Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Received: 13 May 2020 | Revised: 11 August 2020 | Accepted: 25 August 2020

DOI: 10.1002/htj.21933

ORIGINAL ARTICLE

Assessment of heat transfer and flow


characteristics of a two‐phase closed
thermosiphon

Israa S. Ahmed1 | Ayad M. Al Jubori2

1
Department of Electromechanical
Engineering, University of Abstract
Technology–Iraq, Baghdad, Iraq In this study, comprehensive modeling and simula-
2
Department of Communication tions were developed and carried out to perform the
Engineering, University of
Technology–Iraq, Baghdad, Iraq investigation of the thermal performance of the en-
closed thermosiphon through pool boiling in the
Correspondence
evaporator sector and the condensation of the liquid
Ayad M. Al Jubori, Department of
Communication Engineering, University of film in the condenser part. To simulate these phe-
Technology–Iraq, Al‐Wehda nomena, the volume of fluid model was utilized. The
neighbourhood, Al‐Sinaa' Street, P. O.
simulation modeling using the computational fluid
Box 19006, Baghdad, Iraq.
Email: ayad.m.salman@uotechnology. dynamics (CFD) technique was validated with exist-
edu.iq ing experimental results, and a good agreement was
reached. The simulation results were presented and
evaluated in terms of temperature profiles and con-
tours, the volume of fraction contours, and velocity
vector distribution. Moreover, the thermal perfor-
mance (ie, the heat transfer coefficient and thermal
resistance) through the thermosiphon operation was
analyzed. From the simulation results, it is found that
the thermosiphon performance can be improved by
the tilt angle and fill ratio. The results indicated that
the optimal performance (ie, a high heat transfer
coefficient and a low thermal resistance) was attained
at a power input of 250 W, tilt angle of 90°, and fill
ratio of 0.5. The established CFD simulations effec-
tively predicted the formation of two‐phase flow

Abbreviations: 2D, two‐dimensional; CFD, computational fluid dynamics; UDF, user‐defined function; VOF, volume
of fluid.

Heat Transfer. 2020;1–20. wileyonlinelibrary.com/journal/htj © 2020 Wiley Periodicals LLC | 1


2 | AHMED AND AL JUBORI

pattern and boiling and condensation zones with


water at a low power input, termed as geyser boiling.

KEYWORDS
CFD simulation, fill ratios, heat transfer thermosiphon, tilt angles,
volume of fluid

1 | INTRODUCTION

For energy saving and heat recovery in various engineering applications, the heat pipe tech-
nique has been employed in recent years, which plays a main role in enhancing the thermal
performance of various industrial applications such as heat exchangers. In general, thermosi-
phon as a heat transfer device has a high thermal conductivity. The usual thermosiphon device
as a heat pipe consists of a hollow metal shell filled with a working fluid in a small amount,
resulting in a very lightweight device. This is mostly as a result of its simple configuration, good
compactness, special flexibility, high effectiveness, and excellent reversibility. In the heat pipe,
the working fluid is circulating in a sealed tube/container with a very effective heat transfer rate
during the evaporation and condensation process. The working fluid absorbs the heat in the
evaporator and the heat is rejected from the working fluid in the condenser. The heat added in
the evaporator section is transferred to the working fluid and it evaporates, generating vapor,
after it has reached the saturation temperature. Then, the vapor rises toward the condenser. The
working fluid in vapor status condenses in the condenser, transporting its latent heat into the
cooling medium.
For numerous engineering applications, computational techniques show a significant role in
solving complex flow problems due to their effectiveness, accuracy, and flexibility. Therefore, a
number of computational studies on the thermosiphon operation have been carried out. Ma
et al1 evaluated the thermosiphon thermal performance for energy‐saving applications in the
heat recovery system. Their results exhibited that the seasonal temperature has a greater impact
on the system performance in winter as compared with summer. Salehi et al2 predicted the
enhancement of the heat transfer characteristics in a two‐phase closed thermosyphon with a
CuO/water nanofluid. The artificial neural network was used as an optimization algorithm.
Fadhl et al3 presented a computational fluid dynamics (CFD) algorithm of a two‐phase closed
thermosiphon filled by R134a and R404a at a steady‐state condition. Their results showed that
the growth of bubbles was very small. Ababsa and Bougoul4 numerically analyzed the efficiency
of an open thermosiphon unit. In their model, the radiative and natural convection heat
transfer were coupled in an inclined air passage. The radiative heat transfer had a significant
effect on the average Nusselt number. Zhou et al5 proposed and tested a prototype of the
thermosiphon unit for temperature control purposes. They found that the root temperature of a
steel thermosiphon was less by 2°C as compared with a copper thermosiphon. Zhang and Liu6
adopted a numerical solution for the heat transfer coefficient and thermal resistance of the heat
pipe with a constant temperature. There was a reduction in thermal resistance with the increase
in the pipe diameter. Yue et al7 offered a simple CFD model of a microchannel finned heat pipe.
The heat transfer characteristics were investigated and validated under various filling ratios,
ranging from 60% to 100%. They highlighted that the highest cooling capacity, 4087 W, was
achieved at a fill ratio of 0.78. Li and Zhang8 suggested a dynamic steady state of the
AHMED AND AL JUBORI | 3

wall‐implanted heat pipe to its performance in the summer months. Their results showed that
the thermal response of the wall‐implanted heat pipe is faster on the basis of the heat transfer
rate of 1.1 kW/m2, as compared with the conventional wall. Wang et al9 implemented coupling
between the CFD and visualization to investigate geyser boiling in a closed thermosiphon with
a two‐phase flow. The results indicated that geyser boiling started at the cycle time ranging from
80 to 100 seconds, with an evaporator temperature of 95°C and a condenser temperature of
30°C. Akkus et al10 assessed heat pipe performance on the basis of molecular dynamics and
nanogroove. They found that the HP performance was affected by the groove size, filling ratio,
and heat load. The results indicated that the sizing down of the groove dimension led to a
decrease in thermal performance. Liu et al11 optimized the wall implanted with a heat pipe to
enhance its performance. Moreover, they experimented with the equivalent heat transfer
coefficient. Their results exhibited that the equivalent heat transfer coefficient of 1.12 W/m2·K
was achieved. Sarafraz et al12 assessed the closed thermosiphon that was charged by zirconia‐
acetone nanofluid as a working fluid. The parametric studies in terms of thermal performance
were carried out. Wang et al13 proposed an approach for a liquid fuel heat pipe reactor. They
concluded that the offered system possessed a very high power density.
Considerable experimental studies have been accomplished to monitor the performance of
the thermosiphon in different working conditions. Manimaran et al14 utilized a copper heat
pipe with two layers of mesh filament to enhance its performance. Their experimental results
indicated that the optimum performance was achieved at a fill ratio of 0.75. Du et al15 combined
the solar panel with a nanocoated heat pipe to enhance its performance by transfer of heat
from the hot spots on a solar panel. Their experiments showed that the temperature difference
inside the solar panel was reduced up to 2.5°C. Shi et al16 theoretically and experimentally
investigated the influence of nanofluids on the mini heat pipe performance. They found that the
thermal resistance was reduced by around 40%. Tecchio et al17 offered experimental tests of
geyser boiling in the thermosiphon. In the steady‐state region, they found that geyser boiling
takes place at vapor pressure less than 25 kPa and heat flux over 12.5 kW/m2. Smith et al18
offered a design and test to explore the performance of the heat pipe in high‐power battery
applications of 400 W to dissipate the heat load (as a thermal management system). They
concluded that the proposed system can be used for automotive batteries as a thermal man-
agement device, simply and safely. Abdulshaheed et al19 experimentally tested the heat pipe
performance utilizing a copper oxide (CuO) nanowire, where the nanowire was used to improve
the evaporator part by decreasing the thermal resistance. The results revealed that the thermal
resistance was reduced by using the nanowire. Xiong et al20 considered the experimental and
analytical study of the heat pipe using nine SiO2 nanoparticles in three different sizes. They
detected that the nanoparticles' mass concentration and size of the SiO2 led to enhance the heat
fusion and the decomposing point through the heat pipe. Liang et al21 tested the thermal
characteristics of the mini heat pipe, which was rotated at rotational speeds ranging from 1000
to 3000 rpm and power inputs ranging from 2.5 to 30 W. The results showed that the cir-
cumferential temperature variance improved with the increase of the rotational rates and heat
input. Kujawska et al22 determined the impact of several nanofluids on the thermal perfor-
mance of geyser boiling (ie, closed thermosiphon). Their results showed that chemical stabi-
lizers have a greater effect on thermal performance than nanoparticles. Wang et al23 analyzed
the impact of the fill ratio and the working fluid on the heat characteristics through the
miniature heat pipe type of a flat plate. They concluded that the maximum thermal conductivity
(2311 W/m·K) and minimum thermal resistance (0.67 K/W) were obtained in a horizontal
situation. Eltaweel et al24 tested the thermal performance of the flat‐plate solar collector
4 | AHMED AND AL JUBORI

utilizing a twisted tube to enhance the heat transfer. Also, the performance with normal
circular tubes was compared. They showed that the performance improved by 34%, based on
the twisted tube. Almahmoud and Jouhara25 conducted theoretical and experimental analyses
on the heat pipe of the radiating heat exchanger at several heat source temperatures. It was
detected that the back panel improved the rate of the heat transfer by around 330%, which
enabled the heat pipe to recuperate 8.5 kW.
Although the closed‐loop thermosiphon has been used in numerous applications, limited
CFD studies have been presented to investigate its thermal performance. These studies have
been focused on modeling the condensation and evaporation process. However, there are
limited studies concerning the CFD analysis of the design parameters like tilt angle and fill ratio
at various heat inputs that impact the thermosiphon thermal performance, especially the vi-
sualization of the phase‐change process of the inclined closed‐loop thermosiphon at various fill
ratios. Moreover, limited investigations have determined the geyser boiling occurrence inside
the closed‐loop thermosiphon and its characteristics. Also, the geyser boiling impact on the
thermal features of the closed‐loop thermosiphon has not been stated in the literature.
Therefore, numerical simulation modeling is performed to visualize the geyser boiling and
condensation phenomena and thermal performance through the thermosiphon operation with
water. A user‐defined function (UDF) is implemented and integrated with ANSYS Fluent to
simulate the flow phase‐change and heat/mass transfer processes. The influences of different
tilt angles, fill ratios, and power inputs on the thermal performance of the thermosiphon have
been considered in this study. The numerical simulation methodology has been validated with
preceding experiments.

2 | P H Y S I C A L G E O M E T R Y A N D OP E R A T I N G
C ON D I T I O NS

A simplified schematic diagram of the closed two‐phase thermosiphon is displayed in


Figure 1A,B, which is used in this study. The thermosiphon is considered as a closed circuit
made up of an evaporator and condenser sectors. In the evaporator sector, the vapor rises to the
top (ie, in the direction of the condenser), whereas in the condenser section, the vapor con-
denses and falls back toward the evaporator again, and the process continues, as shown in
Figure 1A. The structure and dimensions of the two‐dimensional (2D) closed thermosiphon are

(A) (B)
Cooling

Liquid

Vapour
F I G U R E 1 A simplified schematic
diagram (A) and two‐dimensional geometry
and dimensions (B) of the closed two‐phase
Liquid Heating thermosiphon [Color figure can be viewed at
wileyonlinelibrary.com]
AHMED AND AL JUBORI | 5

shown in Figure 1B. Thermosiphon dryness phenomenon is caused due to the rise in the
evaporator temperature and the drop in the condenser temperature, depending on the length of
the evaporator and condenser sections. Moreover, the dry‐out indicates that the flow of the
working fluid has stopped. Also, the working fluid in the condenser section tends to reach the
coolant temperature by conduction. The variety in the evaporator and condenser lengths affects
the working characteristics inside the thermosiphon, because the heat flux from the evaporator
wall and from the working fluid to the condenser wall varies with the length. Therefore, the
same length of the evaporator and condenser sections has been considered to avoid the dry‐out
phenomenon.26‐28 For the two‐phase closed thermosiphon, the specifications, dimensions, and
operating conditions are listed as below:

Pipe material: copper


Pipe inner diameter: 22 mm
Pipe outer diameter: 23.8 mm
Pipe wall thickness: 0.9 mm
Condenser sector: 200 mm
Evaporator sector length: 200 mm
Tilt angles: 15°, 45°, 75°, and 90°
Fill ratios: 0.1, 0.3, 0.5, 0.75, and 0.9
Power inputs: 50, 100, and 250 W
Upper and lower caps: insulated
Walls: stationary
Working fluids: water

3 | MATHEMATICAL MODEL OF THE THERMOSIPHON

To describe the working fluid motion in the two‐phase closed thermosiphon geometry, the
mass, momentum, and energy equations are employed. These governing equations will be
clarified in the following section:

3.1 | The governing equation for the volume of fluid method

The volume of fluid (VOF) method was used in the most numerical studies to model the
multiphase flows, where the VOF approach showed that it can capture well the analytical and
experimental results. Moreover, the CFD modeling of the phase‐change phenomena can si-
mulate boiling and condensation processes and capture gas/liquid regimes with heat and mass
transfer through the phase‐change progressions. Moreover, the VOF approach can be applied to
simulate the two‐phase flow, because it is able to track the working fluid interfaces effectively
through the phase‐change processes inside the thermosiphon by solving one set of equations
only. Consequently, the VOF method can be utilized to model interface tracking between the
liquid and gas in steady/transient cases, stratified flows, bubbles' motion, and free‐surface flows.

(i) Continuity equation


For two‐phase flow (ie, vapor and liquid), the continuity equation can be presented as
follows:
6 | AHMED AND AL JUBORI

∂α v → Sm,v (1)
+ ∇ × (α v V ) = ,
∂t ρv
∂αl → Sm,l (2)
+ ∇ × (α l V ) = ,
∂t ρl
where Sm,v and Sm,l are mass transfer sources terms in continuity equation (ie, Equations 1
and 2) throughout condensation and evaporation. The relation between those terms is
Sm,v = −Sm,l . The mass source terms can be obtained using the below equations29:

Tv − Tsat (3)
Sm,v = −Sm,l = βevap αl ρ l Tv > Tsat ,
Tsat
Tsat − Tl (4)
Sm,l = −Sm,v = βcond α v ρv Tl < Tsat.
Tsat

The volume fraction (α) should be satisfied with the summation of the vapor and liquid as
follows:

α v + αl = 1. (5)

α Values should be between 0 and 1, where α = 1 refers to the vapor state and α = 0 refers to
the liquid state. Consequently, α for the two‐phase flow (ie, mixture) is 0 < α < 1.
For the mixture phase (ie, two‐phase) flow, the density for vapor and liquid phases can be
calculated as below30:

ρ = α v ρv + αl ρ l . (6)

(ii) The momentum equations


In VOF model, the momentum equation involves the pressure, the gravitational force,
friction, and surface tension as follows:

∂ ⇀ ⇀⇀ ⇀ ⇀T (7)
(ρV ) + ∇ × (ρV V ) = −∇p + ∇ × [μ (∇ V + ∇ V )] + ρg + FCSF.
∂t

The two‐phase dynamic viscosity (μ) is calculates as follows30:

μ = α v μ v + αl μl . (8)

The continuum surface force is derived on the basis of the continuum surface model as
follows31:

α v ρv κ v ∇α v + αl ρ l κl ∇αl (9)
FCSF = σ ,
0.5(ρv + ρ l )
AHMED AND AL JUBORI | 7

where κ refers to surface curvature, which is calculated from the following equation:

⎛ ∇ αl ⎞
κl = − κ v = ∇ × ⎜ ⎟. (10)
⎝ | ∇ αl | ⎠

(iii) The energy equation for the VOF method


In the VOF method, the energy transfer equation through the phase‐change process is
presented by introducing the source term (SE) as follows:

∂ → (11)
(ρE ) + [(ρE + p) V ] = ∇ × (λ∇T ) + SE,
∂t

where E refers to the enthalpy of the mixture flow, which is defined as a mass‐average value
and can be obtained as follows:

α v ρv E v + αl ρ l El (12)
E= ,
α v ρv + αl ρ l

where E v and El are calculated in terms of the shared temperature and specific heat (Cv ) as
follows3:

E v = Cp,v (T − Tsat ), (13)

El = Cp,l (T − Tsat ). (14)

The thermal conductively (λ) and energy source term (SE) can be calculated using the
following equations29,30:

λ = α v λ v + αl λ l , (15)

SE = Sm,v (E v − El ). (16)

4 | T H E R M O S I P H O N G E O M E T R Y AN D M E S H I N G

A 2D sketch of the thermosiphon geometry with a length of 400 mm, wall thickness of 0.9 mm,
and an outer diameter of 23.8 mm, as sketched in Figure 1B, is generated using geometry
modular in ANSYS workbench. In CFD simulations, the grid generation is one of the significant
techniques. The grid quality plays an essential role in the solution stability and the precision of
results. In this study, the quadrilateral mesh type is created using mesh modular in various
mesh densities. The quadrilateral mesh can be applied to the boundary layer, closely bursting in
the direction normal to the wall and more lightly in the tangential direction. This is because the
quadrilateral mesh allows cells clustering in certain areas of the flow domain to capture the
phase‐change phenomena accurately in the boundary layer zone.
8 | AHMED AND AL JUBORI

Various mesh densities/sizes are generated and tested on the basis of the quadrilateral mesh
to achieve mesh independence and sensitivity. As shown in Table 1, different mesh sizes are
monitored with the average temperatures of the evaporator and condenser parts. It was found
from Table 1 that the average evaporator and condenser temperatures are nearly independent of
the mesh size beyond 60 000 cells. The mesh of the thermosiphon consists of 45 000 cells for a
fluid zone and 15 000 cells for the solid zone. Thus, this mesh size is used in all CFD simu-
lations. The mesh format is shown in Figure 2.

5 | S E T U P O F SI M U L A T I O N CF D M E T H O D O L O G Y

In CFD simulations, the governing equations are solved in terms of space and time. To perform
CFD simulation using ANSYS Fluent, one has to select a discretization scheme for each term in
governing equations from a list of CFD schemes. Each scheme based on the discretization
technique has its numerical solution behavior. Consequently, on the basis of the scheme/
technique in use, the highest accuracy in terms of numerical stability and quantitative re-
presentation (ie, physically more accurate) can be accomplished. To simulate the dynamic flow
behavior and heat transfer phenomena for the two‐phase flow in a closed thermosiphon, a 2D
CFD model based on the VOF method was established. For evaporator sector, various heat
inputs were set, changing from 50, 100, and 250 W for each fill ratio (0.1, 0.3, 0.5, 0.7, and 0.9) at
four indication angles, namely 15°, 45°, 75°, and 90°. However, for the condenser sector, a
constant wall temperature was fixed as a boundary condition. A stationary wall and no‐slip
conditions were specified for the momentum and shear conditions, respectively. In CFD si-
mulation, the PISO algorithm (ie, pressure‐velocity scheme) was used to make a balance be-
tween the computational effort and time, where a large time step could be used as compared
with other schemes. PISO scheme is a calculation procedure for velocity‐pressure coupling (ie,
to solve the momentum discreteness) for the unsteady compressible flow.32 Moreover, the
PRESTO and Geo‐Reconstruction schemes were selected for pressure solution and volume
fraction, respectively. The UDF was implemented to link the source terms mass and energy
with the governing equations in ANSYS Fluent, where the UDFs have been performed to
determine the transmission the mass and energy through the condensation and evaporation
phenomena. In the VOF model, the volume fraction has been defined for each working fluid
phase, where two mass sources are required to describe the mass added to the vapor phase and
the process is required for the condensation progression.
To simulate condensation and evaporation through the thermosiphon, unsteady solutions with
the VOF model were conducted with a time step of 5 × 10−3 s. The simulations converged when
the mass, velocity, and temperature residuals were equal or less than 10−5 for mass and velocity
and 10−6 for temperature. In the CFD setup, the liquid is presented as a primary phase, whereas
the vapor is defined as a secondary phase. To calculate the heat and mass transfer through the
evaporation and condensation developments, the latent heat and evaporation/boiling temperature

TABLE 1 Mesh independence study

Cells number 20 000 40 000 50 000 60 000 70 000


Tav,evp,w , °C 121.337 118.683 116.874 116.482 116.315
Tav,cond,w , °C 40.221 38.663 37.961 37.844 37.760
AHMED AND AL JUBORI | 9

F I G U R E 2 Thermosiphon grid
generation [Color figure can be viewed
at wileyonlinelibrary.com]

of the liquid phase of the working fluid are defined in the UDF code. In the evaporator part,
the liquid is heated once the CFD simulation is started. When the saturation temperature of the
working fluid specified in the UDF is reached, the evaporation process begins and the two‐phase
flow (ie, phase‐change process) starts due to the boiling at the inner wall of the evaporator part.
Then, the saturated vapor phase flows up to the condenser part, and it begins to condense along
the cold wall of the condenser part, making a thin liquid film. The surface tension between the
fluid interfaces is considered and obtained as the below equation:

σ = 0.0118 − 1.5175e−5T + 1.5708e−7T 2. (17)

In the UDFs code, the saturation temperature and the working fluid latent heat are
specified.
When the mass flow rate of the vapor phase balanced with the mass flow rate of the liquid
phase and the temperature was relatively stable, the thermosiphon was assumed to reach a
steady state.
The heat transfer coefficient was considered as follows33:

Q (18)
h= ,
As (Tevap − Tcond )

where Q is the rate of transfer rate, As represents the surface area, and Tevap and Tcond present
the average temperatures of the evaporator part and the condenser part, respectively.
Once the CFD solution is run, the evaporator sector (ie, liquid pool) is heated, where the
evaporation process occurs when the saturation temperature is reached (ie, 373 K). The vapor is
then moved into the condenser sector, and the saturated vapor condenses alongside the
10 | AHMED AND AL JUBORI

condenser cold walls, creating a liquid film. To allow heat transfer between the fluid zone and
the inner solid wall, the interfaces are defined as coupled boundary conditions between the
fluid and solid zones. The water as a working fluid is considered with various fill ratios (ie, the
initial ratio of the liquid to the evaporator total size). The inclination angle of the thermosiphon
was defined from the horizontal axis.

6 | C F D SI M UL A T I O NS ' V A LI D A T I O N

The CFD model of the developed thermosiphon is validated using the previous experiment34 in
terms of temperature profile along the thermosiphon length, as presented in Table 2. It de-
monstrates the comparison between the previously reported experimental result and the cur-
rent simulation results. From Table 2, the comparison showed that the maximum deviation (ie,
the relative errors) of the temperature between the simulation results and the reported ex-
perimental results is within 3.55%. Consequently, it revealed that the proposed simulation
methodology fairly delivers the predicted results of the thermosiphon heat transfer perfor-
mance. The deviation in the temperature distribution in the condenser maybe due to the
insulation deficiency producing heat losses, which are not included in the CFD simulations,
where the lower condenser temperature has been delivered from the CFD simulations.

7 | R E S U LTS AN D D I S C U S S I ON

The established mathematical model of the enclosed thermosiphon is applied to visualize the
flow and heat transfer patterns inside the evaporator and condenser sectors. Moreover, the
thermosiphon thermal performance for a wide range of fill ratios, power inputs, and inclination
angles will be discussed. The temperature and thermal resistance profiles along the thermosi-
phon will also be presented.

7.1 | Flow conception of the CFD simulations

To recognize the heat transfer development through the thermosiphon operation, the tem-
perature contours at a fill ratio of 0.5 and vertical positioning with different simulation times are

TABLE 2 Comparison of the numerical simulation results with previous experiments

CFD results Experimental results

Position, mm Tevap, K Tcond, K Tevap, K Tcond, K Deviation %

20 386.76 … 381.32 … 1.405


70 392.49 … 384.64 … 2.001
120 390.35 … 383.05 … 1.871
250 … 300.13 … 311.18 3.55
350 … 300.94 … 309.35 2.721

Abbreviation: CFD, computational fluid dynamics.


AHMED AND AL JUBORI | 11

observed during the heating process (ie, at the strating of the process), reaching steady‐state
operation, as shown in Figure 3. As can be observed from Figure 3, the distribution of the
temperature in the working fluid area throughout the evaporator and condenser sectors is
documented at the power input of 100 W. Due to the applied heat flux on the evaporator wall,
the temperature increased, which allowed heat to transmit into the water pool inside the
evaporator sector, as shown in Figure 3 for different simulation times. It is observed that the
temperatures continue to increase due to the temperature difference between the working fluid
and evaporator wall. This leads to boiling of the working fluid along the walls of the evaporator
sector. With the increase in simulation time, the high‐temperature region expands, due to the
vapor movement upward, and the high‐temperature zone reaches the condenser sector, as
shown in Figure 3. Close to the internal wall of the condenser sector, the vapor condenses and a
lower temperature is observed. Then, the condensed vapor in a liquid form drops down back to
the evaporator sector due to the gravitational force. This process continues until a steady state is
reached.
Figure 4 shows the pool boiling and condensation phenomena in terms of the volume
fraction inside the evaporator sector and condenser sector, respectively, with a fill ratio of 0.5 of
the water at vertical position and power input of 100 W for different simulation times. The red
color denotes the existence of only vapor (ie, the vapor volume = 1.0), and the blue color
denotes the existence of only liquid (ie, the volume fraction of vapor = 0.0). At the starting of

Temp. (K)

0s 0.2s 1s 5s 10s 20s 40s 60s

F I G U R E 3 Heat transfer process contours through the thermosiphon operation at a fill ratio of 0.5
and different simulation times with a vertical orientation [Color figure can be viewed at
wileyonlinelibrary.com]
12 | AHMED AND AL JUBORI

Volume
fraction

0.0s 0.2s 1.0s 5s 10s 20s 40s 60s

F I G U R E 4 Volume fraction contours of the vapor phase through the thermosiphon operation for a
vertical position and a fill ratio of 0.5 at different simulation times [Color figure can be viewed at
wileyonlinelibrary.com]

the process, the evaporator sector is filled half of its volume (ie, fill ratio = 0.5). The phase‐
change occurs when the water begins to boil and the water temperature goes beyond the
saturation temperature. From this time, nucleation sites occur, and vapor bubbles begin to form
at those locations, as displayed in Figure 4. As can be observed in the figures, the volume
fraction of the vapor increases due to continuous boiling of water. This leads to a reduction in
the liquid volume fraction. After the condensation phenomenon, the liquid film is back to the
water pool due to the gravity to start a new cycle.
Figure 5 illustrates the evolution of the velocity vector distribution throughout 60 seconds
for a vertical position at a power input of 100 W and a fill ratio of 0.5. The fast growth of the
vapor phase, 0.801 m/s, becomes more significant with the increase of the heating time. The
maximum vapor velocity was observed at the end of the evaporator sector and the beginning of
the condenser sector. The layer of the thin film of liquid forms on the condenser wall due to the
decrease in the trapped vapor temperature in the condenser sector. Then, the formed liquid
layers go back into the evaporator sector with a velocity of 0.17 m/s . The working fluid velocity
in the liquid status is affected by the growth rate of the bubbles. The liquid‐vapor change after
the recurring detachment of vapor bubbles from the wall allows the cold liquid to reduce the
hot wall temperature. For a low inclination angle, the flow layers are lined up with a small
vapor velocity and small interface tension. Moreover, the flow velocity changes very quickly in
the normal direction to the wall. The working fluid velocity inside the thermosiphon is less,
because the pressure of the working fluid vapor is high.
AHMED AND AL JUBORI | 13

Velocity
(m/s)

0.0s 0.2s 1.0s 5s 10s 20s 40s 60s

F I G U R E 5 Velocity vector (m/s) through the thermosiphon operation for a vertical position and a fill
ratio of 0.5 at different simulation times [Color figure can be viewed at wileyonlinelibrary.com]

At a steady state (ie, at simulation time of 60 seconds), the temperature contour, volume
fraction contour, and velocity vector distribution are presented in Figure 6 for a fill ratio of 0.5
and power input of 100 W at a tilt angle of 75°. It is observed from the temperature contour that
a high temperature is reached at the upper end of the condenser sector as compared with a
vertical position (as shown in Figure 4), which leads to reduce the temperature variance be-
tween the evaporator and condenser. Moreover, the volume fraction of the vapor took a large
space as compared with the volume of liquid, as revealed in Figure 6B. From Figure 6C, it can
be found that the lower zone of pool boiling has a high velocity as compared with other regions
of the thermosiphon. The temperature distribution showed that the thermal boundary in the
liquid pool has a lower temperature as compared with other regions. Also, the vapor tem-
perature is a bit higher than that of saturation in the evaporator. This is due to the fact the
falling droplets of the formed liquid do not wet the whole inner surface of the liquid pool in the
evaporator. It can be marked that the liquid film/liquid droplets are back into the evaporator
part due to gravity, which recharges the liquid pool. Continuous liquid boiling leads to a
reduction in the volume fraction of the liquid and an increase in the volume fraction of the
vapor.

7.2 | The working conditions' impact on the thermosiphon


performance

Figure 7 illustrates the temperature profiles along the enclosed thermosiphon wall at a vertical
position with a power input of 50 W for a varying fill ratio of 0.5. The thermosiphon length
14 | AHMED AND AL JUBORI

(A) (B) (C)


Temp. (K) Volume fraction
Velocity (m/s)

Temperature Volume fraction Velocity vector

F I G U R E 6 Temperature contour (A), volume fraction (B), and velocity vector (C) for a fill ratio of 0.5
and tilt angle of 75° at a steady state [Color figure can be viewed at wileyonlinelibrary.com]

between 0 and 0.2 m specifies the evaporator length, whereas the length from 0.2 to 0.4 m
indicates the condenser length. As shown in Figure 7, at fill ratios of 0.1 and 0.3 in the
thermosiphon, the temperature profiles are approximately uniform along with the evaporator
sector as compared with other fill ratios at a power input of 50 W. In Figures 8 and 9, the
maximum temperature was achieved at mid‐length of the evaporator sector with fill ratios of 0.1
and 0.3, whereas the highest temperatures with fill ratios of 0.5 and 0.9 were obtained in the
upper zone of the evaporator sector. Moreover, Figures 8 and 9 show considerably higher
temperatures with fill ratios of 0.9 and 1.0, as compared with the thermosiphon. It can be
clarified through the fact that the fluid amount in pool boiling (ie, evaporator sector) is not

180
Fill rao=0.1 Fill rao=0.3 Fill rao=0.5
160
Fill rao=0.75 Fill rao=0.9
140
Temperature (°C)

120
100
80
60
40
20 F I G U R E 7 The temperature profile
0 at a vertical position and a power input of
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 50 W [Color figure can be viewed at
Thermosiphon length (m)
wileyonlinelibrary.com]
AHMED AND AL JUBORI | 15

F I G U R E 8 The temperature profile at a 200


Fill rao=0.1 Fill rao=0.3 Fill rao=0.5
vertical position and a power input of 100 W 180
160
Fill rao=0.75 Fill rao=0.9
[Color figure can be viewed at

Temperature (°C)
140
wileyonlinelibrary.com]
120
100
80
60
40
20
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Thermosiphon length (m)

F I G U R E 9 The temperature profile at a 200


Fill rao=0.1 Fill rao=0.3 Fill rao=0.5
vertical position and a power input of 250 W 180
Fill rao=0.75 Fill rao=0.9
[Color figure can be viewed at 160

wileyonlinelibrary.com] 140
Temperature (°C)

120
100
80
60
40
20
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Thermosiphon length (m)

enough to make sufficient liquid in the condenser sector and the liquid return rate to the
evaporator sector will be not enough.
Furthermore, it can be observed from Figures 8 and 9 with power inputs of 100 and 250 W
that the temperature distributions are approximately similar. The above‐described figures il-
lustrate the heat transfer progress through the thermosiphon operation charged with water. For
FR of 0.9, a higher temperature is observed in the upper zone of the evaporator section as
compared with other fill ratio values for the power input of 100 and 250 W, respectively. This is
due to the fact that the increased height of the liquid in the evaporator upper part averts the
large bubbles to reach the liquid surface, making a vapor film on the inner wall of the eva-
porator section, which leads to an increase in the temperature in that zone. There is an increase
in the liquid height with the increase in the power input in the case of a fill ratio of 90%.
Figure 10 shows the change of the temperature distribution along the thermosiphon wall for
four tilt angles (15°, 45°, 75°, and 90°) at a power input of 100 W and fill ratio of 0.5. It displays a
similar development for four tilt angles, where the lowest and highest wall temperature is
present in the upper zone of the evaporator sector, with tilt angles of 90° and 15°, respectively.
At a low tilt angle (15°), the highest temperature may be attributed to the fact that the upper
part of the evaporator sector does not touch the base of the working fluid (ie, liquid phase).
The thermosiphon performance of the CFD simulations is described by the overall thermal
resistance. Figure 11 shows the change in thermal resistance with fill ratios at different power
inputs and vertical positions. The thermal resistance is directly proportional to the difference in
the temperatures between the evaporator and condenser sectors, and inversely proportional to
the overall heat transfer rate. Moreover, this figure indicates that with an increase in the power
16 | AHMED AND AL JUBORI

160 F I G U R E 1 0 The temperature profile at


Tilt angle=15° Tilt angle=45°
140 a power input of 100 W and fill ratio of 0.5
Tilt angle=75° Tilt angle=90°
120 [Color figure can be viewed at
Temperature (°C)

100
wileyonlinelibrary.com]
80

60

40

20

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Thermosiphon length (m)

0.6 F I G U R E 1 1 The thermal resistance


Power input= 50W Power input= 100W profile at a vertical position [Color figure can
Thermal resistance (°C/W)

0.5
Power input= 250W be viewed at wileyonlinelibrary.com]
0.4

0.3

0.2

0.1

0
0.1 0.3 0.5 0.75 0.9
Fill raos

inputs, the thermal resistance decreases, where the minimum thermal resistance can be ob-
tained at a fill ratio of 0.5 for all investigated power inputs. The thermal resistance increases
with the lower power input and decreases with the increase in the power inputs. For high fill
ratios, the thermal resistance is almost constant, thus following the same trend. Figure 12 also
shows the change of thermal resistances with the tilt angles of the water‐charged thermosiphon
of 0.5 and different power inputs. For all power inputs, the lower values of thermal resistance
are achieved in a vertical position.
Figure 13 displays the heat transfer coefficient with the fill ratios at various power inputs in
a vertical position. For all power inputs, the highest values of the heat transfer coefficient are

0.5
Power input=50 W Power input=100 W
0.45
Power input=250 W
Thermal resistance (°C/W)

0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
F I G U R E 1 2 Variation of thermal
0 resistance with tilt angles for different power
15 45 75 90 inputs and a fill ratio of 0.5 [Color figure can
Tilt angle (degree)
be viewed at wileyonlinelibrary.com]
AHMED AND AL JUBORI | 17

F I G U R E 1 3 Variation of heat transfer 450

Heat transfer coefficient (W/m 2.K)


Power input=50 W Power input=100 W
coefficient profile with fill ratios at a vertical 400
Power input=250 W
position [Color figure can be viewed at 350
wileyonlinelibrary.com] 300
250
200
150
100
50
0
0.1 0.3 0.5 0.7 0.9
Fill raos

achieved at a fill ratio of 0.5, and the maximum heat transfer coefficient is obtained at the power
input of 250 W. For each power input, the trends are similar. The fill ratio of 0.5 illustrates good
results in terms of reducing the difference in the temperatures between the evaporator and
condenser sector, thus improving the heat transfer coefficient.

8 | CONCLUSIONS

Describing the liquid‐vapor phase distribution in the two‐phase flow field is still one of the key
problems in modeling. This study has been aimed to develop a computational methodology to
visualize the working fluid phase‐change and heat transfer through the enclosed thermosiphon.
The computational analysis of these processes is one of the essential stages to model the phase
change through the thermosiphon operation by developing the proper source terms in the
governing equations for mass and heat transfer. The computational model of the thermosiphon
was undertaken to examine the effect of a wide range of operating conditions, namely ther-
mosiphon fill ratios, inclination angles, and power inputs. This methodology assisted to reveal
the essential mechanisms of the thermal performance and two‐phase flow of the proposed
thermosiphon for concentrated solar collector applications. The results from the computational
model have been validated with the available experimental results in terms of the average
temperature alongside the thermosiphon surface, indicating that the simulation results were in
good agreement. On the basis of the flow visualization, the simulation results showed that
ANSYS Fluent with the developed VOF scheme can effectively characterize the complicated
phenomena through the thermosiphon, where it is able to simulate the thermosiphon operation
comprising pool boiling in the evaporator sector and vapor condensation film in the condenser
sector. The increase in the power input leads to an increase in the heat transfer coefficient and a
decrease in the thermal resistance. The thermosiphon exhibits better thermal performance as
heat transfer characteristics at a vertical position and a fill ratio of 0.5 for all power inputs. The
proposed computational model has been successfully capturing the fluid phase‐change regimes.
Therefore, this model can be utilized to simplify the experimental investigation procedure and
enhance the two‐phase flow in the enclosed thermosiphon.

ACKNOWLEDGMEN TS
The authors gratefully acknowledge the University of Technology, Iraq, for providing the
facilities that helped in this study.
18 | AHMED AND AL JUBORI

NOMEN C LAT U RE
As surface area, m2
Cp specific heat capacity, J/K·kg
E enthalpy, J/kg
FCSF continuum surface force, N/m3
g gravity, m/s2
h heat transfer coefficient, W/m2·K
p pressure, Pa
Q heat transfer rate, W
SE energy source term, J/m3·s
Sm mass transfer source terms, kg/m3·s
T temperature, K
t time, s
V ⃑ velocity, m/s

GREEK S YMBOLS
α volume fraction
β phase‐change coefficient, s−1
κ interface curvature, m−1
λ thermal conductivity, W/m·K
μ dynamic viscosity, Pa·s
ρ density, kg/m3
σ surface tension, N/m

S U BS C RI P T
av average
cond condenser
evap evaporator
l liquid
sat saturated
v vapor

ORCID
Ayad M. Al Jubori http://orcid.org/0000-0002-7334-554X

REFERENCES
1. Ma G, Zhou F, Liu T, Wang L, Liu Z. Energy‐saving evaluation of a thermosyphon heat recovery unit for an
air‐conditioning system. Heat Transfer. 2013;42(5):377‐388.
2. Salehi H, Zeinali Heris S, Sharifi F, Razbani MA. Effects of a nanofluid and magnetic field on the thermal
efficiency of a two‐phase closed thermosyphon. Heat Transfer. 2013;42(7):630‐650.
3. Fadhl B, Wrobel LC, Jouhara H. CFD modelling of a two‐phase closed thermosyphon charged with R134a
and R404a. Appl Therm Eng. 2015;78:482‐490.
4. Ababsa D, Bougoul S. Modeling of natural convection and radiative heat transfer in inclined thermosyphon
system installed in the roof of a building. Heat Transfer. 2017;46(8):1148‐1157.
5. Zhou F, Ma G, Zhang X. Winter turf growing season elongation characteristics of a thermosyphon unit that
uses a shallow geothermal source. Heat Transfer. 2017;46(7):720‐731.
6. Zhang J, Liu X. Analysis on heat transfer characteristics of constant‐temperature heat pipe in waste heat
utilization for semi‐coke. Heat Transfer. 2017;46(5):434‐446.
AHMED AND AL JUBORI | 19

7. Yue C, Zhang Q, Zhai Z, Ling L. CFD simulation on the heat transfer and flow characteristics of a mi-
crochannel separate heat pipe under different filling ratios. Appl Therm Eng. 2018;139:25‐34.
8. Li Z, Zhang Z. Dynamic heat transfer characteristics of wall implanted with heat pipes in summer. Energy
Build. 2018;170:40‐46.
9. Wang X, Wang Y, Chen H, Zhu Y. A combined CFD/visualization investigation of heat transfer behaviors
during geyser boiling in two‐phase closed thermosyphon. Int J Heat Mass Transfer. 2018;121:703‐714.
10. Akkus Y, Nguyen CT, Celebi AT, Beskok A. A first look at the performance of nano‐grooved heat pipes. Int
J Heat Mass Transfer. 2019;132:280‐287.
11. Liu C, Zhang Z, Shi Y, Ding Y. Optimisation of a wall implanted with heat pipes and applicability analysis in
areas without district heating. Appl Therm Eng. 2019;151:486‐494.
12. Sarafraz M, Pourmehran O, Yang B, Arjomandi M. Assessment of the thermal performance of a thermo-
syphon heat pipe using zirconia‐acetone nanofluids. Renew Energy. 2019;136:884‐895.
13. Wang X, Zhang Q, Zhuang K, He X, Seidl M, Macian‐Juan R. Neutron physics of the liquid‐fuel heat‐pipe
reactor concept with molten salt fuel—static calculations. Int J Energy Res. 2019;43(14):7852‐7865.
14. Manimaran R, Palaniradja K, Alagumurthi N, Hussain J. Experimental comparative study of heat pipe
performance using CuO and TiO2 nanofluids. Int J Energy Res. 2014;38(5):573‐580.
15. Du Y, Le NCH, Chen D, Chen H, Zhu Y. Thermal management of solar cells using a nano‐coated heat pipe
plate: an indoor experimental study. Int J Energy Res. 2017;41(6):867‐876.
16. Shi J, Zhao W, Li J, Liu Z. Heat transfer performance of heat pipe radiator with SiO2/Water nanofluids. Heat
Transfer. 2017;46(7):1053‐1064.
17. Tecchio C, Oliveira J, Paiva K, Mantelli M, Galdolfi R, Ribeiro L. Geyser boiling phenomenon in two‐phase
closed loop‐thermosyphons. Int J Heat Mass Transfer. 2017;111:29‐40.
18. Smith J, Singh R, Hinterberger M, Mochizuki M. Battery thermal management system for electric vehicle
using heat pipes. Int J Therm Sci. 2018;134:517‐529.
19. Abdulshaheed AA, Wang P, Huang G, Li C. High performance copper‐water heat pipes with nanoengi-
neered evaporator sections. Int J Heat Mass Transfer. 2019;133:474‐486.
20. Xiong Y, Wang Z, Wu Y, et al. Performance enhancement of bromide salt by nano‐particle dispersion for
high‐temperature heat pipes in concentrated solar power plants. Appl Energy. 2019;237:171‐179.
21. Liang F, Gao J, Li F, Xu L. An experimental work on thermal features of the miniature revolving heat pipes.
Appl Therm Eng. 2019;146:295‐305.
22. Kujawska A, Zajaczkowski B, Wilde L, Buschmann M. Geyser boiling in a thermosyphon with nanofluids
and surfactant solution. Int J Therm Sci. 2019;139:195‐216.
23. Wang G, Quan Z, Zhao Y, Wang H. Performance of a flat‐plate micro heat pipe at different filling ratios and
working fluids. Appl Therm Eng. 2019;146:459‐468.
24. Eltaweel M, Abdel‐Rehim AA, Hussien H. Indirect thermosiphon flat‐plate solar collector performance
based on twisted tube design heat exchanger filled with nanofluid. Int J Energy Res. 2020;44:4269‐4278.
25. Almahmoud S, Jouhara H. Experimental and theoretical investigation on a radiative flat heat pipe heat
exchanger. Energy. 2019;174:972‐984.
26. Anand AR. Investigations on effect of evaporator length on heat transport of axially grooved ammonia heat
pipe. Appl Therm Eng. 2019;150:1233‐1242.
27. Kim J, Kim SJ. Experimental investigation on the effect of the condenser length on the thermal performance
of a micro pulsating heat pipe. Appl Therm Eng. 2018;130:439‐448.
28. Andrzejczyk R. Experimental investigation of the thermal performance of a wickless heat pipe operating
with different fluids: water, ethanol, and SES36. Analysis of influences of instability processes at working
operation parameters. Energies. 2019;12(1):80.
29. De Schepper SC, Heynderickx GJ, Marin GB. Modeling the evaporation of a hydrocarbon feedstock in the
convection section of a steam cracker. Comput Chem Eng. 2009;33(1):122‐132.
30. Collier JG, Thome JR. Convective Boiling and Condensation. Oxford, UK: Clarendon Press; 1994.
31. Brackbill JU, Kothe DB, Zemach C. A continuum method for modeling surface tension. J Comput Phys.
1992;100(2):335‐354.
32. Parvareh A, Rahimi M, Alizadehdakhel A, Alsairafi A. CFD and ERT investigations on two‐phase flow
regimes in vertical and horizontal tubes. Int Commun Heat Mass Transfer. 2010;37(3):304‐311.
33. Ma H. Oscillating Heat Pipes. New York, NY: Springer; 2015.
20 | AHMED AND AL JUBORI

34. Abdullahi B. Development and Optimization of Heat Pipe based Compound Parabolic Collector[PhD dis-
sertation]. Birmingham, UK: University of Birmingham; 2015.

How to cite this article: Ahmed IS, Al Jubori AM. Assessment of heat transfer and flow
characteristics of a two‐phase closed thermosiphon. Heat Transfer. 2020;1–20.
https://doi.org/10.1002/htj.21933

You might also like