, USA, 1980. [Google Scholar]
  47. Khanafer, K.M.; Chamkha, A.J. Mixed convection flow in a lid-driven enclosure filled with a fluid-saturated porous medium. Int. J. Heat Mass Transf. 1999, 31, 1354–1370. [Google Scholar] [CrossRef]
  48. Iwatsu, R.; Hyun, J.M.; Kuwahara, K. Mixed convection in a driven cavity with a stable vertical temperature gradient. Int. J. Heat Mass Transf. 1993, 36, 1601–1608. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram and coordinate system.
Figure 1. Schematic diagram and coordinate system.
Symmetry 14 01159 g001
Figure 2. Comparison of the present study with R e = 1000 , P r = 0.71 ,   G r = 10 2 ,   ϕ = H a = 0 . Khanafer and Chamkha [47]; Iwatsu et al. [48].
Figure 2. Comparison of the present study with R e = 1000 , P r = 0.71 ,   G r = 10 2 ,   ϕ = H a = 0 . Khanafer and Chamkha [47]; Iwatsu et al. [48].
Symmetry 14 01159 g002
Figure 3. (a) Streamlines and (b) Isotherms for Ha = 10, φ = 0.05, B1 = B2 = 0.5, and Ri = 1.0, Q = 1.0.
Figure 3. (a) Streamlines and (b) Isotherms for Ha = 10, φ = 0.05, B1 = B2 = 0.5, and Ri = 1.0, Q = 1.0.
Symmetry 14 01159 g003
Figure 4. Profiles of the local Nusselt number along the heat source for Ha = 10, φ = 0.05, B1 = B2 = 0.5, Q = 1.0. (a,b): local Nusselt number on the heated parts (B1 and B2) at varying Richard-son number values.
Figure 4. Profiles of the local Nusselt number along the heat source for Ha = 10, φ = 0.05, B1 = B2 = 0.5, Q = 1.0. (a,b): local Nusselt number on the heated parts (B1 and B2) at varying Richard-son number values.
Symmetry 14 01159 g004
Figure 5. (a) Streamlines and (b) Isotherms for φ = 0.05, B1 = B2 = 0.5, Ri = 1.0, Q = 1.0.
Figure 5. (a) Streamlines and (b) Isotherms for φ = 0.05, B1 = B2 = 0.5, Ri = 1.0, Q = 1.0.
Symmetry 14 01159 g005aSymmetry 14 01159 g005b
Figure 6. Profiles of the local Nusselt number along the heat source for φ = 0.05, B1 = B2 = 0.5, Q = 1.0, Ri = 1.0. (a,b): local Nusselt number with several Hartmann numbers.
Figure 6. Profiles of the local Nusselt number along the heat source for φ = 0.05, B1 = B2 = 0.5, Q = 1.0, Ri = 1.0. (a,b): local Nusselt number with several Hartmann numbers.
Symmetry 14 01159 g006
Figure 7. (a) Streamlines and (b) Isotherms for Ha = 10, B1 = B2 = 0.5, Ri = 1.0, Q = 1.0.
Figure 7. (a) Streamlines and (b) Isotherms for Ha = 10, B1 = B2 = 0.5, Ri = 1.0, Q = 1.0.
Symmetry 14 01159 g007aSymmetry 14 01159 g007b
Figure 8. Profiles of the local Nusselt number along the heat source for Ha = 10, B1 = B2 = 0.5, Q = 1.0, Ri = 1.0. (a,b): local Nusselt number with altered φ values.
Figure 8. Profiles of the local Nusselt number along the heat source for Ha = 10, B1 = B2 = 0.5, Q = 1.0, Ri = 1.0. (a,b): local Nusselt number with altered φ values.
Symmetry 14 01159 g008
Figure 9. (a) Streamlines and (b) Isotherms for Ha = 10, φ = 0.05, B1 = B2 = 0.5, Ri = 1.0.
Figure 9. (a) Streamlines and (b) Isotherms for Ha = 10, φ = 0.05, B1 = B2 = 0.5, Ri = 1.0.
Symmetry 14 01159 g009
Figure 10. Profiles of the local Nusselt number along the heat source for Ha = 10, φ = 0.05, B1 = B2 = 0.5, Ri = 1.0. (a,b): local Nusselt number with altered Q values.
Figure 10. Profiles of the local Nusselt number along the heat source for Ha = 10, φ = 0.05, B1 = B2 = 0.5, Ri = 1.0. (a,b): local Nusselt number with altered Q values.
Symmetry 14 01159 g010
Figure 11. (a) Streamlines and (b) Isotherms for Ha = 10, φ = 0.05, Q = 1.0, Ri = 1.0.
Figure 11. (a) Streamlines and (b) Isotherms for Ha = 10, φ = 0.05, Q = 1.0, Ri = 1.0.
Symmetry 14 01159 g011
Table 1. Thermophysical properties of water and nanoparticle materials [26].
Table 1. Thermophysical properties of water and nanoparticle materials [26].
ρ (kg m−3) C p (Jkg−1K−1) k (Wm−1K−1) β (K−1) σ (μS/cm)
Pure water997.141790.61321 × 10−50.05
Copper (Cu)89333854011.67 × 10−55.96 × 107
Table 2. Grid-independency study for Cu-water nanofluid.
Table 2. Grid-independency study for Cu-water nanofluid.
Grid-Size41 × 4161 × 6181 × 8191 × 91101 × 101121 × 121
N u m 1.4646921.7324801.9045521.9848081.9942441.9948425
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nabwey, H.A.; Rashad, A.M.; Mansour, M.A.; Salah, T. Magneto-Nanofluid Flow via Mixed Convection Inside E-Shaped Square Chamber. Symmetry 2022, 14, 1159. https://doi.org/10.3390/sym14061159

AMA Style

Nabwey HA, Rashad AM, Mansour MA, Salah T. Magneto-Nanofluid Flow via Mixed Convection Inside E-Shaped Square Chamber. Symmetry. 2022; 14(6):1159. https://doi.org/10.3390/sym14061159

Chicago/Turabian Style

Nabwey, Hossam A., Ahmed M. Rashad, Mohamed A. Mansour, and Taha Salah. 2022. "Magneto-Nanofluid Flow via Mixed Convection Inside E-Shaped Square Chamber" Symmetry 14, no. 6: 1159. https://doi.org/10.3390/sym14061159

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