Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

International Journal of Heat and Mass Transfer 215 (2023) 124513

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Numerical investigation and experimental verification of topological


optimized double-layer mini-channels
Xing-ping Tang a, Huan-ling Liu a, *, Li-si Wei a, Chuan-geng Tang a, Xiao-dong Shao a, Han Shen a,
Gongnan Xie b, *
a
Shaanxi Key Laboratory of Space Solar Power Station System, School of Mechano-Electronic Engineering, Xidian University, Xi’an 710071, China
b
Research & Development Institute of Northwestern Polytechnical University in Shenzhen, Shenzhen 518057, China

A R T I C L E I N F O A B S T R A C T

Keywords: Double-layer micro/mini channel heat sinks have great potential in the thermal management for integrated
Topology optimization chips. In this work, two bi-objective optimization design of mini-channels is implemented using the variable
Double layer mini-channel heat sink density topology optimization method to get two topological single-layer mini-channel heat sinks (SL-MHSs).
Pressure drop
Among them, the SL-MHS obtained with the minimization of average temperature and power dissipation as the
Nusselt number
optimization objectives is named S1, while the SL-MHS obtained with the minimization of temperature variance
and power dissipation as the optimization objectives is named S2. Then, S1 and S2 are used as the upper/lower
layer of the double layer mini channel heat sink (DL-MHS), respectively. Therefore, four different topological DL-
MHSs including D11, D12, D21 and D22 are achieved by assembling S1 and S2. Subsequently, the flow and
thermal characteristics of the four topological DL-MHSs and conventional straight DL-MHS (C1) are studied
numerically. The results show that when Re = 315, compared with C1, the maximum temperature of the sub­
strate of the topological DL-MHSs (D11, D12, D21 and D22) can be respectively reduced by 17.0,16.6, 14.0 and
13.6 K, while the Nusselt number can be improved by 27.3%, 26.6%, 27.1% and 26.4%, respectively. Addi­
tionally, the total thermal resistance is also reduced at least 20%. However, although topological DL-MHSs
produce higher pressure drop penalty, the PEC values of topological DL-MHSs are always greater than 1,
which indicates that the new topological DL-MHSs is meaningful. Finally, the flow and thermal characteristics of
the best one D11 are investigated experimentally. The simulated results agree well with the experimental results.

reported by Tuckerman and Pease [6]. Since then, to enhance the per­
1. Introduction formance of microchannel heat sinks, researchers have conducted
studies by altering the geometric shapes of the channels, such as rect­
With the rapid development of information technology, the elec­ angular, tapered [7] and corrugated [8] channel structures. Addition­
tronic industry is also developing in a breakthrough way. High- ally, in order to improve the thermal and hydraulic performance,
performance microelectronic devices are increasingly integrated and researchers have also examined the effect of different types of coolant on
miniaturized, which generates high heat flux and a wide range of hot the cooling ability, such as nanofluids [9] and GaInSn coolant [10], as
spots, seriously affecting the service life of electronic devices [1]. well as investigated the effect of channel wall roughness, including
Therefore, reasonable and efficient thermal management technology is channels with porous media [11]. These improvement methods have
particularly important. At present, the high-throughput cooling methods indeed enhanced the performance of the heat sink to a certain extent.
for microelectronic components mainly include axial fan forced air However, higher pressure drop may be also produced, which means that
cooling, heat pipe phase change cooling, micro/mini channel liquid a larger pumping power is required to circulate the coolant. This
cooling technology, porous media cooling technology, spray cooling and increased pressure drop may pose challenges in terms of the encapsu­
jet impingement cooling [2–4]. Among them, micro/mini channel heat lation, potentially leading to leakage issues in heat exchangers [12]. This
sinks have been widely concerned due to their good heat transfer per­ constitutes a threat to the safe and stable operation of electronic devices
formance, compact structure, stable operation and low cost [5]. further. Additionally, the temperature along the flow direction rises
Single-layer liquid-cooled microchannel heat sink was firstly sharply as the coolant temperature becomes warmer and warmer due to

* Corresponding authors.
E-mail addresses: hlliu@xidian.edu.cn (H.-l. Liu), xgn@nwpu.edu.cn (G. Xie).

https://doi.org/10.1016/j.ijheatmasstransfer.2023.124513
Received 14 March 2023; Received in revised form 28 June 2023; Accepted 14 July 2023
Available online 26 July 2023
0017-9310/© 2023 Elsevier Ltd. All rights reserved.
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

Nomenclature u vector velocity of fluid, m/s


uin inlet velocity, m/s
Ab bottom surface area of heat sink, m2 vf volume fraction of fluid
Acon contact area of fluid and channel wall, m2 w width of channel, m
c specific heat capacity, J⋅kg− 1⋅K− 1
Dh hydraulic diameter, m Greek symbols
h height of channel, m α reverse permeability
β projection slope
h average heat transfer coefficient, W⋅m− 2⋅K− 1
γ design variable
k thermal conductivity, W/(m⋅K)
μ dynamic viscosity of fluid
Nu average Nusselt number
ρ density of material, Kg/m3
ΔP pressure drop, Pa
Φ1 average temperature function, K
P pressure, Pa
Φ2 temperature difference function, K2
qw heat flux, W/m2
Ψ fluid dissipation power, W/m
qα penalty factor for drag coefficient
Π bi-objective function
qc penalty factor for specific heat capacity
qρ penalty factor for density Subscripts
qk penalty factor for thermal conductivity avg average
Qv volume flow rate, ml/s f fluid
Qw total input power, W l lower channel
Re Reynolds number s solid
RT thermal resistance, K/W u upper channel
RL total thermal resistance in inlet
T temperature, K out outlet
ΔT temperature difference, K max maximum value
Twavg average temperature of channel wall, K tci thermocouple location
Twtc wall temperature of channel by calculating, K
Tfavg average fluid temperature, K

absorbing the heat from the chip. This leads to uneven thermal stress of it can be categorized as shape, size and topology optimization [26].
the electronic chips, which makes these chips appearing thermal insta­ Conventional optimization methods including shape and size opti­
bility and thermal breakdown and other undesirable phenomena [13]. mization have few design degrees of freedom and make it difficult to
To effectively restrain the continuous temperature rise of the single obtain an optimal structure. Topology optimization is a non-intuitive
layer heat sink in the flow direction, Vafai and Zhu [14] firstly reported approach in the optimal design which does not rely on the re­
a microchannel heat sink with double-layer counter flow layout struc­ searcher’s experience [27], priori hypothetical structures, etc., which
ture in 1999. They found that the double-layer design significantly re­ has high-level of design flexibilities. Topology optimization methods can
duces the temperature rise in the flow direction of the bottom surface, be specifically divided into variable density method, level set method
providing a more uniform cooling effect. Subsequently, extensive and direct explicit method [26]. Among them, the variable density
research works have been done based on the double-layer design method is extensively used because of its high computational efficiency
concept to further improve its performance [15–19]. Sarvar-ardeh et al. and easy to achieve numerical solution. Xie et al. [28] performed the
[20] further explored the influence of using hydrophobic and super­ structural optimization of air radiators using variable density method.
hydrophobic surfaces in bilayer conical microchannels on the hydro­ Zhou et al. [29] proposed a novel small channel heat sink by topology
thermal performance. Ghahremannezhad et al. [21] numerically optimization for a rectangular domain with arc manifold, and conducted
investigated the influence of porous substrates on the hydrothermal numerical and experimental tests on its cooling ability. They found that
performance of bilayer microchannel radiator. They found that using a the comprehensive hydrothermal property of the optimized heat sinks is
porous substrate with a higher porosity can reduce pumping power with, superior to that of the traditional straight heat sink. To further under­
but it will result in higher thermal resistances. Derakhshanpour et al. stand the application status of topology optimization in the field of heat
[22] adopted the channel design structure with cylindrical ribs and exchanger, Fawaz et al. [26] reviewed ninety-one related studies pub­
studied the hydrothermal performance single and double layer micro­ lished in the past 15 years, and comprehensively described the devel­
channel radiators for different fin spacing and radius parameters. Dai opment of topology optimization technology in the field of heat
et al. [23] numerically and experimentally investigated the flow and exchanger.
thermal performance of countercurrent double-layer porous micro­ Most of the previous optimization studies on the flow and thermal
channel radiator using microcapsule phase change material (MPCM) performance of double-layer mini/micro channel heat sinks have
suspension as coolant. Their results showed that the comprehensive focused on the predetermined channel structures [30–33], which have
thermo-hydraulic performance of porous microchannel radiator con­ obtained lots of achievements in the heat sink design. However, the
taining MPCM suspensions is increased by 10% compared to water when optimization of these conventional methods is very limited to get the
the number of channels is twenty. El-dean et al. [24] experimentally best flow and thermal characteristics of the heat sink. Currently, to­
studied the thermal characteristics of two-phase flow in a double-layer pology optimization methods based on the variable density method is
microchannel. Shen et al. [25] introduced a deflector plate into the advantageous in the design of heat sink structures, which can generate
traditional straight double-layer micro-channel heat sink. They found the best channel configurations [28]. However, this method has mostly
that the use of deflector plates can promote the mixing of upper and been applied to the design of single-layer mini/microchannel heat sinks,
lower fluids, leading to highly efficient heat dissipation ability. Struc­ while it has rarely been used to the design of double-layer mini/­
tural optimization is an essential means of improving the comprehensive microchannel heat sinks. Therefore, in order to further improve the
hydrothermal performance of heat sinks. For the structure optimization, comprehensive hydrothermal performance of the double-layer heat

2
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

sink, the topology optimization method based on variable density is Table 1


applied to design double-layer heat sink. There is no report on this Geometric parameters.
design method of double-layer heat sinks in literature available. This Parameters Values (mm)
work provides a new idea for the design of double-layer heat sinks.
Lx, Ly, Lz 96, 96, 16
Firstly, we designed two single-layer topological heat sinks with excel­ δu, δb, δm 2, 2, 4
lent performance using two constructed bi-objective optimization w, h, H 15, 4, 12
functions as topological objects. Then, four double-layer topological d1, d2 10.5, 15
heat sinks are obtained by cross-assembling two single-layer topological
heat sinks together. Finally, to find the optimum combination between
the upper and lower layers, in terms of thermal characteristics, param­ ∇⋅u = 0 (1)
eters such as temperature, thermal resistance and Nusselt number of the
bottom plate of these DL-MHSs are studied in detail, and the influence of ρ(u⋅∇)u = − ∇P + μ∇2 u + F (2)
heat transfer between the upper and lower layers of the DL-MHS has also
been investigated. For the flow characteristics, the paper focuses on the where ρ, u and μ represent the density, velocity vector and dynamic
pressure drop and friction factor. viscosity coefficient of the fluid, respectively. P is the pressure of the
fluid, and F represents the friction force of the fluid during the flow.
2. Model descriptions Energy equation for thermal field:
ρc(u⋅∇T) = ∇⋅(k∇T) + Q (3)
The three-dimensional (3D) model straight DL-MHS named C1 is
shown in Fig. 1(a). The length, width and height of C1 are denoted by Lx, where c is the specific heat capacity at constant pressure, k is the thermal
Ly and Lz, respectively. See Table 1 for detailed parameters. Fig. 1(b) conductivity, and Q is the heat source.
shows the two-dimensional (2D) schematic of the same fluid-solid
domain for the top and bottom layers of the DL-MHS (x-y plane), 3.3. Variable density method
named R0. From Table 1, the percentage of the fluid domain in R0 is
46.875%. There are three channels with the same size in each layer of The variable density method is one of the most popular numerical
the DL-MHS. The fluid inlet and outlet positions have been marked with methods used in topology optimization. The essence is to discretize the
arrows on each channel. The bottom wall of the DL-MHS is tightly entire design domain into finite elements by introducing the design
connected with the heating source with uniform heat flux qw. variable γ as the pseudo-density value of the material, and to establish a
nonlinear relationship between the pseudo-density values of the ele­
3. Topology optimization for 2D models ments and the physical properties of the material. The design variable γ
varies from 0 to 1. When γ = 0, it denotes the solid domain. When γ = 1,
3.1. Design domain description it denotes the fluid domain.
In order to pursue a better distribution of the fluid-solid material in
The purple area in Fig. 2 is the initial design domain to be optimized. the design domain, it is necessary to simplify the friction term F in Eq.
It can be observed that three staggered inlets and outlets are set on both (2). In this work, the Brinkmann penalty model is applied to simulate the
sides of the design domain. Both of the length (L) and width (W) of the friction of a fluid when it passes through an ideal porous medium [34]:
design domain are 90 mm. Moreover, the width of the inlet/outlet is w,
F = − αu (4)
the distance between the two adjacent inlets and outlets is d2, and the
distance between the inlet/outlet on both sides of the design domain and
where α represents the local inverse permeability or friction coefficient
the upper/lower edges is d3, which is 7.5 mm. The inlet fluid tempera­
of porous media, which is dimensionless and determined by the design
ture Tin = 293 K, the outlet pressure P0 is standard atmosphere. The
variable γ.
uniform heat flux with qw of 1 × 105W/m2 is applied to the entire design
domain.
3.4. Material property interpolation

3.2. Governing equations for 2D models The SIMP (Solid Isotropic Material with Penalization) [35] interpo­
lation model is used to interpolate the material physical property pa­
The fundamental governing equations need to be satisfied in the fluid rameters in the design domain. The design variable γ is the independent
flow and heat transfer process. For laminar incompressible flow, the variable of the interpolation function, which is used for regulating the
continuity and momentum equations can be expressed as follows: distribution of fluid-solid materials in the design domain. Material

Fig. 1. Straight DL-MHS (a)3D model C1; (b) 2D fluid-solid diagram R0.

3
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

Fig. 2. Initial design domain.

properties to be interpolated include inverse permeability α, thermal


conductivity k, density ρ, and specific heat capacity at constant pressure
c. The corresponding interpolation functions are as follows:
( ) ( )
q (1 − γ)
α γ = αf + αs − αf α (5)
qα + γ
( ) ( )
qk (1 − γ)
k γ = kf + ks − kf (6)
qk + γ
( ) ( )
qρ (1 − γ)
ρ γ = ρf + ρs − ρf (7)
qρ + γ
( ) ( )
qc (1 − γ)
c γ = cf + cs − cf (8)
qc + γ

where the subscripts f and s denote the fluid and solid, respectively. The
material physical properties parameters are listed in Table 2.qα ,qk ,qρ
andqc are the penalty factors for reverse permeability, thermal conduc­
tivity, density and specific heat capacity of the material, respectively.
The penalty factor is used to adjust the shape of the interpolation Fig. 3. The influence of different penalty factor q on interpolation function.
function, which is critical in the value selection. If it is not selected
properly, more gray cells may be appeared in the optimization results choose the value of q by experience and then adjust the q value to obtain
[27]. According to the method of Yu et al. [36], we study the effect of better topological results. Therefore, in this work, the penalty factor is
penalty factor q on the interpolation of properties. Fig. 3 shows the selected based on previous studies, and then appropriately adjusted to
relationship between the penalty factor q and the penalty term q obtain reasonable topological results. The selected value of q is shown in
(1-γ)/(1+γ) of the interpolation function. It can be seen that the con­ Table 3.
vexity of the penalty term gradually becomes larger as q decreases,
indicating an increasing level of penalization on material properties. For
the selection of the penalty factor q, we refer to the parameters of these
publications [27,37–39]. These references have been showed that there
is no specific standard procedure for the choice of q. Researchers always Table 3
Values of penalty factors in previous studies and this work.
Reference qα qk qρ qc
Table 2
basic parameters of materials. Zhou et al. [27] 0.015–0.03 0.015–0.03 0.015–0.03 100–150
Han et al. [37] 0.02 0.01 0.01 100
Material k[W/(m⋅K)] c [J/Kg⋅K] ρ [kg/m3] α μ [Pa⋅s]
Koga et al. [38] 0.01–0.1 – – –
3
Water 0.6 4182 998.2 0 10− Hu et al. [39] 0.02 0.01 0.01 100
Aluminum 202.4 871 2713 106 – This work 0.01 0.01 0.01 100

4
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

3.5. Density filtering and projection


Φi Ψ
Πi = ω + (1 − ω) (i = 1, 2) (14)
Φ0i Ψ0
In order to avoid the influence of grid dependency and checkerboard
problems in topology optimization, Helmholtz density filters are used to where ω is the weight coefficient of the bi-objective function, and ω =
filter design variables. Its essence is to replace the real design variable 0.5. Φ0i and Ψ0 are the initial values of the thermal and flow charac­
value with the filtered design variable to participate in the optimization teristics functions, respectively. The purpose of introducing initial
iteration process and improve the numerical stability of the solution values is to achieve dimensionless of the objective function to facilitate
process. The formula of Helmholtz density filters can be written as fol­ analysis.
lows: The final optimization mathematical model can be determined as
− r2 ∇2̃γ + ̃γ = γ (9) follows [37]:
Find γj (j = 1, 2...n)
where γ and ̃γ are the design variables before and after filtering,
respectively. r is the filtering radius. The minimum mesh size is the Minimize Πi = ω
Φi
+ (1 − ω)
ψ
default value for filter radius. Φ0i ψ0
∫ ∫ (15)
Since a large number of gray units (units with density values between {
γdΩ ≤ vf 1dΩ
0 and 1) may be generated after filtering by the filter, a smooth pro­ Subject to Ω Ω
jection ladder function is used to adjust the density value to zero or unit 0 ≤ γj ≤ 1
in order to eliminate the gray units and obtain a clear topology. The
function can be obtained by [40]: wherevf represents the volume fraction occupied by the fluid in the
( ( )) (
tanh β ̃γ − γβ + tanh βγ β
) design domain, and vf = 0.5. n represents the number of discrete grid
̂γ = ( ( )) ( ) (10) cells in the design domain.
tanh β 1 − γ β + tanh βγ β

wherêγ denotes the post-projection design variable, γ β denotes the hy­ 3.7. Optimization methods and results
perbolic tangent projection point with the value of 0.5, β is the projec­
tion slope with the value of 8 [41]. The optimization solution software used in this work is COMSOL
Muitiphysics 5.4. Firstly, the design variable field, the thermal field and
3.6. Mathematical model of optimization the flow field are discretized by Lagrangian linear elements. Then, the
finite element analysis is performed based on the governing equations
For the liquid cooling system, its performance evaluation criteria are for the different physical fields, and the objective function values are
mainly considered by two aspects: one is the power dissipation, the evaluated to determine if convergence is achieved. If the convergence
other is the heat dissipation ability of the system. Therefore, this work accuracy or maximum step length is achieved, the result is output. On
aims at these two points to determine the optimization objective the other hand, if convergence is not achieved, sensitivity analysis is
function. carried out based on the adjoint method [42], which has the advantages
The first objective to be optimized is the power dissipation, whose of high computational efficiency, high solution accuracy and low storage
characteristics are reflected in the pressure drop in the channels. Hence, requirements. Specifically, according to references [42–44], in contrast
the objective function is defined as follows [37,39]: to the direct method (such as, the finite difference method), the adjoint
method avoids multiple calculations due to a large number of parame­
∫ [ ∑( ) ]
1 ∂ui ∂uj 2 ∑ ters and calculates all the sensitivities in a single backward propagation
Ψ= μ + + α(γ)u2i dΩ (11)
Ω 2 i,j
∂xj ∂xi i process. The adjoint method chooses analytical derivatives rather than
numerical approximations, which reduces the effect of rounding errors
where Ω represents the design domain and α(γ) is the inverse perme­ and improves the accuracy of the results. In addition, compared with the
ability of the interpolated porous medium. direct method, the adjoint method only needs to store the solution of the
The second objective to be optimized is the heat dissipation ability. original problem and the corresponding adjoint variables, and does not
In general, there are two main types of requirements for the temperature need to store a huge sensitivity matrix. Then, the design variables are
uniformity of heat sinks in engineering applications: one is to reduce the updated by the global convergence method of moving asymptotes
temperature gradient of the bottom surface of the heat sink. Therefore, it (GCMMA) [45], which has the advantages of global convergence, fast
is necessary to use the minimum temperature variance as the optimi­ convergence speed and good stability. That is, according to references
zation objective. The other requirement is to ensure that the average [46–48], the GCMMA algorithm guarantees global convergence in to­
temperature of bottom surface of the heat sink is minimum, so it is pology optimization, i.e., finding the global optimum across the entire
appropriate to use the minimum average temperature as the optimiza­ design space rather than just a local optimum. Furthermore, GCMMA
tion objective. Therefore, two objective functions are set up as follows employs the strategy of moving asymptotes to guide the search direc­
[37]: tion, resulting in faster convergence to the optimal solution. By gradu­
∫ ally adjusting the position and slope of the asymptotes, GCMMA can
1
Φ1 = TdΩ (12) adapt to changes in the design space, which enhances the stability and
|Ω| Ω
convergence speed of the algorithm. Subsequently, the design variables
∫ ( ∫ )2 are filtered and projected, and finally returned to the initial stage for
1 1
Φ2 = T− TdΩ dΩ (13) recalculation. The whole topology optimization procedure is shown in
Ω |Ω| |Ω| Ω Fig. 4.
To guarantee the accuracy and robustness of the topology optimi­
where Φ1 and Φ2 are functions that characterize the average tempera­
zation results, a grid independence test is required for the 2D model. In
ture and temperature variance of the design domain, respectively.
this work, an unstructured triangular mesh is used to discretize the
Considering the flow and heat dissipation performance of the system,
computational domain. The results of grid independence tests are listed
two bi-objective optimization functions are determined as follows:
in Table 4. It can be seen that the objective values of both optimization
schemes Π1 and Π2 gradually stabilize as the number of grids increases,
with very little change in the objective values for both when the number

5
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

Fig. 4. Flowchart of the topology optimization procedure.

optimization needs to be extended to 3D models for subsequent nu­


Table 4
merical analysis and experimental validation, so it is necessary to ensure
The results of grid independence tests for 2D models.
that the inlet and outlet positions are clear and the boundaries are
Grid numbers Π1 Relative error Π2 Relative error continuous. Therefore, we need to modify the boundary of the topology
25,346 0.36148 1.41% 0.04105 9.97% result. According to the method of reference [49], we make appropriate
37,834 0.35906 0.72% 0.03896 3.30% incremental processing of the boundaries on the topological optimized
56,096 0.35756 0.31% 0.03784 1.37%
results (M1 and M2) to maintain the connectivity of the structure by
64,828 0.35647 – 0.03733 –
75,426 0.35589 0.16% 0.03698 0.94%
adding thin frames to act as solid/fluid materials, and set the locations of
the inlets and outlets. These thin frames are marked with yellow lines in
Fig. 6. The width of the frame d0 is 3 mm, the width of the inlet and
of grids reaches 64,828. Therefore, considering the computational re­ outlet are the same as the initial setting w, and the length Lx and width Ly
sources, time cost and computational accuracy, the grid number of of the frame are 96 mm. M1 and M2 are respectively named R1 and R2
64,828 is used to complete the topology calculation. after correction. The volume fraction of fluid in M1 and M2 before
Fig. 5 shows the distribution of fluid-solid materials obtained after correction is 50%. After correction, the volume fraction of fluid in R1
topology optimization, where the red and blue areas refer to solids and and R2 can be calculated as 46.875%. It can be seen from the above data
fluids, respectively. M1 is the result obtained with the average temper­ that R0, R1, and R2 have the same overall geometric parameters, outlets,
ature and pressure drop as the optimization objectives, and M2 is the and inlets, as well as the same volume fraction of the fluid, which en­
result obtained with the temperature variance and pressure drop as the sures the fairness of the performance comparison between the DL-MHSs
optimization objectives. achieved by topology optimization and the traditional straight DL-MHS.

4. Numerical investigation of 3D models


3.8. Correction of topology results
4.1. 3D model heat sinks
According to the 2D topology optimization results M1 and M2 shown
in Fig. 5, the positions of inlet and outlet are not clear and the bound­
In order to obtain the DL-MHS with 3D topological structure, firstly,
aries referring to the areas except the inlets and outlets on both sides are
2D models R1 and R2 are stretched to generate 3D models S1 and S2, as
not completely closed. The 2D optimized structure obtained by topology

Fig. 5. Topology optimization results.

6
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

Fig. 6. Correction of topology optimization results.

shown in Fig. 7(a). The total height of S1 and S2 is 8 mm, respectively.


∇⋅u = 0 (16)
The channel height is 4 mm, and both of the thickness of top and bottom
walls are 2 mm. Then, four types of DL-MHSs can be obtained by taking Momentum equation:
S1 and S2 as the lower or upper layers of the DL-MHS, respectively. With
different combinations, we obtain four different DL-MHSs as shown in ρ(u⋅∇u) = − ∇P + μ⋅∇2 u (17)
Fig. 7(b), namely D11 (S1 lower, S1 upper), D12 (S1 lower, S2 upper), Energy equation for fluid:
D21 (S2 lower, S1 upper) and D22 (S2 lower, S2 upper). All geometric
parameters and inlet and outlet positions of D11, D12, D21, and D22 are ρc(u⋅∇T) = kf ∇2 T (18)
the same as the C1. Energy equation for solid:

4.2. Governing equations for 3D models ks ∇2 T = 0 (19)

Before numerical simulation, it is necessary to make some assump­ 4.3. Boundary conditions
tions about the models:
The inlet boundary conditions:
• The fluid is an incompressible Newtonian fluid; For lower channel:
• The flow is laminar; ( )⊤
• The thermal physical parameters for solid and fluid are constant. u = uin,l , 0, 0 , T = Tin,l = 293K (20)

In terms of upper channel:


The governing equation can be expressed as follows:
( )⊤
Continuity equation: u = uin,l , 0, 0 , T = Tin,u = 293K (21)

The outlet boundary conditions:

Fig. 7. 3D heat sinks (a) SL-MHSs S1 and S2; (b) four combination DL-MHSs D11, D12, D21 and D22.

7
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

The outlets for all channels are pressure outlets, which can be
Qv = N⋅uin ⋅w⋅h (33)
specified as:
Pout,l = Pout,u = 1atm (22) where N is the number of channels.
The total thermal resistance (RT) and the local thermal resistance
For the fluid-solid wall interfaces:
(RL) can be expressed as follows [50,51]:
u = 0, Tf = Ts (23) Tb,max − Tin(l/u)
RT = (34)
qw ⋅Lx ⋅Ly
kf ∇Tf = ks ∇Ts (24)

Considering the experimental conditions, that the maximum total RL =


Tb − Tin(l/u)
(35)
volume flow of the peristaltic pump should not exceed 1700 ml/min. qw ⋅Lx ⋅Ly
Therefore, the inlet velocity should not exceed 0.08 m/s due to six inlets
in 3D heat sinks. This velocity is small that the heat flux applied to the where Tb is the local temperature of the bottom surface.
bottom surface of the 3D heat sinks cannot reach too large, otherwise, The flow performance of the double-layer radiator can be charac­
the higher temperature of the heat sink may be found. For the bottom terized by the pressure drop. In this work, the average pressure drop ΔP
surfaces of DL-MHS, the same uniform heat flux qw of 1 × 105 W/m2 are of a single inlet and outlet can be defined as:
applied to the bottom plate of the 3D heat sinks: (∑
Pin,l −
∑ ) (∑
Pout,l + Pin,u −

Pout,u
)
/ ΔP = (36)
∂Ts N
qw = − ks = constant = 1 × 105 W m2 (25)
∂n ∑ ∑
where Pin,l and Pout,l respectively represent the sum of pressures of

The rest surfaces: all inlets and outlets for the lower layer of the DL-MHS, and Pin,u and

Pout,u respectively represent the sum of pressures for all inlets and
∂Ts
=0 (26) outlets at the upper layer of the DL-MHS.
∂n
The friction factor can be gotten [50]:
where u represents velocity; subscripts in,l and in,u represent the inlet of 2ΔPDh
lower and upper channels, respectively; subscripts out,l and out, u f = (37)
ρLx u2in
represent the outlet of lower and upper channels, respectively; n is the
normal unit vector of the related surface. To comprehensively analyze the heat flow performance of the DL-
MHS, we use the performance evaluation criterion PEC to evaluate the
performance. The PEC is expressed as:
4.4. Data reduction
Nut /Nu0
PEC = (38)
The Reynolds number can be determined:
1
(ΔPt /ΔP0 )3
ρuin Dh
Re = (27) where subscripts t and 0 represent topology and straight-channel
μ
structure, respectively.
where uin is the inlet velocity. Dh is the hydraulic diameter, which can be
expressed as follows: 4.5. Numerical method

Dh =
2w⋅h
(28) The 3D models C1, D11, D12, D21 and D22 are numerically exam­
w+h ined by using the commercial software ANSYS FLUENT 2021R1. The
The average Nusselt number: SIMPLE algorithm is applied to cope with the coupling of pressure and
velocity fields. The second order upwind difference is taken to discretize
Nu =
h⋅Dh
(29) the energy and momentum formulas. The convergence residual is set
kf according to the default parameters of the software. Table 5 shows the
inlet flow velocity and its corresponding volume flow and Reynolds
where h represents the average convective heat transfer coefficient, number.
defined as [22]:
Aq 4.6. Grid independence tests
h= ( b w ) (30)
Acon Twavg − Tfavg
In this work, the commercial software ICEM CFD 2021 R1 is used to
where Ab is the bottom surface area. Acon is the contact area of fluid and create the required grid system. The meshes of the flow channel walls as
channel wall. Twavg is the average temperature of all channel walls. Tfavg well as the outlet and inlet faces of the model are densified. A boundary
is the average fluid temperature. layer mesh system is used near the walls of the flow channel. The size of
The average temperature of the channel wall Twavg and the average the first cell near the wall is determined by the Reynolds number,
temperature of the fluid Tfavg can be calculated as: satisfying y+=1, and the growth factor of the grid spacing is set to 1.1. To

Twavg,l + Twavg,u Tfavg,l + Tfavg,u


Twavg = , Tfavg = (31) Table 5
2 2
Inlet velocity, Reynolds number and volume flow rate.
The average temperature of the fluid can be defined as: uin(m/s) Qv(ml/s) Re
Tin,(l/u) + Tout,(l/u) 0.02 7.2 126
Tfavg,(l/u) = (32)
2 0.03 10.8 189
0.04 14.4 252
where Tin and Tout represent the temperature of the fluid at the inlet and 0.05 18 315
outlet of the channel, respectively. 0.06 21.6 378
0.07 25.2 441
The volumetric flow rate can be calculated as follows:

8
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

exclude the influence of grid numbers on the numerical results, grid bottom layers.
independence tests are conducted for the five DL-MHSs with six grid The variation of the friction factor as a function of Reynolds number
systems. Fig. 8 reports the effect of the number of grids on Tmax and ΔP is shown in Fig. 10(b). With the increase of Reynolds number, the fric­
for the five simulation models at uin = 0.05 m/s and qw=105W/m2. It can tion factors of the five DL-MHSs show an overall decreasing trend, while
be noticed that when the number of grid magnitudes reaches 6 million, it is shown the fastest decrease rate of C1. When Re > 189, the friction
the fluctuations for both of Tmax and ΔP tend to stabilize, with the factor of C1 is always the lowest of the five DL-MHSs. This can be
fluctuation range of Tmax within 0.4% and that of ΔP below 1.1%. This explained that the friction factor is positively related to the pressure
indicates that the influence of the number of grids here on the numerical drop for the same uin, namely, the lower the pressure drop, the lower the
results is small and can be ignored. Considering the computational re­ friction factor, and C1 is at the lowest pressure drop. When Re < 189, the
sources, time cost and computational accuracy, the six million grid friction factor of D11 consistently remains the lowest. This is because at
system is chosen to complete the subsequent numerical calculations. lower Reynolds numbers, the velocity is lower for optimal flow chan­
nels, the fluid primarily flows along the main wider flow channels, while
for the branching fine flow channels, the fluid may be in a stagnant state.
4.7. Results and discussion
4.7.2. Analysis of thermal characteristics
4.7.1. Analysis of flow characteristics The temperature contours of the bottom surface of the five DL-MHSs
The velocity distributions of the three channel structures at uin = are shown in Fig. 11 for uin = 0.05 m/s. The hot spots of C1 are
0.05 m/s are shown in Fig. 9. Fig. 9(a) shows the velocity distribution of concentrated in the middle and downstream positions of the channels,
the straight channel. After the inflow of fluid from the inlet, the flow while the temperature distribution in other areas is relatively uniform.
boundary layer in the channel develops stably and the flow velocity The hot spot area at the outlet of the channels on both sides is large due
distribution is uniform. Fig. 9(b) and Fig. 9(c) show the velocity distri­ to the gradual increase of the temperature of the fluid along the flow
bution of topological SL-MHSs S1 and S2, respectively. It is noted that direction. Furthermore, the gradual decrease of the temperature differ­
the velocity at the inlets and outlets is higher, while the velocity in the ence between the wall and the fluid, leading to the reduction of the heat
middle part is lower owing to the shunt of the secondary flow channel. dissipation efficiency. The hot spot area at the outlet of the middle
However, the velocity remains relatively uniform and stable. channel is relatively small, which is due to the influence of staggered
The relationship between the pressure drop and Reynolds number is inlets and outlets. Compared to C1, it is easy to see that the bottom
shown in Fig. 10(a). It is shown that the pressure drop and Reynolds surface temperature of topological DL-MHSs (D11, D12, D21 and D22) is
number are positively correlated. When Re <189, the pressure drop of significantly reduced leading to better temperature uniformity. Mean­
the five models have little difference. It can be explained that the effect while, it can also be found that the hot area of topological DL-MHSs only
of the flow channel structure on the pressure drop is very slight due to appears near the outlets on both sides, and the lower temperature area
lower velocity. When Re >189, it is obvious that the rise of pressure drop appears near the entrance and the central area. The lower temperature is
gradually becomes remarkable leading to distinct difference in terms of due to the scouring of fluid in multiple directions, which results in
pressure drop among these models. The pressure drop of C1 is the stronger convective heat transfer than in the other areas. When the
smallest among the five DL-MHSs at the same Reynolds number. This is surrounding fluid hits the central solid wall, the flow path changes,
because the three straight flow channels of each layer are mutually in­ accompanied by a series of vortices, resulting in a thinning of the flow
dependent. Although the outlets and inlets are staggered, the flow in boundary layer. This cause the conductive thermal resistance is reduced,
each channel does not interfere with each other, and the fluid flow path which improving the heat transfer capability. Additionally, it is notice­
is single, resulting in small pressure drop loss. However, for topology able that the substrate temperature distributions of D11 and D12 are
optimized heat sinks, to maximize heat dissipation, topology optimiza­ similar, while they are somewhat different. The similarity is due to the
tion generates discrete special-shaped structures, which leads to the fact that they have the same channel structure of the lower layer.
growth of on-way resistance and the increase of the local resistance. For However, the flow channels of the upper layer are different. This cause
the four topological DL-MHSs, the pressure drop of D11 is the minimum different velocities in the upper layer for D11 and D12, which generates
and that of D22 is the maximum, while those of D12 and D21 are the different temperature distributions for D11 and D12. The substrate
medium and completely equal. This is because D12 and D21 are both temperature distributions of D21 and D22 are similar to that of D11 and
composed of S1 and S2, while their positions are different in the top and

Fig. 8. Effect of the number of grids on (a) maximum temperature (Tmax) and (b) pressure drop (ΔP) for five models at uin = 0.05 m/s and qw=105W/m2.

9
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

Fig. 9. Velocity distribution of the three channel structures at uin =0.05 m/s (a) C1; (b) S1; (c) S2.

Fig. 10. Variations of (a) pressure drop and (b) friction factor with Reynolds number for five DL-MHSs.

D12. The maximum temperature differences for D11 and D12 are 12.80 temperature variation trend for the five DL-MHSs is similar to that in
and 12.87 K, while those for D21 and D22 are 15.63 K and 14.93 K, Fig. 12(a). However, the amplitude of temperature fluctuations of to­
respectively. In addition, the temperature of the hot zone for D11 and pological DL-MHSs (D11, D12, D21 and D22) are larger than that on the
D12 is lower than that for D21 and D22, which indicates that D11 and line y = 48 mm, because the outlet on both sides is the area where the
D12 have better heat dissipation uniformity. hot spots are concentrated. Fig. 12(c) and (d) show the variation of the
To explore the heat transfer characteristics of the bottom surfaces of local thermal resistance for y = 48 mm and y = 18 mm, respectively. It
the five DL-MHSs in more depth, Fig. 12 shows the variations of tem­ can be observed that the local thermal resistance of four topological DL-
perature and local thermal resistance along the lines y = 48 mm and y = MHSs is much lower than that of the conventional DL-MHS. The local
18 mm for the five DL-MHSs at uin = 0.05 m/s. As is evident from Fig. 12 thermal resistance directly reflects the local heat transfer capability,
(a), the temperature of C1 gradually increases along the flow direction, which indicates that the heat transfer capacity of the topological DL-
reaching the peak value at the distance of 50 mm from the inlet, and MHSs is better. Additionally, it can be observed that the variation
then slowly decreasing. In contrast, for the four topological DL-MHSs, trend of local thermal resistance is similar to the corresponding tem­
the temperature variation is fluctuating, reaching peak values at dis­ perature change trend. This is because local thermal resistance is line­
tances of 23 and 70 mm from the inlet. Within the range of 30 to 65 mm arly related to temperature.
from the inlet, the temperature is lower and exhibits a stable variation. From the temperature contours in Fig. 11, it can be observed that
The comparison shows that the four topological DL-MHSs have smaller areas with significant temperature gradients along the x-direction
temperature fluctuation range and better temperature uniformity appear near the outlets on both sides. Therefore, in this study, the
compared to C1. The advantages of topological DL-MHSs are mainly temperature distribution at the cross-section (x-z plane) at y = 13 mm is
showed in two aspects: Firstly, the topology generates multiple chosen to analyze the heat transfer performance of the upper and lower
branching flow paths, which results in discrete special-shaped struc­ layers of the DL-MHSs. Fig. 13 shows the temperature distribution of the
tures. This produces more uniform fluid distribution and stronger fluid cross-section at y = 13 mm for the five DL-MHSs at uin = 0.05 m/s. It can
mixing, which enhancing heat transfer; Secondly, the cooling effect of be seen that, compared to the other four topological DL-MHSs, the
the upper counter-flowing channels becomes better, which weakens the traditional DL-MHS C1 has no obstruction disturbance in the flow
rise of the temperature of the bottom surface. The temperature variation channel, the thermal boundary layer develops steadily. This leads to the
for y = 18 mm is depicted in Fig. 12(b). It can be seen that the fluid cooling effect weaken gradually along the flow direction. It is

10
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

Fig. 11. Temperature contours of bottom surface of five DL-MHSs at uin = 0.05 m/s.

worth noting that the bottom surface temperature gradient of topolog­ not only conducive to the concentration of dispersed hot spots, but also
ical DL-MHSs (D11, D12, D21 and D22) in the y direction is smoother can effectively reduce the temperature of hot spots. As can be observed
than that of the traditional DL-MHS C1. This is because complex sec­ in Fig. 14(b), the DL-MHSs have the lowest temperatures at both ends of
ondary flow channels are generated by the topology, and flow the x-axis, and the temperatures at both ends of each DL-MHS are almost
obstruction disturbance occurs in the flow process of the fluid, resulting the same. The reason is that the middle plane is affected by the counter-
in vortex generation, which damages the flow boundary layer resulting flow which is generated by the upper layer and lower layer. Meanwhile,
in improving the heat transfer efficiency. This can also be seen from the it can be found that the temperature rise trend of C1 is upward convex,
temperature of the fluid in the channel. The fluid temperature in the and the hot area remains concentrated in the middle section. This in­
topology channel is higher than that in the straight channel, indicating dicates that the upper fluid plays a smaller role in reducing the tem­
that more heat can be carried away by the fluid in the topology channel. perature rise along the flow direction at the bottom surface. For the four
Therefore, the topology heat sink exhibits better cooling performance. topological DL-MHSs, there is less temperature fluctuation along the x-
Furthermore, it can be observed that D11 and D12 have the same direction and the temperature is relatively low. This means that the use
structure in the lower channel, and the difference in temperature dis­ of topological channels in the upper and lower layers is more effective in
tribution is attributed to the influence of the upper heat sink on the cutting the temperature rise in the bottom plate flow direction.
thermal performance. Due to the longer flow path on the inlet side of the Furthermore, it can be observed that D11 exhibits the least temperature
upper channel in D11 compared to D12, it provides better cooling effect fluctuation. Therefore, the optimization combination of D11 is relatively
to the high temperature fluid at the outlet of the lower channel. As a better.
result, the temperature near the outlet of the lower channel in D11 is The average temperature versus Reynolds number for the bottom
lower than that in D12.The comparative analysis between D21 and D22 surfaces of the DL-MHSs is shown in Fig. 15(a). At the same Reynolds
is similar to that between D11 and D12, while the difference of tem­ number, the average temperatures of the bottom surface of the topo­
perature gradient of their bottom along y direction is smaller. logical DL-MHSs (D11, D12, D21 and D22) are significantly lower than
Comparatively speaking, the combination of upper and lower layers of the traditional DL-MHS C1, while D11 have the lowest average
D11 is the best. temperature.
Fig. 14 shows the temperature variations along the lines l1 (x,13,0) The temperature difference (the difference between the maximum
and l2 (x,13,8) for the five DL-MHSs at uin = 0.05 m/s. The two lines are and minimum temperature of the bottom surface) versus Reynolds
located at the bottom and middle planes of DL-MHS respectively. As can number for the bottom surfaces of the DL-MHSs is shown in Fig. 15(b).
be seen in Fig. 14(a), in the positive direction of the x-axis, the tem­ As the Reynolds number increases, the temperature difference de­
perature of C1 on the line l1 firstly increases, then tends to stabilize, and creases. Among them, D11 and D12 show the fastest rate of decrease,
finally decreases slowly. It indicates that the upper counter-flow fluid while C1 exhibits the slowest rate of decrease. This indicates that as the
inhibits the continuous rise of substrate temperature along the flow di­ flow rate increases, the fluid mixes more vigorously with the wall in the
rection, but the inhibition effect is small. For the topology DL-MHSs, topological flow channel than in the straight channel. Additionally, the
D11 and D12 have similar temperature rise trends, and as well as D21 temperature difference of D11 remains minimal throughout the process,
and D22. Their temperatures increase in fluctuating patterns with indicating that D11 has the best temperature uniformity.
increasing x. It indicates that the topological flow channel structure is Fig. 16(a) shows the influence of Reynolds number on total thermal

11
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

Fig. 12. Variations of temperature and local thermal resistance along the lines y = 48 mm ((a) and (c)) and y = 18 mm ((b) and (d)) when uin = 0.05 m/s.

resistance for the five DL-MHSs. It is apparent that the thermal resistance conductive thermal resistance remains constant. Therefore, the total
is negatively correlated with the Reynolds number. For the same DL- thermal resistance decreases with the increase in Reynolds number. It
MHS, with increasing Reynolds number, the fluid velocity increases, can also be found that the total thermal resistance of D11 is always lower
leading to the improvement of convective heat transfer efficiency and than that of the other four DL-MHSs. When Re = 315, compared with C1,
the decrease of convective thermal resistance. Meanwhile, the the total thermal resistances of D11, D12, D21 and D22 can be reduced

12
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

by 27.6%, 26.9%, 22.7% and 22.0%, respectively. To sum up, D11 has
the best heat transfer ability.
Fig. 16(b) shows the variation of average Nusselt number versus
Reynolds number for the DL-MHSs. It is evident that the average Nusselt
number is positively correlated with the Reynolds number. At the same
Reynolds number, the average Nusselt number of the four topological
DL-MHSs is significantly higher than C1. For example, when Re = 315,
the average Nusselt numbers of D11, D12, D21 and D22 are improved
27.3%, 26.6%, 27.1% and 26.4% compared to C1, respectively. This
indicates that the topologically optimized design has higher heat
transfer capacity than the conventionally designed DL-MHS. Meanwhile,
among the four topological DL-MHSs, D11 exhibits larger average
Nusselt number than the others, indicating that D11 has the highest heat
transfer efficiency.

4.7.3. Comprehensive performance analysis


According to Sections 4.7.1 and 4.7.2, although the thermal perfor­
mance of the topological DL-MHSs is superior to the traditional straight
Fig. 13. Temperature contours of the cross-section (x-z plane) for y = 13 mm
DL-MHSs, the topology structure may result higher pressure drop loss.
for the five DL-MHSs when uin = 0.05 m/s.
To comprehensively assess their advantages and disadvantages, the
performance evaluation criterion (PEC) is adopted to evaluate the
comprehensive hydrothermal performance of the five DL-MHSs. The

Fig. 14. Temperature variations along the lines (a) l1 (x,13,0) and (b) l2 (x,13,8) at uin=0.05 m/s.

Fig. 15. Variations of (a) average temperature and (b) temperature difference with Reynolds number on the bottom surface of the DL-MHSs.

13
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

Fig. 16. Variations of (a) total thermal resistance and (b) average Nusselt number with Reynolds number for the DL-MHSs.

definition of PEC has been provided in Section 4.4. of resin. The heat source on the bottom surface is achieved by using film
Fig. 17 illustrates the variation of the performance evaluation cri­ resistors. Fig. 18(b) is the physical picture of the test model D11. Fig. 18
terion (PEC) for the five DL-MHSs with Reynolds numbers. In this (c) shows the internal structure of the DL-MHS D11.
analysis, the PEC value of C1 is used as the reference, which is main­
tained with constant value of 1 in the range of Reynolds numbers. As can 5.2. Measurement method of relevant parameters
be seen, the PEC of the topological DL-MHSs (D11, D12, D21 and D22)
initially increase and subsequently slowly decrease with increasing 5.2.1. Measurement of channel wall temperature
Reynolds number, while it is always greater than 1. This indicates that Since the temperature of the inner wall of the channel is not easy to
the heat transfer advantage of these topological DL-MHSs is greater than be measured directly, this work refers to the test method of Al-Neama
the disadvantage of pressure drop loss of compared to the traditional DL- et al. [2] to solve this difficulty, as shown in Fig. 18(a). Firstly, four
MHS C1. It is noticeable that the PEC value of D11 always remains the temperature measuring holes are made at z = 1 mm away from the base
highest, indicating that D11 has the best comprehensive performance. wall of the channels. Then, the thermocouples are inserted to these holes
to measure the temperature of corresponding points Tztci,l and Tztci,u.
5. Experimental design Subsequently, the local temperature of the channel wall Twtci,l for lower
heat sink and Twtci,u for the upper heat sink can be calculated by
5.1. Test model one-dimensional heat conduction method. The calculation formula is as
follows:
To validate the credibility of the numerical results, the D11 model is
Qw z
selected as sample to verify the calculated results in this paper. Fig. 18 Twtci,(l/u) = Tztci,(l/u) − (39)
Lx Ly ks
(a) shows the diagram of the D11 model and temperature measurement
points. The DL-MHS is processed by 3D printing technology, which is where Qw is the total heating power, Tztci,(l/u) represents the temperature
made of aluminum. The auxiliary flow collector at the outlet/inlet on data measured by the thermocouples, the subscript i represents the
both sides of the DL-MHS is manufactured by machining, which is made number of temperature measuring holes, and l/u represents the lower/
upper position of the DL-MHS.
Based on the data measured at four local temperature measuring
points of each layer, the average temperature of the channel wall of DL-
MHS can be calculated:
∑4
Twtci,(l/u)
Twavg,(l/u) = i=1 (40)
4

5.2.2. Measurement of pressure drop


As the upper and lower channel structures of D11 model are the
same, they are both composed of SL-MHS S1. Therefore, it is only
necessary to test the pressure drop at the inlet and outlet of one layer of
flow channel, and in this work, only the upper layer is tested. Since the
flow channel structure of S1 is axially symmetric with respect to the
centerline y = 48 mm (x-y plane). Therefore, simply select the inlet and
outlet of one of the sides to test the pressure drop. In addition, the
pressure drop at the inlet and outlet where the centerline is located also
needs to be tested.

Fig. 17. Variation of PEC with Reynolds number.

14
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

Fig. 18. Test model D11.

Fig. 19. Schematic diagram of the experiment.

15
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

5.3. Experimental process According to Eq. (41), the relative uncertainty of (UNu/Nu) can be ob­
tained (see Appendix).
Fig. 19 shows the experimental schematic. The whole experimental Based on Eq. (A1) - (A9) in the appendix and the parameters in
test system can be divided into: circulating cooling, data acquisition and Table 5, the relative uncertainty of Nu is 1.4%.
heating systems. The circulating cooling system is composed of ther­
mostatic water bath box (BAIHUI-HH420), digital peristaltic pump
(MASTERFLEX GY7792175), allocation model and silicone hose. The 5.5. Result analysis
thermostatic water bath box ensures the temperature of the fluid at the
inlet of the heat sink constant and keeps at 293 K, and the flow rate of the Table 7 lists the comparison of simulation and experimental data at
fluid at the inlet is adjusted by the digital peristaltic pump. The splitter different flow velocities for D11. It is worth noting that the relative error
on the left distributes the fluid output from the pump evenly into the of ΔP is within 8.70%, and the maximum relative error of Nusselt
heat sink through each inlet, while the splitter on the right collects the number does not exceed 5.38%, which is within a reasonable range.
fluid output back into the thermostatic water bath. The silicone hose is Fig. 20(a) shows the comparison between the simulated and exper­
used to build the cooling circulation circuit. Data acquisition part in­ imental pressure drop. With the increase of flow rate, the numerical and
cludes thermometer (Agilent34972A), K-type thermocouple and digital experimental pressure drop are in good agreement, which indicates that
pressure gage (Commark C9551). The thermometer is used to record the correctness of the numerical results is verified. Fig. 20(b) is the com­
temperature data collected by thermocouples. There are 14 temperature parison of the numerical Nusselt number and the experimental Nusselt
points to be tested in this experiment. Eight small circular holes are number. It is obvious that the numerical Nusselt number agree well with
machined on the side wall of the DL-MHS to measure the temperature of the experimental Nusselt number, which verifies the correctness nu­
the upper and lower channel walls. See Fig. 18(a) for details. The merical thermal characteristic of D11. In summary, the simulation re­
remaining 6 points need to be tested are the temperature of the fluid at sults are reliable.
each outlet of the upper and lower channels for the DL-MHS. The digital
pressure gage is for measuring the pressure drop at the inlet and outlet of 6. Conclusions
the DL-MHS. The heating system consists of the DC power supply
(Tektronix/2230 G-30–6) and eight film resistors (LXP100 TO-247). In In this work, based on the variable density topology optimization
this experiment, the film resistors are connected in series with the DC method, the optimal 2D channel structures with two bi-objective fuc­
power supply. The resistance value of the film resistor is 10 Ω, the rated tions are obtained. Then, the two 2D topological structures are stretched
power is 100 W, and the section area is 16 mm × 20.96 mm, therefore, to form 3D SL-MHSs. Subsequently, these two topological SL-MHSs are
the available heat flux range is 0~30 W/cm2. To minimize contact then assembled by self-stacking and interactive stacking to obtain four
thermal resistance, we use the thermal conductive adhesive with good topological DL-MHSs D11, D12, D21, and D22. The flow and thermal
thermal conductivity to mount the film resistor on the bottom plate of characteristics of these topological DL-MHSs and traditional straight DL-
the DL-MHS. MHS C1 are studied by numerical simulation. Finally, the flow and
thermal performance of D11 is verified experimentally. Several con­
clusions can be drawn:
5.4. Uncertainty analysis
• The heat transfer performance of the topological DL-MHSs is signif­
There are mainly two reasons of uncertainty in the experiment, one is icantly better than that of the traditional straight DL-MHS C1. The
caused by the accuracy of the experimental instrument itself, and the maximum bottom surface temperature of D11 can be reduced by 17
other is caused by the machining accuracy of the model. Table 6 shows K compared to C1.
the uncertainty of the experimental parameters. In this work, according • When uin = 0.05 m/s, the growth of pressure drops of D11, D12, D21
to the method of Coleman et al. [52] and ASME [53], the uncertainty can and D22 are 18.20%, 27.90%, 27.90% and 37.26% compared to C1,
be determined by: respectively. This means that the topological DL-MHSs improve the
[ )2 ]12 thermal performance at the expense of increased pressure drop.
θ (

UR =
∂R
U Vi (41) • Compared with C1, the average Nusselt numbers for D11, D12, D21
i=1
∂Vi and D22 can be increased by 27.3%, 26.6%, 27.1% and 26.4%,
respectively. Furthermore, the thermal resistances can be decreased
where UR is the total absolute uncertainty, Vi is the type of parameter, by 27.6%, 26.9%, 22.7% and 22.0%, respectively.
UVi is the absolute error of a single parameter, and θ is the number of • The PEC values for four topological DL-MHSs are greater than 1,
parameters. which indicates that the DL-MHSs designed by topological optimi­
From Eq. (29) and (30), it can be deduced that: zation is meaningful. Among them, the PEC value of D11 is the
Aq D Qw Dh highest, which indicates that its comprehensive hydrothermal per­
Nu = ( b w h ) = (42) formance of D11 is the best.
Acon Twavg − Tfavg kf Acon ΔTkf
• For pressure drop and Nusselt number, the maximum error between
Therefore, the uncertainty of Nu is related to the uncertainty of Qw, the experimental and numerical results for D11 are 8.70% and
Dh, Acon, kf and ΔT (temperature difference between the average tem­ 5.78%, which adequately verifies the accuracy of the numerical
perature of the channel wall and the average temperature of the fluid). results.

Table 6 Table 7
Uncertainty of experimental data. . Comparison of numerical and experimental data.
Parameters Absolute Uncertainty Relative Uncertainty uin (m/s) 0.02 0.03 0.04 0.05 0.06 0.07

w (m) ±0.05mm – ΔP(sim) (Pa) 1.84 3.17 4.85 6.82 9.09 11.61
h (m) ±0.05mm – ΔP(exp) (Pa) 2.00 3.33 5.24 7.33 9.67 12.40
T (K) ±0.2K – ΔP error 8.70% 5.05% 8.04% 7.48% 6.38% 6.80%
ΔP (Pa) – ±0.3% Nu(sim) 6.00 7.81 9.41 10.83 12.09 13.21
Qv (ml/s) – ±0.1% Nu(exp) 6.24 8.21 9.79 11.37 12.74 13.80
Qw (W) – ±0.2% Nu error 4.00% 5.12% 4.04% 4.99% 5.38% 4.47%

16
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

Fig. 20. Comparison of numerical results with experimental results for (a) ΔP; (b) Nu.

CRediT authorship contribution statement interests or personal relationships that could have appeared to influence
the work reported in this paper.
Xing-ping Tang: Methodology, Software, Investigation, Writing –
original draft. Huan-ling Liu: Methodology, Supervision, Writing – re­ Acknowledgments
view & editing. Li-si Wei: Data curation. Chuan-geng Tang: Software.
Xiao-dong Shao: Resources. Han Shen: Investigation. Gongnan Xie: This work is supported by the General Program of National Natural
Supervision, Writing – review & editing. Science Foundation of China (52275160). The last author (Dr. Xie)
wishes to thank the support of Shenzhen Science and Technology Pro­
Declaration of Competing Interest gram (JCYJ20210324142812032).

The authors declare that they have no known competing financial

Appendix

Appendix
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ( )2 ( )2 ( )2 ( )̅
UNu U Qw UDh UAcon UΔT Ukf 2
= + + − + − + − (A1)
Nu Qw Dh Acon ΔT kf
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ( )2̅
UΔT = UTwavg + UTfavg (A2)
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ( )2̅
UTwavg = 0.5 UTwavg,l + UTwavg,u (A3)
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ( )2 ( )2 ( )2̅
UTwavg,l = 0.25 UTwtc1,l + UTwtc2,l + UTwtc3,l + UTwtc4,l (A4)
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ( )2 ( )2 ( )2̅
UTwavg,u = 0.25 UTwtc1,u + UTwtc2,u + UTwtc3,u + UTwtc4,u (A5)
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ( )2̅
UTfavg = 0.5 UTfavg,l + UTfavg,u (A6)
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ( )2
UTfavg,l = 0.5 UTin,l + UTout,l (A7)
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ( )2
UTfavg,u = 0.5 UTin,u + UTout,u (A8)
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
)2̅
√( )2 (
√ 2h − 1 2h − 1
U Dh =√ Uw + 2 whUw + Uh + 2 whUh (A9)
w+h (w + h)2 w+h (w + h)2

17
X.-p. Tang et al. International Journal of Heat and Mass Transfer 215 (2023) 124513

References [26] A. Fawaz, Y. Hua, S. Le Corre, Y. Fan, L. Luo, Topology optimization of heat
exchangers: a review, Energy 252 (2022), 124053.
[27] J. Zhou, M. Lu, Q. Zhao, D. Hu, H. Qin, X. Chen, Thermal design of microchannel
[1] K. Lim, J. Lee, Experimental study on single-phase convective heat transfer of
heat sinks using a contour extraction based on topology optimization (CEBTO)
interlocking double-layer counterflow mini-channel heat sink, Energy Convers.
method, Int. J. Heat Mass Transf. 189 (2022), 122703.
Manag. 243 (2021), 114415.
[28] L. Xie, Y. Zhang, M. Ge, Y. Zhao, Topology optimization of heat sink based on
[2] A.F. A.l-neama, Z. Kapur, et al., An experimental and numerical investigation of
variable density method, Energy Rep. 8 (2022) 718–726.
chevron fin structures in serpentine minichannel heat sinks, Int. J. Heat Mass
[29] T. Zhou, C. Guo, X. Shao, A novel mini-channel heat sink design with arc-type
Transf. 120 (2018) 1213–1228.
design domain by topology optimization, Micromachines 13 (2) (2022) 180
[3] P. Smakulski, S. Pietrowicz, A review of the capabilities of high heat flux removal
(Basel).
by porous materials, microchannels and spray cooling techniques, Appl. Therm.
[30] K. Kulkarni, A. Afzal, K.Y. Kim, Multi-objective optimization of a double-layered
Eng. 104 (2016) 636–646.
microchannel heat sink with temperature-dependent fluid properties, Appl. Therm.
[4] A.Ü. Tepe, Y. Yetişken, Ü. Uysal, K. Arslan, Experimental and numerical
Eng. 99 (2016) 262–272.
investigation of jet impingement cooling using extended jet holes, Int. J. Heat Mass
[31] Y. Zhai, G. Xia, Z. Li, H. Wang, A novel flow arrangement of staggered flow in
Transf. 158 (2020), 119945.
double-layered microchannel heat sinks for microelectronic cooling, Int. Commun.
[5] G. Lu, J. Zhao, L. Lin, X.D. Wang, W.M. Yan, A new scheme for reducing pressure
Heat Mass Transf. 79 (2016) 98–104.
drop and thermal resistance simultaneously in microchannel heat sinks with wavy
[32] H. Shen, Y. Zhang, C.C. Wang, G. Xie, Comparative study for convective heat
porous fins, Int. J. Heat Mass Transf. 111 (2017) 1071–1078.
transfer of counter-flow wavy double-layer microchannel heat sinks in staggered
[6] D.B. Tuckerman, R.F.W. Pease, High-performance heat sinking for VLSI, IEEE
arrangement, Appl. Therm. Eng. 137 (2018) 228–237.
Electron Device Lett. 2 (5) (1981) 126–129.
[33] S.L. Wang, X.Y. Li, X.D. Wang, G. Lu, Flow and heat transfer characteristics in
[7] T.C. Hung, W.M. Yan, Effects of tapered-channel design on thermal performance of
double-layered microchannel heat sinks with porous fins, Int. Commun. Heat Mass
microchannel heat sink, Int. Commun. Heat Mass Transf. 39 (9) (2012) 1342–1347.
Transf. 93 (2018) 41–47.
[8] L. Lin, J. Zhao, G. Lu, X.D. Wang, W.M. Yan, Heat transfer enhancement in
[34] T. Borrvall, J. Petersson, Topology optimization of fluid in Stokes flow, Int. J.
microchannel heat sink by wavy channel with changing wavelength/amplitude,
Numer. Meth. Fluids 41 (2003) 77–107.
Int. J. Therm. Sci. 118 (2017) 423–434.
[35] D.J. Lohan, E.M. Dede, J.T. Allison, Topology optimization for heat conduction
[9] S. Nawaz, H. Babar, H.M. Ali, M.U. Sajid, M.M. Janjua, Z. Said, A.K. Tiwari, L.
using generative design algorithms, Struct. Multidiscip. Optim. 55 (2017)
S. Sundar, C. Li, Oriented square shaped pin-fin heat sink: performance evaluation
1063–1077.
employing mixture based on ethylene glycol/water graphene oxide nanofluid,
[36] M. Yu, S. Ruan, J. Gu, M. Ren, Z. Li, X. Wang, C. Shen, Three-dimensional topology
Appl. Therm. Eng. 206 (2022), 118085.
optimization of thermal-fluid-structural problems for cooling system design, Struct.
[10] H.L. Liu, Y.Q. Shao, Z.T. Chen, Z.L. Xie, Heat transfer and flow performance of a
Multidiscip. Optim. 62 (2020) 3347–3366.
novel T type heat sink with GaInSn coolant, Int. J. Therm. Sci. 144 (2019)
[37] X.H. Han, H.L. Liu, G. Xie, L. Sang, J. Zhou, Topology optimization for spider web
129–146.
heat sinks for electronic cooling, Appl. Therm. Eng. 195 (2021), 117154.
[11] A. Ghahremannezhad, K. Vafai, Thermal and hydraulic performance enhancement
[38] A.A. Koga, E.C.C. Lopes, H.F.V. Nova, C.R. De Lima, E.C.N. Silva, Development of
of microchannel heat sinks utilizing porous substrates, Int. J. Heat Mass Transf. 122
heat sink device by using topology optimization, Int. J. Heat Mass Transf. 64
(2018) 1313–1326.
(2013) 759–772.
[12] T.H. Wang, H.C. Wu, J.H. Meng, W.M. Yan, Optimization of a double-layered
[39] D. Hu, Z. Zhang, Q. Li, Numerical study on flow and heat transfer characteristics of
microchannel heat sink with semi-porous-ribs by multi-objective genetic
microchannel designed using topological optimizations method, Sci. China
algorithm, Int. J. Heat Mass Transf. 149 (2020), 119217.
Technol. Sci. 63 (1) (2020) 105–115.
[13] C. Leng, X.D. Wang, T.H. Wang, W.M. Yan, Optimization of thermal resistance and
[40] Y.S. See, J.Y. Ho, K.C. Leong, T.N. Wong, Experimental investigation of a topology-
bottom wall temperature uniformity for double-layered microchannel heat sink,
optimized phase change heat sink optimized for natural convection, Appl. Energy
Energy Convers. Manag. 93 (2015) 141–150.
314 (2022), 118984.
[14] K. Vafai, L. Zhu, Analysis of two-layered micro-channel heat sink concept in
[41] A. Zou, R. Chuan, F. Qian, W. Zhang, Q. Wang, C. Zhao, Topology optimization for
electronic cooling, Int. J. Heat Mass Transf. 42 (12) (1999) 2287–2297.
a water-cooled heat sink in micro-electronics based on Pareto frontier, Appl.
[15] K.C. Wong, F.N.A. Muezzin, Heat transfer of a parallel flow two-layered
Therm. Eng. 207 (2022), 118128.
microchannel heat sink, Int. Commun. Heat Mass Transf. 49 (2013) 136–140.
[42] T. Dbouk, A review about the engineering design of optimal heat transfer systems
[16] C. Leng, X.D. Wang, T.H. Wang, An improved design of double-layered
using topology optimization, Appl. Therm. Eng. 112 (2017) 841–854.
microchannel heat sink with truncated top channels, Appl. Therm. Eng. 79 (2015)
[43] M.P. Bendsoe, O. Sigmund, Topology optimization: theory, methods, and
54–62.
Applications, Springer Science & Business Media, 2003.
[17] S. Chamoli, R. Lu, H. Chen, Y. Cheng, P. Yu, Numerical optimization of design
[44] K.K. Choi, N.H. Kim, Structural Sensitivity Analysis and Optimization 1: Linear
parameters for a modified double-layer microchannel heat sink, Int. J. Heat Mass
Systems, Springer Science & Business Media, 2004.
Transf. 138 (2019) 373–389.
[45] K. Svanberg, A class of globally convergent optimization methods based on
[18] A. Kumar, S. Nath, D. Bhanja, Effect of nanofluid on thermo hydraulic performance
conservative convex separable approximations, SIAM J. Optim. 12 (2) (2002)
of double layer tapered microchannel heat sink used for electronic chip cooling,
555–573.
Numer. Heat Transf. A Appl. 73 (7) (2018) 429–445.
[46] G. Allaire, F. Jouve, A.M. Toader, Structural optimization using sensitivity analysis
[19] K.C. Wong, M.L. Ang, Thermal hydraulic performance of a double-layer
and a level-set method, J. Comput. Phys. 194 (1) (2004) 363–393.
microchannel heat sink with channel contraction, Int. Commun. Heat Mass Transf.
[47] F. Wang, B.S. Lazarov, O. Sigmund, On projection methods, convergence and
81 (2017) 269–275.
robust formulations in topology optimization, Struct. Multidiscip. Optim. 43 (6)
[20] S. Sarvar-Ardeh, R. Rafee, S. Rashidi, Effects of convergence and
(2011) 767–784.
superhydrophobicity on the hydrothermal features of the tapered double-layer
[48] M. Zhou, Q. Li, X. Liang, Recent advances in topology optimization for additive
microchannel, Int. J. Therm. Sci. 181 (2022), 107745.
manufacturing: a review, Eng. Struct. 56 (2013) 1454–1471.
[21] A. Ghahremannezhad, H. Xu, M.A. Nazari, M.H. Ahmadi, K. Vafai, Effect of porous
[49] F. Chen, J. Wang, X. Yang, Topology optimization design and numerical analysis
substrates on thermohydraulic performance enhancement of double layer
on cold plates for lithium-ion battery thermal management, Int. J. Heat Mass
microchannel heat sinks, Int. J. Heat Mass Transf. 131 (2019) 52–63.
Transf. 183 (2022), 122087.
[22] K. Derakhshanpour, R. Kamali, M. Eslami, Improving performance of single and
[50] H. Dai, W. Chen, Numerical investigation of heat transfer in the double-layered
double-layered microchannel heat sinks by cylindrical ribs: a numerical
minichannel with microencapsulated phase change suspension, Int. Commun. Heat
investigation of geometric parameters, Int. Commun. Heat Mass Transf. 126
Mass Transf. 119 (2020), 104918.
(2021), 105440.
[51] K.C. Wong, F.N.A. Muezzin, Heat transfer of a parallel flow two-layered
[23] H. Dai, C. Zhu, Y. Liu, Thermal performance of double-layer porous-microchannel
microchannel heat sink, Int. Commun. Heat Mass Transf. 49 (2013) 136–140.
with phase change slurry, Appl. Therm. Eng. 211 (2022), 118457.
[52] H.W. Coleman, W.G. Steele, Experimentation, Validation, and Uncertainty Analysis
[24] A.S. El-dean, O. Hassan, H.M. Shafey, Heat transfer characteristics of two-phase
For Engineers, 3rd ed., John Wiley and Son. Inc., Hoboken, New Jersey, USA, 2009.
flow in a double-layer microchannel heat sink, Int. Commun. Heat Mass Transf.
[53] ASME. PTC 19.1-2013. (Revision of ASME PTC 19.1-2005). Test Uncertainty.
132 (2022), 105899.
[25] H. Shen, G. Xie, C.C. Wang, H. Liu, Experimental and numerical examinations of
thermofluids characteristics of double-layer microchannel heat sinks with
deflectors, Int. J. Heat Mass Transf. 182 (2022), 121961.

18

You might also like