You are on page 1of 11

International Journal of Heat and Mass Transfer 181 (2021) 121899

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Thermal performance enhancement of phase change material heat


sinks for thermal management of electronic devices under constant
and intermittent power loads
Huizhu Yang∗, Yongyao Li, Liang Zhang, Yonggang Zhu
School of Mechanical Engineering and Automation, Harbin Institute of Technology, Shenzhen, Shenzhen 518055, China

a r t i c l e i n f o a b s t r a c t

Article history: Metal fin enhanced phase change material (PCM) has been successfully demonstrated as an effective ap-
Received 30 June 2021 proach for electronic devices cooling, especially in the case of high-power density for only a limited
Revised 14 August 2021
period. Application of PCMs in the heat sink accelerates the heat storage and reduces the heat rejection
Accepted 24 August 2021
resulting in a tradeoff effect. In this study, the effects of PCM filling ratio, intermittent power load with
Available online 1 September 2021
different power off time interval and heat transfer coefficient on the performance of PCM-filled heat sinks
Keywords: are studied for electronic devices cooling. A constant power load is firstly studied to reveal the effect of
Phase change material PCM filling ratio. An intermittent power load with different power off time interval is then investigated to
Heat sink enlarge the operation time of electronic devices. Finally, the effect of heat transfer coefficient is discussed
Intermittent power load to assess the negative and positive effects of PCM. The results show that the effective operation time of
Electronic device PCM-filled heat sinks significantly increases and its peak temperature obviously reduces under intermit-
Passive cooling
tent power load compared to the constant power load, and it becomes more apparent with increasing
power off time interval. The thermal performance of PCM-filled heat sinks significantly strengthens un-
der natural cooling condition, i.e., h = 10 or 20 w•m−2 •K−1 . However, during forced convection cooling,
e.g., h = 100 w•m−2 •K−1 , the heat sink without PCM may have a better performance because the nega-
tive effect of PCM outweighs its positive effect. The results of this paper are of great significance in the
optimal design of PCM-filled heat sink.
© 2021 Elsevier Ltd. All rights reserved.

1. Introduction Organic PCMs such as typically paraffin and n-eicosane, are


characterized by low melting temperature, large latent heat, non-
The greater miniaturization and ongoing performance enhance- corrosiveness and low cost that are well suited for electronic de-
ments of electronic devices have led to extremely high increase in vices cooling. However, the main defect of the organic PCMs is a
power density. Such a high-power density further yields a high low thermal conductivity, which severely restricts the heat dissipa-
heat generation which has adverse effects on the electronic de- tion and causes the temperature rising to exceed the critical value
vices’ reliability and performance. To address this challenge, an ef- of electronic devices. Many efforts were carried out to improve the
fective thermal management approach is very imperative for guar- thermal conductivity of PCMs by adding thermal conductivity en-
anteeing the electronic devices working efficiency and safety. Var- hancers, such as extended metal fins and pins [7–11], porous metal
ious thermal management techniques such as natural/forced air materials [12–16], nanoparticles [17,18], woven metal fiber [19] and
cooling [1,2], liquid cooling [3], boiling [4] and heat pipe [5] have heat pipes [20–22]. Among these, the extended metal fins are an
been reported. The application of Phase Change Materials (PCMs) excellent thermal conductivity enhancer owing to its high thermal
as a passive thermal management technique has gained tremen- conductivity, easy fabrication and low cost.
dously increasing attention in the last decade. PCMs are known as Many researchers have investigated the heat transfer enhance-
a very promising passive cooling technique for effective thermal ment by addition of metal fins [23,24]. Yang et al. [25] reported
management of electronic devices owing to its huge latent heat a composite PCM heat sink to manage the thermal issues of high
during phase change process, relatively constant phase transition power GaN devices. The composite PCM heat sink device includes
temperature and chemical stability [6]. the thermal spreader, thermal buffer (PCM) and fin array. Com-
pared with the copper heat sink, the device junction tempera-
∗ ture was reduced by 10–20 °C in the composite PCM heat sink.
Corresponding author.
E-mail address: yanghuizhu@hit.edu.cn (H. Yang). Kalbasi [26] numerically studied the cooling ability of a novel hy-

https://doi.org/10.1016/j.ijheatmasstransfer.2021.121899
0017-9310/© 2021 Elsevier Ltd. All rights reserved.
H. Yang, Y. Li, L. Zhang et al. International Journal of Heat and Mass Transfer 181 (2021) 121899

[44] employed Kriging model to confirm the optimum design pa-


Nomenclature rameters of PCM-based pin fin heat sink.
From the above literatures, there is no doubt that the metal
Latin symbols fin enhanced PCM is highly effective for improving the thermal
cp heat capacity, J•K−1 •kg−1 management of electronic devices. Many optimization studies have
h convective heat transfer coefficient, W•m−2 •K−1 been done to improve the thermal performance of PCM-filled heat
H total enthalpy, kJ•kg−1 sinks for electronic devices cooling. Besides, combining the advan-
H latent heat, kJ•kg−1 tages of different PCMs, including its large latent heat and low
P pressure, Pa melting temperature, multiple PCMs based heat sink has been
q power load, W•m−2 achieved better performance. However, these studies are mainly
Q total energy storage, kJ focused on using phase change materials to store more heat to
t time, s extend the operating time of electronic devices under constant
T temperature, °C power load. Very few studies combined PCM heat storage with
u, v velocity, m•s−1 natural/forced air cooling to extend the operation time of elec-
tronic devices under intermittent power load condition. In practice,
Greek symbols
this study is even more crucial for electronic devices, especially the
ρ density, kg•m−3
mobile device, in which a high power is applied for only a limited
k Thermal conductivity, W•m−1 •K−1
period. Ren et al. [45] numerically investigated the thermal perfor-
μ dynamic viscosity, kg•m−1 •s−1
mance of PCM-2D stacked metal fiber composite-based heat sink
γ PCM filling ratio
under continuous and intermittent heat flux conditions. The results
β thermal expansion coefficient, K−1
showed that heat sink with anisotropic stacked 2D metal fibers
ξ melt fraction
obviously improves the thermal performance of heat sink under
ψ time extension ratio
intermittent working condition. Baby and Balaji [46] experimental
τ energy storage rate
studied the effect of same heat input with intermittent heat loads
Subscripts on PCM based heat sink. The results indicated that the design size
amb ambient of PCM based heat sink under intermittent operation smaller than
ini initial that of under continuous heat duty condition. Farzanehnia et al.
l liquid [47] experimentally investigated the thermal management of nano-
s solid PCM heat sink, and it presented that the chipset peak tempera-
ture is decreased by applying more on-and-off cycles under in-
termittent heating. Gharbi et al. [48] experimentally investigated
brid heat sink filled with PCM and air separated by inner fins. the cooling performance of PCM with heat sources under intermit-
The results showed that the performance of the novel heat sink tent working regimes. The results presented that fractionate cy-
was better than the air-cooled heat sinks. Mozafari et al. [27] used cle length into 4 or 8 cycles extends respectively nearly 3 and
both single and multiple PCMs heat sinks for cooling the electronic even 5 times the critical time comparing to one cycle. Therefore,
devices. Their studies revealed that the multiple n-eicosane/RT44 a significant reduction in temperature or an extended operation
based heat sink has the best performance where its operational time were confirmed in the PCM-filled heat sink with intermittent
time is increased by 3.3 ∼ 12% comparing to single n-eicosane or power load relative to that with constant power load. A compre-
RT44. Huang et al. [28] reported a dual-PCM heat sink by using hensive study to discuss the effects of PCM filling ratio, intermit-
both Paraffin and low melting point alloy for reducing the weight tent power load with different power off time interval and heat
and cost of low melting point alloy heat sink. The results revealed transfer coefficient on the performance of PCM-based heat sink
that with proper fin structure and low melting point alloy volume for managing the thermal issues of electronic devices deserves fur-
fraction, the performance of dual-PCM heat sink was comparable ther investigation. In this study, a typical and constant power load
with low melting point alloy heat sink. Gharbi et al. [29] experi- q = 10,0 0 0 w•m−2 with natural convective heat transfer coefficient
mental studied the behavior of PCMs for electronic devices cooling h = 10 w•m−2 •K−1 is firstly studied to reveal the effect of PCM
including pure PCM, PCM/silicone matrix, PCM/graphite matrix and filling ratio on the performance of PCM-filled heat sinks. Then, the
PCM with fins. It was found that the combination of PCM and fins effect of an intermittent power load with different power off time
shows an effective means for electronic devices cooling. interval is investigated to enlarge the effective operation time of
In another aspect, many optimization studies have been con- PCM-filled heat sinks. Finally, the effect of convective heat transfer
ducted to further improve and maximize the performance of PCM- coefficient is studied to assess the negative effect of PCM due to
filled heat sinks. Levin et al. [30] numerically researched the ef- its low thermal conductivity and the positive effect of PCM latent
fect of PCM volume fraction, fin number and fin thickness, in heat storage.
which minimizing the heat sink height and maximizing operation
time were taken as the objective functions. Arshad et al. [31] ex- 2. Theoretical model
perimental investigated the effects of fin thickness and PCM vol-
ume fraction for enhancing the operating time of PCM based heat 2.1. Physical model
sinks. A similar parametric approach was presented to optimize the
PCM-based heat sink design, in which the parameters include fin Fig. 1 schematically shows the electronic device cooling unit
shape, fin arrangement, fin height, fin thickness, orientation, num- equipped with PCM-filled heat sink. The length and width of the
ber of fins, PCM volumetric fraction and so on [32–40]. In ad- chip are assumed as 50 mm and 50 mm, respectively. The uni-
dition, Iradukunda et al. [41] used the Topology optimization to form heat flux of the chip is applied to the bottom wall of PCM-
obtain optimal structures for PCM-based heat sink. Kalbasi et al. filled heat sink. In this study, the average bottom temperature of
[42] used the regression model to optimize the design parame- PCM-filled heat sink is adopted to monitor the allowable temper-
ters of PCM-based heat sink for electronic devices cooling. Baby ature of the chip since the thermal contact resistance between
and Balaji [43] combined Genetic algorithm and Artificial Neural the chip and the PCM-filled heat sink is neglected. The maxi-
Network for optimizing PCM based heat sink. Dammak and Hami mum allowable bottom temperature of the chip is set to 80 °C,

2
H. Yang, Y. Li, L. Zhang et al. International Journal of Heat and Mass Transfer 181 (2021) 121899

Table 1.
Thermophysical properties of the PCM and heat sink material [49].

Properties n-eicosane Aluminum

Density ρ (kg•m−3 ) 781 (solid) 871


759 (liquid)
Heat capacity cp (J•K−1 •kg−1 ) 2100 (solid) 2719
2420 (liquid)
Thermal conductivity k (W•m−1 •K−1 ) 0.39 (solid) 202.4
0.16 (liquid)
Dynamic viscosity μ (kg•m −1 • −1
s ) 0.0038
Latent heat H (kJ•kg−1 ) 268.2
Thermal expansion coefficient β (K−1 ) 9.4 × 10−4
PCM solidus temperature Ts (°C) 35
PCM liquidus temperature Tl (°C) 42

The continuity equation:


∂ρ ∂ (ρ u ) ∂ (ρv )
+ + =0 (1)
∂t ∂x ∂y
The momentum equation:
   
∂ (ρ u ) ∂ (ρ uu ) ∂ (ρ uv ) ∂P ∂ ∂u ∂ ∂u
+ + =− + μ + μ + Fx
∂t ∂x ∂y ∂x ∂x ∂x ∂y ∂y
(2)

 
∂ ( ρ u ) ∂ ( ρ uv ) ∂ ( ρ uv ) ∂P ∂ ∂v
+ + =− + μ
∂t ∂x ∂y ∂y ∂x ∂x
 
∂ ∂v  
+ μ + ρ gβ T − Tre f + Fy (3)
Fig. 1. Schematic of the PCM-filled heat sink and computational domain. ∂y ∂y
The energy equation:
   
which is the most encountered electronic devices specification. The ∂ (ρ H ) ∂ (ρ uH ) ∂ (ρvH ) ∂ ∂T ∂ ∂T
+ + = α + α (4)
computational domain of the PCM-filled heat sink is simplified as ∂t ∂x ∂y ∂x ∂x ∂y ∂y
a two-dimensional structure illustrated in Fig. 1. The PCM-filled
heat sink consists of 2-cavity plate fin with interior dimensions of where ρ is the PCM density, u and v are respectively the superficial
50 mm in height, 2 mm in thickness and 14 mm in fin spacing. velocities in the x and y directions, μ is the dynamic viscosity, P is
The heat sink is made from aluminum slabs using milling process. the pressure, g is the gravitational acceleration, T is the tempera-
The PCM used for this study is n-eicosane which is characteris- ture, Tref is the reference temperature taken as 38.5 °C (the average
tic of non-corrosiveness, large latent heat as well as appropriate value of PCM melting point) in this study, β is the thermal expan-
phase change temperature for electronic devices cooling. The ther- sion coefficient, α = k/(ρ cp ) is the thermal diffusivity, Fx , and Fy
mophysical properties of n-eicosane and aluminum are presents in are sources terms, and H is the total enthalpy of PCM.
Table 1 [49]. The source terms Fx and Fy in the momentum equations are
To reveal the effect of PCM filling ratio on the performance of Darcy Law damping terms accounting for the effect of phase
PCM-filled heat sinks, six different cases are investigated in this change on convection. They are given as:
study. The heat sink without PCM is designed as the base case. (1 − ξ )2
The PCM filling ratio (γ = h/H) of Case 1, Case 2, Case 3, Case 4 Fx = − A u (5)
ξ 3 + ε mush
and Case 5 is 10%, 30%, 50%, 70% and 90%, respectively.
( 1 − ξ )2
Fy = − A v (6)
2.2. Mathematical model ξ 3 + ε mush
where Amush is mushy zone constant which basically ranges from
The transient phase change of PCM is simulated as laminar, 104 –107 [51]. An appropriate can be confirmed by comparing the
incompressible and Newtonian fluid. In the momentum equa- experimental data with the numerical values. ε = 0.001 is a small
tions, the Boussinesq approximation is used to study the buoy- number to prevent division by zero for when ξ = 0, and ξ is the
ancy source term caused by the density variation effects. The melt fraction.
density of PCM is assumed constant in all terms of the govern- The melt fraction ξ is calculated in Eq. (7).
ing equations except the body force term. Consequently, volume 
0 if T ≤ Ts
change is ignored during the PCM phase change process. Hence,
the enthalpy-porosity approach can be easily used to model the
ξ= T −Ts
Tl −Ts
if Ts < T ≤ Tl (7)
phase change process of PCM with popular fixed-domain tech- 1 if T > Tl
niques. The enthalpy-porosity approach considers the solid-liquid The total enthalpy H of PCM is the sum of sensible and latent
mushy zone as a porous medium with porosity equal to the liquid enthalpies which is calculated as:
PCM volume fraction [50]. Accordingly, the porosity varies from 0 ⎧
T
to 1 when the PCM transforms from solid to liquid. The governing ⎨
Tini c ps dT if T ≤ Ts
c ps dT + ξ H
T
equations for the transient heat transfer process can be studied as H= if Ts < T ≤ Tl (8)

TTini
T
following. Tini c ps dT + ξ H + Tl c pl dT if T > Tl

3
H. Yang, Y. Li, L. Zhang et al. International Journal of Heat and Mass Transfer 181 (2021) 121899

Fig. 2. Effects of grid and time step sizes on the variation of average bottom temperature versus time for Case 3 (γ = 50%).

(14), respectively.
∂T
−k | =q (13)
∂ y y=0

−k ∂∂Ty |y=0 = q poweron
(14)
−k ∂∂Ty |y=0 = 0 poweroff

Convection heat transfer at other walls of the heat sink is cal-


culated as:
∂T
−k = h(T − Tamb ) (15)
∂n
Convection heat transfer at the top boundary of PCM is set as:
∂T
−k = h(T − Tamb ) (16)
∂n
where Tamb is the ambient temperature taken as 25 °C.

Fig. 3. Comparison of the simulation results and experimental data by Zhao et al.
3. Numerical method and validation
[53].
The governing equations describing the phase change process of
PCM are solved by the Ansys Fluent 2020 R1, in which it is itera-
where Tini is the initial temperature taken as 25 °C in this study, tively solved by the finite-volume-method with SIMPLE pressure–
H is the latent heat of PCM, cps and cpl are the specific heats velocity coupling algorithm. Diffusion and convection terms in the
of the solid and liquid PCM, respectively, and it is set as change momentum and energy equations are discretized by applying the
linearly during the phase change process. third order MUSCL scheme. Pressure correction is settled using the
The initial velocity and temperature in the PCM-filled heat sink PRSTO scheme which is the suggested method for computing nat-
are taken as: ural convection flows [52]. The convergence criterion is that the
 normalized residuals are less than 1 × 10−4 for the flow equations
Tini = 25◦ C t = 0
and 1 × 10−8 for the energy equation.
u=0 t=0 (9)
The computational domain is meshed with structured hexahe-
v t=0
dral grids using ICEM CFD. Before the numerical simulation, Case 3
No-slip assumption at the walls is given as: (γ = 50%) with constant power load q = 10,0 0 0 w•m−2 and heat
transfer coefficient h = 10 w•m−2 •K−1 is studied to inspect the
u = 0, v=0 (10) grid convergence and time-step independence. Five different grid
Coupled boundary conditions at the interface between fin and PCM sizes with 28,924, 36,262, 43,920, 47,114 and 51,737 computational
surfaces are given as: cells are investigated, named as Grid A, Grid B, Grid C, Grid D and
Grid E, respectively. The transient average bottom temperature of
∂ T f in ∂ TPCM Case 3 is shown in Fig. 2(a). The results show that Grids A and B
−k f in = −kPCM (11)
∂n ∂n reach completely liquid faster than Grids C, D and E. After the PCM
complete melting in Grids C, D and E, the deviation between the
T f in = TPCM (12) average bottom temperatures for Grid A and Grid C is 48%, while
it is less than 1% for Grid C and E. Therefore, Grid C can ensure
where n is the normal vector of interfaces. a satisfactory solution in this case. Therefore, the grid production
Constant and intermittent power load boundary conditions at method of Grid C is applied to produce grids for other models. Fur-
the bottom wall of PCM-filled heat sink are given in Eqs. (13) and ther, five different time steps with ࢞t = 0.05, 0.1, 0.2, 0.5 and 1.0 s

4
H. Yang, Y. Li, L. Zhang et al. International Journal of Heat and Mass Transfer 181 (2021) 121899

Fig. 4. Velocity, temperature and liquid fraction contours with superimposed velocity vectors in Case 3 (γ = 50%) at different times.

are tested and illustrated in Fig. 2(b). It is obvious that the time Therefore, it illustrates that the present numerical model is accept-
step of ࢞t = 0.1 s is enough to maintain the solution accuracy able and reliable.
with a reasonable computational time. In this study, the parallel
computations with 8 numbers of processes were performed in a 4. Results and discussion
workstation with CPU frequency of 2.5 GHz, and it usually took
approximately 24 h for each task. 4.1. The performance of PCM-filled heat sinks
For model validation, simulation results and the corresponding
experimental data by Zhao et al. [53] were compared in Fig. 3. In The PCM cooling is suitable for the range of heat flux
the experimental study, the paraffin wax RT58 was placed in rect- between 103 and 104 w/m−2 . So, a constant power load
angular enclosure with 200 × 120 × 25 mm in dimension and a q = 10,0 0 0 w•m−2 with natural convective heat transfer co-
uniform heat flux was imposed on the bottom surface of the rect- efficient h = 10 w•m−2 •K−1 is studied in this section. To facility
angular enclosure. y is the vertical coordinate in the computational better understanding of heat transfer mechanism, Fig. 4 shows the
domain, i.e., the distance from monitor locations to the bottom velocity, temperature and liquid fraction contours with superim-
wall surface. Note that the simulation results were obtained by av- posed velocity vectors for Case 3 (γ = 50%) at t = 30 0, 60 0 and
eraging the cross-section temperature based on the 2D geometry. 900 s during the melting process. In the early stage t = 300 s
As shown, the temperatures obtained by this numerical model are (Fig. 4(a)), heat is conducted from fins to the solid PCM that leads
in excellent agreement with those reported by Zhao et al. [53]. For to solid PCM melting and forms a thin liquid PCM layer around
most of the data, the deviations of temperature are less than 5%. the fins. Some small vortexes are formed near the bottom wall

5
H. Yang, Y. Li, L. Zhang et al. International Journal of Heat and Mass Transfer 181 (2021) 121899

Fig. 5. Variation of average bottom temperature versus time for the base case and Fig. 6. Variation of total energy storage and energy storage rate for the base case
Case 3 (γ = 50%) in constant power load. and Case 3 (γ = 50%) in constant power load.

because of the effects of non-uniform temperature distribution and


buoyancy force. On the vertical fin walls, liquid PCM flows upward,
ascends to top part and then flows downward along melting front
resulting in an inward circulating current in the liquid region,
which demonstrates the natural convection can contribute to heat
transfer and accelerate the melting rate. However, the solid-liquid
interface is almost parallel to the fin walls that the heat transfer is
dominated by heat conduction during this time.
As the melting continues, shown as Fig. 4(b), the velocity of
liquid PCM becomes larger. The flow of liquid PCM forms two
bigger circulating current. Similarly, the liquid PCM flows upward
along the vertical fin walls and then flows downward along the
solid-liquid interface. Moreover, the original small vortexes become
larger and expand to upper region. Furthermore, it can find that
more PCM becomes molten and then the effect of natural convec-
tion is enhanced. Hence, the PCM melts faster, and its temperature
is higher in the upper region than in the lower region. It is proved
that natural convection dominates the heat transfer process at this Fig. 7. Variation of average bottom temperature versus time for different PCM fill-
time. After the PCM melting completely (Fig. 4(c)), the velocity of ing ratio in constant power load.
liquid PCM continues to increase that the strong convective cur-
rents and some separate big vortexes can be found in the lower
region. The temperature of the PCM becomes more uniform and it cantly after PCM melting completely t ≥ 751 s. At t = 751 s, the
is still higher in the upper region than in the lower region due to energy storage rate of Case 3 and base case is 0.8 and 0.016, re-
the buoyancy effect. spectively. Thus, it can be understood that the heat energy of elec-
Fig. 5 presents the average bottom temperature evolution with tronic device is majorly absorbed by the PCM latent heat and a
time in the base case and Case 3 (γ = 50%). The temperature in- small amount released by natural convection though fin walls in
creases rapidly with increasing time and then stabilizes to 140 °C Case 3, while it is majorly released by natural convection in the
at t = 1400 s in the base case. However, the temperature of Case 3 base case. As a result, the temperature of Case 3 is significantly
is significantly lower than that of the base case before PCM melt- lower than that of the base case. In the later period, the heat re-
ing completely (t = 751 s). During the time, the maximum temper- jection by natural convection is smaller in Case 3 than in base case
ature of Case 3 is less than 63 °C, which allows the electronic de- due to the reduction of the heat dissipation area replaced by filled
vice to maintain a great performance. After the PCM melting com- PCM. Besides, the inherently low thermal conductivity of PCM in-
pletely, the temperature of Case 3 increases rapidly and then ex- creases the thermal resistance between PCM’s top surface and the
cess the temperature of the base case. To explain these results, the heat source. Thus, the bottom temperature of Case 3 exceeds the
heat energy analysis is carried out as follows. base case and stabilizes to 196 °C when it is balanced between
Fig. 6 shows the total energy storage Q and energy storage rate heat absorption and rejection.
τ as a function of time in the base case and Case 3. The energy To take into account the effect of PCM filling ratio on the per-
storage rate τ expresses the percentage of heat absorbed relative formance of PCM-filled heat sinks, Fig. 7 illustrates the average
to the heating power. For all the times presented, it is apparent bottom temperature evolution with time at different PCM filling
that the values of the total energy storage and energy storage rate ratio. The temperature increases rapidly with increasing time in
of Case 3 are much higher than that of the base case. Before PCM the base case. However, the temperature of Case 1, Case 2, Case
melting completely < 751 s, the energy storage of Case 3 mainly 3, Case 4 and Case 5 increases first, then stabilizes and where-
depends on the sensible and latent heat of PCM, while after PCM after increases rapidly. The stability temperatures of these curves
melting completely t ≥ 751 s, it depends on the sensible heat of are relatively similar, in which the temperatures are all less than
PCM. Therefore, the energy storage rate of Case 3 reduces signifi- 65 °C and slightly reduce with increasing the PCM filling ratio.

6
H. Yang, Y. Li, L. Zhang et al. International Journal of Heat and Mass Transfer 181 (2021) 121899

Fig. 8. Comparison of operation time and time extension ratio for the base case
and PCM-filled heat sinks in constant power load.

Fig. 10. Variation of average bottom temperature versus time for different PCM fill-
ing ratio in intermittent power load.

Fig. 9. Variation of average bottom temperature versus time for the base case and
Case 3 (γ = 50%) in intermittent power load.

Fig. 11. Comparison of effective operation times in the base case and PCM-filled
heat sink for different power load.
Fig. 8 compares the effective operation time t and time exten-
sion ratio ψ for the base case and PCM-filled heat sinks. The effec-
tive operation time is defined as the time when the average bot-
tom temperature reaches 80 °C. As shown, the effective operation
time increases with the increase of PCM filling ratio. Comparison
of the base case, the time extension ratio of Case 1, Case 2, Case 3,
Case 4 and Case 5 increases 2, 3.9, 5.5, 7.1 and 9.3 times, respec-
tively. Thus, the increased percentage in the time extension ratio
becomes slightly smaller with increasing of PCM filling ratio owing
to the inherently low thermal conductivity of PCM forms a growing
thermal resistance between PCM’s melting front and heat source.

4.2. Performance enhancement with intermittent power load

In this section, an intermittent power load (q = 10,0 0 0 w•m−2 )


with natural convective heat transfer coefficient
h = 10 w•m−2 •K−1 is studied. The average bottom tempera-
ture of the base case reaches 80 °C at t = 160 s. Hence the
intermittent power load of 10,0 0 0 w•m−2 is first set to turn on Fig. 12. Enhancement of operation time due to intermittent power load with dif-
ferent power off time interval.
and off for every 160 s. Then the effect of time interval to turn off
the power load is discussed for every 320 and 480 s.
Fig. 9 shows the average bottom temperature versus time in the fore PCM melting completely (t = 1405 s), which its peak temper-
base case and Case 3 (γ = 50%). The temperature rises oscillat- ature is 57.6 °C. After PCM melting completely, the temperature of
ing with increasing time. The temperature of the base case reaches Case 3 increases rapidly because the energy storage by PCM is its
80 °C at t = 406 s and presents a regular oscillation with peak sensible and heat, and then shows a regular oscillation with peak
temperature Tb = 100 °C at t ≥ 1120 s. As expected, the tempera- temperature Tb = 120 °C. Meanwhile, it can be found that com-
ture of Case 3 is significantly lower than that of the base case be- pared with the constant power load, the operation time of the base

7
H. Yang, Y. Li, L. Zhang et al. International Journal of Heat and Mass Transfer 181 (2021) 121899

Fig. 13. Variation of average bottom temperature versus time in the base case and Case 3 (γ = 50%) for different convective heat transfer coefficient.

case and Case 3 can significantly increase and its peak temperature thermal conductivity and the positive effect of PCM heat storage is
can obviously reduce in intermittent power load condition. the best.
Fig. 10 shows the average bottom temperature for different PCM
filling ratio. Before PCM melting completely, the temperature of 4.3. Effect of convective heat transfer coefficient
Case 1, Case 2, Case 3, Case 4 and Case 5 shows a relatively regu-
lar oscillation, where the peak temperature is less than 58 °C, and In this section, a high-power q = 30,0 0 0 w•m−2 with dif-
the effective operation time increases with increasing the PCM fill- ferent convective heat transfer coefficients h = 10, 20, 50 and
ing ratio. After PCM melting completely, the temperature of Case 1, 100 w•m−2 •K−1 is studied. This is considered cooling a high-
Case 2, Case 3, Case 4 and Case 5 increases rapidly. To better reveal power electronic device combines PCM passive cooling with nat-
the effect of intermittent power load, Fig. 11 compares the effective ural or forced convection cooling technique.
operation time of PCM-filled heat sink with different PCM filling Fig. 13 shows the average bottom temperature evolution with
ratio in constant power load and intermittent power load. Com- time in the base case and Case 3 (γ = 50%) for different convective
pared with the constant power load, the effective operation time heat transfer coefficient. It is observed that by increasing convec-
of PCM-filled heat sink slightly increases more than 2 times un- tive heat transfer coefficient, the peak temperature significantly re-
der intermittent power load condition. Moreover, the time exten- duces in both base case and Case 3. As the convective heat transfer
sion ratio decreases slightly with increasing the PCM filling ratio. coefficient increases to 100 w•m−2 •K−1 , it appears that the peak
Therefore, the extended time is greater than the total time interval temperature is always less than 80 °C in the base case. However,
to turn off the power load. It has been demonstrated the intermit- the peak temperature exceeds 80 °C at t = 1735 s and reaches a
tent power load is ability to enlarge the effective operation time of peak value of 84 °C in Case 3. Therefore, by increasing the con-
electronic device. vective coefficient, the effectiveness of Case 3 becomes weak. The
Fig. 12 shows the enhancement of operation time due to in- performance of the base case even performs better than that of
termittent power load with different time interval to turn off the Case 3 at h = 100 w•m−2 •K−1 .
power load. In Fig. 12, the one, two and three times refer to the in- Fig. 14 shows the average bottom temperature evolution with
termittent power load of 10,0 0 0 w•m−2 turn on for 160 s and then time for different convective heat transfer coefficient. As shown in
turn off for 160, 320 and 480 ss, respectively. It is obvious that Fig. 14(a) and (b), the temperature exceeds the maximum allow-
the operation time will increase as the time interval to turn off able limit during the first cycle for the base case, Case 1, Case 2,
the power load or PCM filling ratio increases. Compared with the Case 3 and Case 4, and it will exceed 80 °C during the second cy-
constant power load, the operation time increases slightly more cle for Case 5. At h = 50 w•m−2 •K−1 (Fig. 14(c)), the effective op-
than 2 times for the one-time intermittent power load, while it eration time significantly increases for Case 3, Case 4 and Case 5,
increases larger than 3 time for the two-time intermittent power while the temperature still exceeds 80 during the first cycle for
load and Case 2 has the biggest increment. When the time inter- the base case, Case 1 and Case 2. At h = 100 w•m−2 •K−1 , the
val to turn off the power load increases to 480 s, the operation peak temperature is always less than 80 °C in the base case, Case
time increases much greater than 4 time. The peak temperature is 1 and Case 2, in which its peak temperatures are 71.1 °C, 75.5 °C
79.9 and 79.4 °C in Case 1 and Case 2, while the peak temperature and 79.5 °C, respectively. However, the temperature of Case 5 is
reaches 84.5 and 83.7 °C in the base case and Case 3. Therefore, by the first to exceed 80 °C. Subsequently, the temperature of Case 3
increasing the time interval to turn off the power load, the ability and Case 4 will also quickly exceed 80 °C. Therefore, the PCM-filled
of PCM-filled heat sinks to enlarge the effective operation time of heat sinks can significantly extend the effective operation times of
electronic devices becomes more apparent, and the PCM-filled heat electronic device at low convective heat transfer coefficient, while
sink is more appropriate for managing the thermal issues of elec- the presence of PCM may be a negative effect at high convective
tronic devices with intermittent power loads. More importantly, as heat transfer coefficient because the negative effect of PCM due to
the time interval to turn off the power load increases to 480 s, the low thermal conductivity outweighs its positive effect of heat stor-
temperature of Case 1 and Case 2 is always less than the maximum age.
allowable temperature limit. In this situation, the PCM-filled heat The effective operation time for the base heat sink and PCM-
sink with 30% PCM filling ratio has the best performance because filled heat sinks at different convective heat transfer coefficient
the tradeoff effect between the negative effect of PCM due to low is summarized in Fig. 15. For natural convection air-cooling, i.e.,
h = 10,20 w•m−2 •K−1 , the effective operation time of the base

8
H. Yang, Y. Li, L. Zhang et al. International Journal of Heat and Mass Transfer 181 (2021) 121899

Fig. 14. Variation of average bottom temperature versus time at different convective heat transfer coefficient.

case is the best. It is understandable that with the convective heat


transfer coefficient increases, the negative effect of PCM due to its
low thermal conductivity is strengthened and the effectiveness of
PCM-filled heat sinks is expected to be reduced. The negative ef-
fect of PCM outweighs its positive effect at h = 100 w•m−2 •K−1 ,
and hence the base case has the best performance.

5. Conclusion

In this study, the thermal performance enhancement of the


PCM-filled heat sinks for electronic devices cooling is investigated
in detail under constant and intermittent power load conditions.
The effects of PCM filling ratio, intermittent power load with dif-
ferent power off time interval and convective heat transfer coef-
ficient are discussed to weigh the negative effect of PCM due to
its low thermal conductivity and its positive effect of heat storage.
The main conclusions are made as follows:
Fig. 15. Effective operation time at different convective heat transfer coefficient.
(1) At the early stage of PCM melting, the melted PCM flows up-
ward along the vertical fin and then flows downward along the
case and PCM-filled heat sinks is all short. It can be found that melting front formed two inward circulating currents in the fin
PCM-filled heat sinks with higher PCM-filled ratio are more effi- cavity. As the melting continues, the effect of natural convec-
cient to extend the effective operation time of electronic device. tion is enhanced that the PCM melts faster, and the tempera-
For forced convection air-cooling h = 50 w•m−2 •K−1 , the effective ture is higher in the upper region than in the lower region.
operation time significantly increases when the PCM filling ratio is (2) The thermal performance of the PCM-filled heat sinks is signifi-
over 50%, and it is not evident as the PCM filling ratio further in- cantly better than the base case under constant power load. The
creases. At h = 100 w•m−2 •K−1 , the PCM-filled heat sinks can no effective operation time of PCM-filled heat sinks increases with
longer compete with the base case. The performance of the base increasing of PCM filling ratio. However, the increased percent-

9
H. Yang, Y. Li, L. Zhang et al. International Journal of Heat and Mass Transfer 181 (2021) 121899

age in the time extension ratio is slightly smaller with increas- [12] Z.Q. Zhu, Y.K. Huang, N. Hu, Y. Zeng, L.W. Fan, Transient performance of a
ing of PCM filling ratio because the melted liquid PCM forms PCM-based heat sink with a partially filled metal foam: effects of the filling
height ratio, Appl. Therm. Eng. 128 (2018) 966–972.
a growing thermal resistance between PCM’s melting front and [13] H.Z. Yang, Y.Y. Li, Y. Yang, D. Chen, Y.G. Zhu, Effective thermal conductivity
heat source. of high porosity open-cell metal foams, Int. J. Heat Mass Transf. 147 (2020)
(3) The effective operation time of PCM-filled heat sinks signif- 118974.
[14] Q. Ren, Y.L. He, K.Z. Su, C.L. Chan, Investigation of the effect of metal foam
icantly increases and its peak temperature obviously reduces characteristic on the PCM melting performance in a latent heat thermal energy
under intermittent power load compared to the constant power storage unit by pore-scale lattice Boltzmann modeling, Numer. Heat Transf. A
load. Moreover, it becomes more apparent as the power off Appl. 72 (2017) 745–764.
[15] H.Z. Yang, Y.Y. Li, B.J. Ma, Y.G. Zhu, Review and a theoretical approach on pres-
time interval increases. Under an intermittent power load of
sure drop correlations of flow through open cell metal foam, Materials 14 (12)
10,0 0 0 w•m−2 turning on for 160 s and then off for 480 s with (2021) 3153.
h = 10 w•m−2 •K−1 , the peak temperature of Case 1 and Case 2 [16] Z.G. Qu, W.Q. Li, W.Q. Tao, Numerical model of the passive thermal manage-
ment system for high-power lithium ion battery by using porous metal foam
is, respectively 79.9 °C and 79.4 °C that is always less than the
saturated with phase change material, Int. J. Hydrog. Energy 39 (8) (2014)
maximum allowable temperature limit of electronic devices. 3904–3913.
(4) The PCM-filled heat sinks can significantly extend the effective [17] R. Qaiser, M.M. Khan, L.A. Khan, M. Irfan, Melting performance enhancement of
operation time of electronic device under natural cooling condi- PCM based thermal energy storage system using multiple tubes and modified
shell designs, J. Energy Storage 33 (2021) 102161.
tion, i.e., h = 10 or 20 w•m−2 •K−1 . However, for forced convec- [18] Q. Ren, P. Guo, J. Zhu, Thermal management of electronic devices using pin-fin
tion cooling h = 100 w•m−2 •K−1 , under an intermittent power based cascade microencapsulated PCM/expanded graphite composite, Int. J.
load of 30,0 0 0 w•m−2 turning on and off for every 160 s, the Heat Mass Transf. 149 (2020) 119199.
[19] Q. Ren, Z. Wang, T. Lai, J.F. Zhang, Z.G. Qu, Conjugate heat transfer in
base case has the best performance because the negative effect anisotropic woven metal fiber-phase change material composite, Appl. Therm.
of PCM outweighs its positive effect. Eng. 189 (2021) 116618.
[20] H. Behi, M. Ghanbarpour, M. Behi, Investigation of PCM-assisted heat pipe for
electronic cooling, Appl. Therm. Eng. 127 (2017) 1132–1142.
Declaration of Competing Interest [21] Z.Y. Jiang, Z.G. Qu, Lithium-ion battery thermal management using heat pipe
and phase change material during discharge–charge cycle: a comprehensive
numerical study, Appl. Energy 242 (2019) 378–392.
None.
[22] Z.Y. Jiang, Z.G. Qu, Lithium-ion battery thermal management using heat pipe
and phase change material during discharge-charge cycle: a comprehensive
numerical study, Appl. Energy 242 (2019) 378–392.
CRediT authorship contribution statement
[23] Z. Ling, Z. Zhang, G. Shi, X. Fang, L. Wang, X. Gao, Y. Fang, T. Xu, S. Wang,
X. Liu, Review on thermal management systems using phase change materials
Huizhu Yang: Conceptualization, Methodology, Validation, In- for electronics component, Li-ion batteries and photovoltaic modules, Renew.
vestigation, Writing – original draft. Yongyao Li: Writing – review Sustain. Energy Rev. 31 (2014) 427–438.
[24] S.K. Sahoo, M.K. Das, P. Rath, Application of TCE-PCM based heat sink for cool-
& editing. Liang Zhang: Investigation, Data curation. Yonggang ing of electronic components: a review, Renew. Sustain. Energy Rev. 59 (2016)
Zhu: Supervision. 550–582.
[25] T. Yang, P.V. Braun, N. Miljkovic, W.P. Kin, Phase change material heat sink for
transient cooling of high-power devices, Int. J. Heat Mass Transf. 170 (2021)
Acknowledgments 121033.
[26] R. Kalbasi, Introducing a novel heat sink comprising PCM and air-adapted to
electronic device thermal management, Int. J. Heat Mass Transf. 169 (2021)
This work was supported by the Guangdong Basic and Applied 120914.
Basic Research Foundation (No. 2020A1515110270) and National [27] M. Mozafari, A. Lee, J. Mohammadpour, Thermal management of single and
Natural Science Foundation of China (NO. 52106004), for which the multiple PCMs based heat sinks for electronics cooling, Therm. Sci. Eng. Prog.
23 (2021) 100919.
authors are thankful.
[28] P. Huang, G. Wei, L. Cui, C. Xu, X. Du, Numerical investigation of a dual-PCM
heat sink using low melting point alloy and paraffn, Appl. Therm. Eng. 189
References (2021) 116702.
[29] S. Gharbi, S. Harmand, S.B. Jabrallah, Experimental comparison between differ-
[1] Y. Kozak, B. Abramzon, G. Ziskind, Experimental and numerical investigation of ent configurations of PCM based heat sinks for cooling electronics components,
a hybrid PCM-air heat sink, Appl. Therm. Eng. 59 (2013) 142–152. Appl. Therm. Eng. 87 (2015) 454–462.
[2] H.Z. Yang, Y.Y. Li, Y. Yang, Y.G. Zhu, J. Wen, Effect of surface efficiency on the [30] P.P. Levin, S. Avraham, H. Gad, Numerical optimization of a PCM based heat
thermal design of plate-fin heat exchangers with passages stack arrangement, sink with internal fins, Int. J. Heat Mass Transf. 61 (2013) 638–645.
Int. J. Heat Mass Transf. 143 (2019) 118494. [31] A. Arshad, H.M. Ali, M. Ali, Thermal performance of phase change material
[3] A. Shahsavar, M.M. Baseri, A.A.A.A. Al-Rashed, M. Afrand, Numerical investi- (PCM) based pin-finned heat sinks for electronics devices: effect of pin thick-
gation of forced convection heat transfer and flow irreversibility in a novel ness and PCM volume fraction, Appl. Therm. Eng. 112 (2017) 143–155.
heat sink with helical microchannels working with biologically synthesized [32] J. Xie, K. Fah Choo, J. Xiang, H.M. Lee, Characterization of natural convection in
water-silver nano-fluid, Int. Commun. Heat Mass Transf. 108 (2019) 104324. a PCM-based heat sink with novel conductive structures, Int. Commun. Heat
[4] L. Yin, P. Jiang, R. Xu, H. Hu, Water flow boiling in a partially modified micro- Mass Transf. 108 (2019) 104306.
gap with shortened micro pin fins, Int. J. Heat Mass Transf. 155 (2020) 119819. [33] B. Kamkari, H. Shokouhmand, Experimental investigation of phase change ma-
[5] S.F. Zu, X.N. Liao, Z. Huang, D.Q. Li, Q.F. Jian, Visualization study on boiling terial melting in rectangular enclosures with horizontal partial fins, Int. J. Heat
heat transfer of ultra-thin flat heat pipe with single layer wire mesh wick, Int. Mass Transf. 78 (2014) 839–851.
J. Heat Mass Transf. 173 (2021) 121239. [34] B. Kamkari, H.J. Amlashi, Numerical simulation and experimental verification
[6] X.S. Hu, X.L. Gong, Experimental and numerical investigation on thermal per- of constrained melting of phase change material in inclined rectangular enclo-
formance enhancement of phase change material embedding porous metal sures, Int. Commun. Heat Mass Transf. 88 (2017) 211–219.
structure with cubic cell, Appl. Therm. Eng. 175 (2020) 115337. [35] H.M. Ali, M.J. Asgraf, A. Giovannelli, M. Irfan, T.B. Irshad, H.M. Hamid, F. Hassan,
[7] R. Kothari, S.K. Sahu, S.I. Kundalwal, S.P. Sahoo, Experimental investigation of A. Arshad, Thermal management of electronics: an experimental analysis of
the effect of inclination angle on the performance of phase change material triangular, rectangular and circular pin-fin heat sinks for various PCMs, Int. J.
based finned heat sink, J. Energy Storage 37 (2021) 102462. Heat Mass Transf. 123 (2018) 272–284.
[8] P. Ping, R. Peng, D. Kong, G. Chen, J. Wen, Investigation on thermal manage- [36] S. Gharbi, S. Harmand, S.B. Jabrallah, Experimental comparison between differ-
ment performance of PCM-fin structure for Li-ion battery module in high-tem- ent configurations of PCM based heat sinks for cooling electronics components,
perature environment, Energy Convers. Manag. 176 (2018) 131–146. Appl. Therm. Eng. 87 (2015) 454–462.
[9] N. Sharifi, T.L. Bergman, A. Faghri, Enhancement of PCM melting in enclosures [37] S.K. Saha, K. Srinivasan, P. Dutta, Studies on optimum distribution of fins in
with horizontally-finned internal surfaces, Int. J. Heat Mass Transf. 54 (2011) heat sinks filled with phase change materials, J. Heat Transf. 130 (3) (2008)
4182–4192. 034505.
[10] K.A.R. Ismail, F.A.M. Lino, Fins and turbulence promoters for heat transfer en- [38] S.F. Hosseinizadeh, F.L. Tan, S.M. Moosania, Experimental and numerical stud-
hancement in latent heat storage systems, Exp. Therm. Fluid Sci. 35 (2011) ies on performance of PCM-based heat sink with different configurations of
1010–1018. internal fins, Appl. Therm. Eng. 31 (17) (2011) 3827–3838.
[11] M.Y. Yazici, M. Avci, O. Aydin, Combined effects of inclination angle and fin [39] G. Setoh, F.L. Tan, S.C. Fok, Experimental studies on the use of a phase change
number on thermal performance of a PCM-based heat sink, Appl. Therm. Eng. material for cooling mobile phones, Int. Commun. Heat Mass Transf. 37 (9)
159 (2019) 113956. (2010) 1403–1410.

10
H. Yang, Y. Li, L. Zhang et al. International Journal of Heat and Mass Transfer 181 (2021) 121899

[40] T. Sathe, A.S. Dhoble, Thermal analysis of an inclined heat sink with finned [48] S. Gharbi, S. Harmand, S.B. Jabrallah, Experimental study of the cooling per-
PCM container for solar applications, Int. J. Heat Mass Transf. 144 (2019) formance of phase change material with discrete heat sources continuous and
118679. intermittent regimes, Appl. Therm. Eng. 111 (2017) 103–111.
[41] A.C. Iradukunda, A. Vargas, D. Huitink, D. Lohan, Transient thermal perfor- [49] G.K. Marri, C. Balaji, Experimental and numerical investigations on the effect
mance using phase change material integrated topology optimized heat sinks, of porosity and PPI gradients of metal foams on the thermal performance of a
Appl. Therm. Eng. 179 (2020) 115723. composite phase change material heat sink, Int. J. Heat Mass Transf. 164 (2021)
[42] R. Kalbasi, M. Afrand, J. Alsarraf, M.D. Tran, Studies on optimum fins number 120454.
in pcm-based heat sinks, Energy 171 (2019) 1088–1099. [50] A.D. Brent, V.R. Voller, K.J. Reid, Enthalpy-porosity technique for modeling con-
[43] R. Baby, C. Balaji, Thermal optimization of PCM based pin fin heat sinks: an vection-diffusion phase change: application to the melting of a pure metal,
experimental study, Appl. Therm. Eng. 54 (1) (2013) 65–77. Numer. Heat Transf. 13 (3) (1988) 297–318.
[44] K. Dammak, A.E. Hami, Thermal reliability-based design optimization using [51] H. Shmueli, G. Ziskind, R. Letan, Melting in a vertical cylindrical tube: numer-
Kriging model of PCM based pin fin heat sink, Int. J. Heat Mass Transf. 166 ical investigation and comparison with experiments, Int. J. Heat Mass Transf.
(2021) 120745. 53 (19–20) (2010) 4082–4091.
[45] Q. Ren, Z. Wang, J. Zhu, Z.G. Qu, Pore-scale heat transfer of heat sink filled with [52] A.M. Abdulateef, S. Mat, K. Sopian, J. Abdulateef, A.A. Gitan, Experimental and
stacked 2D metal fiber-PCM composite, Int. J. Therm. Sci. 161 (2021) 106739. computational study of melting phase-change material in a triplex tube heat
[46] R. Baby, C. Balaji, Thermal performance of a PCM heat sink under different exchanger with longitudinal/triangular fins, Sol. Energy 155 (2017) 142–153.
heat loads: an experimental study, Int. J. Therm. Sci. 79 (2014) 240–249. [53] C.Y. Zhao, W. Lu, Y. Tian, Heat transfer enhancement for thermal energy stor-
[47] A. Farzanehnia, M. Khatibib, M. Sardarabadia, M. Passandideh-Fard, Experi- age using metal foams embedded within phase change materials (PCMs), Sol.
mental investigation of multiwall carbon nanotube/paraffn based heat sink for Energy 84 (2010) 1402–1412.
electronic device thermal management, Energy Convers. Manag. 179 (2019)
314–325.

11

You might also like