Next Article in Journal
Temporal–Spatial Evolution of Kinetic and Thermal Energy Dissipation Rates in a Three-Dimensional Turbulent Rayleigh–Taylor Mixing Zone
Previous Article in Journal
On Leader-Following Consensus in Multi-Agent Systems with Discrete Updates at Random Times
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Constructal Design of Elliptical Cylinders with Heat Generating for Entropy Generation Minimization

1
College of Power Engineering, Naval University of Engineering, Wuhan 430033, China
2
Institute of Thermal Science and Power Engineering, Wuhan Institute of Technology, Wuhan 430205, China
3
School of Mechanical & Electrical Engineering, Wuhan Institute of Technology, Wuhan 430205, China
*
Authors to whom correspondence should be addressed.
Entropy 2020, 22(6), 651; https://doi.org/10.3390/e22060651
Submission received: 14 April 2020 / Revised: 5 June 2020 / Accepted: 8 June 2020 / Published: 12 June 2020

Abstract

:
A heat dissipation model of discrete elliptical cylinders with heat generation on a thermal conduction pedestal cooled by forced convection is established. Constructal design is conducted numerically by taking the distributions of thermal conductivity and heat generating intensity as design variables, the dimensionless entropy generation rate (DEGR) as performance indicator. The optimal designs for discrete elliptical cylinders with heat generating are obtained respectively, i.e., there are optimal distributions of heat generating intensity with its fixed total amount of heat sources, and there are optimal distributions of thermal conductivity with its fixed total amount of heat sources. These optimums for minimum DEGRs are different at different Reynolds numbers of airflow. The heat generating intensity can be decreased one by one appropriately in the fluid flow direction to achieve the best effect. When the Reynolds number of airflow is smaller, the thermal conductivity of heat source can be increased one by one appropriately in the fluid flow direction to achieve the best effect; when the Reynolds number of airflow is larger, the thermal conductivity of each heat source should be equalized to achieve the best effect. The results can give thermal design guidelines for the practical heat generating devices with different materials and heat generating intensities.

1. Introduction

Electronic manufacture technology has developed quickly, and as electronic devices and equipment have been continually and highly miniaturized and integrated, the power per unit volume of the devices has increased continually, which has made their heat dissipation problems increasingly prominent. Research aiming to optimize electronic device and heat sink designs to enhance their heat dissipation has attracted the interest of many scholars [1,2]. In recent years, extensive and in-depth investigations for heat transfer optimization in the light of constructal theory [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24] and entropy generation minimization [25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60] have flourished, which are beneficial for technologic development of electronics cooling.
Bejan, who investigated the universal physical mechanism of various natural dendritic structures [4,5,6,9,10,11,12,13,14,15,16,17,18,19,20,21,22,24], found the constructal law and proposed constructal theory from analyzing the formation and evolution of street network configurations in major cities around the world as an example in 1996 [3]. Constructal theory was firstly applied to design an optimal network of high conductivity material for electronics cooling, and the constructal law was found correspondingly [4]. The constructal law [4] stated that, “For a finite-size flow system to persist in time (to live), its configuration must change in time such that it provides easier and easier access to its currents.” For everything in nature, society and engineering, the configuration evolution in time follows this deterministic physics law [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24]. Constructal theory not only accounts for and predicts the flow pattern/configuration evolution in nature and society, but also provides theoretical guidelines for the flow pattern/configuration design in various disciplines. The statement of constructal law can be further simplified as “the structures of matters come from their optimal performances [6].” The purpose of constructal design is to seek the optimal configuration with the best distribution of imperfections, so it was also named as a new philosophy of geometry.
For heat transfer optimization, the studies following constructal law can be classified as single-objective optimization, multi-objective optimization and multi-disciplinary optimization by optimization objectives; and these also can be classified as heat conduction optimization, convective heat and mass transfer optimization and various device and component optimization by study objects, and so on [4,5,6,9,10,11,12,13,14,15,16,17,18,19,20,21,22,24]. It has also been applied to the designs of cavities and fin heat sinks [24,61,62,63], channel heat sinks [64,65], heat exchangers [66,67,68,69], nonuniform heat generating units [70,71], evaporators [72,73,74], boiler superheater [75], tubular arrangements [76,77], dual-pressure turbine [78,79], etc., in recent years. For the constructal design in engineering, heat sources [80,81,82,83,84,85,86,87,88,89,90,91,92] is one of important and interesting research hotspots.
Da Silva and Gosselin [80] pointed out that increasing the cooling channel number doesn’t necessarily enhance the heat dissipation by investigating the constructal design of square components with heat generation in cooling channels. Gong et al. [81,82,83] investigated the constructal designs for two types of cylinders with heat generating which surrounded with a fin under the conditions of natural cooling, forced cooling and uniform heat transfer coefficient, respectively. Jassim and Muzychka [84] indicated that the non-uniform distribution of heat sources had better heat dissipation performance than that of uniform distribution. Tye-Gingra et al. [85] optimized the initial phase of heat flux density’s function and position of heat source by establishing a heat source model with sinusoidal variation over time of heat generating intensity. Shi and Dong [86] explored the constructal layout of discrete components with heat generating subjected to forced laminar flow. Fan et al. [87] calculated the heat dissipation performance with single and discrete cylindrical heat sources numerically and the results showed that the performances of multi-scale geometry and non-equal distribution of heat generating intensity were better than those of the single-scale geometry and equal distribution of heat generating intensity. Wang et al. [88] carried out the constructal design for several heat source models including discrete stationary and rotating cylinders, singe and discrete elliptical tables. Sarper et al. [89] investigated the constructal design for discrete multi-scale length heat sources in vertical ducts. In addition, Chen et al. [90], Aslan et al. [91] and Armaghani et al. [92] investigated the heat source layout designs.
Practical processes, such as fluid flow with finite pressure difference and heat transfer with finite temperature difference, are non-equilibrium irreversible processes from the thermodynamic point of view. Bejan [25,26] first derived the corresponding entropy generation rate (EGR) formula for fluid flow with finite pressure difference and heat transfer with finite temperature difference, and proposed the principle of entropy generation minimization (EGM). Many scholars [27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46] have performed extensive and in-depth researches on various processes [32,42], cycles [28,29,36,37], devices [30,31,35,40] and systems [44,45,46] based on EGM. Recently, various fluid flows such as Newtonian flow [93], carbon nanotube flow [94] and Darcy-Forchheimer nanofluid flow [95], microchannel heat sinks [96], heat exchangers [97,98,99], and so on, have been investigated with EGM. The conventional metrics, such as the heat transfer enhancement and pressure drop, evaluate convective heat transfer from the view point of the first law of thermodynamics, while the entropy generation analysis is from the view point of the second law of thermodynamics by uniformly characterizing the irreversibility of heat transfer and fluid friction. The purpose of entropy generation minimization is to seek the minimization of thermodynamic irreversibility which is characterized and quantified by the indicators such as entropy generation rate, entropy generation number, Bejan number, and so on [44,45,46].
From the references mentioned above, although the constructal design and entropy generation minimization have respective theoretical connotations, core positions and optimization purposes, one can see that there are internal physical relations between optimal configuration and entropy generation minimization for various heat and mass transfer processes and systems [6,9,10,13,14,18,19,22,24,31,34], and there is a very interesting research content in thermal design employing the method that combining constructal theory with entropy generation minimization for heat source optimization [87,88]. The heat sources mentioned in [80,81,82,83,84,85,86,87,88,89,90,91,92] were usually cylinders, square columns, elliptical tables, embedded plate structures, and so on. Based on the above combining method, this paper will establish a 3-D heat dissipation model of discrete elliptical cylinders with heat generating, and design heat generating intensity distribution and thermal conductivity distribution to reach the minimum total DEGR of system. This method herein can be adapted to the thermal design requirements of heat sources with different cross-section shapes in practical applications, and especially help to provide theoretical guidelines for the optimization of practical electronic components with different materials and heat generating intensities.

2. Heat Source Model and Numerical Method

2.1. Geometric Model

Figure 1 gives a geometric model of discrete elliptical cylinders with heat generating on a thermal conduction pedestal with forced convection in a rectangular channel. The length (L), width (W) and thickness (H) of thermal conduction pedestal is 60, 20 and 1 mm, respectively. Four equal-size discrete heat sources (numbered 1, 2, 3, and 4 from left to right) are evenly arranged along the central axis of pedestal, wherein the ellipse short axis (La) of the heat source bottom surface is 2.5 mm, the ellipse long axis (Lb) is 5 mm, and the height Hs of the heat source is 5 mm. The elliptical cylinders can only be arranged in line to meet some multi-disciplinary requirements for manufacture and performance in a practical engineering case.

2.2. Heat Transfer Model

The reference material of heat conductive pedestal is silicon (cb is 700 J·kg−1·K−1, ρb is 2329 kg·m3, and λb is 130 W·m−1·K−1), the reference material of heat source is copper (cs is 385 J·kg−1·K−1, ρb is 8960 kg·m3, λs is 400 W·m−1·K−1). The cooling medium through the rectangular channel is clean air, and the variable properties and viscosity dissipation of airflow are considered. The contact surfaces between airflow and channel wall, as well as that between airflow and heat source are all set with non-slip boundaries. The outer wall surfaces of channel and the bottom surface of pedestal are all set with adiabatic boundary conditions. The end faces at inlet and exit of the channel are open boundaries, and the ambient air temperature is set as Tin (293.15 K). The range of Prandtl number of air is 0.938–0.973. The forced airflow is a compressible steady-state turbulent flow, and the inlet airflow is perpendicular to the inlet end face.
The total EGR [25,26,44] of overall heat transfer process is generally clarified as the EGR of solid section and the EGR of fluid section, i.e.,:
S g , ϕ = S solid + S fluid
The EGR of solid section [25,26,44] is:
S solid = V 1 T 2 q T d V
The EGR of fluid section [25,26,44] is:
S fluid = V ( 1 T 2 q T + μ T Φ ) d V
where q (W·m−2) is the heat flux vector and Φ is a viscous dissipation function per unit volume. In the brackets of the right side of Equation (3), the first term is the EGR of heat transfer, and the second term is the EGR of fluid viscous dissipation.
According to Equations (1)–(3), the total EGR of system is:
S g , ϕ = V s 1 T 2 q s T d V V b 1 T 2 q b T d V + V f ( 1 T 2 q f T + μ T Φ ) d V
where q s , q b and q f are the heat flux vectors of heat sources, pedestal and fluid, respectively. Vs, Vb and Vf are the volumes of the heat sources, pedestal and fluid, respectively.
According to Equation (4), the dimensionless entropy generation rate (DEGR) of system can be defined as:
S ˜ g , ϕ = S g , ϕ T in P t
The energy equation for solid pedestal with steady-state heat conduction and constant properties is:
2 T = 0
The energy equation for heat sources with steady-state heat conduction and constant properties is:
2 T + q λ s = 0
Under forced convection condition, the governing equations of continuous fluid flow, momentum transfer, energy transfer and turbulence characteristics are as follows:
( ρ U ) = 0
ρ U U = [ p I + ( μ + μ T ) ( U + ( U ) T ) 2 3 ( μ + μ T ) ( U ) I 2 3 ρ k I ] + F
ρ c f U T + q = Q
ρ ( U ) k = [ ( μ + μ T σ k ) k ] + P k ρ ε
ρ ( U ) ε = [ ( μ + μ T σ ε ) ε ] + C ε 1 ε k P k C ε 2 ρ ε 2 k
μ T = ρ C μ k 2 ε
P k = μ T [ U : ( U + ( U ) T ) 2 3 ( U ) 2 ] 2 3 ρ k U
where the parameters with pulsation are all time-averaged. The empirical constants Cε1, Cε2, Cμ, σk and σε are 1.44, 1.92, 0.09 and 1.3, respectively.
The performance indicator for optimization is the total DEGR of system. The smaller the total DEGR of system, the better the thermodynamic performance of system.

2.3. Numerical Method

The governing equations and boundary conditions of the heat dissipation model are solved by the finite element calculation method [100]. The tetrahedral meshing is performed in solid and fluid regions of the model, respectively. For reducing calculation deviations, the grid independence is tested. In the test case, the heat generating intensities (q” = 1.52 × 107 W·m−3) and the thermal conductivities (λs = 200 W·m1·K−1) of four heat sources are equal, and the inlet Reynolds number is set as 5000. There are three types of meshing with different grid numbers for computations, which are 12047, 52054 and 71006. The total DEGRs of system are 0.0170770, 0.0170654 and 0.0170732, and the relative errors are 0.068% and 0.046%, respectively. In order to balance the computation accuracy and efficiency, this paper uses the meshing criterion corresponding to 52054 grids for the following calculations. The general default convergence criteria for the continuity, momentum as well as energy equations are employed [100].
Reynolds Averaged Navier-Stokes (RANS) method is used to perform high-fidelity turbulence simulation calculations by Comsol Multiphysics. To further assess the accuracy of the computational model in this paper, 3-D models of cylindrical heat sources surrounded by fins with different dimensionless height H ˜ and ratio b (the center-to-center distance of the fin and heat source to the radius of fin) in [82] are built for comparative calculations. The numerical results about dimensionless hot spot temperature T ˜ are listed in Table 1.
For the same H ˜ and b, the maximum difference of T ˜ is only 1.27%, which means that the results herein agree well with the calculation results in [82]. The model of discrete elliptical cylinders with heat generating in this paper is established by the same method, so the effectiveness of the simulation method is verified.

3. Results and Analyses

Discrete electronic components with fixed positions always have various heat generating intensities and are made of different materials. Under forced convection conditions, the thermal conductivity distribution and the heat generating intensity distribution are chosen as design variables with the fixed total thermal conductivity and the fixed total heat generating intensity of heat sources, respectively. The constructal design of discrete elliptical cylinders with heat generating is conducted by taking the total dimensionless entropy generation rate minimization as performance indicator.

3.1. Effects of Heat Generating Intensity on Heat Dissipation Performance

The heat generating intensities from top to bottom are set as q 1 , q 2 , q 3 , and q 4 , respectively. The distribution is expressed as:
q 2 q 1 = q 3 q 2 = q 4 q 3 = Δ q
where Δ q   is the difference of heat generating intensity.
The total heat generating intensity of heat sources is fixed, the total heating rate of the heat sources is Pt (48 W), so the heat generating intensity of each heat source is:
q i = 15.2 × 10 6 + i 2.5 Δ q
The influence of Δ q on S ˜ g , ϕ of the system is shown in Figure 2 with the fixed equal thermal conductivity (400 W·m1·K−1) for each heat source. The results show that the higher the Reynolds number (Re), the smaller the system temperature gradient, which makes the total DEGR ( S ˜ g , ϕ ) reduce with the increase of airflow Re for specified heat generating intensity distribution. At the same Re, as the Δ q increases, the S ˜ g , ϕ decreases firstly and then increases, the optimal difference ( Δ q opt ) of heat generating intensity makes S ˜ g , ϕ minimum. S ˜ g , ϕ at Δ q = 10 × 10 6 W·m1·K−1 is larger than that at Δ q = 1010 6 W·m−3, which means that the heat generating intensity is preferably distributed from large to small in the airflow direction, the device with higher heat generating intensity should be arranged near the channel inlet. Furthermore, numerical results listed in Table 2 show that optimal differences ( Δ q opt   s ) for the same optimization objective are different under different fluid flow conditions. From Table 2, all Δ q opt   s are a little bit less than 0. There is a very thin boundary layer improves the heat transfer at the upstream body and thus reduces temperature gradients. The three downstream bodies are affected by the weakening of upstream body. That is the upstream body could tolerate a higher heat flow rate than its downstream fellows. When Δ q is close to the Δ q opt   vaue, the cooling capacity of air on the part with high heat load is also strong, and it is relatively easy to avoid insufficient cooling of the part with high heat load. When Δ q is less than Δ q opt   ,   S ˜ g , ϕ increases when Δ q decreases. This is because the effect of heat generating intensity distribution on temperature gradient is more significant compared with that of boundary layer.
The influence of Δ q on the average Nusselt number ( N u ¯ ) of the system is shown in Figure 3. It can be seen from Figure 3 that N u ¯ increases with the increases of the Re and Δ q .
The color maps of the temperature gradient distribution on upper surfaces of heat sources and thermal conduction pedestal at Re = 5000 are shown in Figure 4. From the figure, as the difference of heat generating intensity increases, the heat generating intensity in the flow direction which is from large to small changes to the distribution which is small to large, the temperature gradient distribution gradually changes from decreasing in flow direction to distributing uniformly around the heat sources. From Figure 4a, although both the heat generating intensity and the temperature gradient decrease in flow direction, the heat generating intensity at channel inlet is relatively larger which makes the heat source near inlet is not well cooled with Δ q = −6 × 106 W·m−3 the total EGR of system is higher. From Figure 4b, the temperature gradients around the heat sources gradually decrease in flow direction with Δ q = −1 × 106 W·m−3. In this case, the cooling of each heat source is optimally balanced, so that the total DEGR of system is the lowest. From Figure 4c, the temperature gradients around the four heat sources are substantially equal, but the heat generating intensity gradually increases one by one in flow direction with Δ q = 4 × 106 W·m−3, which leads to poor cooling of the heat source near channel outlet, so that the total DEGR of system increases. Figure 5 shows the pressure distributions in channel for Δ q = 0 W·m−3. From Figure 5, there is little change in the situation of pressure distribution with changes of airflow Reynolds numbers, but the pressure drop in channel rises with the increases of airflow Reynolds numbers.

3.2. Effects of Thermal Conductivity of Heat Source on Heat Dissipation Performance

Figure 6 shows the influences of thermal conductivity (λs) of each heat source on total DEGR (   S ˜ g , ϕ ) of the system under forced convection with fixed equal heat generating intensity of each heat source. As can be seen from the figure, the   S ˜ g , ϕ decreases directly and then gradually becomes gentle with the increase of λs at the same Reynolds number of airflow. When λs < 150, the decreasing magnitude is larger, when λs rises to a certain value, the temperature gradients of heat sources are relatively uniform because the thermal resistances of heat sources are smaller. Then, the effect becomes little that improving the heat dissipation performance and reducing the   S ˜ g , ϕ of the system by simultaneously improving the heat conductivities of heat sources. The improvement effect is more obvious when the thermal conductivity of the heat source is lower.
To study the change of DEGR of system by changing the difference of thermal conductivity with the fixed equal heat generating intensity of each heat source, the thermal conductivities of heat sources are set as λs,1, λs,2, λs,3 and λs,4, respectively, and distributed with equal difference along flow direction:
λ s , 2 λ s , 1 = λ s , 3 λ s , 2 = λ s , 4 λ s , 3 = Δ λ s
where Δλs is the difference of thermal conductivity between adjacent heat sources.
Considering the sum of thermal conductivities and the constraints for practical thermal conductivities of materials, it is assumed that the thermal conductivities of heat sources satisfy 30 ≤ λs,I ≤ 450 (i = 1, 2, 3, 4).
λ s ,   j = 240 + i 2.5 Δ λ s   ( 140 Δ λ s ,   j 140 )
Figure 7 shows the influences of thermal conductivity difference (Δλs) on total DEGR (   S ˜ g , ϕ ) of system when the heat generating intensities are constants. The results show that the total   S ˜ g , ϕ of the system decreases first and then rises with the rise of Δλs and   S ˜ g , ϕ at Δλs = −140 W·m1·K−1 is larger than that at Δλs = 140 W·m1·K−1, which means that the thermal conductivity is preferably distributed from small to large in the airflow direction, the device with higher thermal conductivity should be arranged near the channel outlet. Further numerical results listed in Table 3 show that (Δλs)opt for the same optimization objective are different under different fluid flow conditions. From Table 3, when the Re of airflow are 2500 and 3000, the optimal difference of thermal conductivity ((Δλs)opt > 0) makes the total   S ˜ g , ϕ reach minimums; when Re of airflow is 4000, 5000 and 6000, the optimal difference of thermal conductivity ((Δλs)opt = 0 W·m1·K−1) makes the total   S ˜ g , ϕ reach minimums.
This occurs because the discrete elliptical cylinders with heat generating reinforce the flow resistance in the channel but cause secondary vortexes at the same time. When the Re of airflow is 3000 and Δλs is larger than 0, the thermal conductivity of heat source increases one by one appropriately to reduce the thermal resistance for thermal conduction in corresponding regions along flow direction, and the temperature gradients decrease obviously resulted from the secondary vortexes which enhance heat transfer locally, so the cooling requirements of all parts can be met better and appropriately, and the total DEGR of system decreases. With the Re of airflow increasing, the forced convective condition changes, so the heat transfer including heat conduction and convective heat transfer should be adapted coordinately to benefit heat flow for EGM and the total   S ˜ g , ϕ of system are the smallest when the thermal conductivities of heat sources are equal.
From Figure 2, Figure 6 and Figure 7, the impacts of distributions about heat generating intensity and thermal conductivity is 6% and 1% roughly. The higher the Reynolds number, the smaller the maximum temperature (Tmax) of electrical device. At the same Re, as the Δ q and Δλs increase, the Tmax decreases firstly and then increases, the Δq”opt and the (Δλs)opt make Tmax minimum, and the Δq”opt and (Δλs)opt are different, respectively, for different performance indicators.

4. Conclusions

In this work, a 3-D heat dissipation model of discrete elliptical cylinders with heat generation on a thermal conduction pedestal with forced convection is established. The effects of heat generation intensity distribution and thermal conductivity distribution on the total DEGR of system are investigated, respectively. The results bring to light that when the total heat generating intensity of heat sources is specified, there is an optimal distribution of the heat generation intensity that makes the total DEGR smallest. The minimums of total DEGR corresponding to optimal differences of heat generating intensity are different at different airflow velocities. The heat generating intensity can be decreased one by one appropriately in flow direction to achieve the best effect.
When the thermal conductivity of each heat source is equal, the total DEGR of system decreases with the simultaneous rise of the thermal conductivity of each heat source and then gradually becomes equilibrium. The thermal conductivity of each heat source can be appropriately increased to enhance heat dissipation.
When the sum of thermal conductivities of heat sources is fixed, there are optimal distributions of thermal conductivity that minimize the DEGR of system. When the Reynolds number of airflow is smaller, the thermal conductivity of heat source can be increased one by one appropriately in flow direction to achieve the best effect; while the Reynolds number of airflow is larger, the thermal conductivity of each heat source should be equalized to achieve the best effect.

Author Contributions

R.W., Z.X., Y.Y., L.C. prepared the manuscript together. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China grant number 51579244.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grant Nos. 51579244, 51979278 and 51506220). The authors wish to thank the reviewers for their careful, unbiased and constructive suggestions, which led to this revised manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

bRatio
Cε1, Cε2, CμEmpirical constants
cConstant pressure specific heat capacity, J·kg−1·K−1
FVolume force vector, N
HThickness of pedestal, m
HsHeight of heat source, m
H ˜ Dimensionless height
IUnit matrix
kTurbulent kinetic energy, J
LLength of pedestal, m
LaShort axis of ellipse, m
LbLong axis of ellipse, m
PkGeneration term of turbulent kinetic energy
PtThe total heating power of heat sources
QHeat source item including viscous dissipation and pressure work, W·m−3
qHeat flux vector, W·m−2
qHeat generating intensity, W·m−3
S g , ϕ Entropy generation rate, W·K−1
  S ˜ g , ϕ Dimensionless entropy generation rate
TMatrix transpose operator symbol
TTemperature, K
T ˜ Dimensionless hot spot temperature
UVelocity vector, m/s
WWidth of pedestal, m

Greek Symbols

εTurbulent dissipation rate, %
λThermal conductivity, W·m−1·K−1
μDynamic viscosity coefficient, Pa·s
μTTurbulent viscosity coefficient, Pa·s
ρDensity, kg·m−3
σk, σεEmpirical constants

Subscripts

bHeat conductive pedestal
fFluid
iNumber of heat source
inInlet
optOptimum
sHeat source

Abbreviations

EGMEntropy generation minimization
EGREntropy generation rate
DEGRdimensionless entropy generation rate

References

  1. Yousef, A.; Mohammad, Z.T.; Mohammad, M.H.; Nima, G. Effect of a micro heat sink geometric design on thermo-hydraulic performance: A review. Appl. Therm. Eng. 2020, 170, 114974. [Google Scholar]
  2. Khattak, Z.; Ali, H.M. Air cooled heat sink geometries subjected to forced flow: A critical review. Int. J. Heat Mass Transf. 2019, 130, 141–161. [Google Scholar] [CrossRef]
  3. Bejan, A. Street network theory of organization in nature. J. Adv. Transp. 1996, 30, 85–107. [Google Scholar] [CrossRef]
  4. Bejan, A. Constructal-theory network of conducting paths for cooling a heat generating volume. Int. J. Heat Mass Transf. 1997, 40, 799–816. [Google Scholar] [CrossRef]
  5. Reis, A. Constructal theory: From engineering to physics, and how flow systems develop shape and structure. Trans. ASME Appl. Mech. Rev. 2006, 59, 269–282. [Google Scholar] [CrossRef]
  6. Chen, L.G. Progress in study on constructal theory and its applications. Sci. China Tech. Sci. 2012, 55, 802–820. [Google Scholar] [CrossRef]
  7. Bejan, A. Constructal thermodynamics. Int. J. Heat Tech. 2016, 34, S1–S8. [Google Scholar] [CrossRef] [Green Version]
  8. Bejan, A.; Errera, M.R. Complexity, organization, evolution, and constructal law. J. Appl. Phys. 2016, 119, 074901. [Google Scholar] [CrossRef]
  9. Chen, L.G.; Feng, H.J.; Xie, Z.H. Generalized thermodynamic optimization for iron and steel production processes: A theoretical exploration and application cases. Entropy 2016, 18, 353. [Google Scholar] [CrossRef] [Green Version]
  10. Feng, H.J.; Chen, L.G.; Xie, Z.H. Multi-disciplinary, multi-objective and multi-scale constructal optimizations for heat and mass transfer processes performed in Naval University of Engineering, a review. Int. J. Heat Mass Transf. 2017, 115, 86–98. [Google Scholar] [CrossRef]
  11. Lorente, S. The constructal law: From microscale to urban-scale design. Annu. Rev. Heat Transf. 2017, 19, 335–368. [Google Scholar] [CrossRef]
  12. Chen, L.G.; Xiao, Q.H.; Feng, H.J. Constructal optimizations for heat and mass transfers based on the entransy dissipation extremum principle, performed at the Naval University of Engineering: A review. Entropy 2018, 20, 74. [Google Scholar] [CrossRef] [Green Version]
  13. Bejan, A. Constructal law, twenty years after. Proc. Rom. Acad. Ser. A Math. Phys. Tech. Sci. Inform. Sci. 2018, 18, 309–311. [Google Scholar]
  14. You, J.; Feng, H.J.; Chen, L.G.; Xie, Z.H. Constructal optimization for cooling a nonuniform heat generating disc-shaped area by conduction. Entropy 2018, 20, 685. [Google Scholar] [CrossRef] [Green Version]
  15. Lorente, S.; Bejan, A. Current trends in constructal law and evolutionary design. Heat Transf. Asian Res. 2019, 48, 357–389. [Google Scholar] [CrossRef]
  16. Bejan, A.; Gunes, U.; Sahin, B. The evolution of air and maritime transport. Appl. Phys. Rev. 2019, 6, 021319. [Google Scholar] [CrossRef]
  17. Bejan, A. The evolving design of our life. Interdiscip. J. Lands. Archit. 2019, 9, 8–13. [Google Scholar]
  18. Chen, L.G.; Feng, H.J.; Xie, Z.H.; Sun, F.R. Progress of constructal theory in China over the past decade. Int. J. Heat Mass Transf. 2019, 130, 393–419. [Google Scholar] [CrossRef]
  19. Zhang, F.Y.; Feng, H.J.; Chen, L.G.; You, J.; Xie, Z.J. Constructal design of an arrow-shaped high thermal conductivity channel in a square heat generation body. Entropy 2020, 22, 475. [Google Scholar] [CrossRef]
  20. Bejan, A.; Lorente, S. Design with Constructal Theory; Wiley: New Jersey, NJ, USA, 2008. [Google Scholar]
  21. Rocha, L.A.O.; Lorente, S.; Bejan, A. Constructal Law and the Unifying Principle of Design; Springer: Berlin, Germany, 2013. [Google Scholar]
  22. Chen, L.G.; Feng, H.J. Multi-objective Constructal Optimization for Flow and Heat and Mass Transfer Processes; Science Press: Beijing, China, 2016. (In Chinese) [Google Scholar]
  23. Bejan, A. Freedom and Evolution; Springer Science and Business Media LLC: Berlin, Germany, 2020. [Google Scholar]
  24. Chen, L.G.; Yang, A.B.; Feng, H.J.; Xia, S.J. Constructal designs for eight types of heat sinks. Sci. China Techno. Sci. 2020, 63, 879–911. [Google Scholar] [CrossRef]
  25. Bejan, A. General criterion for rating heat exchanger performance. Int. J. Heat Mass Transf. 1978, 21, 655–658. [Google Scholar] [CrossRef]
  26. Bejan, A. Two thermodynamic optima in the design and operation of thermal energy storage systems. Trans. ASME, J. Heat Transf. 1978, 100, 708–712. [Google Scholar] [CrossRef]
  27. Torabi, M.; Zhang, K.; Karimi, N.; Peterson, G.P. Entropy generation in thermal systems with solid structures—A concise review. Int. J. Heat Mass Transf. 2016, 97, 917–931. [Google Scholar] [CrossRef] [Green Version]
  28. Ge, Y.L.; Chen, L.G.; Sun, F.R. Progress in finite time thermodynamic studies for internal combustion engine cycles. Entropy 2016, 18, 139. [Google Scholar] [CrossRef] [Green Version]
  29. Wu, Z.X.; Chen, L.G.; Feng, H.J. Thermodynamic optimization for an endoreversible Dual-Miller cycle (DMC) with finite speed of piston. Entropy 2018, 20, 165. [Google Scholar] [CrossRef] [Green Version]
  30. Roach, T.N.F.; Salamon, P.; Nulton, J.; Andresen, B.; Felts, B.; Haas, A.; Calhoun, S.; Robinett, N.; Rohwer, F. Application of finite-time and control thermodynamics to biological processes at multiple scales. J. Non-Equilib. Thermodyn. 2018, 43, 193–210. [Google Scholar] [CrossRef] [Green Version]
  31. Feng, Y.L.; Chen, L.G.; Meng, F.K.; Sun, F.R. Influences of Thomson effect on performance of thermoelectric generator-driven thermoelectric heat pump combined device. Entropy 2018, 20, 29. [Google Scholar] [CrossRef] [Green Version]
  32. Pourkiaei, S.M.; Ahmadi, M.H.; Sadeghzadeh, M.; Moosavi, S.; Pourfayaz, F.; Chen, L.G.; Yazdi, M.A.; Kumar, R. Thermoelectric cooler and thermoelectric generator devices: A review of present and potential applications, modeling and materials. Energy 2019, 186, 115849. [Google Scholar] [CrossRef]
  33. Li, W.S.; Zhao, Y.X.; Wang, Q.; Zhou, J. Twenty years of entropy research: A bibliometric overview. Entropy 2019, 21, 694. [Google Scholar] [CrossRef] [Green Version]
  34. Zhu, F.L.; Chen, L.G.; Wang, W.H. Thermodynamic analysis of an irreversible Maisotsenko reciprocating Brayton cycle. Entropy 2018, 20, 167. [Google Scholar] [CrossRef] [Green Version]
  35. Zhang, L.; Chen, L.G.; Xia, S.J.; Wang, C.; Sun, F.R. Entropy generation minimization for reverse water gas shift (RWGS) reactor. Entropy 2018, 20, 415. [Google Scholar] [CrossRef] [Green Version]
  36. Chen, L.G.; Zhang, L.; Xia, S.J.; Sun, F.R. Entropy generation minimization for hydrogenation of CO2 to light olefins. Energy 2018, 147, 187–196. [Google Scholar] [CrossRef]
  37. Li, P.L.; Chen, L.G.; Xia, S.J.; Zhang, L. Entropy generation rate minimization for in methanol synthesis via CO2 hydrogenation reactor. Entropy 2019, 21, 174. [Google Scholar] [CrossRef] [Green Version]
  38. Chen, L.G.; Xia, S.J. Progresses in generalized thermodynamic dynamic-optimization of irreversible processes. Sci. Sin. Technol. 2019, 49, 981–1022. (In Chinese) [Google Scholar] [CrossRef] [Green Version]
  39. Chen, L.G.; Xia, S.J.; Feng, H.J. Progress in generalized thermodynamic dynamic-optimization of irreversible cycles. Sci. Sin. Technol. 2019, 49, 1223–1267. (In Chinese) [Google Scholar] [CrossRef] [Green Version]
  40. Zhang, L.; Xia, S.J.; Chen, L.G.; Ge, Y.L.; Wang, C.; Feng, H.J. Entropy generation rate minimization for hydrocarbon synthesis reactor from carbon dioxide and hydrogen. Int. J. Heat Mass Transf. 2019, 137, 1112–1123. [Google Scholar] [CrossRef]
  41. Chen, L.G.; Ge, Y.L.; Liu, C.; Feng, H.J.; Lorenzini, G. Performance of universal reciprocating heat-engine cycle with variable specific heats ratio of working fluid. Entropy 2020, 22, 397. [Google Scholar] [CrossRef] [Green Version]
  42. Li, P.L.; Chen, L.G.; Xia, S.J.; Zhang, L.; Kong, R.; Ge, Y.L.; Feng, H.J. Entropy generation rate minimization for steam methane reforming reactor heated by molten salt. Energy Rep. 2020, 6, 685–697. [Google Scholar] [CrossRef]
  43. Meng, Z.W.; Chen, L.G.; Wu, F. Optimal power and efficiency of multi-stage endoreversible quantum Carnot heat engine with harmonic oscillators at the classical limit. Entropy 2020, 22, 397. [Google Scholar] [CrossRef] [Green Version]
  44. Bejan, A. Entropy Generation through Heat and Fluid Flow; Wiley: New York, NY, USA, 1982. [Google Scholar]
  45. Sieniutycz, S.; Szwast, Z. Optimizing Thermal, Chemical and Environmental Systems; Elsevier: Amsterdam, The Netherlands, 2018. [Google Scholar]
  46. Haseli, Y. Entropy Analysis in Thermal Engineering Systems; Elsevier: London, UK, 2019. [Google Scholar]
  47. Chen, L.G.; Wu, C.; Sun, F.R. Finite time thermodynamic optimization or entropy generation minimization of energy systems. J. Non-Equilib. Thermodyn. 1999, 24, 327–359. [Google Scholar] [CrossRef]
  48. Chen, L.G.; Sun, F.R. Advances in Finite Time Thermodynamics: Analysis and Optimization; Nova Science Publishers: New York, NY, USA, 2004. [Google Scholar]
  49. Kingston, D.; Razzitte, A.C. Entropy generation minimization in Dimethyl Ether synthesis: A case study. J. Non-Equilib. Thermodyn. 2018, 43, 111–120. [Google Scholar] [CrossRef]
  50. Paéz-Hernández, R.T.; Chimal-Eguía, J.C.; Sánchez-Salas, N.; Ladino-Luna, D. General properties for an Agrawal thermal engine. J. Non-Equilib. Thermodyn. 2018, 43, 131–140. [Google Scholar] [CrossRef]
  51. Schwalbe, K.; Hoffmann, K.H. Novikov engine with fluctuating heat bath temperature. J. Non-Equilib. Thermodyn. 2018, 43, 141–150. [Google Scholar] [CrossRef]
  52. Feidt, M.; Costea, M. From finite time to finite physical dimensions thermodynamics: The Carnot engine and Onsager’s relations revisited. J. Non-Equilib. Thermodyn. 2018, 43, 151–162. [Google Scholar] [CrossRef]
  53. Schwalbe, K.; Hoffmann, K.H. Optimal control of an endoreversible solar power plant. J. Non-Equilib. Thermodyn. 2018, 43, 255–272. [Google Scholar] [CrossRef]
  54. Zaeva, M.A.; Tsirlin, A.M.; Didina, O.V. Finite time thermodynamics: Realizability domain of heat to work converters. J. Non-Equilib. Thermodyn. 2019, 44, 181–191. [Google Scholar] [CrossRef]
  55. Raman, R.; Kumar, N. Performance analysis of Diesel cycle under efficient power density condition with variable specific heat of working fluid. J. Non-Equilib. Thermodyn. 2019, 44, 405–416. [Google Scholar] [CrossRef]
  56. Schwalbe, K.; Hoffmann, K.H. Stochastic Novikov engine with Fourier heat transport. J. Non-Equilib. Thermodyn. 2019, 44, 417–424. [Google Scholar] [CrossRef]
  57. Marsik, F.; Weigand, B.; Tomas, M.; Tucek, O.; Novotny, P. On the efficiency of electrochemical devices from the perspective of endoreversible thermodynamics. J. Non-Equilib. Thermodyn. 2019, 44, 425–438. [Google Scholar] [CrossRef]
  58. Chen, L.G.; Li, J. Thermodynamic Optimization Theory for Two-Heat-Reservoir Cycles; Science Press: Beijing, China, 2020. [Google Scholar]
  59. Boikov, S.Y.; Andresen, B.; Akhremenkov, A.A.; Tsirlin, A.M. Evaluation of irreversibility and optimal organization of an integrated multi-stream heat exchange system. J. Non-Equilib. Thermodyn. 2020, 45, 155–171. [Google Scholar] [CrossRef]
  60. Ding, Z.M.; Ge, Y.L.; Chen, L.G.; Feng, H.J.; Xia, S.J. Optimal performance regions of Feynman’s ratchet engine with different optimization criteria. J. Non-Equilib. Thermodyn. 2020, 45, 191–207. [Google Scholar] [CrossRef]
  61. Yang, A.B.; Chen, L.G.; Xie, Z.H.; Feng, H.J.; Sun, F.R. Constructal operation cost minimization for in-line cylindrical pin-fin heat sinks. Int. J. Heat Mass Transf. 2019, 129, 562–568. [Google Scholar] [CrossRef]
  62. Mustafa, A.W. Constructal design of multi-scale diamond-shaped pin fins cooled by mixed convection. Int. J. Therm. Sci. 2019, 145, 106018. [Google Scholar] [CrossRef]
  63. Hazarika, S.A.; Deshmukhya, T.; Bhanja, D.; Nath, S. A novel optimum constructal fork-shaped fin array design for simultaneous heat and mass transfer application in a space-constrained situation. Int. J. Therm. Sci. 2020, 150, 106225. [Google Scholar] [CrossRef]
  64. Shi, H.N.; Xie, Z.H.; Chen, L.G.; Sun, F.R. Constructal optimization for line-to-line vascular based on entropy generation minimization principle. Int. J. Heat Mass Transf. 2018, 126, 848–854. [Google Scholar] [CrossRef]
  65. Lu, Z.H.; Zhang, K.; Liu, J.X.; Li, F. Effect of branching level on the performance of constructal theory based Y-shaped liquid cooling heat sink. Appl. Therm. Eng. 2020, 168, 114824. [Google Scholar] [CrossRef]
  66. Feng, H.J.; Chen, L.G.; Xia, S.J. Constructal design for disc-shaped heat exchanger with maximum thermal efficiency. Int. J. Heat Mass Transf. 2019, 130, 740–746. [Google Scholar] [CrossRef]
  67. Feng, H.J.; Chen, L.G.; Wu, Z.X.; Xie, Z.J. Constructal design of a shell-and-tube heat exchanger for organic fluid evaporation process. Int. J. Heat Mass Transf. 2019, 131, 750–756. [Google Scholar] [CrossRef]
  68. Wu, Z.X.; Feng, H.J.; Chen, L.G.; Ge, Y.L. Performance optimization of a condenser in ocean thermal energy conversion (OTEC) system based on constructal theory and multi-objective genetic algorithm. Entropy 2020, 22, 641. [Google Scholar] [CrossRef]
  69. Ariyo, D.O.; Bello-Ochende, T. Constructal design of subcooled microchannel heat exchangers. Int. J. Heat Mass Transf. 2020, 146, 118835. [Google Scholar] [CrossRef]
  70. Chen, L.G.; You, J.; Feng, H.J.; Xie, Z.H. Constructal optimization for “disc-point” heat conduction with nonuniform heat generating. Int. J. Heat Mass Transf. 2019, 134, 1191–1198. [Google Scholar] [CrossRef]
  71. You, J.; Feng, H.J.; Chen, L.G.; Xie, Z.H. Constructal design of nonuniform heat generating area based on triangular elements: A case of entropy generation minimization. Int. J. Therm. Sci. 2019, 139, 403–412. [Google Scholar] [CrossRef]
  72. Cai, C.G.; Feng, H.J.; Chen, L.G.; Wu, Z.X.; Xie, Z.J. Constructal design of a shell-and-tube evaporator with ammonia-water working fluid. Int. J. Heat Mass Transf. 2019, 135, 541–547. [Google Scholar] [CrossRef]
  73. Xie, Z.J.; Feng, H.J.; Chen, L.G.; Wu, Z.X. Constructal design for supercharged boiler evaporator. Int. J. Heat Mass Transf. 2019, 138, 571–579. [Google Scholar] [CrossRef]
  74. Wu, Z.X.; Feng, H.J.; Chen, L.G.; Xie, Z.J.; Cai, C.G. Pumping power minimization of an evaporator in ocean thermal energy conversion system based on constructal theory. Energy 2019, 181, 974–984. [Google Scholar] [CrossRef]
  75. Feng, H.J.; Xie, Z.J.; Chen, L.G.; Wu, Z.X.; Xia, S.J. Constructal design for supercharged boiler superheater. Energy 2020, 191, 116484. [Google Scholar] [CrossRef]
  76. You, J.; Feng, H.J.; Chen, L.G.; Xie, Z.H.; Xia, S.J. Constructal design and experimental validation of a non-uniform heat generating body with rectangular cross-section and parallel circular cooling channels. Int. J. Heat Mass Transf. 2020, 148, 119028. [Google Scholar] [CrossRef]
  77. Pedroti, V.A.; de Escobar, C.C.; dos Santos, E.D.; Souza, J.A. Thermal analysis of tubular arrangements submitted to external flow using constructal theory. Int. Comm. Heat Mass Transf. 2020, 111, 104458. [Google Scholar] [CrossRef]
  78. Wu, Z.X.; Feng, H.J.; Chen, L.G.; Xie, Z.J.; Cai, C.G.; Xia, S.J. Optimal design of dual-pressure turbine in OTEC system based on constructal theory. Energy Convers. Manage. 2019, 201, 112179. [Google Scholar] [CrossRef]
  79. Wu, Z.X.; Feng, H.J.; Chen, L.G.; Tang, W.; Shi, J.Z.; Ge, Y.L. Constructal thermodynamic optimization for ocean thermal energy conversion system with dual-pressure organic Rankine cycle. Energy Convers. Manage. 2020, 210, 112727. [Google Scholar] [CrossRef]
  80. Da Silva, A.K.; Gosselin, L. Constructal peripheral cooling of a rectangular heat-generating area. Trans. ASME, J. Electron. Pack. 2006, 128, 432–440. [Google Scholar] [CrossRef]
  81. Gong, S.W.; Chen, L.G.; Xie, Z.H.; Feng, H.J.; Sun, F.R. Constructal optimization of cylindrical heat sources with forced convection based on entransy dissipation minimization. Sci. China Techhol. Sci. 2016, 59, 631–639. [Google Scholar] [CrossRef]
  82. Gong, S.W.; Chen, L.G.; Feng, H.J.; Xie, Z.H.; Sun, F.R. Constructal optimization of cylindrical heat sources surrounded with a fin based on minimization of hot spot temperature. Int. Comm. Heat Mass Transf. 2015, 68, 1–7. [Google Scholar] [CrossRef]
  83. Gong, S.W. Optimal Design of Cylindrical Heat Sources Based on Constructal Theory. Master′s Thesis, Naval University of Engineering, Wuhan, China, 2015. (In Chinese). [Google Scholar]
  84. Jassim, E.; Muzychka, Y.S. Optimal distribution of heat sources in convergent channels cooled by laminar forced convection. Trans. ASME, J. Heat Transf. 2010, 132, 011701. [Google Scholar] [CrossRef]
  85. Tye-Gingras, M.; Gosselin, L.; da Silva, A.K. Synchronizing time-varying discrete heating elements in forced convection for minimal thermal resistance. Int. J. Heat Mass Transf. 2010, 53, 4917–4926. [Google Scholar] [CrossRef]
  86. Shi, Z.Y.; Dong, T. A synthetic layout optimization of discrete heat sources flush mounted on a laminar flow cooled flat plate based on the constructal law. Energy Convers. Manage. 2015, 106, 300–307. [Google Scholar] [CrossRef]
  87. Fan, X.D. Constructal Study on Multi-scale Heat Source and Heat Sink. Master′s Thesis, Naval University of Engineering, Wuhan, China, 2015. (In Chinese). [Google Scholar]
  88. Wang, G. Heat Dissipation Analysis and Constructal Design of Columnar Heat Generation Components. Master′s Thesis, Naval University of Engineering, Wuhan, China, 2015. (In Chinese). [Google Scholar]
  89. Sarper, B.; Saglam, M.; Aydin, O. Constructal placement of discrete heat sources with different lengths in vertical ducts under natural and mixed convection. Trans. ASME, J. Heat Transf. 2018, 140, 121401. [Google Scholar] [CrossRef]
  90. Chen, K.; Xing, J.W.; Wang, S.F.; Song, M.X. Heat source layout optimization in two- dimensional heat conduction using simulated annealing method. Int. J. Heat Mass Transf. 2017, 108 Pt A, 210–219. [Google Scholar] [CrossRef]
  91. Aslan, Y.; Puskely, J.; Yarovoy, A. Heat source layout optimization for two-dimensional heat conduction using iterative reweighted L1-norm convex minimization. Int. J. Heat Mass Transf. 2018, 122, 432–441. [Google Scholar] [CrossRef] [Green Version]
  92. Armaghani, T.; Rashad, A.M.; Vahidifar, O.; Mishra, S.R.; Chamkha, A.J. Effects of discrete heat source location on heat transfer and entropy generation of nanofluid in an open inclined L-shaped cavity. Int. J. Numer. Methods Heat Fluid Flow 2019, 29, 1363–1377. [Google Scholar] [CrossRef]
  93. Khan, I.; Khan, W.A.; Qasim, M.; Afridi, I.; Alharbi, S.O. Thermodynamic analysis of entropy generation minimization in thermally dissipating flow over a thin needle moving in a parallel free stream of two Newtonian fluids. Entropy 2019, 21, 74. [Google Scholar] [CrossRef] [Green Version]
  94. Alsaadi, F.E.; Muhammad, K.; Hayat, T.; Alsaedi, A. Numerical study of melting effect with entropy generation minimization in flow of carbon nanotubes. J. Therm. Anal. Calorim. 2020, 140, 321–329. [Google Scholar] [CrossRef]
  95. Hayat, T.; Qayyum, S.; Alsaedi, A.; Ahmad, B. Entropy generation minimization: Darcy-Forchheimer nanofluid flow due to curved stretching sheet with partial slip. Int. Comm. Heat Mass Transf. 2020, 111, 104445. [Google Scholar] [CrossRef]
  96. Japar, W.M.A.A.; Sidik, N.A.C.; Asako, Y. Entropy generation minimization in sinusoidal cavities-ribs microchannel heat sink via secondary channel geometry. CFD Lett. 2019, 11, 1–10. [Google Scholar]
  97. Mohapatra, T.; Rout, S.K. Experimental investigation and performance optimization of a cross flow heat exchanger by entropy generation minimization approach. J. Therm. Eng. 2019, 5, 1–12. [Google Scholar]
  98. Han, Y.; Wang, X.S.; Zhang, H.N.; Chen, Q.Z.; Zhang, Z. Multi-objective optimization of helically coiled tube heat exchanger based on entropy generation theory. Int. J. Therm. Sci. 2019, 147, 106150. [Google Scholar] [CrossRef]
  99. Nemati, H.; Moghimi, M.A.; Sapin, P.; Markides, C.N. Shape optimisation of air-cooled finned-tube heat exchangers. Int. J. Therm. Sci. 2020, 148, 106233. [Google Scholar] [CrossRef]
  100. COMSOL Multiphysics User’s Guide, Version 5.2a; COMSOL Incorporated: Stockholm, Sweden, 2012; pp. 103–147.
Figure 1. Geometric model of heat sources.
Figure 1. Geometric model of heat sources.
Entropy 22 00651 g001
Figure 2. The influence of Δ q on   S ˜ g , ϕ .
Figure 2. The influence of Δ q on   S ˜ g , ϕ .
Entropy 22 00651 g002aEntropy 22 00651 g002b
Figure 3. The influence of Δ q on N u ¯ .
Figure 3. The influence of Δ q on N u ¯ .
Entropy 22 00651 g003
Figure 4. The color maps of temperature gradient distribution on the upper surfaces of discrete elliptical cylinders with heat generating and thermal conduction pedestal (Re = 5000).
Figure 4. The color maps of temperature gradient distribution on the upper surfaces of discrete elliptical cylinders with heat generating and thermal conduction pedestal (Re = 5000).
Entropy 22 00651 g004
Figure 5. Contour plot for pressure in channel ( Δ q = 0 W·m−3).
Figure 5. Contour plot for pressure in channel ( Δ q = 0 W·m−3).
Entropy 22 00651 g005
Figure 6. The influence of λs on   S ˜ g , ϕ .
Figure 6. The influence of λs on   S ˜ g , ϕ .
Entropy 22 00651 g006
Figure 7. The influences of Δλs on   S ˜ g , ϕ .
Figure 7. The influences of Δλs on   S ˜ g , ϕ .
Entropy 22 00651 g007
Table 1. The influences of H ˜ and b on T ˜ .
Table 1. The influences of H ˜ and b on T ˜ .
H ˜
0.10.2
b 0.46520.4835
T ˜ [82]7.233913.6509
T ˜ of this work7.180513.8243
Table 2. The optimal distributions of heat generating intensity for different Re.
Table 2. The optimal distributions of heat generating intensity for different Re.
Re3000400050006000
Δ q opt   /106 W·m−3−0.6−0.9−1.0−1.1
  S ˜ g , ϕ 0.029190.022010.017020.01355
Table 3. The optimal distributions of the thermal conductivity for different Re.
Table 3. The optimal distributions of the thermal conductivity for different Re.
Re250030003500400050006000
(Δλs)opt / W·m−1·K−1460000
  S ˜ g , ϕ 0.033820.029190.025300.022010.017020.01355

Share and Cite

MDPI and ACS Style

Wang, R.; Xie, Z.; Yin, Y.; Chen, L. Constructal Design of Elliptical Cylinders with Heat Generating for Entropy Generation Minimization. Entropy 2020, 22, 651. https://doi.org/10.3390/e22060651

AMA Style

Wang R, Xie Z, Yin Y, Chen L. Constructal Design of Elliptical Cylinders with Heat Generating for Entropy Generation Minimization. Entropy. 2020; 22(6):651. https://doi.org/10.3390/e22060651

Chicago/Turabian Style

Wang, Rong, Zhihui Xie, Yong Yin, and Lingen Chen. 2020. "Constructal Design of Elliptical Cylinders with Heat Generating for Entropy Generation Minimization" Entropy 22, no. 6: 651. https://doi.org/10.3390/e22060651

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