You are on page 1of 14

Thermal Science and Engineering Progress 46 (2023) 102170

Contents lists available at ScienceDirect

Thermal Science and Engineering Progress


journal homepage: www.sciencedirect.com/journal/thermal-science-and-engineering-progress

Thermal enhancement of heat sinks with bio-inspired textured surfaces


Esteban Foronda a, b, Francisco J. Ramírez-Gil a, Álvaro Delgado-Mejía a, Luis M. Ballesteros a, b, c,
J.S. Rudas a, b, Luis C. Olmos-Villalba a, b, *
a
Institución Universitaria Pascual Bravo, Faculty of Engineering, Department of Mechanics, Medellín, Colombia
b
Universidad Nacional de Colombia, Faculty of Mines, Medellín, Colombia
c
Centro de Ingeniería y Desarrollo Industrial, Av. Playa Pie de la Cuesta No. 702, Desarrollo San Pablo, C.P. 76125 Querétaro, Mexico

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

Keywords: The current requirements of the electronics industry for heat sink design are to increase heat dissipation and to
Electronic cooling reduce the occupied space and weight. Research to improve the thermal efficiency and reduce the size of heat
Thermal numerical analysis sinks has mainly been conducted by modifying their geometry to increase the heat transfer area. In the field of
Finite element method (FEM)
tribology research, one way to modify a surface is to apply texture, which involves reducing the contact area but
Bio-inspired design
increasing the external area, which in turn could improve the performance of heat dissipation devices. In this
Flat fin heatsink
Surface texturing paper, bio-textured surfaces are fabricated on straight fin heat sinks and numerically evaluated by thermal finite
element analysis. The simulation results are validated by experimental measurements on a single-fin heat sink.
The potential of surface texturing in heat sinks is demonstrated here through a numerical parametric analysis of a
full-scale heat sink. The results show a reduction in operating temperatures of more than 26 %, reduction in
thermal resistance around 34 % and increase between 21 % and 40 % of heat sink effectiveness for textured
surface heat sinks compared to smooth fin heat sinks, depending on the convection coefficient. These results
suggest that the design methodology proposed here is more relevant for heat sinks used in high power dissipation
applications such as computer CPUs and GPUs.

sinks becomes the most important problem in this developing industry


[6].
1. Introduction There is a wide variety of heat sinks on the market that can be
classified by characteristics such as application, manufacturing process,
Due to the increasing demand for electronic devices in different and geometry [7]. Heat sinks are used to control the temperature rise in
applications like gaming computers, smartphones, smart TVs, tablets, electronic components for reliability and performance, as overheating
video consoles, and others, microprocessor manufacturers have had to can reduce their lifetime. Many of current devices can control their
focus on innovation and improve their performance, especially in terms behavior to avoid reaching the point of irreversible damage, typically
of cooling and temperature control. Modern electronic processors between 75 ◦ C and 150 ◦ C [8]. This temperature control ensures that
generate substantial amounts of heat, usually over 100 W/cm2 [1]. This critical electronic components operate within temperature limits by
excess heat can lead to overheating, significantly reduce their useful life, transferring thermal energy (heat) from the component to a lower
or even cause permanent damage [2]. In this sense, the use of heat sinks temperature region. Heat transfer occurs primarily by convection and
has been the widespread alternative to remedy this problem. Therefore, radiation, with the most common heat sink design incorporating assisted
heat sinks are widely used in electronic thermal management. Heat sinks ventilation systems (forced convection). However, in recent years there
are heat exchangers characterized by ease of fabrication and adapt­ has been great interest in the use of natural convection heat sinks due to
ability to any available space, which are two distinctive features their many advantages, such as the elimination of additional mechanical
required for cooling systems in electronics [3–5]. The increasing devices (fans, blowers, compressors), which reduces noise, power con­
complexity of current electronic devices and their systems allows them sumption, the probability of failure due to the absence of moving parts,
to achieve higher data processing capacity, which requires higher power and consequently the overall cost [9–11].
consumption and operating temperature, and can lead to overheating. To the best of the authors’ knowledge, the case of heat sinks with
Hence, the need to improve the heat dissipation characteristics of heat

* Corresponding author at: Institución Universitaria Pascual Bravo, Faculty of Engineering, Department of Mechanics, Medellín, Colombia.
E-mail address: luis.olmos@pascualbravo.edu.co (L.C. Olmos-Villalba).

https://doi.org/10.1016/j.tsep.2023.102170
Received 30 May 2023; Received in revised form 23 September 2023; Accepted 1 October 2023
Available online 2 October 2023
2451-9049/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

Nomenclature TavgE experimental average temperature (◦ C)


TpreS prescribed temperature in smooth surface configuration
Symbol (◦ C)
A surface area TpreT prescribed temperature in textured surface configuration
Dhe Hydraulic diameter (mm) (◦ C)
ξf fin effectiveness Tmin minimum operating temperature (◦ C)
∊ the emissivity Tpre prescribed temperature (◦ C)
σ Stefan-Boltzmann constant x position vector
H Fin height (mm)
Abbreviations
h convective coefficient (W/m2 K)
3D three-dimensional
hm average convective coefficient (W/m2 K)
AFW average fibril width (mm)
k thermal conductivity of air (W/m K)
AFL average fibril length (mm)
kij the conductivity tensor
ADR average distance between the rows of fibrils (mm)
L Fin length (mm)
BVP boundary value problem
Num Average Nusselt number (dimensionless)
CFD computational fluid dynamics
n normal vector to the surface
CPHS cold plate heat sinks
qb heat transfer rate without fins
CPU central processing unit
q″c convection heat flux
EDM electrical discharge machining
qf heat transfer rate with fins
FAR fibril aspect ratio (dimensionless)
q″r radiation heat flux FC forced convection
qs heat dissipation for smooth surface FEA finite element analysis
qT Heat dissipation for textured surface FEM finite element method
Re Reynolds number (dimensionless) FFHS flat fin heat sinks
Rth thermal resistance GPU graphics processing unit
Γq regions in which the Neumann boundary conditions are IC integrated circuit
applied MCHS microchannel heat sinks
ΓT regions in which the Dirichlet boundary conditions are NC natural convection
applied PFHS pin fin heat sinks
s fin spacing (mm) P. regius python regius snake
sk the location of a point on the surface of the domain SEM Scanning Electron Microscopy
t Fin thickness (mm) TIM thermal interface material
T temperature WHS wave heat sinks
Tc reference temperature for convection heat flux WMCHS wavy microchannel heat sinks
Tr reference temperature for radiation heat flux
TavgN numerical average temperature (◦ C)

textured surfaces has not received enough attention in the scientific 1.1. Literature review
community, which motivated the present study. According to the sci­
entific literature review, geometry changes on heat sinks corresponding Research on the thermal performance of heat sinks has been pre­
to textured surfaces have a substantial impact on thermal efficiency and dominantly dedicated to determining the optimal geometry for several
represent a promising option to improve the relationship between heat decades. This has resulted in the development of various heat sink ge­
dissipation, size and weight of cooling systems for electronics. This work ometries such as flat fin heat sinks (FFHS), pin fin heat sinks (PFHS),
aims to characterize bio-inspired textured surfaces applied to flat fin wave heat sinks (WHS), cold plate heat sinks (CPHS), microchannel and
heat sinks (FFHS), by numerical simulation with experimental valida­ cooled heat sinks, and their many variations [12,7]. Geometric optimi­
tion. The thermal performance of the proposed heat sink is measured by zation of a heat sink to efficiently dissipate heat from a device to its
its temperature distribution, the thermal resistance, effectiveness and surroundings typically focuses on maximizing the surface area exposed
compared to that of a reference smooth FFHS. The geometric features of to airflow. In particular, microchannel heat sinks (MCHS) have
the bio-inspired deterministic textured lamellar surfaces are inspired by demonstrated higher cooling capacity due to their higher surface-to-
the topographical features found in the skin of a Python Regius (P. volume ratio and the flow of air, water, ethylene glycol, and other
regius) snake. The numerical model is based on thermal finite element nanofluids through the microchannels [13–15]. MCHS can increase both
analysis (FEA) and was conducted using commercial software. The FEA the contact area between the fins and the coolant as well as the local
model has been experimentally validated and used to explore a wider mixing velocity of the coolant, resulting in better heat dissipation [16].
range of operating conditions. This parametric study allows us to The concept of MCHS cooling was first introduced by Tuckerman and
conclude that textured heat sinks improve heat dissipation without Pease in 1981 [17,18]. The advantages of MCHS have motivated re­
increasing geometric dimensions once the challenges of texture fabri­ searchers to investigate their thermal performance in various situations
cation are solved. Therefore, the novelty of this work lies in the vali­ by changing geometric parameters and introducing materials with a
dation of the main research hypothesis, which combines two different more favorable weight-to-thermal properties ratio. This has led to
conceptual frameworks: thermal analysis through mathematical several experimental and numerical works [11,16,19–21]. A typical
modelling solved computationally and texture design through MCHS comprises small and straight channels and fins arranged in par­
bioinspiration. allel. The streamlines of the coolant are also straight. Several MCHS
variations have been proposed to enhance the effective convective heat
transfer area while limiting the pressure drop to a reasonable level. Peles

2
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

et al. [22] conducted a study on the heat transfer and pressure drop in 2. Materials and methods
tubular microfin heat sinks. The study found that this type of geometry
has a low thermal resistance, similar to that of microchannel heat sinks. Fig. 1 shows the methodology used in this study, which includes
Several researchers have demonstrated through numerical simulations texture design and manufacturing, and numerical simulation and vali­
that wavy microchannel heat sinks (WMCHS) have better heat transfer dation. We decided to identify objects in nature with specific geometric
performance than straight microchannels with a small pressure drop patterns using the conceptual framework of biomimetics, particularly
penalty [23–25]. In addition, Kadir et al. [26] evaluated smooth surfaces the bio-inspiration from nature [51]. In this work, topographical char­
against perforated fins and found that the latter resulted in a 40 % higher acterization of the Python Regius snake skin is performed to define the
net energy gain and an energy efficiency of approximately 77 %. The bio-inspired texture geometry to enhance the thermal response of heat
pressure drop was not significantly affected. The study of heat sinks with sinks by increasing the area of dissipation. This geometry is used as input
porous media and perforated fins has also shown excellent heat transfer to evaluate the thermal performance of textured heat sinks through both
results, but pressure drop must be considered, especially for dense fluids experimental and numerical approaches. The experimental setup is first
such as high concentration nanofluids [1]. presented to clarify the thermal phenomena involved in this application,
There are several passive and active techniques used to improve the the challenges associated with texture fabrication in straight fin heat
efficiency of thermal systems [11]. One of these techniques is the sinks, the assumptions adopted for the boundary conditions, and the
modification of heat sink surfaces to enhance convective heat transfer limitations of the experimental setups in terms of robustness and pre­
[27], which has been widely explored in the literature, for example, cision of results. Numerical simulations are subsequently conducted to
using dimpled surfaces [28], creating surfaces with protrusions [29], explore the response of the heat sinks under a wide range of operating
manufacturing artificial surface roughness [30], and texturing surfaces. scenarios, obtain the temperature distribution of the heat sink over the
In this work we focus on improving the performance of heat sinks with entire geometry, and demonstrate the potential of the proposed texture
textured surfaces, which basically consists of adding small grooves to the to enhance thermal performance. The experimental and numerical ap­
fin surfaces. Actually, a textured surface with a continuous trajectory proaches are linked through a simplified case study of a single fin heat
can be considered as a microchannel [31], but with an external air flow, sink, which allows to validate the numerical methodology and to
which is the main difference with traditional MCHSs, which usually use demonstrate the technical recommendations made by the authors for the
closed channels with liquids as the coolant. In this field, Refai-Ahmed texture fabrication. The parameters used in the simulations of the
and co-workers [32–34] improve existing air-cooled heat sinks by simplified single fin model (such as convection coefficient, nominal
incorporating a textured surface on the bottom of the heat sink to trap operating temperature of the bottom face of the heat sink and air tem­
and expel gas bubbles that remain in contact between the integrated perature) are determined based on the experimental measurements. In
circuit (IC), thermal interface material (TIM) and heat sink. This reduces order to design textures for the FFHS, the skin of a Python Regius snake
thermal resistance and increases thermal conductivity, which in turn was characterized using Scanning Electron Microscopy (SEM).
increases heat transfer from the IC die to the heatsink. In addition,
Saravanakumar and Kumar [35] experimentally investigated several 2.1. Topographical characterization
types of threaded surface textures on PFHS and found that the heat
transfer was 18.13 % higher than that of the plain PFHS. This perfor­ With the purpose of having a geometrical inspiration for the
mance enhancement is due to increased surface area, flow turbidity and designed textures, a topographic microscale information of the ventral
delays flow separation of the threaded texture. Moreover, Dhadda et al.
[36] reported a 500 % increase in heat transfer coefficient for pool
boiling using textured surfaces fabricated with electrical discharge
machining (EDM) compared to a polished surface, mainly due to the
increased roughness. Similarly, Liang and Mudawar [37] reviewed the
passive enhancement of pool boiling using surface modification tech­
niques and Zhu et al. presented the boiling enhancement using bio-
inspired surfaces [38–40]. Finally, Chang et al. [41] proposed a scale-
roughened surface similar to the skin of a fish.
The surface control and surface texturing processes have been a
widely studied topic by materials and tribology science [42]. In the same
way, biological inspiration has influenced the design and implementa­
tion of several mechanisms as texturing processes, achieving better
performance than the traditional stochastic surface design [43]. Toro
et al. [44] studied the anisotropic friction of some reptiles and Borma­
shenko the superhydrophobic effect of rose petals [45]. Ballesteros et al
[46] studied the tribological response in dry conditions of some specific
texture patterns obtained by 3D printing polymer against metal, inspired
by the ventral scale of the Python Regius snake. Their results showed
that the coefficient of friction was consistently lower for the tests with
the textured samples compared to those with stochastic surfaces. In
terms of heat transfer, lower friction values can positively affect the
performance of heat-dissipating devices. Moreover, several researchers
have reported a direct relationship between surface texture and tem­
perature reduction [47–50]. In cutting tools, texture imposition has
shown a temperature reduction between 12 % and 65 % [49], in plain
bearing performance, thermal analysis has shown a temperature
reduction of 60 % [47,50]. Consequently, studies of textured surfaces for
heat dissipation applications are needed to explore their benefits, and
this is the aim of this article. Fig. 1. Methodology for design, texture manufacturing, numerical simulation
and validation.

3
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

scale of the Python Regius skin snake is characterized. The shed skin of Table 1
P. Regius is coated with gold using a sputtering system to obtain a good Description of the main components used in the experimental setup.
quality SEM image. The scanning electron microscope used is a JEOL Item number Element Specifications
5910LV.
1 Fan Generated pressure: 3.94 mm/H2O
Generated air flow: 121.80 CFM
Size: 92 × 25 mm
2.2. Experimental setup for the single straight fin heat sink
2 Thermal plate Temperature range: up to 100 ◦ C
3 Arduino module Microcontroller: Atmega 328
An experimental setup is constructed to determine the operating Supply voltage: 5 V
conditions of a single fin heat sink subjected to a heat source at the Digital signals: 8
Analog signals: 6
bottom face and heat transfer to the environment by convection (natural
4 Thermocouple Accuracy ± 1 ◦ C
and forced) and radiation. The experimental setup is also used to iden­
tify the challenges and limitations of discrete temperature measure­
ments and demonstrate a possible manufacturing technique for heat sink temperature field can be obtained for the entire heat sink, so no other
texturing. Fig. 2 presents the experimental setup used to evaluate the experimental temperature measurement methods (such as infrared im­
thermal response of the reference single-fin heat sink (with and without aging) were implemented. Measurements are made with thermocouples
texturing), which consists of a structural chassis, the thermal hotplate located in small perforations (see Fig. 2c) and using thermal paste
(item 2) used to apply prescribed temperatures, an adjustable computer (Halnziye HY610) to ensure reliable contact between each sample and
fan (item 1) to control the level of convection, and an Arduino module the thermocouple tips (yellow wires in Fig. 2b) to mitigate some factors
(item 3) that controls the fan speed and power supply to the hotplate. affecting thermocouple temperature measurement error [52,53]. The air
Table 1 provides relevant technical specifications of the main compo­ temperature was measured with a thermocouple placed between the fan
nents of the experimental setup. The heat sink is placed on the hotplate and the single fin heat sink.
using thermal paste (Halnziye HY610) as a TIM. Forced convection (FC)
is imposed as shown in Fig. 2b) by placing the fan at a predefined dis­
tance from the heat sink to obtain horizontal and vertical air flows. Four 2.3. Machining of textured surface
probe locations are defined for each sample, two near the top surface
and the other two near the bottom surface. In addition, the purpose of The surface quality of the textured surface is the main advantage of
the experimental setup is to define the temperature difference between using micromachining techniques [54]. In this work, microdrilling is
the base and the tip of the single fin heat sink. Since the boundary used to create the grooves for texturing, as this method is advantageous
conditions are uniform along the length of the heat sink, four probe in terms of material removal rate compared to electrical discharge
locations (two for the tip and two for the base of the fin) are considered machining (EDM) [36]. This technique, with appropriate cutting pa­
sufficient to compensate for a possible temperature gradient along the rameters, allows control of dimensional tolerance and top/bottom burr
length. After validating the results of the numerical model, the [55]. A Haas UMC 750 machining center at the Computer Numerical

Fig. 2. Experimental setup for evaluating the thermal response of heat sinks.

4
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

Control Laboratory of Pascual Bravo University is used to machine the where hc is the convective coefficient, which is assumed to be con­
textured surfaces in conjunction with an end mill with a 3.175-20-0.2 stant in this work. Tc and Tr are the reference temperatures for con­
reference drill. The cutting parameters successfully used to obtain the vection and radiation heat flux, respectively. ∊ is the emissivity of the
textured surface are presented in Table 2. surface and σ is the Stefan-Boltzmann constant. Surface-to-surface ra­
diation that may occur between the heat sink surfaces is not included.
2.4. Numerical simulation setup
2.4.2. Fin performance
In traditional numerical modeling of heat sinks, the focus is on Fin performance can be quantified in terms of thermal resistance
calculating the temperature field and measuring the heat dissipation (Rth ) and effectiveness (ξf ). The Rth and ξf of the surfaces were calculated
through radiation and convection under steady-state nominal operating according to Eq. (6) and Eq. (7), respectively.
conditions. Since the main objective of this work is to assess the overall
1
performance of heat sinks and investigate the difficulties related to Rth = (6)
hc A
texture fabrication and experimental measurements. Thus, the models
used in this work are limited to the structural field. This means that qf
although the heat dissipation effect of air is taken into account, air itself ξf = (7)
qb
is not modeled explicitly [6]. The decision to limit the models to the
structural field leads to the attribution of the performance difference where A is the surface area for both smooth and textured surfaces, qf
between smooth and textured heat sinks to the increase in area resulting is the heat transfer rate with smooth or textured surface fins, qb is the
from the textured surface. While computational fluid dynamics (CFD) heat transfer rate without fins.
analysis is an alternative method for evaluating the local phenomena of
interaction between the air surrounding the heat sinks [28], and the 2.4.3. Numerical implementation
grooves of the textured surfaces, this work is limited to solving the The thermal performance of heat sinks with and without surface
conduction equation in a solid. This limitation is intentional because the texturing is evaluated by numerical simulation. A 3D steady-state ther­
results under these assumptions are relevant for the purposes of a robust mal FEA is performed using the commercial software Ansys™ v17.
design and the evaluation of multiple scenarios through numerical Although radiation to the environment is a phenomenon present in the
design of experiments. physical model, radiation boundary conditions were not applied to
simplify the calculations. This decision was made because convection is
2.4.1. Governing equations and boundary conditions more relevant than radiation for the selected temperature range.
The heat transfer equation in its strong (differential) form is pre­ Furthermore, this decision does not affect the results due to the
sented in Cartesian coordinates in Equation (1), which represents parameterization of the convection coefficient, which can be interpreted
steady-state conduction in a solid with no internal energy generation as an effective heat transfer coefficient combining both convection and
[56]. radiation. Two types of numerical heat sink models are proposed: a
( ) simplified model to validate with experimental data and a full-size heat

kij
∂T
=0 (1) sink to study the influence of textures on heat sink surfaces.
∂xi ∂xj Fig. 3 shows the full-scale heat sink design used for the numerical
simulations and the respective boundary conditions. The bottom of the
where T is the temperature, kij is the conductivity tensor, and the
heat sink has a uniform temperature which is the source of the power to
terms i, j = 1, 2, 3 are subindices for a three-dimensional (3D) domain.
be dissipated. To save computational time, only a quarter of the geom­
The position vector is x. Equation (1) is a second-order differential
etry (two planes of symmetry) was considered in these simulations. The
equation that is part of a boundary value problem (BVP) with boundary
remaining surfaces of the heat sink are exposed to the environment
conditions: Dirichlet T = Tk in ΓT (the temperature is imposed according
through convection. Convective heat transfer is imposed on the entire
to Eq. (2)) and Neumann, according to the type of heat flux specified in
geometry through a Neumann boundary condition with a constant heat
Γq (see Eq. (3)).
transfer coefficient and reference temperature (hc and Tc are constant in
T = f T (sk ) on ΓT (2) Eq. (4)) because the air is not explicitly modeled. This assumption is
deemed suitable as the focus of this work is on the overall performance
( )
∂T rather than on the local phenomena resulting from the interaction be­
− kij ni = q″c + q″r = f q (sk ) on Γq (3)
∂xj tween the heat sink and the ambient air.
where sk is the location of a point on the surface of the domain, ΓT
2.4.4. Simplified numerical model for simulation validation
and Γq are the boundaries in which the essential and natural boundary
A first thermal simulation using a simplified model of a heat sink is
conditions are applied, respectively. The normal vector to the surface is
used to validate the following numerical results. The overall setup of this
n. q″c and q″r correspond to the heat flux by convection (see Eq. (4)) and simplified model is equivalent to Fig. 3 for a single-fin heat sink. In this
radiation (see Eq. (5)), respectively. In addition, the nonlinearities of the case, the single-fin heat sink is subjected to a fixed temperature on the
material or of the boundary condition parameters are not included, i.e. bottom surface and convection on the other external surfaces. The
they do not depend on the temperature. temperature field is obtained for comparison with the experimental re­
q″c = hc (T − Tc ) (4) sults. This analysis is performed for both the smooth and the textured
geometry. The material model for a steady-state thermal analysis,
( ) ( ) assuming isotropy, is defined by the thermal conductivity with a con­
q″r = ∊ σ T 4 − Tr 4
= ∊ σ (T − Tr ) (T + Tr ) T 2 + Tr2 (5)
stant value of 160 W/m ◦ C, which corresponds to the 6061-aluminum
alloy, one of the most common materials for extruded heat sinks. The
Table 2 meshes for both heat sinks (smooth and textured) are constructed with
Cutting parameters for textured surface machining. second-order tetrahedral finite elements, where mesh convergence is
Parameter Value guaranteed.
Spindle speed 10,000 rpm
Forward speed 70 mm/min 2.4.5. Convective heat transfer coefficient for numerical models
Coolant Liquid synthetic coolant The convective heat transfer coefficient (h) is calculated using the

5
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

Fig. 3. Geometry of the heat sinks and boundary conditions for the numerical simulations.

correlations for finned heat sinks proposed by Sato et al [57]: 3.1. Bio-inspired texture design
( s )0.47 ( t )0.36 ( L )0.13
Num = Re0.5 (8) Fig. 4 shows the morphology and some specific features of a ventral
H L Dhe scale of P. Regius skin snake. The repetitive and oriented arrangements
features are called fibrils and its control the locomotion of the snake. It
hm =
Num k
(9) can be seen that the fibrils have a deterministic distribution and some
Dhe measurements can be obtained from the micrographs. The measured
where Num is the average Nusselt number, Re is the Reynolds number topographical parameters of the fibrils are: average fibril width (AFW)
defined at the inlet of the heat sink flow, (s/H) is the fin aspect ratio, (t/ (see Fig. 4a), average fibril length (AFL) (see Fig. 4b), average distance
L) is the inlet width ratio, (L/Dhe) is the dimensionless channel length, hm between fibril rows (ADR) (see Fig. 4c). A detailed description of these
is the average convective coefficient, and k is the thermal conductivity of parameters can be found elsewhere [43,44,58,59]. Table 4 shows the
air. The Re is calculated using the air flow (AF) shown in Table 1, which fibril dimensions.
was measured with a Model 9671 Anemometer Air Flow Datalogger. The Using the reported fibril dimensions, it is possible to calculate some
thermal properties of air are estimated at the air temperature measured geometric characteristics that can help in the design of fin heat sink
by the anemometer. Therefore, by applying Equations (1) and (2) and textures. The Fibril Aspect Ratio (FAR) is defined as the ratio of the
substituting the values of the flow conditions and geometric character­ average fibril width (AFW) to the average fibril length (AFL):
istics of the model considered here, h can be estimated and used as input AFL
for the numerical simulations. The parameters for the calculations cor­ FAR = (10)
AFW
responding to the models for the experimental tests are given in Table 3.
The geometric surface design was chosen considering the FAR of the
2.4.6. Full-size numerical model for heat sinks P. Regius skin snake between 1.33 < FAR < 1.65. The trajectory
After validating the numerical model, the potential of the proposed implemented in this work is taken from the work of Ballesteros et al.
texture to improve heat sink performance is evaluated through numer­ [46] and is inspired by the topography of the fibril. Furthermore, Ma
ical simulations by exploring a wider range of scenarios in terms of et al. [60] have shown that the zigzag channel can improve the uni­
operating conditions (convection and power source) and geometry. A formity of the temperature distribution. The trajectory parameters of the
65.8 × 52 × 86 mm heat sink with 17 fins (1.8 mm thick) is used. A zigzag texture considered in this work are: AFW ≈ 1.77mm, AFL ≈
parametric analysis is performed by varying the values of the boundary 1.1mm and ADR ≈ 2mm and a result of FAR = 1.54. Fig. 5 shows the
conditions to evaluate the effect of the textured surface under different single-fin heat sink texturing with the corresponding parameters.
operating scenarios (see Fig. 3).
3.2. Textured surface machining
3. Results and discussion
Once the design parameters have been selected, the manufacturing
The methodology explained in the previous section was used to process is carried out. The following factors are considered: i) the ge­
obtain the following results: topographical characterization, experi­ ometry of the profile using a commercially available cutting tool of a
mental, and numerical evaluation of reference single-fin heat sink, and conventional milling process, ii) the equipment and assembly to perform
thermal analysis of full-scale textured heat sinks using FEA. Further­ the machining of the textured surface, iii) the optimization of the area to
more, a parametric study was carried out to determine the impact of be subjected to the texturing process to obtain more patterns and thus
distinct design parameters on the possible performance improvement increase the heat transfer area, and iv) the patterns and the selected FAR.
achievable with textured heat sink surfaces. The dimensions of the final R-profile for texturing are shown in Fig. 6, as
well as the manufactured geometry for experimental testing.

Table 3 3.3. Numerical model validation


Geometrical characteristics of the single-fin heat sink model.
Symbol Parameter Value Unit The experiments included three runs for each type of surface and
s Fin spacing 9.5 mm convection. Three types of surfaces are evaluated for the single-fin heat
L Fin length 41 mm sink reference geometry: smooth, vertically textured, and horizontally
t Fin thickness 1.8 mm textured. Two levels of convection are considered: natural convection
H Fin height 20.15 mm (without fan) and forced convection (with fan). The Nusselt number and
Dhe Hydraulic diameter 5.22 mm
convective heat transfer coefficient are calculated using Equation (8)

6
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

Fig. 4. The topographical parameters of the fibrils.

Table 4
Topographical parameters of the skin snake.
Parameter Value (um) Average

AFW 0.69 0.59 0.51 0.60 0.60 0.54 0.65 0.71 0.81 0.61 0.82 0.73 0.66
AFL 1.33 1.35 1.10 0.85 0.95 0.81 1.11 0.98 1.06 0.98 0.89 1.06 1.04
ADR 5.10 4.65 4.15 4.65 3.96 3.96 4.78 4.12 4.17 4.89 5.16 4.14 4.48

and Equation (9) and the information given in Table 1 and Table 3. The for NC and FC. The temperature corresponding to the measured value
result of the average value of the convection coefficient is h = 12 W/ during the experiments on the bottom surface is imposed in the FEA
m2 ◦ C for natural convection (NC) and h = 50 W/m2 ◦ C for forced model. The average temperature is calculated at the four probe loca­
convection (FC). tions. The temperature patterns are consistent with the experimental
The thermal hotplates are expected to reach up to 100 ◦ C if the test results, and it is observed that the textured surface produces a more
assembly is sufficiently insulated to prevent significant energy losses to uniform temperature distribution. Due to the convection boundary
the environment by convection and radiation. However, the experi­ condition, a lower temperature is obtained at the top face of the heat
mental setup shown in Fig. 2 is open and the temperature of the hotplate sink. For the sake of brevity, and because the overall phenomena
is affected by energy losses due to ambient conditions. Therefore, the modeled in the case studies of this work do not depend on local thermal
temperature of the plate was found to be between 84 ◦ C and 92 ◦ C for the gradients, a detailed analysis of mesh convergence is not included here.
natural convection case and between 47 ◦ C and 50 ◦ C for the forced The numerical models for the single-fin heat sink with 184,000 s-order
convection case. The surface temperature at different locations of the tetrahedral finite elements guaranteed mesh independence of the
heat sink was measured every five minutes until a stable condition was results.
reached. The experiment included 4 runs for each level of surface texture Table 6 presents the raw data for the comparison between numerical
and type of convection, where the surface temperature on the heat sink and experimental results for the simplified models, while Table 7 shows
and air temperature was measured every 6 min, the average temperature the processed percentages for the comparison. First, the difference in the
results are presented in the Table 5. average temperature of the probe locations is at most 1.1 % between the
Since the complexity of the physical model of the full-size textured numerical models (Tavg N ) and the experimental results (Tavg E ), which
heat sink is similar for all cases, only the horizontal pattern texture is validates the FEA model. This supports the decision to build another
implemented for the numerical simulations, which is considered suffi­ more complex numerical model to compare full-scale smooth and
cient to validate the proposed methodology. The first set of numerical textured heat sinks without the need for additional experimental vali­
results is the temperature field for the single-fin models shown in Fig. 7 dation and/or manufacturing. On the other hand, although the smooth

7
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

Fig. 5. Horizontal and vertical zigzag trajectories of the textured surface on the flat fin.

Fig. 6. Manufactured single-fin heat sink with textured surfaces and R-profile groove dimensions (units: mm).

more heat to be dissipated at lower operating temperatures.


Table 5
Average hotplate temperature and air temperature of the experimental tests for
the single-fin heat sink.
3.4. Performance evaluation of textured heat sinks

Texture type Hotplate temperature Air temperature


The validation of the numerical approach by the experimental results
NC FC NC FC of the single-fin geometry allows the focus of the analysis to be shifted to
Smooth 90.8 ◦ C 49.5 ◦ C 35.6 27.1 the full-scale heat sink. The thermal conductivity of the heat sink ma­
Vertical pattern 83.8 ◦ C 47.3 ◦ C 31.7 24.6 terial for the parametric analysis is changed to 229 W/m ◦ C, assuming
Horizontal pattern 88.4 ◦ C 48.0 ◦ C 27.74 24.4 mass production of this component with a controlled extrusion
manufacturing process that guarantees a higher conductivity than the
(Tpre S ) and textured (Tpre T ) heat sinks are not prescribed with the same material used for the single-fin heat sink of the experimental tests. Mesh
independence is verified by running simulations with up to 2.5 million
temperature (Tpre ) due to the experimental results, it can be observed
finite elements (see Fig. 8). Since the implemented structural approach
that the use of textured surface enhances thermal performance as a
does not generate significant local gradients on the geometry (no explicit
higher power dissipation is achieved, 28.9 % in natural convection and
modeling of the air surrounding the heatsink), no local mesh refinement
25.9 % in forced convection more than the smoothed surfaces.
was required to achieve mesh convergence.
Furthermore, it is also observed that even the textured heat sink is
Fig. 9 shows the temperature field for a quarter of the geometry of
imposed with lower temperature compared to the smoothed surfaces
the two heat sinks shown in Fig. 3, where a temperature of 90 ◦ C is
(Tpre /Tpre S , − 2.5 % and − 1.4 % for natural and forced convection,
imposed on the bottom surface and a convection coefficient of 50 W/
respectively), a minimum operating temperature is obtained using
m2 ◦ C is used on the other surfaces (the assumed ambient temperature is
textured heat sinks. In fact, in natural convection, textured surfaces
22 ◦ C). As expected, when the geometry of the heat sink allows a sig­
reduce the temperature by up to 3 % (Tmin /Tpre ), while smoothed sur­
nificant increase in surface area through the use of textured surfaces, the
faces reduce the temperature by only 2.1 %. In forced convection, the
temperature of the heat sink is reduced. In particular, the numerical
percentages are higher, with textured surfaces reducing the heat sink
model shows that the minimum temperature obtained for the heat sink
temperature by up to 8.5 %, while smooth surfaces reduce the temper­
can be reduced to more than 12 ◦ C (17 %) by including the texture on
ature by only up to 4.4 %. This is an important result because it shows
both surfaces of the fins. In this case, the total surface area is increased
that the improvement in thermal performance of textured heat sinks is
from 0.145 m2 for the plain heat sink to 0.218 m2 for the textured heat
more significant at higher values of the convection coefficient, allowing
sink, which is a net surface area gain of 50 %. Moreover, the weight of

8
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

Fig. 7. Temperature field in the smooth and textured single-fin heat sinks for the natural convection (NC) and forced convection (FC) simulations.

Table 6
Raw thermal results for the single-fin heat sink.
Convection type Surface type Tpre [◦ C] Tmin [◦ C] Tavg [◦ C] (Numerical) Tavg [◦ C] (Experimental) Heat dissipation [W]

Natural Smooth 91.9 90.5 91.13 90.75 2.04


Textured 89.6 86.9 88.14 88.35 2.63
Forced Smooth 50.4 48.2 49.16 49.48 3.36
Textured 49.7 45.5 47.44 47.97 4.23

Table 7
Percentage results comparing smoothed and textured heat sink surfaces.
Convection Surface Tpre Tmin Tavg N − Tavg E qT − qS
[ [ [ [
type type Tpre S Tpre TavgE qS
%] %] %] %]

Natural Smooth – 97.9 0.42 –


Textured 97.5 97.0 − 0.24 28.9
Forced Smooth – 95.6 − 0.65 –
Textured 98.6 91.5 − 1.10 25.9

the textured heat sink is reduced by ~12 % (384.57 g versus 436.86 g,


for the full size textured and smoothed heat sinks, respectively), which is
a significant gain especially for mobile application. The power dissi­
pated is 98.9 W for the smooth heat sink and 127.2 W for the textured
heat sink (only a quarter of the model), representing a 28.6 % increase in
power dissipation.
Fig. 8. Finite element mesh for the full-size textured heat sink – 2.5 million
As mentioned above, the results obtained suggest that the textured
finite elements.
heat sink can dissipate more thermal energy at a lower operating tem­
perature. These results are consistent with the increase in thermal

9
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

Fig. 9. Temperature field and power dissipation for a quarter of the full-size heat sink.

performance found by other authors who have used surface modification and is consistent with the results described for the previous simulation.
in heat sinks [28–30]. However, a direct quantitative comparison with The main conclusion to be drawn from Fig. 10 is that the distance
the literature on the thermal effectiveness of texturing is not possible between the curves is not constant, which means that for a given ge­
because it is a factor that is highly dependent on the operating condi­ ometry, the performance enhancement of the textured surface heat sink
tions of the heat sink of each case study, and improvements can be found is greater for thermal scenarios consisting of high convection and
from 18 % [35] to 500 % [36]. elevated power source temperature. For example, at a convection co­
efficient of 80 W/m2 ◦ C, the average temperature of the top faces can be
3.4.1. Parametric study of the textured heat sink performance reduced by 26 % (from 106.8 ◦ C to 79.1 ◦ C) by using the textured surface
The effect of the textured surface on the performance of the heat sink for a 150 ◦ C heat source, compared to a 15 % reduction (from 40.5 ◦ C to
depends on the operating conditions. Therefore, a parametric study is 34.5 ◦ C) for a 50 ◦ C heat source. This result suggests that textured sur­
performed numerically for different values of the temperature imposed faces on heat sinks have greater potential in the most demanding ap­
on the bottom surface (50 ◦ C–150 ◦ C) and the convection coefficient plications such as CPUs and GPUs in the computer industry. This
(20–80 W/m2 ◦ C), using the same heat sinks presented in Fig. 9. Here, 30 behavior is also demonstrated in Fig. 11 for total power dissipated as a
configurations were evaluated for each heat sink (smooth and textured). function of convection coefficient and operating temperature. Textured
The results are analyzed in terms of the average temperature of the top heat sinks outperform smoothed heat sinks in all cases. The difference
faces, which is the coldest region of the heat sink as shown in Fig. 9. between the dissipated power of textured and non-textured heat sinks is
Fig. 10 shows the average temperature obtained from the simula­ nearly constant for the same imposed temperatures, regardless of the
tions for the top faces as a function of the convection coefficient and for convection coefficient, suggesting that the dissipated power of textured
three values of the imposed temperature in the bottom face. Although heat sinks will outperform traditional heat sink designs in all convection
the curves are not perfectly lineal, their curvature is small, which sup­ conditions (natural or forced).
ports the selection of the number of samples. The higher surface area of The metrics thermal resistance and effectiveness have been used in
the textured heat sink with respect to the smooth heat sink allows an this work to quantify the thermal performance of elements with
increase in heat dissipation capabilities and a lower overall operating extended surfaces. Fig. 12 illustrates the convective thermal resistance
temperature, which is evidenced by the position of the curves for the for both the smooth and textured types of surfaces as a function of the
textured heat sink (always below the curves for the smooth heat sink) convection coefficient. Based on the observed trend, the textured surface

Fig. 10. Average temperature of the top faces of the heat sinks as a function of the convection coefficient and different temperatures of the heat source – Numerical
simulations.

10
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

Fig. 11. Heat sink power dissipation as a function of convection coefficient and different heat source temperatures – Numerical simulations on a quarter model.

Fig. 12. Heat sink thermal resistance as a function of convection coefficient– Numerical simulations on a quarter model.

demonstrates a reduction of 34.4 % in thermal resistance compared to Similarly, Fig. 13 shows the effect of texture of the heat sink effec­
the smooth surface, giving it an advantage throughout the simulated tiveness, in this case it was possible to obtain between 21 % and 40 % in
convective coefficient range. This trend, which is supported by existing the performance of the fin when evaluated with this term. It can be seen
literature [32–34], further reinforces the convenience of this method. that the effect of the heat sink effectiveness is more representative when
Similarly, the effect is determined by the higher surface area of the working with low convection coefficients, i.e. in natural convection.
textured surface under the same convective coefficient.

Fig. 13. Heat sink effectiveness as a function of convection coefficient – Numerical simulations on a quarter model.

11
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

3.4.2. Sensitivity analysis study experimentally and numerically examines the impact of surface
One of the main advantages of the parametric analysis by numerical texturing on heat dissipation in heat sinks, with the goal of achieving
design of experiments described above is the definition of a response higher dissipation without increasing size or operating temperature.
surface adapted to the selected operating conditions. The response sur­ The parametric analysis indicates that the textured heat sink out­
face allows to quantify the sensitivity of the input factors on the response performs the plain heat sink since lower operating temperatures (a dif­
variable, i.e. the sensitivity of the minimum temperature of the heat sink ference of more than 12 ◦ C or 17 % lower temperature level), and
to the convection coefficient and the input temperature. Fig. 14 shows increased heat dissipation (up to 28.6 % of increase in power dissipa­
the local sensitivity at 150 ◦ C and two values of forced convection (50 tion), which is evidenced by a reduction of 34.4 % in thermal resistance,
W/m2 ◦ C and 80 W/m2 ◦ C). The first relevant aspect of the sensitivity and a heat sink effectiveness ranging from 21 % to 40 %. The heat sink
analysis is the difference in the effect of the input temperature and the textures offer the additional benefit of reducing the weight by approx­
convection coefficient on the minimum temperature of the heat sink. For imately 12 %, which is especially useful for mobile applications that
the case of the smooth heat sink, the magnitude of the local sensitivity of currently cannot utilize liquid cooling. Texturing the heat sink surfaces
the input temperature (68.4 %) is more than twice the sensitivity of the improves all of these performance metrics by increasing the surface area
convection coefficient (31.6 %). This difference suggests that the effect by up to 50 % compared to plain heat sinks of the same size. The nu­
of an increase in the input temperature cannot be compensated by the merical design of experiments allowed for a sensitivity analysis,
value of the convection coefficient alone but requires changes in other revealing the potential of textured heat sinks for applications requiring
variables such as the geometry. high levels of heat dissipation. Thus, textured surfaces provide greater
Conversely, the difference between the two input factors is smaller benefits at higher temperatures and convection coefficients. This design
for the textured heat sink. The difference between the sensitivity of the methodology can be used to design heat sinks for specific applications,
input temperature and the convection coefficient is 29 % for 50 W/m2 ◦ C where expanding the heat sink’s size or the cooling fluid’s flow velocity
and 1 % for 80 W/m2 ◦ C for the textured heat sink, while this difference is not practically feasible due to weight, space and/or energy
is 151 % for 50 W/m2 ◦ C and 116 % for 80 W/m2 ◦ C for the smooth heat constraints.
sink. Therefore, if the source temperature and the overall dimensions of Moreover, statistical analysis of the surface temperatures for each
the heat sink are kept constant, the textured heat sink allows a signifi­ texture and convection type revealed that the horizontally textured
cant increase in performance by changing the convection coefficient. In surface was the most appropriate for both natural and forced convec­
fact, the sensitivity of the convection coefficient is higher than the tion. However, the study of pressure drop and other related fluid­
sensitivity of the input temperature for the textured heat sink when the –structure interactions is essential for heat dissipation applications (flow
highest convection coefficient is used. Finally, this result was found to be shadowing, turbulence generation by the groove edges, thermal resis­
relevant for high values of the input source temperature (the highest tance of the fin relative to the air boundary, effective heat transfer co­
value of the explored design space), confirming that the potential of efficient, inclusion of radiation effect, etc.). Further research is needed
textured heat sinks increases for applications with high energy dissipa­ to understand in detail the mechanisms of textured heat sinks that
tion requirements. Furthermore, this demonstrates that textured fin enhance heat dissipation. In addition, optimizing the textured heat sink
surfaces of air-cooled heat sinks can extend their capabilities for high parameters would be appropriate (geometric patterns, as well as their
performance thermal applications, as previous research has shown for depth and width). Moreover, the disturbance of the measurement field
texturing at the base of the heat sink [32–34]. due to the presence of the thermocouples and the implications it has on
temperature could be improved and infrared thermography will be
4. Conclusions considered in future works. Finally, the development of a mass pro­
duction process could alleviate the time and cost difficulties associated
Advancements in the electronics industry have increased the demand with the manufacture of textured heat sinks.
for more efficient heat dissipation from cooling devices. Higher energy
density of transistors results in a necessity for more efficient heat sinks, Authors’ contributions
where efficiency is determined by the ratio of heat transfer to size. Thus,
the design of heat sinks usually focuses on improving the performance E.F.O.: Software, Visualization, Formal analysis, Writing – original
by modifying the geometry, the flow of the cooling fluid, by proposing draft, Writing – review & editing. F.J.R.G.: Contributed to the analysis
innovative materials, or a combination of these factors. The present and writing of the paper. A.D.M: Validation, Methodology, Formal

Fig. 14. Local sensitivity analysis of convection coefficient and input temperature on the temperature of the cold face of the heat sink – Input temperature 150 ◦ C
(results on a quarter model).

12
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

analysis, Writing – original draft and Writing – review & editing. L.M.B.: [15] X.F. Peng, G.P. Peterson, Convective heat transfer and flow friction for water flow
in microchannel structures, Int. J. Heat Mass Transf. 39 (12) (1996) 2599–2608,
Methodology and Formal analysis. J.S.R.: Methodology, Formal anal­
https://doi.org/10.1016/0017-9310(95)00327-4.
ysis and Writing – review & editing. L.C.O.V.: Conceptualization, [16] I. Hassan, P. Phutthavong, M. Abdelgawad, Microchannel heat sinks: an overview
Validation, Methodology, Formal analysis, Writing – original draft and of the state-of-the-art, Microscale Thermophys. Eng. 8 (3) (2004) 183–205, https://
Writing – review & editing. doi.org/10.1080/10893950490477338.
[17] D.B. Tuckerman, R.F.W. Pease, High-performance heat sinking for VLSI, IEEE
Electron Device Lett. EDL-2 (5) (1981) 126–129, https://doi.org/10.1109/
Declaration of Competing Interest EDL.1981.25367.
[18] D.B. Tuckerman, Heat-Transfer Microstructures for Integrated Circuits, University
of California, Livermore, California, 1984. Ph.D. Thesis.
The authors declare the following financial interests/personal re­ [19] R.J. Phillips, Microchannel heat sinks, Lincoln Lab. J. 1 (1) (1988) 31–48.
lationships which may be considered as potential competing interests: [20] A. Mohammed Adham, N. Mohd-Ghazali, R. Ahmad, Thermal and hydrodynamic
analysis of microchannel heat sinks: a review, Renew. Sustain. Energy Rev. 21
Esteban Foronda Obando reports financial support was provided by
(2013) 614–622, https://doi.org/10.1016/J.RSER.2013.01.022.
Pascual Bravo University Institution. [21] S.V. Garimella, C.B. Sobhan, Transport in microchannels - a critical review, Ann.
Rev. Heat Transfer 13 (13) (2003) 1–50, https://doi.org/10.1615/
AnnualRevHeatTransfer.v13.30.
Data availability
[22] Y. Peles, A. Koşar, C. Mishra, C.-J. Kuo, B. Schneider, Forced convective heat
transfer across a pin fin micro heat sink, Int. J. Heat Mass Transf. 48 (17) (2005)
The datasets generated or analyzed during the study are available 3615–3627, https://doi.org/10.1016/j.ijheatmasstransfer.2005.03.017.
from the corresponding author upon a reasonable request. [23] H.M. Metwally, R.M. Manglik, Enhanced heat transfer due to curvature-induced
lateral vortices in laminar flows in sinusoidal corrugated-plate channels, Int. J.
Heat Mass Transf. 47 (10–11) (2004) 2283–2292, https://doi.org/10.1016/j.
Acknowledgments ijheatmasstransfer.2003.11.019.
[24] Y. Sui, C.J. Teo, P.S. Lee, Y.T. Chew, C. Shu, Fluid flow and heat transfer in wavy
microchannels, Int. J. Heat Mass Transf. 53 (13–14) (2010) 2760–2772, https://
The authors would like to express their gratitude for the financial doi.org/10.1016/j.ijheatmasstransfer.2010.02.022.
support provided by the Institución Universitaria Pascual Bravo with the [25] H.A. Mohammed, P. Gunnasegaran, N.H. Shuaib, Numerical simulation of heat
project IN202003“Desarrollo de un código de elementos finitos usando transfer enhancement in wavy microchannel heat sink, Int. Commun. Heat Mass
Transfer 38 (1) (2011) 63–68, https://doi.org/10.1016/J.
Python y computación paralela en GPU para la evaluación de disipadores de ICHEATMASSTRANSFER.2010.09.012.
calor de alta complejidad geométrica”. We express our gratitude to the [26] K. Bilen, U. Akyol, S. Yapici, Heat transfer and friction correlations and thermal
Manufacturing and CNC Laboratory at Institución Universitaria Pascual performance analysis for a finned surface, Energy Convers. Manag. 42 (9) (2001)
1071–1083, https://doi.org/10.1016/S0196-8904(00)00119-9.
Bravo, and particularly to Javier Mejía and Esteban Rave. [27] E.K. Kalinin, G.A. Dreitser, Heat transfer enhancement in heat exchangers, Adv.
Heat Transf. 31 (C) (1998) 159–332, https://doi.org/10.1016/S0065-2717(08)
References 70242-9.
[28] S. Rashidi, F. Hormozi, B. Sundén, O. Mahian, Energy saving in thermal energy
systems using dimpled surface technology – a review on mechanisms and
[1] M.M. Sarafraz, V. Nikkhah, M. Nakhjavani, A. Arya, Thermal performance of a heat
applications, Appl. Energy 250 (2019) 1491–1547, https://doi.org/10.1016/J.
sink microchannel working with biologically produced silver-water nanofluid:
APENERGY.2019.04.168.
experimental assessment, Exp. Therm. Fluid Sci. 91 (2018) 509–519, https://doi.
[29] S.V. Garimella, P.A. Eibeck, Heat transfer characteristics of an array of protruding
org/10.1016/j.expthermflusci.2017.11.007.
elements in single phase forced convection, Int. J. Heat Mass Transfer 33 (12)
[2] F.J. Ramírez-Gil, C.M. Pérez-Madrid, E.C.N. Silva, W. Montealegre-Rubio, Parallel
(1990) 2659–2669, https://doi.org/10.1016/0017-9310(90)90202-6.
computing for the topology optimization method: performance metrics and energy
[30] A. Sohal, K. Kumar, R. Kumar, Heat transfer enhancement with channel surface
consumption analysis in multiphysics problems, Sustain. Comput. Inf. Syst. 30
roughness: a comprehensive review, Proc. Inst. Mech. Eng. C: J. Mech. Eng. Sci.
(2021), 100481, https://doi.org/10.1016/J.SUSCOM.2020.100481.
236 (11) (2022) 6308–6334.
[3] D. Borba Marchetto, D. Carneiro Moreira, G. Ribatski, A review on polymer heat
[31] D. Patel, V. Jain, J. Ramkumar, Micro texturing on metallic surfaces: state of the
sinks for electronic cooling applications, in: Proceedings of the 17th Brazilian
art, Proc. Inst. Mech. Eng. B: J. Eng. Manuf. 232 (6) (2018) 941–964, https://doi.
Congress of Thermal Sciences and Engineering, ABCM, São Paulo, 2018, doi:
org/10.1177/0954405416661583.
10.26678/ABCM.ENCIT2018.CIT18-0394.
[32] G. Refai-Ahmed, B. Philofsky, V. Gektin, B. Sammakia, H. Do, S. Rangarajan, Best
[4] Z. Zhang, X. Wang, Y. Yan, A review of the state-of-the-art in electronic cooling, e-
engineering practices to establish cooling limit for 375W Add-in PCI-e center
Prime – Adv. Electr. Eng. Electron. Energy 1 (2021), 100009, https://doi.org/
accelerator card with active optical, in: 2019 IEEE 21st Electronics Packaging
10.1016/J.PRIME.2021.100009.
Technology Conference, EPTC 2019, Dec. 2019, pp. 341–348, doi: 10.1109/
[5] A. Benallou, Heat exchangers for electronic equipment cooling, in: L. Castro
EPTC47984.2019.9026618.
Gómez, V. Manuel Velázquez Flores, M. Navarrete Procopio (Eds.), Heat
[33] G. Refai-Ahmed et al., Establishing thermal air-cooled limit for high performance
Exchangers, IntechOpen, 2022.
electronics devices, in: 2020 IEEE 22nd Electronics Packaging Technology
[6] F.J. Ramírez-Gil, Á. Delgado-Mejía, E. Foronda-Obando, L.C. Olmos-Villalba,
Conference, EPTC 2020, Dec. 2020, pp. 347–354, doi: 10.1109/
Thermal finite element analysis of complex heat sinks using open source tools and
EPTC50525.2020.9315139.
high-performance computing, Revista Facultad de Ingeniería Universidad de
[34] G. Refai-Ahmed, S. Ramalingam, B.D. Philofsky, Thermal management device with
Antioquia (2022), https://doi.org/10.17533/udea.redin.20220888.
textured surface for extended cooling limit, U.S. Patent No. 9,812,374, 2017
[7] Z. Khattak, H.M. Ali, Air cooled heat sink geometries subjected to forced flow: a
Accessed: Jul. 25, 2023 [Online], Available: https://patents.google.com/pa
critical review, Int. J. Heat Mass Transf. 130 (2019) 141–161.
tent/US9812374B1/en.
[8] I.U. Perera, N. Narendran, Analysis of a remote phosphor layer heat sink to reduce
[35] S. T, S.K. D, Heat transfer study on different surface textured pin fin heat sink, Int.
phosphor operating temperature, Int. J. Heat Mass Transf. 117 (2018) 211–222,
Commun. Heat Mass Transfer 119 (2020), 104902.
https://doi.org/10.1016/j.ijheatmasstransfer.2017.10.016.
[36] G. Dhadda, M. Hamed, P. Koshy, Electrical discharge surface texturing for
[9] S. Husain, S.A. Khan, A review on heat transfer enhancement techniques during
enhanced pool boiling heat transfer, J. Mater. Process. Technol. 293 (2021),
natural convection in vertical annular geometry, Clean Eng. Technol. 5 (2021),
117083, https://doi.org/10.1016/j.jmatprotec.2021.117083.
100333, https://doi.org/10.1016/J.CLET.2021.100333.
[37] G. Liang, I. Mudawar, Review of pool boiling enhancement by surface
[10] R.V Unni, V.S. Majali, A review on rectangular heat sinks under natural convection,
modification, Int. J. Heat Mass Transf. 128 (2019) 892–933, https://doi.org/
Int. J. Adv. Sci. Technol. 27(7) (2019) 144–149, Accessed: Aug. 16, 2022. [Online],
10.1016/j.ijheatmasstransfer.2018.09.026.
Available: http://sersc.org/journals/index.php/IJAST/article/view/434.
[38] Y. Zhu, D.S. Antao, E.N. Wang, Bioinspired surfaces for enhanced boiling, in:
[11] N.A.C. Sidik, M.N.A.W. Muhamad, W.M.A.A. Japar, Z.A. Rasid, An overview of
Bioinspired Engineering of Thermal Materials, Wiley-VCH Verlag GmbH & Co.
passive techniques for heat transfer augmentation in microchannel heat sink, Int.
KGaA, Weinheim, Germany, 2018, pp. 47–71, doi: 10.1002/9783527687596.ch3.
Commun. Heat Mass Transfer 88 (2017) 74–83, https://doi.org/10.1016/J.
[39] I.A. Ghani, N.A.C. Sidik, R. Mamat, G. Najafi, T.L. Ken, Y. Asako, W.M.A.A. Japar,
ICHEATMASSTRANSFER.2017.08.009.
Heat transfer enhancement in microchannel heat sink using hybrid technique of
[12] H.E. Ahmed, B.H. Salman, A.S. Kherbeet, M.I. Ahmed, Optimization of thermal
ribs and secondary channels, Int. J. Heat Mass Transf. 114 (2017) 640–655.
design of heat sinks: a review, Int. J. Heat Mass Transf. 118 (2018) 129–153,
[40] G. Wang, D. Niu, F. Xie, Y. Wang, X. Zhao, G. Ding, Experimental and numerical
https://doi.org/10.1016/j.ijheatmasstransfer.2017.10.099.
investigation of a microchannel heat sink (MCHS) with micro-scale ribs and
[13] C.T. Nguyen, G. Roy, C. Gauthier, N. Galanis, Heat transfer enhancement using
grooves for chip cooling, Appl. Therm. Eng. 85 (2015) 61–70, https://doi.org/
Al2O3–water nanofluid for an electronic liquid cooling system, Appl. Therm. Eng.
10.1016/j.applthermaleng.2015.04.009.
27 (8) (2007) 1501–1506, https://doi.org/10.1016/j.
[41] S.W. Chang, T.M. Liou, M.H. Lu, Heat transfer of rectangular narrow channel with
applthermaleng.2006.09.028.
two opposite scale-roughened walls, Int. J. Heat Mass Transf. 48 (19–20) (2005)
[14] X.Q. Wang, A.S. Mujumdar, A review on nanofluids - part II: experiments and
3921–3931, https://doi.org/10.1016/J.IJHEATMASSTRANSFER.2005.04.015.
applications, Braz. J. Chem. Eng. 25 (4) (2008) 631–648, https://doi.org/10.1590/
S0104-66322008000400002.

13
E. Foronda et al. Thermal Science and Engineering Progress 46 (2023) 102170

[42] C. Greiner, M. Schäfer, Bio-inspired scale-like surface textures and their [51] B. Bhushan, Biomimetics: lessons from nature -an overview, Philos. Trans. R. Soc.
tribological properties, Bioinspir. Biomim. 10 (4) (2015), 044001, https://doi.org/ A: Math. Phys. Eng. Sci. 367 (1893) (2009) 1445–1486, https://doi.org/10.1098/
10.1088/1748-3190/10/4/044001. RSTA.2009.0011.
[43] P. Cuervo, D.A. López, J.P. Cano, J.C. Sánchez, S. Rudas, H. Estupiñán, A. Toro, H. [52] I. Pope, J.P. Hidalgo, R.M. Hadden, J.L. Torero, A simplified correction method for
A. Abdel-Aal, Development of low friction snake-inspired deterministic textured thermocouple disturbance errors in solids, Int. J. Therm. Sci. 172 (2022), 107324,
surfaces, Surf. Topogr. 4 (2) (2016), 024013, https://doi.org/10.1088/2051-672X/ https://doi.org/10.1016/J.IJTHERMALSCI.2021.107324.
4/2/024013. [53] D. Ross-Pinnock, P.G. Maropoulos, Identification of key temperature measurement
[44] A. Toro, H.A. Abdel-Aal, E. Zuluaga, P. Cuervo, L.M. Ballesteros, J.C. Sánchez, J. technologies for the enhancement of product and equipment integrity in the light
S. Rudas, C. Isaza, W.Z. Misiolek, Influence of surface morphology and internal controlled factory, Proc. CIRP 25 (2014) 114–121.
structure on the mechanical properties and tribological response of Boa Red Tail [54] T. Masuzawa, State of the art of micromachining, CIRP Ann. 49 (2) (2000)
and Python Regius snake skin, J. Mech. Behav. Biomed. Mater. 119 (2021), 473–488, https://doi.org/10.1016/S0007-8506(07)63451-9.
104497, https://doi.org/10.1016/j.jmbbm.2021.104497. [55] T. Pratap, K. Patra, Micro–nano surface texturing, characterization, and their
[45] E. Bormashenko, Wetting transitions on biomimetic surfaces, Philos. Trans. R. Soc. impact on biointerfaces, in: Advanced Machining and Finishing, Elsevier, 2021, pp.
A: Math. Phys. Eng. Sci. 369 (1941) (2011) 1712, https://doi.org/10.1098/ 577–610, doi: 10.1016/B978-0-12-817452-4.00008-7.
rsta.2011.0038. [56] J.N. Reddy, D.K. Gartling, The Finite Element Method in Heat Transfer and Fluid
[46] L.M. Ballesteros, E. Zuluaga, P. Cuervo, J.S. Rudas, A. Toro, Tribological behavior Dynamics, third edition, 2010, doi: 10.1201/9781439882573.
of polymeric 3D-printed surfaces with deterministic patterns inspired in snake skin [57] A.I. Sato, C.A.C. Altemani, V.L. Scalon, Mean Nusselt number correlation for tise
morphology, Surf. Topogr. 9 (1) (2021), 014002, https://doi.org/10.1088/2051- heatsink thermal design, Revista de Engenharia Térmica 19 (1) (2020) 24, https://
672x/abe211. doi.org/10.5380/reterm.v19i1.76427.
[47] S.S. Banwait, H.N. Chandrawat, Study of thermal boundary conditions for a plain [58] J.C. Sánchez, H. Estupiñán, A. Toro, Friction response of bioinspired AISI 52100
journal bearing, Tribol. Int. 31 (6) (1998) 289–296, https://doi.org/10.1016/ steel surfaces texturized by photochemical machining, Surf. Topogr. 9 (1) (2021),
S0301-679X(98)00029-2. 014001, https://doi.org/10.1088/2051-672X/abe090.
[48] N. Tala-Ighil, M. Fillon, P. Maspeyrot, Effect of textured area on the performances [59] H.A. Abdel-Aal, R. Vargiolu, H. Zahouani, M. El Mansori, Preliminary investigation
of a hydrodynamic journal bearing, Tribol. Int. 44 (3) (2011) 211–219, https://doi. of the frictional response of reptilian shed skin, Wear 290–291 (2012) 51–60,
org/10.1016/J.TRIBOINT.2010.10.003. https://doi.org/10.1016/j.wear.2012.05.015.
[49] A. Fatima, P.T. Mativenga, On the comparative cutting performance of nature- [60] D.D. Ma, G.D. Xia, Y.F. Li, Y.T. Jia, J. Wang, Effects of structural parameters on
inspired structured cutting tool in dry cutting of AISI/SAE 4140, Proc. Inst. Mech. fluid flow and heat transfer characteristics in microchannel with offset zigzag
Eng. - Part B: J. Eng. Manuf. 231 (11) (2017) 1941–1948, https://doi.org/ grooves in sidewall, Int. J. Heat Mass Transf. 101 (2016) 427–435, https://doi.org/
10.1177/0954405415617930. 10.1016/J.IJHEATMASSTRANSFER.2016.04.091.
[50] N. Tala-Ighil, P. Maspeyrot, M. Fillon, A. Bounif, Effects of surface texture on
journal-bearing characteristics under steady-state operating conditions, Proc. Inst.
Mech. Eng., Part J: J. Eng. Tribol. 221 (6) (2007) 623–633.

14

You might also like