Next Article in Journal
Multivariant Analysis of Photovoltaic Performance with Consideration of Self-Consumption
Previous Article in Journal
Impacts of Photovoltaics in Low-Voltage Distribution Networks: A Case Study in Malta
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Forced Convection Analysis of Water Alumina Nanofluid Flow through a 3D Annulus with Rotating Cylinders via κε Turbulence Model

1
Department of Mathematics, College of Sciences, King Khalid University, Abha 61413, Saudi Arabia
2
Department of Mathematics and Social Sciences, Sukkur IBA University, Sukkur Sindh 65200, Pakistan
3
Department of Electrical, Electronics, and Telecommunication Engineering, UET Lahore New Campus, Lahore 39161, Pakistan
4
Department of Computer Science, National University of Sciences and Technology, Balochistan Campus (NBC), Quetta 87300, Pakistan
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(18), 6730; https://doi.org/10.3390/en15186730
Submission received: 13 July 2022 / Revised: 27 August 2022 / Accepted: 6 September 2022 / Published: 14 September 2022

Abstract

:
We investigated the dynamics of nanofluid and heat transfer in a three-dimensional circular annular using the κ ε turbulence model and energy equations. The pipe contained two concentric and rotating cylinders with a constant speed in the tangential direction. A heat flux boundary condition was executed at the inner cylinder of the annular. The pipe was settled vertically, and water alumina nanofluid was allowed to enter, with the initial velocity depending on the Reynolds number, ranging from 30,000 to 60,000. The volume fraction of the solid particles was tested from 0.001 to 0.1. The speed of the rotation of the cylinders was tested in the range from 0.5 to 3.5. The simulations were developed using COMSOL Multiphysics 5.6, adopting the finite element procedure for governing equations. The results were validated using the mesh independent study and the average Nusselt number correlations. We found that the average Nusselt number in the middle of the channel decreases linearly with the increase in the volume fraction of the water alumina nanofluid. The novelty of the present work is that various correlations between the average Nusselt number and volume fraction were determined by fixing the Reynolds number and the rotation of the inner cylinder. We also found that fixing the Reynolds number and the volume fraction improves the average Nusselt number at the outlet linearly. In addition, it was stated that the increase in the total mass of the nanofluid would decrease the average temperature at the outer cylinder of the annular. Moreover, the maximum average improvement percentage in the average Nusselt number, which is about 21%, was observed when the inner cylinder rotation was changed from 1.5 to 2.5 m/s.

1. Introduction

It has been more than 100 years since the field of fluid mechanics expanded the tactics to broaden heat transfer in industrial applications. Therefore, when employing fluids with lower thermal conductivity, such as water, glycols, and oil, it is difficult to overcome the requirements on the industrial side. Researchers, through various experiments, affirm that this difficulty can be rectified by using a fluid with an outstanding level of thermal conductivity. The experimental study [1] indicates that utilizing solid particles in the base fluid with some volume fraction will improve the thermal conductivity of the subsequent combination. These particles can even be microscopic and sized in millimeters or nanometers. However, they can be influenced by dilemmas such as pressure expense, clogging, and resistance.
Many decades have passed to promote the technology of rendering such components using a mixture of solid and fluid [2]. These nano-sized particles are adjourned in equilibrium in base fluids, which can transmit them to expand in the domain. Mixing these particles in the base fluid, called conventional fluid, will make the thermal conductivity of the resulting mixture superb. Various fluid dynamics experts have put their effort into checking or measuring the thermal conductivity of the nanofluid by changing the volume fraction of the nanoparticles in the base fluid. They found that the exhibition of such a mixture would heighten the thermal conductivity tremendously. In [3], it was demonstrated that when an oxide ceramic is mixed in either water or glycol, it will increase the thermal conductivity of the resulting nanofluid. It was explained in [4] that when Al2O3 particles of 13 nm in diameter are mixed in a base fluid such as water in a 0.043 volume fraction, this will stimulate the thermal conductivity by about 30% under normal conditions. In [5,6], it was illustrated that a justification for increased nanofluid thermal conductivity might be that mixing nanoparticles will heighten the thermal extent instigated by the wild motion of the nanoparticles and raise the energy exchange in the fluid. Such a procedure will improve the thermal conductivity of nanofluid. Four methods were explained in [7] that boost heat transfer when nanoparticles are added to the base fluid. The researchers explained that improved heat transfer could be caused by Brownian motion, multi-level layering, and ballistic heat transfer of solid particles and clusters. Furthermore, in [8], it was explained that adding a liquid molecule interface to a nanoparticle cluster would improve heat transfer rapidly.
Numerous hypothetical and experimental investigations have been expanded to compute the thermal conductivity of nanofluids. The thermal conductivity observed in experiments was found to be more significant than the hypothetical explorations [9,10]. Various studies are based on the theoretical investigation of the nanofluid’s thermal conductivity to enhance the channel’s heat transfer rate [11,12,13,14,15]. Moreover, many researchers explained the relationship between temperature, the size of nanoparticles, and volume fraction with the thermal conductivity of the nanofluids [16,17]. Most researchers have worked with laminar as well as turbulent fluid, and heat transfer to develop critical correlations for future predictions. Among them, a well-known study was carried out by [18,19]. Their work established the first correlation of the average Nusselt number with the Reynolds number and the Prandtl number while examining the different types of nanofluids. A laminar nanofluid flow was investigated to analyze the convection and the pressure loss due to viscous influences in a vertical test tube [20]. Later, the trapezoidal channel was used to evaluate the thermal execution [21]. It was determined that an enhancement in the thermal performance could be achieved by increasing the nanofluid’s pumping power from the channel’s entrance. Moreover, the volume fraction has a favorable role in heightening the thermal performance, but some extra benefits were lost due to the decline in pressure loss. A square cavity was investigated [22] with a cooled water copper nanofluid with laminar and natural convection conditions. It was determined that the square cavity’s heat distribution decreases with the volume fraction increase when the Rayleigh number is fixed. In addition, the heat transfer is positively influenced by increasing the Rayleigh number for a fixed volume fraction. The mean diameter of the solid particles plays a vital role in heat transmission. It was revealed by [23] that the heat transfer coefficient is improved when the mean diameter of the nanomaterials is increased in the mixed convection case. Applications of nanofluids are also presented in the field of thermoelectric sciences. A thermoelectric module was investigated using a nanofluid-based heat exchanger [24]. It was found that there was a time break for the thermal connection between the two components. There are two ways to observe the heat transfer with the nanofluid dynamics; the first is the single-phase flow, and the second is the two-phase flow. Because the single-phase flow is easy and takes less computation time to give outstanding precision, various theoretical studies [25,26,27,28,29] have been conducted with a single-phase approach. However, as indicated in [30], two-phase fluid can be assumed when the Peclet number is greater than 10, and the mixing can be assumed as non-uniform or two-phase. With a numerical scheme of the finite element method, the boundary layer flow was investigated for mass and heat transfer in the presence of a chemical reaction and unsteady magnetic effect [31]. The results illustrated that the Nusselt number increases the Brownian motion and the thermophoresis parameter. With the implication of Cattaneo–Christov characteristics, the observation of temperature fluctuation in the domain was studied with microbes engrossed in the water-based nanofluid [32]. Mainly, it was elaborated that increasing the melting factor significantly enhances the density of the resultant fluid. The consequence of the magnetic dipole movement for the heat transfer was investigated when two nanoparticles, Fe and Fe3O4, were mixed in a water–ethylene glycol solution [33]. It was found that the thermal conductivity of Fe was more improved than that of Fe3O4. A comparative study to enhance the heat transfer rate with rotational flow under the influence of the magnetic effect was conducted to transport two separate hybrid nanofluids of Al2O3-Cu and Al2O3-TiO over a plane sheet [34]. It was indicated that the temperature and concentration of nanoparticles are higher in the case of Al2O3-Cu than Al2O3-TiO.
Conserving large applications in the heat exchanger, annular geometry is widely explored for the heat transfer phenomenon and is often called Isothermal Furnace Liner (IFL). An experiment [35] was performed to observe the thermodynamics between the unheated length of the annular and the annular ratio by changing the heat transfer coefficient from the inlet of the pipe. An annular was also investigated using an implicit finite difference scheme [36]. The research revealed substantial changes in the heat distribution with a relatively small annular ratio. In [37], laminar forced convection was examined at the annular entrance. With distinct initial thermal conditions, it was depicted that exploring the heat transfer in a pipe-like annular is more significant than circular geometry. A three-dimensional cavity was also investigated to observe the convection effect via two rotating cylinders embedded in the cavity [38]. Three different types of nanofluids were observed to obtain the correlation for the average Nusselt number relating to the Rayleigh number and the angular rotation of the cylinder.

2. Problem Formulation, Boundary Condition, and Thermo-Physical Properties

For many decades, the heat transfer process with nanofluid has been observed for various two-dimensional channels. In this research article, we observe a three-dimensional annular constructed by fitting two cylinders. One of the cylinders is an inner cylinder, and the other is an outer cylinder. To make the observed geometry interesting, we produced both cylinders rotating in a tangential direction/along the y-axis with the same velocity ω (see Figure 1). These two cylinders are fixed so that their centerlines coincide with each other. We are interested in settling these cylinders so that the aspect ratio from the radius of the inner cylinder to the outer cylinder will be 0.5. Let R1 be the radius of the inner cylinder and R2 be the radius of the outer cylinder, then R1/R2 = 0.5.
The length of these two cylinders is the same as tradition. This annular pipe is settled vertically, and the top of this channel is recognized as the inlet channel, whereas the opening from the bottom is selected as an outlet. The inner cylinder is heated with a constant heat flux magnitude of 100 [W/m2]. The initial or reference temperature in this tube is 293.15 K.
Let Uin be the average inlet velocity entered from the top of the pipe; then, it must be the function of the Reynolds number. In the current problem, the Reynolds number ranged from 30,000 to 60,000. The nanofluid volume fraction is chosen in the range of 0.001–0.1. We develop the simulations for the current problem with only ϕ = 0.001, 0.009, 0.01, 0.09, and 0.1. Moreover, the nanofluid flow inside the tube will be dispersed further when the inner and outer cylinders rotate in the same tangential direction of the channel at the same speed. The current simulations are developed using the speeds of these cylinders in the range of 0.5 to 3.5 m/s. The fluid entering the tube is assumed to be laminar and turbulent flow with κ ε , and the turbulent dissipation rate is kept from 0.23403 to 9.2401 m2/s3. The Prandtl number is used between 0.85 and 0.97. Taking Tref = 293.15 as the cool temperature and the updated temperature T as the hot temperature, the Grashof number in the current problem varies from 40.368 to 1391.6.
With a wide range of applications on the industrial side to enhance the heat transfer rate in the domain of interest, nanofluid is traditional. It will follow the same steps in almost all research articles. When a mixture of the base fluid and the nanomaterial occurs in the channel, the empirical formulas to determine the density ρ n f , thermal expansion β n f , and heat capacitance ( ρ c p ) n f are uniquely used and are given in Table 1. Moreover, to determine the effective viscosity and thermal conductivity, many models are available. We used the models described in [8,39]. Further details of the selected parameters and the thermophysical properties of the nanofluid are given in Table 1.
Let u = < u r , u θ , u z > be the velocity field; then, governing Navier–Stokes equations and the κ ε RANS model of turbulence in the vector form are displayed as (1)–(5) and the boundary conditions as (6)–(9).
  • Continuity equation:
    ρ n f ( · u ) = 0
  • Momentum equation:
    ρ n f ( u · ) u = · [ p I + ( μ + μ T ) ( u + ( u ) T ]
  • κ ε turbulence model:
    ρ n f ( u · ) κ = · [ ( μ + μ T σ κ ) κ ] + p κ ρ n f ε
    ρ n f ( u · ) ε = · [ ( μ + μ T σ κ ) ε ] + c ε 1 ε κ c ε 2 ρ n f ε 2 κ
    where
    μ T = ρ n f c μ κ 2 ε   and   p κ = μ T [ u : ( u + ( u ) T ]
  • Boundary conditions:
    Inlet :   z = L ,   0 r 2 R 1 ,   0 θ 2 π : u = U i n n ,   κ = 3 2 ( U i n I T ) ,   ε = c μ 3 / 2 κ L T ,   T n = 0
    Outlet :   z = 0 ,   0 r 2 R 1 ,   0 θ 2 π :   u · n = 0 ,   p = 0 ,   k · n = 0   and   ε · n = 0
    At   outer   cylinder :   r = R 2 ,   0 z L ,   0 θ 2 π :   u · n = ω · n , k · n = 0 ,   ε · n = 0 ,   T n = 0
At the inner cylinder:
r = R 1 ,   0 z L ,   0 θ 2 π :   u · n = ω · n , k · n = 0 ,   ε · n = 0 ,   q 0 = k n f T = 100
where n is the normal vector to the selected boundary, IT = 0.05 cm, and LT = L represents the turbulence intensity and length of the pipe, respectively.
  • Turbulence variable:
The turbulence variables are given in Table 2.
  • Computational Parameters:
Finally, we enlist the computational parameters (10)–(16):
Grashof number:
G r = g β f ( T h T c ) D h 3 μ 2
Prandtl number:
Pr = μ n f ( c p ) n f κ n f
Heat flux:
Q = κ n f κ f T
Heat transfer coefficient:
h = Q A ( T T b )
Bulk temperature:
T b = Ω u T   d Ω Ω u   d Ω
Local Nusselt number along the ith direction:
N u = h   D h k n f
Average Nusselt number:
N u a v g = 1 A Ω N u   d A
Rayleigh number:
R a = G r Pr
  • COMSOL Workflow:
The current physical problem based on the governing equations subject to boundary conditions is solved using the software COMSOL Multiphysics 5.6, which applies the methodology of the finite element method to obtain the simulation. In the current investigation, we developed about 80 different simulations using parameters such as Reynolds number, volume fraction, and the rotational speed of the cylinders. The COMSOL Wagon Wheel contained the following steps:
 Step 1:
To make the observed channel and material used in the channel, we will first select the parameters;
 Step 2:
Develop geometry using the tools of COMSOL Multiphysics 5.6;
 Step 3:
Apply the boundary conditions by selecting the premises of the selected channel;
 Step 4:
Perform the grid or mesh independency test to achieve high accuracy;
 Step 5:
Validate or compare the results with the available results or experimental correlations;
 Step 6:
Compute the numerical results by post-processing.

3. Mesh Independent Study and Validation

To scrutinize the computational results, we execute the independent mesh study in this section. We also attempt to validate the results with the correlation for the average Nusselt number at the exit of this geometry. Since a finite element method is functional to solve the current problem with COMSOL Multiphysics 5.6, we used tetrahedral mesh elements for the current channel (see Figure 2). The present channel is meshed by using various elements from 10,000 to 70,000. The objective of the independent mesh study is to fix the required number of elements to achieve high accuracy for the designated derived variable(s). It is acknowledged that for the numerical methods, good precision for one element will give good precision for the next element. Therefore, using a large and sufficient number of elements is necessary to achieve suitable simulation accuracy. We performed the independent mesh study for the current problem for the average Nusselt number in the middle of the channel z = L/2. We plotted the graph in Figure 3a for the average Nusselt number and the increasing number of elements. It is obvious from the graph that increasing the number of elements will improve the accuracy of the computational results. An independent mesh study is achieved when the number of elements exceeds 30,000 for the computational result, correct to two decimal places.
For the current simulation, the computational results for the average Nusselt number against the increasing Reynolds number at the exit of the channel were validated by various correlations [8,38,39,40]. The results are displayed in Figure 3b and Table 3. It can be seen that the numerical results for the average Nusselt number at the outlet of the channel are comparable to the results obtained by the Dittus–Boelter equation [41].

4. Results Discussion

In this section, the numerical results are displayed using the graphs and tables for the average Nusselt number at the exit of the channel by fixing the Reynolds number and the inner cylinder rotation, mass flow rate, and average temperature. In this study, we sought to establish the correlation between the average Nusselt number and nanofluid volume fraction by fixing the inner cylinder’s rotation. We successfully demonstrated the correlation between the total mass flow rate against the volume fraction of the nanofluids.

4.1. Average Nusselt Number at the Exit of the Channel

Figure 4a–e demonstrates the variation in the average Nusselt number against the increasing Reynolds number at the channel outlet by fixing the volume fraction. The present channel is a three-dimensional annular constructed by two concentric cylinders. The inner and upper cylinders are rotated along the tangential direction with the same speed ω . A heat flux condition is applied to the inner cylinder. We developed the graphs in Figure 4a–e for the average Nusselt number with the fixed volume fraction of the nanomaterial, and the pattern was checked by varying the rotational speed of the inner cylinder. These graphs reveal that the average Nusselt number at the outlet of the channel increases linearly with the increasing Reynolds number for a fixed rotation and volume fraction. The average Nusselt number significantly improved by increasing the rotation of both cylinders. By definition, the Nusselt number is the ratio from the convection to conduction process. By increasing the Nusselt number, it can be determined that the convection process is conquered over the conduction process and vice versa. The increment in the average Nusselt number at the outlet of the channel against the Reynolds number was found in a linear pattern. Moreover, with an increase, the volume fraction rate of the increment is slightly affected. This may be the reason why the mass or density of the nanofluid increased due to an increment in the volume fraction. Therefore, the fluid cannot be dispersed within the same rotation value. It can be noted that when the volume fraction has increased, the difference in the average Nusselt number is reduced by changing the rotation from 0.5 to 1.5 m/s. This behavior can be seen in Figure 4c–e.
Table 4 details the precise impacts of altering the Reynolds number, cylinder rotation, and nanofluid volume fraction.
Table 4 shows that for a fixed rotation speed and Reynolds number, the average Nusselt number at the outlet of the channel decreases with the increase in volume fraction. Moreover, it can be concluded that the average Nusselt number increases with the increase in Reynolds number for fixed rotation and volume fraction of the nanofluids. Moreover, with the fixed Reynolds number and volume fraction, the average Nusselt number is improved by increasing the rotation of the cylinders. Furthermore, in Table 4, the percentage change is calculated when the simulation is obtained by altering the speed of the cylinders by one unit. These percentages are visualized in the fifth, seventh, and ninth columns. By fixing any other parameters, the percentage change in the average Nusselt number decreases with the increase in volume fraction, decreases with the increase in rotation of cylinders, and decreases with the increase in Reynolds number. The maximum percentage change in the Nusselt number can be seen when Re = 30,000, ϕ = 0.001, and the cylinders’ rotation is from 0.5 to 1.5. Moreover, the minimum percentage can be seen when Re = 60,000 and ϕ = 0.001 when the rotation turns from 0.5 m/s to 1.5 m/s.

4.2. Average Nusselt Number, Volume Fraction, and the Correlations

We examined the fluid flow and heat transfer in a three-dimensional annular with water alumina nanoparticles, where the solid particles are present in the base fluid in a ratio from 0.001 to 0.1. The inner and outer cylinders are rotated tangibly at the same speed. We tested the impact of volume fraction on the average Nusselt number by fixing the rotation speed of the cylinders. In addition, we endeavored to clarify the correlations between them.
In Figure 5a–d, we show the numerical results for the average Nusselt number at the end of the annular against the increasing volume fraction by fixing the rotation of the cylinder at ω = 0.5 and the Reynolds number, Re. A clear message is that the average Nusselt number decreases with the increasing volume fraction of the nanoparticles in the base fluid for all the cases. An increment in the average Nusselt number demonstrates the domination of the conduction process over the conduction. Moreover, at the volume fraction’s zero value, we found that the maximum Nusselt number increases with an increase in the Reynolds number. In Figure 6a–d, the same pattern of a decreasing average Nusselt number with an increasing volume fraction was found by fixing the cylinders’ rotational speed at ω = 3.5 and the Reynolds number. In both graphs, a clear linear pattern appears. Therefore, to predict the behavior of the average Nusselt number against the volume fraction, we developed several correlations by using the present conditions of the problem.
Table 5 details several correlations made with the present data.
In these correlations, the multiple constants of the volume fraction ϕ are slopes. It can be interpreted that for a fixed rotation of the inner cylinder, these slopes decrease with the increase in Reynolds number. Moreover, for a fixed rotation and Reynolds number, the rate of these slopes increases by the rotational speed of the cylinders.

4.3. Mass Flow Rate and the Average Temperature on the Outer Cylinder

In the present work, we investigated water alumina nanofluid and heat transfer through an annular with two concentric cylinders. In this three-dimensional annular, the inner and outer cylinders rotate in the tangential direction. A heat flux condition was applied to the inner cylinder. We measured the average temperature at the outer cylinder related to mass entering the channel’s inlet. The reference temperature is Tref = 273.15 K. We can obtain the mass entering the annular with the volume integration of the fluid density.
Mass and volume fraction correlation:
M n f = 0.004698632063 + 0.01400918897 ϕ
In Figure 7, the graph is plotted with an increasing volume fraction and mass of the nanofluid. It can be seen that there is a direct relationship between the mass and the volume fraction. Therefore, applying a linear regression process [45], we can drive a linear regression equation that relates the average mass and the volume fraction of the nanofluid, given as (17).
Mass is the physical quantity that depends on the volume fraction and the density of the nanofluid. It cannot be altered with the other parameters used in the problem, such as the Reynolds number and the inner and outer cylinder rotation. We produced a graph which presents the average temperature against the mass of the nanofluids. From Figure 8a,b, it can be understood that the average temperature decreases with an increase in the mass of the nanofluid up to a certain limit, and then increases. The mass flow rates where an alteration in the average temperature can be determined are about 0.006 and 0.0025 kg/s. Moreover, the alteration can be controlled by increasing or decreasing the Reynolds number, which suggests that increasing the mass flow rate will not always support the average temperature in the domain. It may depend on other factors such as the Reynolds number or the rotation of the inner cylinder. However, it can be understood that keeping a high mass flow rate compared to the low mass flow rate always favored decreasing the average temperature at the outer surface in our study. Additionally, the average temperature can be further improved against the mass flow rate when the rotation of the inner cylinder is increased (see Figure 8a,b). Observing these values for the same Reynolds number can be checked for the average temperature against the mass flow rate.

5. Conclusions

We investigated the heat transfer through water alumina oxide nanofluid using the κ ε turbulence model. A three-dimensional annular consisting of two concentric inner cylinders was kept under observation. A heat flux condition of 100 W/m2 was applied to the inner cylinder of the annular, and both cylinders were movable at the same speed ω . The three-dimensional annular was settled vertically, and the water alumina nanofluid was allowed to enter with inlet Uin, the Reynolds number function. The nanoparticle fraction to the base fluid was taken at 0.001, 0.009, 0.01, 0.09, and 0.1. The Reynolds number was tested in the range of 30,000 to 60,000. The cylinders were moved with the same speed from 0.5 to 3.5 m/s in the tangential direction. The Prandtl number, turbulent dissipation rate, and Grashof number were 0.85–0.97, 0.23403–9.2401 m2/s3, and 40.368–1391.6, respectively. All the simulations were performed using COMSOL Multiphysics 5.6. The outcomes were validated using the mesh independent study and by comparing the average Nusselt number at the outlet of the channel with the Dittus–Boelter equation. The results are displayed in graphs and tables. We conclude with the following points.
The average Nusselt number at the exit of the channel is increased in a linear pattern against the increasing Reynolds number when the rotation of the inner and outer cylinders is fixed. Moreover, by increasing the rotation of the cylinders, the same behavior can be seen between the average Nusselt number and the Reynolds number, but the increment rate is gradually increased.
From the tables, it is apparent that by fixing the Reynolds number and volume fraction, the average Nusselt number at the exit of the channel is increased by increasing the rotation of the circular cylinders. By fixing both the Reynolds number and the rotation of the circular cylinder, the average Nusselt number is reduced by increasing the volume fraction.
The average Nusselt number in the middle of the channel decreases linearly with the increased volume fraction for the fixed rotation rate and the Reynolds number. Therefore, we determined several correlations by applying a linear regression process. These linear regression equations can be used for future predictions based on the conditions.
The mass of the nanofluid entering the region was determined based on the computational results. The mass increased linearly with the increase in volume fraction. Moreover, a linear regression was given between the mass and volume fraction of the nanofluid.
The average temperature against the mass flow rate decreased under certain limitations. Two critical values were found for the mass flow rate where the average temperature increases or decreases. Therefore, we suggest that this increment or decrement could be controlled by using the Reynolds number or the rotation of the inner cylinder, or both.

Author Contributions

Conceptualization, M.S.A. and U.; Data curation, M.S.A. and A.A.M.; Funding acquisition, M.S.A.; Investigation, U.; Project administration, T.M.; Software, A.A.M.; Validation, H.A.; Visualization, A.A.M. and H.A.; Writing—original draft, T.M.; Writing—review & editing, U. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University, Abha, Saudi Arabia for funding this work through Large Groups Project under grant number RGP.2/206/43.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors affirm that they have no known competing financial interest or personal relationship that could have appeared to influence the scientific work reported in this paper.

Future Work

The problem presented in this work can be extended by replacing the inner circular cylinder with another cylinder, such as one with an elliptical shape or a polygonal shape. We expect more developments can be achieved by referring to the present problem.

References

  1. Maxwell, J.C. A Treatise on Electricity and Magnetism; Clarendon Press: Oxford, UK, 1873; Volume 1. [Google Scholar]
  2. Choi, S.U.; Eastman, J.A. Enhancing Thermal Conductivity of Fluids with Nanoparticles; No. ANL/MSD/CP-84938; CONF-951135-29; Argonne National Lab (ANL): Argonne, IL, USA, 1995. [Google Scholar]
  3. Lee, S.; Choi, S.S.; Li, S.A.; Eastman, J.A. Measuring thermal conductivity of fluids containing oxide nanoparticles. J. Heat Transfer. 1999, 121, 280–289. [Google Scholar] [CrossRef]
  4. Masuda, H.; Ebata, A.; Teramae, K. Alteration of thermal conductivity and viscosity of liquid by dispersing ultra-fine particles. Dispersion of Al2O3, SiO2 and TiO2 ultra-fine particles. Netsu Bussei 1993, 7, 227–233. [Google Scholar] [CrossRef]
  5. Xuan, Y.; Li, Q. Heat transfer enhancement of nanofluids. Int. J. Heat Fluid Flow 2000, 21, 58–64. [Google Scholar] [CrossRef]
  6. Xuan, Y.; Roetzel, W. Conceptions for heat transfer correlation of nanofluids. Int. J. Heat Mass Transf. 2000, 43, 3701–3707. [Google Scholar] [CrossRef]
  7. Keblinski, P.; Phillpot, S.R.; Choi, S.U.S.; Eastman, J.A. Mechanisms of heat flow in suspensions of nano-sized particles (nanofluids). Int. J. Heat Mass Transf. 2002, 45, 855–863. [Google Scholar] [CrossRef]
  8. Wang, X.; Xu, X.; Choi, S.U. Thermal conductivity of nanoparticle-fluid mixture. J. Thermophys. Heat Transf. 1999, 13, 474–480. [Google Scholar] [CrossRef]
  9. Choi, S.U.S.; Zhang, Z.G.; Yu, W.; Lockwood, F.E.; Grulke, E.A. Anomalous thermal conductivity enhancement in nanotube suspensions. Appl. Phys. Lett. 2001, 79, 2252–2254. [Google Scholar] [CrossRef]
  10. Zhang, X.; Gu, H.; Fujii, M. Effective thermal conductivity and thermal diffusivity of nanofluids containing spherical and cylindrical nanoparticles. J. Appl. Phys. 2006, 100, 044325. [Google Scholar] [CrossRef]
  11. Xue, Q.Z. Model for effective thermal conductivity of nanofluids. Phys. Lett. A 2003, 307, 313–317. [Google Scholar] [CrossRef]
  12. Xuan, Y.; Li, Q.; Hu, W. Aggregation structure and thermal conductivity of nanofluids. AIChE J. 2003, 49, 1038–1043. [Google Scholar] [CrossRef]
  13. Jang, S.P.; Choi, S.U. Role of Brownian motion in the enhanced thermal conductivity of nanofluids. Appl. Phys. Lett. 2004, 84, 4316–4318. [Google Scholar] [CrossRef]
  14. Chon, C.H.; Kihm, K.D.; Lee, S.P.; Choi, S.U. Empirical correlation finding the role of temperature and particle size for nanofluid (Al 2 O 3) thermal conductivity enhancement. Appl. Phys. Lett. 2005, 87, 153107. [Google Scholar] [CrossRef]
  15. Izadi, M.; Behzadmehr, A.; Jalali-Vahida, D. Numerical study of developing laminar forced convection of a nanofluid in an annulus. Int. J. Therm. Sci. 2009, 48, 2119–2129. [Google Scholar] [CrossRef]
  16. Li, C.H.; Peterson, G.P. The effect of particle size on the effective thermal conductivity of Al2O3-water nanofluids. J. Appl. Phys. 2007, 101, 044312. [Google Scholar] [CrossRef]
  17. Mintsa, H.A.; Roy, G.; Nguyen, C.T.; Doucet, D. New temperature dependent thermal conductivity data for water-based nanofluids. Int. J. Therm. Sci. 2009, 48, 363–371. [Google Scholar] [CrossRef]
  18. Pak, B.C.; Cho, Y.I. Hydrodynamic and heat transfer study of dispersed fluids with submicron metallic oxide particles. Exp. Heat Transf. Int. J. 1998, 11, 151–170. [Google Scholar] [CrossRef]
  19. Xuan, Y.; Li, Q. Investigation on convective heat transfer and flow features of nanofluids. J. Heat Transf. 2003, 125, 151–155. [Google Scholar] [CrossRef]
  20. Yu, W.; France, D.M.; Timofeeva, E.V.; Singh, D.; Routbort, J.L. Comparative review of turbulent heat transfer of nanofluids. Int. J. Heat Mass Transf. 2012, 55, 5380–5396. [Google Scholar] [CrossRef]
  21. Li, J.; Kleinstreuer, C. Thermal performance of nanofluid flow in microchannels. Int. J. Heat Fluid Flow 2008, 29, 1221–1232. [Google Scholar] [CrossRef]
  22. Santra, A.K.; Chakraborty, N.; Sen, S. Prediction of heat transfer due to presence of copper–water nanofluid using resilient-propagation neural network. Int. J. Therm. Sci. 2009, 48, 1311–1318. [Google Scholar] [CrossRef]
  23. Mirmasoumi, S.; Behzadmehr, A. Effect of nanoparticles mean diameter on mixed convection heat transfer of a nanofluid in a horizontal tube. Int. J. Heat Fluid Flow 2008, 29, 557–566. [Google Scholar] [CrossRef]
  24. Agwu Nnanna, A.G.; Rutherford, W.; Elomar, W.; Sankowski, B. January. Assessment of thermoelectric module with nanofluid heat exchanger. In Proceedings of the ASME International Mechanical Engineering Congress and Exposition, Washington, DC, USA, 11–15 November 2007; Volume 43025, pp. 663–672. [Google Scholar]
  25. Nguyen, C.T.; Roy, G.; Gauthier, C.; Galanis, N. Heat transfer enhancement using Al2O3—Water nanofluid for an electronic liquid cooling system. Appl. Therm. Eng. 2007, 27, 1501–1506. [Google Scholar] [CrossRef]
  26. Khanafer, K.; Vafai, K.; Lightstone, M. Buoyancy-driven heat transfer enhancement in a two-dimensional enclosure utilizing nanofluids. Int. J. Heat Mass Transf. 2003, 46, 3639–3653. [Google Scholar] [CrossRef]
  27. Koo, J.; Kleinstreuer, C. Laminar nanofluid flow in microheat-sinks. Int. J. Heat Mass Transf. 2005, 48, 2652–2661. [Google Scholar] [CrossRef]
  28. Akbari, M.; Behzadmehr, A.; Shahraki, F. Fully developed mixed convection in horizontal and inclined tubes with uniform heat flux using nanofluid. Int. J. Heat Fluid Flow 2008, 29, 545–556. [Google Scholar] [CrossRef]
  29. Mirmasoumi, S.; Behzadmehr, A. Numerical study of laminar mixed convection of a nanofluid in a horizontal tube using two-phase mixture model. Appl. Therm. Eng. 2008, 28, 717–727. [Google Scholar] [CrossRef]
  30. Ding, Y.; Wen, D. Particle migration in a flow of nanoparticle suspensions. Powder Technol. 2005, 149, 84–92. [Google Scholar] [CrossRef]
  31. Ali, L.; Ali, B.; Liu, X.; Iqbal, T.; Zulqarnain, R.M.; Javid, M. A comparative study of unsteady MHD Falkner–Skan wedge flow for non-Newtonian nanofluids considering thermal radiation and activation energy. Chin. J. Phys. 2022, 77, 1625–1638. [Google Scholar] [CrossRef]
  32. Ali, L.; Ali, B.; Ghori, M.B. Melting effect on Cattaneo–Christov and thermal radiation features for aligned MHD nanofluid flow comprising microorganisms to leading edge: FEM approach. Comput. Math. Appl. 2022, 109, 260–269. [Google Scholar] [CrossRef]
  33. Ali, L.; Liu, X.; Ali, B.; Mujeed, S.; Abdal, S.; Khan, S.A. Analysis of magnetic properties of nanoparticles due to a magnetic dipole in micropolar fluid flow over a stretching sheet. Coatings 2020, 10, 170. [Google Scholar] [CrossRef] [Green Version]
  34. Ali, B.; Naqvi, R.A.; Ali, L.; Abdal, S.; Hussain, S. A comparative description on time-dependent rotating magnetic transport of a water base liquid H2O with hybrid nanomaterials Al2O3-Cu and Al2O3-TiO2 over an extending sheet using Buongiorno model: Finite element approach. Chin. J. Phys. 2021, 70, 125–139. [Google Scholar] [CrossRef]
  35. Srivastava, N.C.; Bakhtar, F.; Bannister, F.K. An investigation into thermal boundary layer growth in the entrance region of an annulus. Int. J. Heat Mass Transf. 1973, 16, 49–59. [Google Scholar] [CrossRef]
  36. Gupta, S.C.; Garg, V.K. Developing flow in a concentric annulus. Comput. Methods Appl. Mech. Eng. 1981, 28, 27–35. [Google Scholar] [CrossRef]
  37. Lu, G.; Wang, J. Experimental investigation on heat transfer characteristics of water flow in a narrow annulus. Appl. Therm. Eng. 2008, 28, 8–13. [Google Scholar] [CrossRef]
  38. Selimefendigil, F.; Öztop, H.F. Mixed convection of nanofluids in a three dimensional cavity with two adiabatic inner rotating cylinders. Int. J. Heat Mass Transf. 2018, 117, 331–343. [Google Scholar] [CrossRef]
  39. Maxwell-Garnett, J.C. Colours in metal glasses and in metallic films. Philos. Trans. R. Soc. A 1904, 203, 385–420. [Google Scholar]
  40. Hazeri-Mahmel, N.; Shekari, Y.; Tayebi, A. Three-dimensional analysis of forced convection of Newtonian and non-Newtonian nanofluids through a horizontal pipe using single-and two-phase models. Int. Commun. Heat Mass Transf. 2021, 121, 105119. [Google Scholar] [CrossRef]
  41. Dittus, F.W.; Boelter, L.M.K. Heat transfer in automobile radiators of the tubular type. Int. Commun. Heat Mass Transf. 1985, 12, 3–22. [Google Scholar] [CrossRef]
  42. Sieder, E.N.; Tate, G.E. Heat transfer and pressure drop of liquids in tubes. Ind. Eng. Chem. 1936, 28, 1429–1435. [Google Scholar] [CrossRef]
  43. Li, Q.; Xuan, Y. Convective heat transfer and flow characteristics of Cu-water nanofluid. Sci. China Ser. E Technol. Sci. 2002, 45, 408–416. [Google Scholar]
  44. Taler, D.; Taler, J. Simple heat transfer correlations for turbulent tube flow. In E3S Web of Conferences; EDP Sciences: Paris, France, 2017; Volume 13, p. 02008. [Google Scholar]
  45. Available online: https://keisan.casio.com/exec/system/14059929550941 (accessed on 12 July 2022).
Figure 1. A view of the 3-dimensional, vertically settled annular parallel to the axial direction.
Figure 1. A view of the 3-dimensional, vertically settled annular parallel to the axial direction.
Energies 15 06730 g001
Figure 2. (ac) The meshing process of the three-dimensional annular tube with tetrahedral elements.
Figure 2. (ac) The meshing process of the three-dimensional annular tube with tetrahedral elements.
Energies 15 06730 g002
Figure 3. (a) Mesh independent study for average Nusselt number with an increasing number of elements. (b) Comparison of Nusselt number in the inner cylinder with available correlations.
Figure 3. (a) Mesh independent study for average Nusselt number with an increasing number of elements. (b) Comparison of Nusselt number in the inner cylinder with available correlations.
Energies 15 06730 g003
Figure 4. Average Nusselt number at the outlet of the channel with increasing Reynolds number and rotation of cylinders: (a) ϕ = 0.001; (b) ϕ = 0.009; (c) ϕ = 0.01; (d) ϕ = 0.09; (e) ϕ = 0.1.
Figure 4. Average Nusselt number at the outlet of the channel with increasing Reynolds number and rotation of cylinders: (a) ϕ = 0.001; (b) ϕ = 0.009; (c) ϕ = 0.01; (d) ϕ = 0.09; (e) ϕ = 0.1.
Energies 15 06730 g004
Figure 5. The average Nusselt number against increasing volume fractions with ω = 0.5: (a) Re = 30,000, (b) Re = 40,000, (c) Re = 50,000, and (d) Re = 60,000.
Figure 5. The average Nusselt number against increasing volume fractions with ω = 0.5: (a) Re = 30,000, (b) Re = 40,000, (c) Re = 50,000, and (d) Re = 60,000.
Energies 15 06730 g005
Figure 6. The average Nusselt number against increasing volume fractions with ω = 3.5: (a) Re = 30,000, (b) Re = 40,000, (c) Re = 50,000, and (d) Re = 60,000.
Figure 6. The average Nusselt number against increasing volume fractions with ω = 3.5: (a) Re = 30,000, (b) Re = 40,000, (c) Re = 50,000, and (d) Re = 60,000.
Energies 15 06730 g006
Figure 7. The increasing mass of the nanofluid and volume fraction of the outer cylinder.
Figure 7. The increasing mass of the nanofluid and volume fraction of the outer cylinder.
Energies 15 06730 g007
Figure 8. Decreasing average temperature of the nanofluid against the mass of the nanofluid at the outer cylinder: (a) ω = 0.5; (b) ω = 3.5.
Figure 8. Decreasing average temperature of the nanofluid against the mass of the nanofluid at the outer cylinder: (a) ω = 0.5; (b) ω = 3.5.
Energies 15 06730 g008
Table 1. Parameters and thermophysical properties [40].
Table 1. Parameters and thermophysical properties [40].
DescriptionProperty NameValue/Range
Density of water ρ f 997.1 [kg/m3]
Density of aluminum oxide ρ s 3970 [kg/m3]
Heat capacity of water at constant pressure ( c p ) f 4179 [J/(mol K)]
Heat capacity of aluminum oxide ( c p ) s 765 [J/(mol K)]
Thermal conductivity of water κ f 0.613 [W/(m K)]
Thermal conductivity of alumina κ s 40 [W/(m K)]
Thermal expansion of water β f 21 × 10−5 [1/K]
Thermal expansion of water β s 0.85 × 10−5 [1/K]
Volume fraction ϕ 0.001, 0.01, 0.1, 0.2
Density of nanofluid ρ n f ( 1 ϕ ) ρ f + ϕ ρ s
Viscosity of water μ f 0.00081 [Pa s]
Dynamic viscosity of nanofluid [8] μ n f μ f ( 1 + 7.3 ϕ + 123 ϕ 2 )
Acceleration due gravityg9.8 m/s2
Thermal expansion of the nanofluids β n f ( 1 ϕ ) β f + ϕ β s
Thermal conductivity of nanofluid Maxwell Garnet model [39] κ n f κ f κ s + 2 κ f 2 ϕ ( κ f κ s ) κ s + 2 κ f + ϕ ( κ f κ s )
Heat capacitance of fluid ( ρ c p ) f ρ f ( c p ) f
Heat capacitance of solid ( ρ c p ) s ρ s ( c p ) s
Heat capacitance of nanofluid ( ρ c p ) n f ( 1 ϕ ) ( ρ c p ) f + ϕ ( ρ c p ) s
Thermal diffusivity of nanofluid α n f κ n f ( ρ c p ) n f
Reference temperature of the square cavity T r e f 293.15 [K]
Hydraulic diameter D h 4 Area of the channel/Perimeter
Inlet velocity u i n Re μ n f D h ρ n f
Reynolds numberRe35,000
Radius of cylinderAr0.4, 0.5, 0.6
Table 2. Turbulence variables.
Table 2. Turbulence variables.
c ε 1 c ε 2 c μ σ κ σ ε
1.441.920.0911.3
Table 3. Comparison table for the average Nusselt number in inner cylinder with ω = 0.5 and ϕ = 0.001.
Table 3. Comparison table for the average Nusselt number in inner cylinder with ω = 0.5 and ϕ = 0.001.
RePresent WorkDittus and Boelter [41]
0.023 Re 0.8 Pr 1 / 3
Sieder and Tate [42]
0.027 Re 0.8 Pr 1 / 3
Li and Xuan [43]
0.0059 Re 0.9238 Pr 0.4
Dawid and Jan [44] 0.02155 Re 0.8018 Pr 0.7095
3 × 10481.60683.27297.75475.73674.886
4 × 104102.674104.770122.99098.73494.216
5 × 104122.712125.217146.994121.303112.620
6 × 104141.964144.861170.055143.533130.312
Table 4. Average Nusselt number at the outlet of the channel and with percentage increments when the rotation of the cylinders changes by one unit.
Table 4. Average Nusselt number at the outlet of the channel and with percentage increments when the rotation of the cylinders changes by one unit.
Re ϕ ω = 0.5 ω = 1.5 % 1 ω = 2.5 % 2 ω = 3.5 % 3
30,0000.00162.1791.1646.63120.9332.66151.0624.92
30,0000.00961.9089.3644.37117.8331.86146.6624.47
30,0000.0161.8689.1144.04117.3931.74146.0424.41
30,0000.0959.7266.6011.5282.3323.6196.2016.85
30,0000.159.6065.059.1379.3221.9592.1416.15
40,0000.00176.47100.8431.87128.4427.37156.4621.82
40,0000.00976.3099.0229.78125.5626.80152.3921.37
40,0000.0176.2898.7629.48125.1526.72151.8221.31
40,0000.0974.7578.935.5991.3515.74105.7615.77
40,0000.174.6678.124.6388.0612.73101.9115.72
50,0000.00190.48109.2320.72136.8625.30163.3619.36
50,0000.00990.34107.2618.72134.1825.10159.5018.87
50,0000.0190.33106.9918.45133.8125.07158.9618.80
50,0000.0989.1192.233.49100.168.60114.7914.61
50,0000.189.0591.682.9697.836.71110.4812.93
60,0000.001104.08117.4612.86145.8024.13170.9217.23
60,0000.009103.97115.8611.44143.1823.58167.2416.80
60,0000.01103.95115.6511.26142.8123.48166.7216.74
60,0000.09102.95105.492.47110.945.17122.6310.54
60,0000.1102.90105.062.10109.424.15118.558.34
%1 = When rotation changes from ω = 0.5   to   ω = 1.5 . %2 = When rotation changes from ω = 1.5   to   ω = 2.5 . %3 = When rotation changes from ω = 2.5   to   ω = 3.5 .
Table 5. Correlations for average Nusselt number against the increasing volume fraction of water alumina nanofluid.
Table 5. Correlations for average Nusselt number against the increasing volume fraction of water alumina nanofluid.
CorrelationsConditions
N u a v g = 62.14770875 26.1240635 ϕ ω = 0.5, Re = 30,000
N u a v g = 76.47099849 18.56012862 ϕ ω = 0.5, Re = 40,000
N u a v g = 90.47993473 14.728676 ϕ ω = 0.5, Re = 50,000
N u a v g = 104.0795444 12.1610582 ϕ ω = 0.5, Re = 60,000
N u a v g = 91.6519204 271.375263 ϕ ω = 1.5, Re = 30,000
N u a v g = 101.0916353 237.0862955 ϕ ω = 1.5, Re = 40,000
N u a v g = 109.0113258 179.3960469 ϕ ω = 1.5, Re = 50,000
N u a v g = 117.1541 125.0397317 ϕ ω = 1.5, Re = 60,000
N u a v g = 121.5389985 428.083536 ϕ ω = 2.5, Re = 30,000
N u a v g = 129.1209901 414.4750866 ϕ ω = 2.5, Re = 40,000
N u a v g = 137.6074661 405.7243489 ϕ ω = 2.5, Re = 50,000
N u a v g = 146.4162574 380.5897152 ϕ ω = 2.5, Re = 60,000
N u a v g = 151.9178705 607.143654 ϕ ω = 3.5, Re = 30,000
N u a v g = 157.2588622 561.667193 ϕ ω = 3.5, Re = 40,000
N u a v g = 164.1914757 542.241229 ϕ ω = 3.5, Re = 50,000
N u a v g = 171.8464556 538.982427 ϕ ω = 3.5, Re = 60,000
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alqarni, M.S.; Memon, A.A.; Anwaar, H.; Usman; Muhammad, T. The Forced Convection Analysis of Water Alumina Nanofluid Flow through a 3D Annulus with Rotating Cylinders via κε Turbulence Model. Energies 2022, 15, 6730. https://doi.org/10.3390/en15186730

AMA Style

Alqarni MS, Memon AA, Anwaar H, Usman, Muhammad T. The Forced Convection Analysis of Water Alumina Nanofluid Flow through a 3D Annulus with Rotating Cylinders via κε Turbulence Model. Energies. 2022; 15(18):6730. https://doi.org/10.3390/en15186730

Chicago/Turabian Style

Alqarni, Marei Saeed, Abid Ali Memon, Haris Anwaar, Usman, and Taseer Muhammad. 2022. "The Forced Convection Analysis of Water Alumina Nanofluid Flow through a 3D Annulus with Rotating Cylinders via κε Turbulence Model" Energies 15, no. 18: 6730. https://doi.org/10.3390/en15186730

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