You are on page 1of 15

International Journal of Advanced Research in Engineering Innovation

e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023


http://myjms.mohe.gov.my/index.php/ijarei

Numerical Investigation of the Hydrothermal Characteristics of


Water Flow in Compound Microchannel Heat Sinks
Sarab Salih Shiko1, Ahmed Fouad Al-Neama1*
1
Mechanical Engineering Department, Faculty of Engineering, University of Mosul, Mosul, Iraq

*Corresponding Author: ahmedfalneama@uomosul.edu.iq

Received: 15 March 2023 | Accepted: 1 May 2023 | Published: 1 June 2023

DOI: https://doi.org/10.55057/ijarei.2023.5.2.1
___________________________________________________________________________
Abstract: In the present work, three-dimensional numerical simulations of laminar forced
convection flow of water in unique compound microchannel heat sinks (MCHSs) were
investigated using computational fluid dynamics (CFD) modeling. The newly proposed MCHS
is made up of circular microchannels slotted from the top within a trapezoidal shape, and its
cooling effectiveness was compared to that of traditional rectangular MCHS. Each MCHS
under consideration has the same hydraulic diameter and heat transfer surface area. Water
volumetric flow rates (Q_in) with a wide range of values are used, ranging from 40 to 90
ml/min, with the fluid inlet temperature set to 20 oC. A constant heat flux boundary condition
of 100 W/cm2 is supplied on the MCHS bottom. The results demonstrated that the inclusion of
reentrant trapezoidal shapes can disrupt both hydrodynamic (δ_hy) and thermal boundary
layers (δ_th), as well as accelerate fluid flow and mixing in the main flow, resulting in a
significant heat transfer enhancement. Furthermore, at Q_in=90 ml/min, the average Nusselt
number (〖Nu〗_avg) of compound MCHS increased by 2.16%, while total pressure drop
(∆P) and total thermal resistance (R_th) decreased by 1.73% and 1.57%, respectively, when
compared to the straight rectangular counterpart.

Keywords: Compound MCHSs, Forced Convection, Conjugate Heat Transfer


__________________________________________________________________________

1. Introduction

Micro manufacturing techniques have advanced rapidly over the last three decades. These
developments have played a major role in improving microchip performance by cramming a
huge number of components into a smaller space. As a result, microchips are becoming more
powerful and synchronously compact. This increment in performance does have one drawback:
intense heat is generated during the process as the number of components increases and the
size of a microchip decreases. The miniaturization of integrated circuits (ICs) packages and
increase in power density has put forward an urgent demand to offer an alternative to the
traditional air-cooling approach, which has proven grossly inadequate to remove extreme high
heat flux [1]. To fulfill this demand, various technological methods for high heat flux removal
have been extensively investigated and attained greater attention, including micro-cooling cold
water. Because of the high heat transfer surface-area-to-volume ratio, micro-pin fin and
microchannel heat sinks (MCHSs) can be considered an excellent alternative to air-cooling
techniques [2]. They are capable of dissipating extremely high heat fluxes generated by
compact electronic devices.

1
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

Researchers have accomplished various studies to gain access to the forced convection
characteristics of MCHSs with traditional shapes, for instance, circular (micro-tubes),
rectangular, trapezoidal, or triangular ones. When liquid flows through straight microchannels,
both the thermal 𝛿𝑡ℎ and hydraulic boundary layers 𝛿𝑡ℎ grow simultaneously along the
microchannel length, and fully developed flow tends to be dominated inside the microchannels.
This may significantly reduce convective heat transfer performance. Thus, several flow channel
modification strategies for promoting flow instabilities and improving microchannel forced
convection performance have been proposed, including wavy or corrugated, zigzag, interrupted
or cross-linked, and reentrant shaped microchannels.

On flow disruption through microchannel, many heat transfer augmentation approaches have
been documented. These strategies encourage free stream separation at the leading edge, which
leads to boundary layer development and greater mixing, resulting in higher heat transfer rate.
Flow disruption can be obtained through passive surface modifications, such as channel shape,
ribs, cavities, groove structures, dimple surfaces, offset strip fins, and porous media. Extensive
researches are being undertaken in conventional channels utilizing various disturbance
promoters to enhance flow mixing, and Dewan and Srivastava [3] have presented a
comprehensive review of heat transfer improvement through microchannels.

The use of ribs and grooves or cavities in the MCHS can provide significant passive heat
transfer augmentation by reducing 𝛿𝑡ℎ or increasing both flow disruptions and velocity gradient
near the heated surface. Heat transfer augment in the MCHS was investigated by Xie et al. [4]
using different vertical crescent ribs protruding from the bottom wall. Straight ribs, as well as
crescent ribs that are concave and convex to the stream-wise direction, are all considered to
improve the cooling channel's thermal performance. Wang et al. [5] suggested a high-
performance MCHS with bidirectional ribs that are made up of vertical and spanwise ribs. Both
experimental and numerical studies were carried out at Reynolds numbers (Re) varied from
100 to 1000, and the results revealed that the microchannel with bidirectional ribs' averaged
Nusselt number (𝑁𝑢𝑎𝑣𝑔 ) is up to 1.4–2 and 1.2–1.42 times that of the microchannels with
vertical and spanwise ribs, respectively. This enhancement is due to the inclusion of
bidirectional ribs in MCHS, which effectively interrupts the 𝛿𝑡ℎ and induces circulation in both
the vertical and spanwise directions. Recently, a novel cross-rib MCHS has been presented
numerically by Chen et al. [6] to make fluid self-rotate. Compared with both the rectangular
and horizontal ribs MCHS, the cross-rib MCHS improved cooling capability by 28.6 % and
14.3 %, respectively, but at the expense of total pressure drop (∆P), which rose by 10.7-fold
and 5.5-fold, respectively.

Many studies have recently focused on curved microchannel research due to these geometries'
high thermal and flow performance. Flow passage curvature characteristics (such as wavy,
zigzag, and serpentine MCHSs) are favourable for microfluidic cooling applications. Because
of the efficient mixing in the curved channel, the curvature MCHS outperformed the straight
MCHS in terms of heat transfer rate. Al-Neama et al. [7,8] performed both experimental and
numerical study using water-cooled single-path serpentine rectangular passage MCHS to
dissipate a high heat flux generated by gallium nitride (GaN) high-electron-mobility transistor
(HEMT) devices. Chai et al. [9] investigated experimentally and numerically the ∆P and heat
transfer characteristics of two MCHSs with periodic expansion-constriction cross-sections,
triangular reentrant and fan-shaped cavities. In comparison to the straight rectangular MCHS,
the Nu_avg for the suggested periodic expansion-constriction cross-sections MCHSs with
triangular reentrant cavities increased by about 1.8 times with moderate ∆P, this was due to the
disruption of δ_th formation and the provision of additional surface area. A numerical and

2
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

experimental work is carried out by Deng et al. [10] to explore the forced convective heat
transfer of single-phase laminar flow on an innovative Ω-shaped reentrant copper MCHS. In
comparison to the conventional MCHS, the suggested design improved heat transfer by 10-
30%, reduced total thermal resistance (R_th) by 22%, and increased frictional factor ratios by
10% overall. This improved heat transfer is due to fluid mixing and flow separation caused by
throttling effects.

Previous studies have shown that using curved microchannels or adding cavities on the
microchannel sidewalls is a viable and effective method of improving forced convective heat
transfer by reinitializing both the 𝛿𝑡ℎ and 𝛿𝑡ℎ , while ∆P can invertibly increase. In this research,
a passive technique is utilized to augment the MCHS heat transfer performance, in which
copper microchannels with a novel trapezoidal-shaped reentrant cross-section configuration
are numerically simulated. To the best of the authors' knowledge, no research study has been
published that considers circular microchannels with slotted trapezoidal shapes, which has
essentially motivated the current work.

2. Problem formulation

2.1 Physical model


Both proposed MCHS are made up of 13 copper microchannels with a footprint area (W × L)
of 17mm × 20mm and a thickness (H) of 1.35mm. Furthermore, the copper substrate thickness
(𝐻𝑏 ) is fixed at 0.25mm. The conventional straight rectangular microchannel has been assigned
as the reference (represented as RMC), and the rectangular microchannel's detailed geometric
parameters are as follows: The height (𝐻𝑐ℎ ), width (𝑊𝑐ℎ ), and length (𝐿𝑐ℎ ) of the microchannels
are each 1.1mm, 0.6mm, and 20mm. The newly proposed compound microchannel is designed
to have circular cavities inside with a diameter of 0.8mm and an exit trapezoidal slot upside in
the cross section to create a reentrant microchannel (referred to as CMC). Both proposed
MCHSs are depicted in Fig. 1. To make a fair comparison, both microchannels have nearly the
same hydraulic diameter (𝐷ℎ ) and heat transfer surface areas (𝐴ℎ𝑡 ), as demonstrated in Table
1.

3
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

Figure 1: Schematic diagrams of computational domain: (a) conventional rectangular microchannel


(RMC); (b) compound microchannel (CMC). (All dimensions in mm).

Table 1: Specification of two MCHSs' geometric parameters.


𝐷𝑖𝑚𝑒𝑛𝑠𝑖𝑜𝑛
𝐴𝑐ℎ 𝐴ℎ𝑡 𝑃𝑐ℎ 𝐷ℎ
𝑀𝑜𝑑𝑒𝑙 𝑊 × 𝐿 × 𝐻 𝑁𝑐ℎ
𝑚𝑚2 𝑚𝑚2 𝑚𝑚 𝑚𝑚
(𝑚𝑚)
RMC 17 × 20 × 1.35 13 0.660 56 3.400 0.7765
CMC 17 × 20 × 1.35 13 0.664 56.24 3.392 0.7835

2.2 Simulation assumptions and governing equations


In this study, three-dimensional (3D) numerical simulation calculations of microchannel flow
and heat transfer characteristics were performed using CFD Fluent 2022 R1 software under the
following assumptions:
i. The fluid flow in the microchannel is characterized as 3D steady laminar flow with a single
phase;
ii. In this work, the fluid is water, which is an incompressible Newtonian fluid. Meanwhile,
the thermal physical properties (i.e., density, specific heat, viscosity and thermal
conductivity) of the water is temperature-dependent, while temperature independent is
considered for the MCHS material;
iii. Thermal radiation, volume force, surface tension, and gravity are not considered.
iv. The model's governing equations are simplified based on the assumptions stated above as
follows:

Continuity equation:
𝜕𝑢 𝜕𝑣 𝜕𝑤
+ + =0 (1)
𝜕𝑥 𝜕𝑦 𝜕𝑧
where 𝑢, 𝑣 and 𝑤 are the velocity components in 𝑥-, 𝑦- and 𝑧-directions, respectively.
Momentum equation:
𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜇𝑓 𝜕 2 𝑢 𝜕 2 𝑢 𝜕 2 𝑢 1 𝜕𝑝
𝑢 +𝑣 +𝑤 = ( 2 + 2 + 2) − (2a)
𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜌𝑓 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜌𝑓 𝜕𝑥

4
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜇𝑓 𝜕 2 𝑣 𝜕 2 𝑣 𝜕 2 𝑣 1 𝜕𝑝
𝑢 +𝑣 +𝑤 = ( 2 + 2 + 2) − (2b)
𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜌𝑓 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜌𝑓 𝜕𝑦
𝜕𝑤 𝜕𝑤 𝜕𝑤 𝜇𝑓 𝜕 2 𝑤 𝜕 2 𝑤 𝜕 2 𝑤 1 𝜕𝑝
𝑢 +𝑣 +𝑤 = ( 2+ 2
+ 2)− (2c)
𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜌𝑓 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜌𝑓 𝜕𝑧
where 𝜌𝑓 and 𝜇𝑓 are the coolant's density and dynamic viscosity, respectively, whilst 𝑝 is the
coolant's pressure.
Energy equation in the fluid domain:
𝜕𝑇𝑓 𝜕𝑇𝑓 𝜕𝑇𝑓 𝑘𝑓 𝜕 2 𝑇𝑓 𝜕 2 𝑇𝑓 𝜕 2 𝑇𝑓
𝑢 +𝑣 +𝑤 = ( + + ) (3)
𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜌𝑓 𝐶𝑝𝑓 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2
where 𝑘𝑓 , 𝐶𝑝𝑓 and 𝑇𝑓 denote the thermal conductivity, the specific heat, and the absolute
temperature of the coolant fluid, respectively.
Heat transport in solids is solely by conduction and satisfies:
𝜕 2 𝑇𝑠 𝜕 2 𝑇𝑠 𝜕 2 𝑇𝑠
0 = 𝑘𝑠 ( 2 + + ) (4)
𝜕𝑥 𝜕𝑦 2 𝜕𝑧 2
where 𝑇𝑠 is the solid temperature, and copper with a thermal conductivity (𝑘𝑠 ), specific heat
(𝐶𝑝𝑠 ) and density (𝜌𝑠 ) of 387.6 W/(m.K), 381 J/(kg.K) and 8978 kg/m3, respectively are
selected in the current numerical simulation.

2.3 Boundary conditions


1) At the microchannel's entrance region, uniform fluid velocity (𝑢𝑖𝑛 ) and constant fluid
temperature (𝑇𝑓,𝑖𝑛 ) are assigned;
𝐴𝑡 𝑥 = 0: 𝑢 = 𝑢𝑖𝑛 𝑎𝑛𝑑 𝑣 = 𝑤 = 0
𝐹𝑜𝑟 𝑓𝑙𝑢𝑖𝑑: 𝑇𝑓 = 𝑇𝑖𝑛 = 20 ℃ (293 𝐾)
2) At the microchannel outlet, a pressure outlet BC was specified;
𝐴𝑡 𝑥 = 𝐿: 𝑃𝑓 = 𝑃𝑜𝑢𝑡 = 1 𝑎𝑡𝑚 (𝑜𝑟 0 𝑔𝑎𝑢𝑔𝑒 𝑝𝑟𝑒𝑠𝑠𝑢𝑟𝑒)
𝜕𝑇𝑓
𝐹𝑜𝑟 𝑓𝑙𝑢𝑖𝑑: −𝑘𝑓 ( ) = 0
𝜕𝑥
3) The symmetry condition is assigned in the microchannel's two side walls;
𝜕𝑇𝑠
𝐴𝑡 𝑧 = 0 𝑎𝑛𝑑 𝑧 = 1.3𝑚𝑚: − 𝑘𝑠 ( ) = 0 (𝑠𝑦𝑚𝑚𝑒𝑡𝑟𝑦)
𝜕𝑧
4) The microchannel bottom surface is designated as the uniform heat flux boundary, while the
top surface is insulated;
𝜕𝑇𝑠 𝑊
𝐴𝑡 𝑦 = 0: 𝑢 = 𝑣 = 𝑤 = 0 , − 𝑘𝑠 ( ) = 𝑞 = 106 2
𝜕𝑦 𝑚
𝜕𝑇𝑠
𝐴𝑡 𝑦 = 1.35𝑚𝑚: 𝑢 = 𝑣 = 𝑤 = 0 , −𝑘𝑠 ( ) = 0
𝜕𝑦
5) In the coupled boundary condition, the fluid is coupled with solid walls, and on the fluid-
solid interface, the no-slip BC is assigned.
𝜕𝑇𝑠 𝜕𝑇𝑓
𝑢 =𝑣 =𝑤 =0, 𝑇𝑠 = 𝑇𝑓 , −𝑘𝑠 ( ) = −𝑘𝑓 ( )
𝜕𝑛 𝜕𝑛
where 𝑛 is the coordinate normal to the wall. The coupled algorithm (𝑝 − 𝑣 coupling) is
employed to solve the governing equations with the finite volume-based CFD solver FLUENT.
For the pressure equation, a second-order discretization scheme is used, while the momentum
and energy equations are discretized using a second-order upwind scheme. The residual criteria
of the continuity equation and velocity are set to 10-6, and that of the energy equation is 10-9.

5
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

2.4 Water temperature dependent thermophysical properties


In the current work, special correlations presented by Abad et al. [11] are employed to
calculate water thermophysical properties as temperature changes. For the following
correlations, user-defined function (UDF) files were created in C++ to account for the
temperature-dependence of the thermophysical properties of water, which were then
interpreted in the ANSYS-fluent software.
(𝑇 − 4)2
𝜌𝑤 = 1000 [1 − ] (5)
119000 + 1365𝑇 − 4𝑇 2
𝐶𝑝𝑤 = 4217.629 − 3.20888𝑇 + 0.09503𝑇 2 − 0.00132𝑇 3 + 9.415 × 10−6 𝑇 4
− 2.5479 × 10−8 𝑇 5 (6)
𝑘𝑤 = 0.56112 + 0.00193𝑇 − 2.60152749 × 10−6 𝑇 2 − 6.08803 × 10−8 𝑇 3 (7)
𝜇𝑤 = 0.00169 − 4.25263 × 10−5 𝑇 + 4.9255 × 10−7 𝑇 2 − 2.09935 × 10−9 𝑇 3 (8)
where 𝜌𝑤 , 𝐶𝑝𝑤 , 𝑘𝑤 and 𝜇𝑤 are density, specific heat, thermal conductivity and dynamic
viscosity of the water, respectively. The unit of 𝑇 is K in Eqs. (5 – 8).

2.5 Data Reduction


The inlet Reynolds number (𝑅𝑒) is defined as:
𝜌𝑤 𝑢𝑖𝑛 𝐷ℎ
𝑅𝑒 = (9)
𝜇𝑤
where 𝑢𝑖𝑛 is the microchannel inlet velocity while 𝐷ℎ denotes the microchannel hydraulic
diameter based on the microchannel section, and can be expressed as:
4𝐴𝑐ℎ 2(𝑊𝑐ℎ 𝐻𝑐ℎ )
𝐷ℎ = = (10)
𝑃𝑐ℎ 𝑊𝑐ℎ + 𝐻𝑐ℎ
where 𝐴𝑐ℎ and 𝑃𝑐ℎ denote the cross-sectional area and wetted perimeter of the microchannel,
respectively.
To account for the friction and the developing region effects, the pressure drop (∆𝑃) equations
can be expressed in terms of an apparent friction factor (𝑓𝑎𝑝𝑝 ). It is the friction factor's average
value over the flow length between the entrance section and the location under consideration.
2
2𝑓𝑎𝑝𝑝 𝜌𝑤 𝐿𝑢𝑚
∆𝑃 = (11)
𝐷ℎ
where 𝐿 is the microchannel length and 𝑢𝑚 is the mean flow velocity in the microchannel.
The average convective heat transfer coefficient (ℎ𝑎𝑣𝑔 ) is expressed as:
𝑞𝐴𝑏
ℎ𝑎𝑣𝑔 = (12)
𝐴𝑐𝑜𝑛 (𝑇𝑤,𝑎𝑣𝑔 − 𝑇𝑓,𝑎𝑣𝑔 )
where 𝑞 is the constant heat flux per unit surface, 𝐴𝑏 is the microchannel bottom wall area,
𝑇𝑤,𝑎𝑣𝑔 and 𝑇𝑓,𝑎𝑣𝑔 are the average bottom wall and fluid temperatures, respectively, while 𝐴𝑐𝑜𝑛
is the convection heat transfer area.
The average Nusselt number (𝑁𝑢𝑎𝑣𝑔 ) of the MCHS is written as:
ℎ𝑎𝑣𝑔 𝐷ℎ
𝑁𝑢𝑎𝑣𝑔 = (13)
𝑘𝑓

6
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

The thermal enhancement factor (𝜂) was introduced to assess the overall performance of the
compound microchannel by evaluating the heat transfer enhancement versus the ∆𝑃 increment
for single-phase laminar flow given by [12]:
𝑁𝑢/𝑁𝑢𝑜
𝜂= 3
(14)
√∆𝑃/∆𝑃𝑜
The baseline Nusselt number (𝑁𝑢𝑜 ) and total pressure drop (∆𝑃𝑜 ) are obtained from the cases
using conventional microchannels (which is RMC in the present study). The larger 𝜂, the better
the microchannel's overall performance.

2.6 Grid Independence test and CFD Simulations


A structured mesh for the solver was generated using the ANSYS software 2022R1. For every
MCHS, a grid independence test is carried out with various mesh sizes. The 𝑁𝑢𝑎𝑣𝑔 was chosen
as the grid independent verification parameter. Meantime, the relative error (𝑒) between the
finest grids (𝐽1 ) and other grids (𝐽2 ) can be expressed as [8]:
𝐽2 − 𝐽1
𝑒% = | | × 100 (15)
𝐽1
Five different grid quantities were chosen for verification after gradually refining the meshes,
and the results are shown in Table 2. When the mesh count is 1,666,166 elements, the 𝑒% is
less than 1%. To reduce computation time, the following simulation calculation used 1,666,166
element grids for the CMC model.

Table 2: Mesh independence test at 𝑸𝒊𝒏 = 𝟔𝟎 𝒎𝒍/𝒎𝒊𝒏 (𝒐𝒓 𝟏 𝒄𝒎𝟑 /𝒔𝒆𝒄).

3. Results

3.1 Verification of Numerical Models


To validate the simulation execution, the ∆𝑃 and 𝑁𝑢 results obtained from a straight
rectangular MCHS were compared and verified versus the data taken away by the available
correlation equations in references [13, 14]. The parameters length (𝐿), width (𝑊), height (𝐻),
microchannel height (𝐻𝑐ℎ ), microchannel width (𝑊𝑐ℎ ), fin width (𝑊𝑓 ), and substrate thickness
(𝐻𝑏 ) of the conventional microchannel used for validation are 10mm, 0.4mm, 0.55mm, 0.4mm,
and 0.2mm, 0.1mm, and 0.15mm, respectively. A uniform heat flux of 106 𝑊 ⁄𝑚2 is applied
to the microchannel bottom wall, which is a common value for electronics with high power
and heat flux. Fluid inlet temperature (𝑇𝑓,𝑖𝑛 ) was set at 20 ℃. The heat generated in electronic
components is captured by the silicon MCHS's bottom surface and transferred to the coolant
fluid. Table 3 lists both the water and silicon thermos-physical properties.

7
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

Table 3: Water and silicon thermo-physical properties.


Material 𝜌 (𝑘𝑔⁄𝑚3 ) 𝐶𝑝 (𝐽⁄𝑘𝑔. 𝐾 ) 𝑘 (𝑊 ⁄𝑚. 𝐾 ) 𝜇 (𝑘𝑔⁄𝑚. 𝑠)
Water Eq. (5) Eq. (6) Eq. (7) Eq. (8)
Silicon 2329 712 148 ̶̶̶̶

In a laminar developing flow zone, ∆𝑃 in the traditional microchannel (RMC) can be expressed
by the incremental pressure drop as:
2(𝑓𝑎𝑝𝑝 𝑅𝑒)𝜇𝑤 𝑢𝑚 𝐿 2(𝑓𝑅𝑒)𝜇𝑤 𝑢𝑚 𝐿 2
𝜌𝑤 𝑢𝑚
∆𝑃 = = + 𝐾(∞) (16)
𝐷ℎ2 𝐷ℎ2 2
The Poiseuille number (𝑃𝑜 = 𝑓𝑅𝑒) is known to be constant for hydrodynamically fully
developed laminar flow and is only dependent on the flow-channel geometry. According to
Shah and London's rectangular channel model [15], 𝑃𝑜 can be expressed as:
𝑃𝑜 = 24(1 − 1.3553𝛼 + 1.9467𝛼 2 − 1.7012𝛼 3 + 0.9564𝛼 4 − 0.2537𝛼 5 ) (17)
where 𝑓 is the Fanning friction factor while α represents the microchannel aspect ratio (𝛼 =
𝑊𝑐ℎ ⁄𝐻𝑐ℎ ). When the laminar flow is fully developed, the Hagenbach factor in Eq. (16) stays
constant. Steinke and Kandlikar [13] proposed the following curve-fit and 𝛼 dependent
equation for the Hagenbach's factor in rectangular channels:
𝑘(∞) = 0.6796 + 1.2197𝛼 + 3.3089𝛼 2 − 9.5921𝛼 3 + 8.9089𝛼 4 − 2.9959𝛼 5 (18)
In addition, the local Nusselt number (𝑁𝑢𝑥 ) variation for thermal entry region is also validated
with the correlation proposed by Kandlikar et al. [14] as:
𝑁𝑢𝑓𝑑,3
𝑁𝑢𝑥,3 = 𝑁𝑢𝑥,4 ( ) (19)
𝑁𝑢𝑓𝑑,4
In Eq. (19), 𝑁𝑢𝑥,3 and 𝑁𝑢𝑥,4 stand for the Nusselt number of laminar flows at a distance 𝑥 from
the entrance region for the three- and four-sided heating configurations, respectively.
Furthermore, 𝑁𝑢𝑓𝑑,3 and 𝑁𝑢𝑓𝑑,4 denote the Nusselt number of fully developed laminar flows
for three- and four-sided heating configurations, respectively. For the rectangular microchannel
with 𝛼 of 0.5, the value of the 𝑁𝑢𝑓𝑑,3, 𝑁𝑢𝑓𝑑,4 and 𝑁𝑢𝑥,4 can be estimated as follows [14]:
8.2321 + 1.2771𝛼 + 2.2389𝛼 2
𝑁𝑢𝑓𝑑,3 = (20)
1 + 2.0263𝛼 + 0.2981𝛼 2 + 0.0065𝛼 3
8.2313 − 2.295𝛼 + 7.928𝛼 2
𝑁𝑢𝑓𝑑,4 = (21)
1 + 1.939𝛼 + 0.938𝛼 2 + 0.0034𝛼 3
2
28.315 + 27038𝑥 ∗ + 1783300𝑥 ∗
𝑁𝑢𝑥,4 = 2 3 (22)
1 + 3049𝑥 ∗ + 472520𝑥 ∗ − 35714𝑥 ∗
𝑥
𝑥∗ = (23)
𝑅𝑒𝑃𝑟𝐷ℎ
In the above equations, 𝑃𝑟 is the Prandtl number while 𝑥 and 𝑥 ∗ represent distance and
dimensionless distance along microchannel length, respectively. Using Eqs. (20 and 21), the
values of the 𝑁𝑢𝑓𝑑,3 and 𝑁𝑢𝑓𝑑,4 are found to be 4.505 and 4.111, respectively. The numerical
simulation results of 𝑓𝑎𝑝𝑝 and 𝑁𝑢𝑥,3 for the traditional MCHS are compared with correlations
of Eqs. (16 and 19), respectively. As shown in Fig. 2, the numerical results and theoretical
correlations agree well, with a maximum deviation of less than 3%. As a result, the current
numerical method is now being used to assess the thermal-hydrodynamic performance of
compound microchannels.

8
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

Figure 2: A comparison of simulation results and theoretical correlation predictions for: (a) 𝒇𝒂𝒑𝒑 ; (b)
𝑵𝒖𝒙,𝟑 .

3.2 Effect of MCHS geometry on pressure drop and fluid flow


Fig. 3 illustrates the influence of water flow rates (40 ≤ 𝑄𝑖𝑛 ≤ 90 𝑚𝑙/𝑚𝑖𝑛) on the total
pressure drop (∆𝑃) for two different configurations of MCHS designs. As can be seen, ∆𝑃
increases rapidly as 𝑄𝑖𝑛 increases for two MCHS models. It is worth noting that the newly
proposed MCHS (CMC) has the lowest ∆𝑃 when compared to the traditional MCHS (RCM),
which may be attributed to the reentrant trapezoidal-shaped slot mounted on the top of the
circular cavity. At the highest volumetric flowrate (𝑄𝑖𝑛 = 90 𝑚𝑙/𝑚𝑖𝑛), ∆𝑃 in CMC design
decreased by 1.73% compared to the RCM model.

Figure 3: Effect of 𝑸𝒊𝒏 on the ∆𝑷 for two different MCHS configurations.

The pressure drop contours for the two MCHS models using water as coolant are shown in Fig.
4. For each MCHS case, six cross-sectional planes were presented at x = 0, 1, 2, 3, 4, and 5mm
from the entrance region. These contours are taken with Q_in=80 ml/min, fluid inlet
temperature of 20 oC, and an input power (Q) of 26 W (q=100 W/cm^2) applied on the MCHS
bottom side. As can be seen, pressure decreases along the flow direction for both MCHSs, but
pressure increases locally along the flow direction in the reentrant part. It is clearly seen that
CMC design create a slightly lower ∆P than the traditional MCHS.

9
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

Figure 4: Comparison of pressure drop contours for two different geometric structures on the 𝒚 − 𝒛 plane
(𝒙 = 𝟎, 𝟏, 𝟐, 𝟑, 𝟒, 𝒂𝒏𝒅 𝟓𝒎𝒎) at the entrance region at 𝑸𝒊𝒏 = 𝟖𝟎 𝒎𝒍/𝒎𝒊𝒏.

Fig. 5 depicts the velocity distribution on the 𝑥 − 𝑧 plane (𝑦 = 0.65𝑚𝑚) in the longitudinal
direction of two different configurations of microchannels at 𝑄𝑖𝑛 = 80 𝑚𝑙/𝑚𝑖𝑛. As soon as
the fluid enters the microchannels, stable velocity streamlines form quickly; the 𝛿ℎ𝑦 is then
fully developed and maintained along the flow direction. In contrast, as apparent in Fig. 5,
compound microchannels exhibit a gradual decrease in velocity streamline from the top to the
center, indicating the continuous development of the 𝛿ℎ𝑦 . The stable velocity contour had not
been reached at the microchannel exit region, and the 𝛿ℎ𝑦 was still growing. Because the
developing region induces much more intense fluid mixing than the fully developed region,
compound microchannels should perform better in terms of heat transfer.

Figure 5: Comparison of 2D velocity contours for two different geometric structures on the 𝒙 − 𝒛 plane
(𝒙 = 𝟎 − 𝟐𝟎𝒎𝒎) at mid-width plane (𝒚 = 𝟎. 𝟔𝟓𝒎𝒎) at 𝑸𝒊𝒏 = 𝟖𝟎 𝒎𝒍/𝒎𝒊𝒏.

10
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

3.3 Effect of MCHS geometry on heat transfer


Fig. 6 demonstrates how water flow rates (40 ≤ 𝑄𝑖𝑛 ≤ 90 𝑚𝑙/𝑚𝑖𝑛) affect maximum wall
temperature (𝑇𝑚𝑎𝑥 ) for two different MCHS model configurations. As can be seen, when 𝑄𝑖𝑛
is increased, 𝑇𝑚𝑎𝑥 decrease monotonically, and using CMC model has superior effect on
decreasing the microchannel base temperature to which the IC is adhered. It is obvious from
Fig. 6 that CMC design has the lowest 𝑇𝑚𝑎𝑥 than the traditional MCHS. At the lowest
volumetric flowrate (𝑄𝑖𝑛 = 40 𝑚𝑙/𝑚𝑖𝑛), 𝑇𝑚𝑎𝑥 in CMC (89.77 oC) decreased by 1.23%
compared to RMC (90.88 oC).

Figure 6: Effect of 𝑸𝒊𝒏 on the maximum wall temperatures for two different MCHS configurations.

Fig. 7 depicts the temperature profiles at 𝑦 = 0.65𝑚𝑚 from the MCHS base of two MCHSs
along the flow streamwise (𝑥 = 0 − 20𝑚𝑚) at 𝑄𝑖𝑛 = 80 𝑚𝑙/𝑚𝑖𝑛, 𝑇𝑓,𝑖𝑛 = 20℃, and 𝑞 =
100 𝑊 ⁄𝑐𝑚2 . For both microchannels, it can be claimed that the temperature gradient between
wall/fluid interface and core fluid increased as the flow moved downstream. Notwithstanding,
the distribution of side wall temperature along the flow length of compound MCHSs was lower
than that of the straight rectangular MCHS, indicating that the temperature of the compound
microchannels' side walls in the simulations was about 0.6–2.56 oC lower than the RMC model
at the same 𝑥 location.
Flow direction

RMC

CMC

Figure 7: Comparison of 2D temperature contours for two different geometric structures on the 𝒙 − 𝒛
plane (𝒙 = 𝟎 − 𝟐𝟎𝒎𝒎) at mid-width plane (𝒚 = 𝟎. 𝟔𝟓𝒎𝒎) at 𝑸𝒊𝒏 = 𝟖𝟎 𝒎𝒍/𝒎𝒊𝒏.

11
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

The comparison of the 3-D temperature distributions of compound (CMC) and traditional
(RMC) MCHSs reveals that compound microchannels have lower wall temperatures, as shown
in Fig 8. For two different microchannel configurations, the six cross-sectional planes at the
longitudinal flow's second half region on the 𝑦 − 𝑧 plane (𝑥 = 10, 12, 14, 16, 18 and 20mm)
from the entrance region at 𝑄𝑖𝑛 = 80 𝑚𝑙/𝑚𝑖𝑛 were presented here. From 𝑥 = 10mm to 𝑥 =
20mm, it is obvious that the compound microchannels had lighter and smaller color variations,
indicating a lower value of wall temperature rise along the flow direction. Furthermore, the
liquid temperature profile inside compound MCHSs was affected by the unique reentrant
configurations of the microchannels.

Figure 8: Comparison of 3D temperature distribution for two different geometric structures on the 𝒚 − 𝒛
plane (𝒙 = 𝟏𝟎, 𝟏𝟐, 𝟏𝟒, 𝟏𝟔, 𝟏𝟖 𝒂𝒏𝒅 𝟐𝟎𝒎𝒎) from the entrance region at 𝑸𝒊𝒏 = 𝟖𝟎 𝒎𝒍/𝒎𝒊𝒏.

Fig. 9 portrays temperature distribution at the outlet's cross-section plane for two MCHS
models at 𝑄𝑖𝑛 = 80 𝑚𝑙/𝑚𝑖𝑛. In contrast to the straight rectangular microchannel (RMC), the
protrusion of compound microchannel disrupted the continuous temperature contours of fluid
in CMC design. The abrupt protrusion at the intersection of the circular cavity and the upside
trapezoidal slot disrupted the normal development of the 𝛿𝑡ℎ from bottom to top in the cross
section of the flow passages, which resulted in two segments of both hydrodynamic and thermal
boundary zones. The flow separation caused by the trapezoidal slot resulted in a thinner 𝛿𝑡ℎ of
the main flow in the circular cavity, particularly at the transition zone from the circular cavity
to the slot. The disruption of the 𝛿𝑡ℎ , the fluid flow acceleration and mixing in the main flow,
caused a significant increase in ℎ for compound microchannels. As a result, the compound
MCHS dissipated more heat from its substrate while maintaining lower wall temperatures.

Figure 9: Comparison of 2D temperature contours for different geometric structures on the 𝒚 − 𝒛 plane
(𝒙 = 𝟐𝟎𝒎𝒎) at the exit region at 𝑸𝒊𝒏 = 𝟖𝟎 𝒎𝒍/𝒎𝒊𝒏.

12
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

Fig. 10 elucidates the impact of different MCHS configurations on the 𝑁𝑢𝑎𝑣𝑔 over a wide
volumetric flow rate range (40 ≤ 𝑄𝑖𝑛 ≤ 90 𝑚𝑙/𝑚𝑖𝑛). As can be seen, that increasing 𝑄𝑖𝑛 has
a superior effect on enhancing 𝑁𝑢𝑎𝑣𝑔 . It has been stated that as flow velocity increases, so does
turbulent intensity and consequently flow mixing, resulting in improved heat transfer
performance. The CMC design clearly has a higher 𝑁𝑢𝑎𝑣𝑔 than the traditional MCHS. This is
due to the reentrant microchannel's effect, which resulted in the reinitialization of thermal
boundary layers and improved fluid mixing, resulting in higher heat transfer. At 𝑄𝑖𝑛 =
90 𝑚𝑙/𝑚𝑖𝑛, it is observed that the 𝑁𝑢𝑎𝑣𝑔 for the CMC model improved by 2.16% compared
to the RMC one.

Figure 10: Effect of 𝑸𝒊𝒏 on the 𝑵𝒖𝒂𝒗𝒈 for two different MCHS configurations.

3.4 Thermal resistance


MCHS's thermal resistance (𝑅𝑡ℎ ) can be employed to evaluate its cooling performance as [8]:
𝑇𝑚𝑎𝑥 − 𝑇𝑓,𝑖𝑛
𝑅𝑡ℎ = (24)
𝑞
As shown in Fig. 11, the 𝑅𝑡ℎ of the CMC model can be reduced because the maximum wall
temperature of the CMC is lower than that of the RMC design. This is due to the presence of
protrusion, which results in a higher heat transfer rate and, as a result, a lower wall temperature.
At 𝑄𝑖𝑛 = 90 𝑚𝑙/𝑚𝑖𝑛, it is revealed that the 𝑅𝑡ℎ for the CMC model declined by 1.57%
compared to the RMC one.

Figure 11: Effect of 𝑸𝒊𝒏 on the total thermal resistance for two different MCHS configurations.

13
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

3.5 Performance analysis


Performance evaluation criteria (𝜂) as defined by Eq. (14) can be employed to compare the
compound microchannel performance (CMC) with traditional microchannel (RMC); which is
regarded as an indicator of thermal enhancement to hydrodynamic performance. For 𝜂 values
in excess of one, it is reasonable to conclude that passive microstructures, like reentrant-shaped
microchannels, would enhance total MCHS performance over conventional microchannels. As
shown in Fig. 12, it is revealed that 𝜂 for CMC model is larger than one, which confirms the
efficacy of the discussed passive microstructures. Indeed, this 𝜂 fluctuated as 𝑄𝑖𝑛 increased,
owing to fluctuations in both the averaged Nusselt number ratio (𝐸𝑁𝑢 ) and the total pressure
drop ratio (𝐸∆𝑃 ). At 𝑄𝑖𝑛 = 50 𝑚𝑙/𝑚𝑖𝑛, albeit higher 𝐸∆𝑃 was found, but a higher 𝜂 value was
obtained, owing to 𝐸𝑁𝑢 , which outperforms 𝐸∆𝑃 .

Figure 12: Variation of 𝜼, 𝑬𝑵𝒖 and 𝑬∆𝑷 with 𝑸𝒊𝒏 .

4. Conclusion
The thermal-hydrodynamic performance of a new design for compound microchannels is
investigated using CFD. The configuration investigated includes a circular cavity microchannel
slotted from the top by a reentrant trapezoidal shape. The following are the major findings of
this study:
i. It has been observed that including a reentrant trapezoidal shape is more effective than
traditional MCHS in improving overall performance.
ii. Compound MCHS can reduce the heated surface temperature by reinitializing both the 𝛿ℎ𝑦
and 𝛿𝑡ℎ , and enhancing fluid mixing.
iii. It is worth noting that the newly proposed design can reduce ΔP while also improving heat
transfer significantly.
iv. By utilizing CMC configurations, 𝑁𝑢𝑎𝑣𝑔 increases by 2.16%, ΔP decreases by 1.73%, 𝑅𝑡ℎ
decreases by 1.57%, and 𝜂 increases by 1.0315% in comparison with RMC configuration.

14
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved
International Journal of Advanced Research in Engineering Innovation
e-ISSN: 2682-8499 | Vol. 5, No. 2, 1-15, 2023
http://myjms.mohe.gov.my/index.php/ijarei

References

Ghani, I. A., Sidik, N. A. C., & Kamaruzaman, N. (2017). Hydrothermal performance of


microchannel heat sink: The effect of channel design. International Journal of Heat
and Mass Transfer, 107, 21–44.
Steinke, M. E., & Kandlikar, S. G. (2004). Single phase heat transfer enhancement technique
in micro-channel and mini-channel flows, in: international conference on Micro-
channel and Mini-channels, New York, ASME, 141–148, ICMM 2004-2328.
Dewan, A. & Srivastava, P. (2015). A review of heat transfer enhancement through flow
disruption in microchannel. Journal of Thermal Science, 24, 203–214.
Xie, G. N., Liu, X. T., Yan, H. B., & Qin, J. (2017). Turbulent flow characteristics and heat
transfer enhancement in a square channel with various crescent ribs on one wall, Int.
J. Heat Mass Tran, 115, 283–295.
Wang, G., Qian, N., & Ding, G. (2019). Heat transfer enhancement in microchannel heat sink
with bidirectional rib. Int. J. Heat Mass Transf, 136, 597–609.
Chen, H., Chen, C., Zhou, Y., Yang, C., Song, G., Hou, F., Jiao, B., & Liu, R. Evaluation and
Optimization of a Cross-Rib Micro-Channel Heat Sink. (2022). Micromachines, 13,
132.
Al-Neama, A. F., Kapur, N., Summers, J., & Thompson, H. M. (2018). Thermal management
of GaN HEMT devices using serpentine minichannel heat sinks. Applied Thermal
Engineering, 140, 622–636.
Al-Neama, A. F. M. (2018). Serpentine minichannel liquid-cooled heat sinks for electronics
cooling applications, Dissertation submitted to the University of Leeds, school of
mechanical engineering.
Chai, L., Xia, G., Wang, L., Zhou, M., & Cui, Z. (2013). Heat transfer enhancement in
microchannel heat sinks with periodic expansion-constriction cross-sections.
International Journal of Heat and Mass Transfer, 62, 741–751.
Deng, D., Wan, W., Tang, Y., Shao, H., & Huang, Y. (2015). Experimental and numerical
study of thermal enhancement in reentrant copper microchannels. International
Journal of Heat and Mass Transfer, 91, 656–670.
Abad, J. M. N., Alizadeh, R., Fattahi, A., Doranehgard, M. H., Alhajri, E., & Karimi, N. (2020).
Analysis of transport processes in a reacting flow of hybrid nanofluid around a bluff-
body embedded in porous media using artificial neural network and particle swarm
optimization. Journal of Molecular Liquids, 313, 113492.
Al-Neama, A. F., Khatir, Z., Kapur, N., Summers, J., & Thompson, H. M. (2018). An
experimental and numerical investigation of chevron fin structures in serpentine
minichannel heat sinks. International Journal of Heat and Mass Transfer, 120, 1213-
1228.
Steinke, M.E. & Kandlikar, S.G. [2006]. Single-phase liquid friction factors in microchannels.
Int. J. Therm. Sci, 45, 1073–1083.
Kandlikar, S.G. Chapter 3 - Single-phase Liquid Flow in Minichannels and Microchannels,
Heat Transfer and Fluid Flow in Minichannels and Microchannels, second ed.,
Butterworth-Heinemann, Oxford, 2014, pp. 103–174.
Shah, R.K. & London, A.L. [1978]. Laminar Flow Forces Convection in Ducts; Academic
Press: New York, NY, USA.
Phillips, R.J. Microchannel Heat Sinks. [1987]. Ph.D. Thesis, Massachusetts Institute of
Technology, Cambridge, MA, USA.

15
Copyright © 2023 ASIAN SCHOLARS NETWORK - All rights reserved

You might also like