You are on page 1of 37

A CFD analysis on the effect of tube curvature, hot flow

control valve profile, and inlet swirl on the thermal


performance of curved vortex tubes

S.Y. Khan1, U. Allauddin1, S.M.F. Hasani2*, R. Khan2, M. Arsalan3


1
Mechanical Engineering Department, NED University of Engineering & Technology, Karachi,
Pakistan
2
Mechanical & Industrial Engineering Department, College of Engineering, Imam Mohammad Ibn
Saud Islamic University, Riyadh 11432, Saudi Arabia
3
Production Technology Division, EXPEC Advanced Research Center, Saudi Aramco, Dhahran,
Saudi Arabia
*
Corresponding author: smhasani@imamu.edu.sa

ABSTRACT

The influence of tube curvature, conical valve geometry, and the initial swirl on the thermal
performance of vortex tubes is investigated through numerical simulations. Numerical models
of straight and curved vortex tubes are developed in a commercial computational fluid
dynamics package Ansys-fluent®. Multiple models are generated with different curvature
angles, hot nozzle truncated-cone profiles, and the number of inlet nozzles. The effect of each
parameter on the flow and temperature fields inside the vortex tube is analyzed in detail
carrying out three-dimensional simulations using the standard 𝑘𝑘 − 𝜀𝜀 turbulence model. An
unstructured mesh is generated which has randomly distributed discretized cells of the
geometry at the junction of inlet and cold exits of the fluid domain. The cold stream mass
fraction is varied by controlling hot exit pressure. The numerical results are validated through
published experimental values and are found to be in good agreement with the allowable degree
of tolerance. It is observed that truncated-cone hot control valves perform better than non-
truncated valves. However, the optimized truncation length would depend on the type of
application for which the vortex tube is to be used. The tube curvature also has a positive effect
in increasing the hot and cold stream temperature differences and temperature separation as
well as the isentropic efficiencies specifically for curvature angles larger than 150o with cold
mass fraction ranging between 0.3 and 0.7. The number of vortices in the multi-circulation loop
also increases with tube curvature but becomes constant for curvature angles larger than 150o.
The performance of curved tubes can be improved further when curvature effect is combined
with other design changes such as the use of truncated-cone hot flow control valve and an

1
optimum number of inlet nozzles. Three inlet nozzles have been found to produce the initial
swirl that gives the maximum cold stream temperature difference as well as maximum end-to-
end temperature separation for the 110o curved vortex tube. The optimum parameter values for
truncation length, curvature angle and number of inlet nozzles are determined and applied to
simulate the flow through the optimized tube. It is found that for the optimized tube, the cold
stream temperature difference increased by about 23.43 percent and the end-to-end temperature
separation improved by about 37.29 percent when compared with the straight vortex tube. This
novel work is believed to provide insight into the energy separation mechanisms in vortex tubes
and lead the way to their applications in thermo-electric power generation especially in
unaccessible locations such as downhole well applications.

Keywords: Curved vortex tubes, power generation, thermal energy separation, hot nozzle
profile, swirl effect, CFD

1. INTRODUCTION

In an era of constantly rising energy demand, researchers and field engineers are persistently
striving for developing energy harvesting techniques. Energy harvesting is a process of
deriving energy from external sources (e.g., solar power, thermal energy, wind energy, salinity
gradients, and kinetic energy, also known as ambient energy), captured, and stored for small
wireless autonomous devices like wearable electronics and sensor networks. It also has great
potential in power generation and non-renewable energy industry. More specifically in hard to
reach, low accessibility locations such as downhole well applications where localized energy
extraction and utilization is of great importance [1] [2]. At high depths and adverse
underground environment, a significant magnitude of energy is required for running
underground equipment and cooling well-monitoring electronic devices specially lithium
batteries. To meet this energy demand, thermo-electric generators, turbine generators,
piezoelectric devices, and vortex tubes are potentially utilized for efficient energy harvesting
[3] [4] [5] [6].

The Ranque-Hilsch vortex tube that splits a single compressed air stream into two separate hot
and cold streams had been successfully used in the past for cooling, refrigeration, spot cooling,
gas conditioning, and gas separation applications primarily due to its long life and low
operating and maintenance costs [7]. However, its potential in small power generation has not
been fully explored and tested. A significant temperature gradient exists between hot and cold
stream ends of the tube which could be utilized for multiple purposes including power

2
generation using thermo-electric generators. The distance between hot and cold ends of the
tube could be vital for small inaccessible down-hole well locations which may require the use
of curved vortex tubes. The efficiency of the vortex tube depends on the temperature difference
between the hot and cold ends.

Several researchers have studied different types of vortex tubes which are classified into two
categories depending on their flow pattern, one is called uni-flow vortex tube and the other is
the counterflow vortex tube [8]. The counterflow vortex tube is found to be more effective
since the cold and hot stream exits are located at opposite ends which results in better energy
separation. The vortex tube was first introduced by Ranque [8] during an experiment with a
pump where a conical valve was inserted for separating the steam into cold and hot streams.
However, the temperature separation was not that efficient for industrial applications which
was later improved by Hilsch [9].

In Fig. 1, a schematic of straight vortex tube is shown that comprises of the following non-
moving parts: (1) inlet nozzles, (2) vortex chamber or diaphragm, (3) cold exit orifice, (4) hot
exit opening, and (5) conical control valve. The energy separation phenomenon occurs when a
pressurized gas is injected through the inlet nozzle and moves toward the hot annular outlet
forming a free vortex and the backflow from the conical valve towards the cold outlet forming
a forced vortex. The interaction of free and forced vortices creates a stagnation point with zero
local velocity as explained by Fulton [10]. The working of vortex tube is also studied by
Maxwell [11] in a famous thought experiment known as Maxwell’s demon.

The cooling effect from vortex tube is relatively less efficient as compared to conventional air
conditioning but it can be applicable to spot cooling which makes it useful for many low-budget
industrial applications [12]. The spot cooling by vortex tube can be applied to laser cutting,
personal air suit, cooling of high-speed machining tools, cooling soldered parts, food
production, and many more. The vortex tube is simple and robust, requires less to no
maintenance due to a lack of moving parts [13].

Kassner and Knoernschild [14] and Scheper [15] also supported Fulton’s theory where the
friction on or by the elements at the tube core lowers the angular velocity, while the periphery
attains energy transfer with the same free and forced vortices concept. The turbulent swirling
flow follows the free vortex toward the other end and expands to a forced vortex at the core of
the vortex tube.

3
The analytical solutions for velocity and temperature profiles were reported by Deissler and
Perlmutter [16] to support Hilsch’s theory where they concluded that the turbulent shear work
affected the total temperature significantly but the temperature difference was somewhat
overpredicted. Reynolds [17] investigated the vortex tube analytically using an order of
magnitude analysis technique applied to the energy equation. He found that large temperature
and pressure gradients existed at the core because of higher energy transfer due to turbulent
mixing and shear work. He also discovered that the radial work flux was in opposite direction
to the radial mass flux, which suggested that radial flow should mainly be inwards.

(a) Inlet Stream

𝐒 : Stagna�on point
Hot Stream

Cold Stream 𝐒

Hot Stream
Forced Vortex

Inlet Stream Free Vortex

(b) Inlet Stream

𝐒 : Stagna�on point
Hot Stream

Cold Stream 𝐒

Hot Stream

FIGURE 1: Schematic of flow pattern (a) streamlines and (b) swirling vortices in a
counterflow vortex tube.

The energy separation mechanism based on expansion and contraction was discussed by Brunn
[18] and Hinze [19]. In these works, adiabatic expansion and contraction of turbulent eddies
that caused fluctuation of the particles in the flow field were investigated. The adiabatic

4
expansion occurred as the particles moved from a high pressure to a low pressure region and
the temperature of the particles decreased near the core. Similarly, temperature increased
rapidly near the periphery of the vortex tube due to the contraction of turbulent eddies.

Gao [20] built a simple vortex tube and used nitrogen as a working fluid to determine the flow
fields. The secondary circulation was observed which was due to the backflow from the outlet.
Earlier, Ahlborn et al. [21] observed a secondary circulation pattern when using a probe inside
the vortex tube. It was found that the mass flow rate at the cross-section was different than the
mass flow rate exiting the cold orifice. The vortex tube was sectioned at the center where the
mass flow rate was observed to be larger than the outlet. Nimbalkar et al. [22] studied straight
vortex tubes with different cold orifice diameters and found that secondary circulation was a
major performance degrading factor in the analysis of these tubes.

Saidi et al. [23] carried out an experimental study using different geometrical and
thermophysical parameters. They found the optimum value of length to diameter ratio of the
vortex tube diaphragm for maximum efficiency and cold temperature difference by varying the
cold mass fraction. The optimum length to diameter ratio was found to be between 20 and 55.5.
Sadeghiseraji et al. [24] investigated temperature separation inside a vortex tube using both
experimental and computational methods. They tested several turbulence models numerically
and found that the standard 𝑘𝑘 − 𝜀𝜀 model provided the best fit to their experiments which were
conducted at an inlet pressure of 7 bar and it simulated the flow inside the vortex tube
reasonably well.

A number of researchers tried to optimize the performance of the straight vortex tube by
varying its geometry. Rafiee et al. [25] used a truncated cone throttle valve to improve the
energy separation phenomenon and used the cone angle of the control valve as an optimizing
parameter. Rafiee and a few other researchers [26] also analyzed the effect of the number of
inlet nozzles on the efficiency of vortex tube. The results of these studies revealed that it was
difficult to make a conclusive argument in this respect as the optimum number of nozzles
largely depended on the inlet conditions and the geometry of the tube. However, a general
observation was made that more nozzles increase turbulence and reduce temperature separation
which could be controlled by modification of the tube or the control valve geometry.

Most of the prior research was done on straight vortex tubes and little work was done on curved
vortex tubes. As pointed out earlier that the use of curved vortex tubes may be necessary for
some specialized applications. Valipour et. al. [27] experimentally studied the performance of

5
curved vortex tubes (CVT) and determined the isentropic efficiencies from temperature
differences at the outlets. Bovand et al. [28] [29] performed numerical investigations on CVT
based on Valipour’s experimental data, using curvature angles of 30°, 60°, 90°, 110°, and
150° together with straight vortex tube (SVT). It was found that the 150° CVT had almost
similar performance as the SVT.

To the knowledge of the authors, no research study is available that combines the effects of
initial inlet swirl and conical valve geometry with the curvature of the vortex tube.
Furthermore, no data for CVT is found in the literature beyond curvature angles of 150o. In this
paper, a numerical model is presented and validated using published experimental data of SVT
and CVT (110o). The effects of the inlet nozzle count that generate the initial swirl and
truncation length of the conical control valve on the isentropic efficiencies and temperature
separation in CVT are investigated in detail and an optimized model is presented. Simulated
results of CVTs with curvature angles larger than 150o have also been included.

2. GOVERNING EQUATIONS

The swirling turbulent flow inside the counterflow vortex tube is modeled using the steady
state form of the continuity, the momentum, and the energy equations, given by Eqs. (1), (2),
and (3), respectively. Air is used as the working fluid and its compressibility is modeled
employing the ideal gas equation, Eq. (4).

𝜕𝜕(𝜌𝜌𝑣𝑣𝑖𝑖 ) (1)
=0
𝜕𝜕𝑥𝑥𝑖𝑖

𝜕𝜕(𝜌𝜌𝑣𝑣𝑖𝑖 𝑣𝑣𝑗𝑗 ) 𝜕𝜕 𝜕𝜕𝑣𝑣𝑖𝑖 𝜕𝜕𝑣𝑣𝑗𝑗 2 𝜕𝜕𝑣𝑣𝑘𝑘 𝜕𝜕𝜕𝜕 𝜕𝜕 (2)


= �𝜇𝜇 � + − 𝛿𝛿𝑖𝑖𝑖𝑖 �� − + ������
�−𝜌𝜌𝑣𝑣 ′ ′
𝚤𝚤 𝑣𝑣𝚥𝚥 �
𝜕𝜕𝑥𝑥𝑖𝑖 𝜕𝜕𝑥𝑥𝑗𝑗 𝜕𝜕𝑥𝑥𝑗𝑗 𝜕𝜕𝑥𝑥𝑖𝑖 3 𝜕𝜕𝑥𝑥𝑘𝑘 𝜕𝜕𝑥𝑥𝑗𝑗 𝜕𝜕𝑥𝑥𝑗𝑗

𝜕𝜕 1 𝜕𝜕 𝜕𝜕𝑇𝑇� 𝑐𝑐𝑝𝑝 𝜇𝜇𝑡𝑡 (3)


�𝜌𝜌𝑣𝑣𝑖𝑖 �ℎ + 𝑣𝑣𝑖𝑖 𝑣𝑣𝑗𝑗 �� = �𝑘𝑘𝑒𝑒𝑒𝑒𝑒𝑒 + 𝑣𝑣𝑖𝑖 �𝜏𝜏𝑖𝑖𝑖𝑖 �𝑒𝑒𝑒𝑒𝑒𝑒 � , 𝑘𝑘𝑒𝑒𝑒𝑒𝑒𝑒 = 𝐾𝐾 +
𝜕𝜕𝑥𝑥𝑖𝑖 2 𝜕𝜕𝑥𝑥𝑗𝑗 𝜕𝜕𝑥𝑥𝑗𝑗 𝑃𝑃𝑟𝑟𝑡𝑡

𝑃𝑃 = 𝜌𝜌𝜌𝜌𝜌𝜌 (4)

The standard 𝑘𝑘 − 𝜀𝜀 model is applied to solve the RANS equations for the simulation of flow
inside the computational domain of the vortex tube. The turbulent kinetic energy (𝑘𝑘) and the
rate of dissipation (𝜀𝜀) are represented by Eqs. (5) and (6) respectively:

6
𝜕𝜕(𝜌𝜌𝑣𝑣𝑖𝑖 𝑘𝑘) 𝜕𝜕 𝜇𝜇𝑡𝑡 𝜕𝜕𝜕𝜕 (5)
= ��𝜇𝜇 + � � + 𝐺𝐺𝑘𝑘 + 𝐺𝐺𝑏𝑏 − 𝜌𝜌𝜌𝜌 − 𝑌𝑌𝑀𝑀
𝜕𝜕𝑥𝑥𝑖𝑖 𝜕𝜕𝑥𝑥𝑗𝑗 𝜎𝜎𝑘𝑘 𝜕𝜕𝑥𝑥𝑗𝑗

𝜕𝜕(𝜌𝜌𝑣𝑣𝑖𝑖 𝜀𝜀) 𝜕𝜕 𝜇𝜇𝑡𝑡 𝜕𝜕𝜕𝜕 𝜀𝜀 𝜀𝜀 2 (6)


= ��𝜇𝜇 + � � + 𝐶𝐶1𝜀𝜀 (𝐺𝐺𝑘𝑘 + 𝐶𝐶3 𝐺𝐺𝑏𝑏 ) − 𝐶𝐶2𝜀𝜀 𝜌𝜌
𝜕𝜕𝑥𝑥𝑖𝑖 𝜕𝜕𝑥𝑥𝑗𝑗 𝜎𝜎𝜀𝜀 𝜕𝜕𝑥𝑥𝑗𝑗 𝑘𝑘 𝑘𝑘

where 𝐺𝐺𝑘𝑘 , 𝐺𝐺𝑏𝑏 , and 𝑌𝑌𝑀𝑀 each represent the turbulent kinetic energy generation due to the mean
velocity gradient, the turbulent kinetic energy generation due to buoyancy, and the fluctuating
dilation turbulence contribution in the overall dissipation rate, respectively. 𝐶𝐶1𝜀𝜀 and 𝐶𝐶2𝜀𝜀 are
constants in the dissipation rate equation. Turbulent Prandtl numbers for 𝑘𝑘 and 𝜀𝜀 are
represented as 𝜎𝜎𝑘𝑘 and 𝜎𝜎𝜀𝜀 respectively. The eddy viscosity is formulated using Eq. (7) as:

𝑘𝑘 2 (7)
𝜇𝜇𝑡𝑡 = 𝜌𝜌𝐶𝐶𝜇𝜇
𝜀𝜀

where 𝐶𝐶𝜇𝜇 is a constant. The default values of the model constants are:

𝐶𝐶1𝜀𝜀 = 1.44, 𝐶𝐶2𝜀𝜀 = 1.92, 𝐶𝐶𝜇𝜇 = 0.09, 𝜎𝜎𝑘𝑘 = 1.0, 𝜎𝜎𝜀𝜀 = 1.3.

3. MODEL DESCRIPTION

The vortex tube dimensions employed to generate the computational model are provided in
Table 1. A computer-aided model of the 110o curved vortex tube with dimensions is shown in
Fig. 2. Similarly, the computational models of a straight vortex tube together with two other
curved vortex tube models with curvature angles of 180° and 270° are created. The cold outlet
is extruded by 1 mm near the inlet of the vortex tube which is half the inner tube diameter as
suggested by Saidi et al. [23].

TABLE 1: Design specifications of the counterflow curved vortex tube

Geometric Parameter Dimension

Working tube length (𝐿𝐿) 400 mm


Inner tube diameter (𝐷𝐷) 19.05 mm

Cold exit diameter (𝑑𝑑𝑐𝑐 ) 9.53 mm


Nozzle diameter (𝑑𝑑𝑛𝑛 ) 4.00 mm
Hot exit area (𝐴𝐴ℎ ) 58.17 mm2
Nozzle total inlet area (𝐴𝐴𝑛𝑛 ) 25.13 mm2

7
2
Inlet Nozzle:
Pressure Inlet
9.53 Adiaba�c Walls,
No-slip condi�on
Cold Exit:
Pressure Outlet

Hot control valve

Hot Exit:
Pressure Outlet

FIGURE 2: CAD drawing and dimension of the 110o curved vortex tube.

The geometry of the hot conical valve, that is used to control the flow inside the vortex tube is
defined by its height 𝐿𝐿ℎ = 18.15 𝑚𝑚𝑚𝑚 and diameter 𝐷𝐷ℎ = 17 𝑚𝑚𝑚𝑚 and is shown in Fig. 3. The
truncated length 𝑙𝑙ℎ of the conical valve is varied from 0 to 12 mm to study the effect of control
valve profile on the thermal performance of the vortex tube while the cone apex angle 𝛼𝛼 is kept
constant at 50o.

FIGURE 3: Schematic of (a) non-truncated and (b) truncated hot conical valve.

The thermal performance of a vortex tube is assessed through several parameters such as the
cold stream temperature difference, the hot stream temperature difference, the isentropic
efficiency, and the coefficient of performance (COP).

The cold stream temperature difference Δ𝑇𝑇𝑐𝑐 is the difference between the temperature of the
inlet stream 𝑇𝑇𝑖𝑖 and the temperature of cold stream 𝑇𝑇𝑐𝑐 . It is represented using Eq. (8).

8
Δ𝑇𝑇𝑐𝑐 = 𝑇𝑇𝑖𝑖 − 𝑇𝑇𝑐𝑐 (8)

The hot stream temperature difference Δ𝑇𝑇ℎ is the difference between the temperature of hot
stream 𝑇𝑇ℎ and the temperature of the inlet stream 𝑇𝑇𝑖𝑖 . It is given by Eq. (9).

Δ𝑇𝑇ℎ = 𝑇𝑇ℎ − 𝑇𝑇𝑖𝑖 (9)

The thermal performance is determined mainly by the cold stream temperature difference if the
device is to be used for cooling purposes. However, the hot stream temperature difference can
be used to assess the performance in heating applications. When vortex tubes are used in power
generation, their performance will be judged by the total temperature difference (Δ𝑇𝑇ℎ𝑐𝑐 )
between the hot and the cold streams, given by Eq. (10).

Δ𝑇𝑇ℎ𝑐𝑐 = 𝑇𝑇ℎ − 𝑇𝑇𝑐𝑐 (10)

The isentropic cooling efficiency 𝜂𝜂𝑐𝑐 determines the thermal efficiency of the cold end of the
vortex tube with the assumption that the gas expands ideally with no heat loss. Therefore, the
isentropic efficiency is the ratio of the actual enthalpy change and the isentropic enthalpy
change. It can be calculated by Eq. (11).

𝑚𝑚̇ 𝑐𝑐 𝑐𝑐𝑝𝑝 (𝑇𝑇𝑖𝑖 − 𝑇𝑇𝑐𝑐 ) 𝜉𝜉Δ𝑇𝑇𝑐𝑐 (11)


𝜂𝜂𝑐𝑐 = =
𝑚𝑚̇ 𝑖𝑖𝑖𝑖 𝑐𝑐𝑝𝑝 �𝑇𝑇𝑖𝑖 − 𝑇𝑇′𝑐𝑐 � 𝑃𝑃 𝛾𝛾−1
𝑇𝑇𝑖𝑖 �1 − ( 𝑐𝑐 ) 𝛾𝛾 �
𝑃𝑃𝑖𝑖
where 𝑃𝑃𝑐𝑐 , 𝑃𝑃𝑖𝑖 , 𝜉𝜉 and 𝛾𝛾 are the cold stream exit pressure, the inlet pressure, cold mass fraction
and the specific heat ratio (𝛾𝛾 = 𝑐𝑐𝑝𝑝 /𝑐𝑐𝑣𝑣 ) respectively. The cold mass fraction is defined as the
ratio of the mass flow rate of the cold stream to the mass flow rate of the inlet stream, which
may be mathematically represented by Eq. (12).

𝑚𝑚̇ 𝑐𝑐 (12)
𝜉𝜉 =
𝑚𝑚̇ 𝑖𝑖𝑖𝑖
Similarly, the isentropic heating efficiency 𝜂𝜂ℎ is formulated from the hot stream exit pressure
𝑃𝑃ℎ , the inlet pressure 𝑃𝑃𝑖𝑖 , and hot temperature difference Δ𝑇𝑇ℎ , which is given by Eq. (13).

𝑚𝑚̇ ℎ 𝑐𝑐𝑝𝑝 (𝑇𝑇ℎ − 𝑇𝑇𝑖𝑖 ) (1 − 𝜉𝜉)Δ𝑇𝑇ℎ (13)


𝜂𝜂ℎ = =
𝑚𝑚̇ 𝑖𝑖𝑖𝑖 𝑐𝑐𝑝𝑝 �𝑇𝑇𝑖𝑖 − 𝑇𝑇′ℎ � 𝑃𝑃ℎ 𝛾𝛾−1
𝑇𝑇𝑖𝑖 �1 − ( ) 𝛾𝛾 �
𝑃𝑃𝑖𝑖
The isentropic energy separation efficiency 𝜂𝜂ℎ𝑐𝑐 is calculated from hot to cold temperature and
pressure values that is given by Eq. (14).

9
(1 − 𝜉𝜉)Δ𝑇𝑇ℎ + 𝜉𝜉Δ𝑇𝑇𝑐𝑐 (14)
𝜂𝜂ℎ𝑐𝑐 =
𝑃𝑃 𝛾𝛾−1 𝑃𝑃 𝛾𝛾−1
𝑇𝑇𝑖𝑖 �( ℎ ) 𝛾𝛾 − ( 𝑐𝑐 ) 𝛾𝛾 �
𝑃𝑃𝑖𝑖 𝑃𝑃𝑖𝑖

4. COMPUTATIONAL DOMAIN AND BOUNDARY CONDITIONS

An unstructured mesh is created with randomly distributed discretized cells of the geometry at
the junction of inlet and cold exit of the fluid domain as shown in Fig. 4. The hexahedrons are
created across the swept area of both tubes. The quality of the mesh and the wall 𝑦𝑦+ are two
important parameters for generating the mesh where the default 𝑦𝑦+ requirement is set to more
than 30 for the standard 𝑘𝑘 −𝜀𝜀 turbulence model.

The cold stream mass fraction, 𝜉𝜉 is varied by controlling the hot exit pressure using the
parametric analysis in ANSYS Workbench®. The mass flow rates at both outlets are
determined and then formulated by taking the ratio of cold and inlet mass flow rates. Improper
selection of boundary conditions and/or wrong choice of turbulence model may result in the
under-prediction of the temperature profile. The RNG turbulence model was used by Bovand
et. al. [28] but in the present work, it was found that its convergence was slow due to steep
gradients and the sensitive nature of the swirling flow. The boundary conditions for the
computational domain are provided in Table 2 for the nozzle inlet, cold and hot outlets, and the
enclosure walls.

TABLE 2 Boundary conditions for curved vortex tube (Bovand et. al [28])

Boundary Type Boundary Conditions Value

Inlet Pressure inlet 2 × 105 Pa (total)


Cold exit Pressure outlet 101325 Pa (static)
Hot exit Pressure outlet Varied with different 𝜉𝜉
Walls Adiabatic with no-slip Heat flux = 0 𝑊𝑊 𝑚𝑚−2

The pressure inlet condition is selected for nozzle entrance with a total pressure of 2 bar at an
inlet temperature of 300 K. The pressure outlet condition is chosen for both cold and hot outlets
with a total backflow temperature of 300 K. The static pressure ranges from 1.19 to 1.61 bar at
the hot annular outlet to control the cold mass fraction. The operating pressure is set to zero so
that absolute pressure values are used for density calculation. The walls of the vortex tube are

10
assumed completely adiabatic with the no-slip condition for velocity. The atmospheric pressure
is assumed to be 1.0125 bar.

FIGURE 4 Computational grid of curved vortex tube consisting of tetrahedral and


hexahedral cells with y+ near-wall grid spacing.

5. SOLVER SETUP

The governing equations are solved numerically for the computational models developed for
the straight and curved vortex tubes using ANSYS FLUENT® software. The grid refinement
is performed based on Richardson’s technique and mesh quality is evaluated using the standard
criteria [30]. The finite volume solver, Semi-Implicit Method for Pressure-Linked Equations
(SIMPLE) is used for the pressure velocity coupling. The second-order upwind scheme is
applied for the discretization of all the governing equations. The residuals are limited to 10−6
as the convergence criterion for all the governing equations. Default values are used for the
under-relaxation factors in all the governing equations.

6. GRID INDEPENDENCE STUDY

The discretization errors are lessened by a grid convergence method called the grid
convergence index (GCI) approach proposed by Celik et al. [30] and originally based on
Roache’s uncertainty computational technique. The grid independence study is performed for
the straight vortex tube with three different grid sizes as shown in Table 3. A similar study is

11
carried out for the curved vortex tube that is shown in Table 4. The grid refinement factor is
set to a constant value of 1.45 in both cases. The inflation layer is set for the standard wall
function, with 𝑦𝑦 + between 30 and 100. The solution is obtained for the cold stream
temperature difference for coarse, medium, and fine grids with cold mass fractions (𝜉𝜉) values
ranging from 0.1 to 0.85. The results are plotted in Fig. 5 and Fig. 6 respectively.

TABLE 3 Grid Convergence Index study for 𝚫𝚫𝑻𝑻𝒄𝒄 of SVT at 𝝃𝝃 = 𝟎𝟎. 𝟑𝟑

Grid No. of elements 𝒚𝒚 + 𝚫𝚫𝑻𝑻𝒄𝒄 (𝑲𝑲)

Coarse 114217 64 13.76


Medium 355483 63 14.67
Fine 1084245 60 14.88

(a) (b)

FIGURE 5 A comparison of the cold temperature difference distribution for different


grids of SVT at different cold mass fractions.

TABLE 4 Grid Convergence Index study for 𝚫𝚫𝑻𝑻𝒄𝒄 of 𝟏𝟏𝟏𝟏𝟏𝟏° CVT at 𝝃𝝃 = 𝟎𝟎. 𝟑𝟑

Grid No. of elements 𝒚𝒚 + 𝚫𝚫𝑻𝑻𝒄𝒄 (𝑲𝑲)

Coarse 114217 64 13.31


Medium 355483 63 13.57
Fine 1084245 60 13.60

An average grid size is selected for the vortex tubes for which the solution is independent of
the grid at 0.5 mm. This generates 1 million elements for both the SVT and the CVT
geometries. The GCI value is found to be 1.0135 percent for the vortex tube and the expression

12
for asymptotic convergence is equal to 1.177 which is close to 1. The error bands are formulated
using the GCI values of different data points of the temperature difference and cold mass
fraction as shown in Fig. 5 and Fig. 6. The results show good convergence for different grid
resolutions.

(a)

FIGURE 6 A comparison of the cold temperature difference distribution for different


grids of CVT at different cold mass fractions.

7. MODEL VALIDATION

The computational models of the straight vortex tube (SVT) and the curved vortex tube (CVT)
with a curvature of 110° are validated using the experimental data of Valipour and Niazi [27].
The temperature differences at cold and hot outlets for both models are obtained by varying
the cold mass fraction (𝜉𝜉). The numerical results are plotted in Fig. 7 which shows good
agreement with the experimental data. In Fig. 7 simulated results are presented for the cold
stream temperature difference, Δ𝑇𝑇𝑐𝑐 and hot stream temperature difference, Δ𝑇𝑇ℎ for the SVT
and the 110° CVT as a function of the cold mass fraction, 𝜉𝜉. The standard 𝑘𝑘 −𝜀𝜀 turbulence
model is employed to model turbulence. Earlier Skye et al. [31] and Pourmahmoud et al. [32]
also demonstrated the effective use of the standard 𝑘𝑘 −𝜀𝜀 turbulence model in their work. The
minimum temperature for the cold outlet is obtained at a cold mass fraction of 𝜉𝜉 = 0.3 and the
maximum temperature for hot outlet is obtained at cold mass fraction of 𝜉𝜉 = 0.85. The
minimum cold outlet temperatures are found to be 285.11 K and 286.40 K for the SVT and the
110° CVT while the maximum hot outlet temperatures are 305.18 K and 304.82 K,
respectively.

13
(a) (b)

FIGURE 7 A comparison of the predicted temperature difference distribution of SVT and


110o CVT at different cold mass fractions with the experimental data (Valipour [27]).

8. RESULTS AND DISCUSSION

The flow through a counterflow CVT is numerically analyzed, and the cold stream temperature
difference, the hot stream temperature difference, and the energy separation are determined.
The performance of the CVT is optimized by modifying its design, which includes (i) changing
the hot conical valve profile, (ii) varying the number of inlet nozzles, and (iii) changing the
curvature angle of the tube. The results of the optimized CVT are compared with those of the
SVT.

8.1. EFFECT OF HOT CONICAL VALVE PROFILE

The effect of changing the hot conical valve profile is studied by truncating the cone. Results
are obtained for five different truncation lengths 𝑙𝑙ℎ = 0, 3, 6, 8, 12 𝑚𝑚𝑚𝑚 in the 110° CVT, 𝑙𝑙ℎ =
0 represents the non-truncated valve. The maximum and minimum temperatures at the hot and
cold outlets are determined as a function of the cold mass fraction (𝜉𝜉) and the results are
summarized in Table 5. In Fig. 8, the variation of the cold stream temperature difference Δ𝑇𝑇𝑐𝑐 ,
the hot stream temperature difference Δ𝑇𝑇ℎ , and the hot to cold stream temperature difference
Δ𝑇𝑇ℎ𝑐𝑐 are plotted against the cold mass fraction (𝜉𝜉) for different values of truncation lengths
(𝑙𝑙ℎ ). The minimum temperature obtained at the cold outlet is 286.40 K at a 𝜉𝜉 value of 0.3 for
the 110° CVT with non-truncated conical valve. The truncation of the conical valve further
decreased the cold outlet temperature to 283.46 K for a truncation length (𝑙𝑙ℎ ) of 8 mm, thus
causing a reduction of about 21.62 percent. For 𝑙𝑙ℎ values greater than 8 mm, the cold outlet
temperature again increased causing turbulent mixing and a decrease in the energy separation.

14
The reason for this increase is the formation of a vortex at the flat face of the truncated conical
valve which can be seen in the enlarged view of the hot stream end of the CVT in Fig. 9.

TABLE 5 Total Temperature variation with 𝝃𝝃 at hot and cold exits of the 110o CVT at
different truncated lengths of hot conical valve

𝑙𝑙ℎ = 0 𝑚𝑚𝑚𝑚 𝑙𝑙ℎ = 3 𝑚𝑚𝑚𝑚 𝑙𝑙ℎ = 6 𝑚𝑚𝑚𝑚 𝑙𝑙ℎ = 8 𝑚𝑚𝑚𝑚 𝑙𝑙ℎ = 12 𝑚𝑚𝑚𝑚
𝜉𝜉 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐
0.1 301.49 289.48 301.68 288.90 302.55 287.45 302.79 285.69 301.39 290.29
0.2 303.19 287.62 303.84 285.35 305.09 284.23 305.29 284.08 303.20 285.40
0.3 304.82 286.40 304.90 284.78 307.63 283.68 307.79 283.46 305.02 284.97
0.4 306.64 286.67 307.13 285.03 310.17 284.46 310.29 284.60 307.54 285.68
0.5 309.14 287.07 309.34 285.29 312.72 286.02 312.79 285.65 309.67 287.15
0.6 311.64 288.26 311.56 286.88 315.26 287.57 315.30 287.13 312.00 288.61
0.7 314.21 289.59 314.37 288.47 317.80 289.69 317.80 289.06 313.67 290.56
0.85 316.78 292.99 317.19 292.57 321.62 291.80 321.75 291.73 315.34 292.50

(a) (b)

(c)

FIGURE 8 Comparison of (a) the cold temperature difference, (b) the hot temperature
difference, and (c) the hot to cold stream temperature difference of 110° CVT for different
truncated lengths of the hot conical valve as a function of cold mass fraction.
15
FIGURE 9 A comparison of velocity streamlines in 110° CVT with truncated and a non-
truncated conical valve at a cold mass fraction of 𝜉𝜉 = 0.3.

A comparison of the static temperature distribution and the velocity contours in 110o CVT with
truncated and non-truncated conical valve is shown in Fig. 9. The comparison is made at a cold
mass fraction (𝜉𝜉) value of 0.3 while a truncated length (𝑙𝑙ℎ ) of 8 𝑚𝑚𝑚𝑚 is used that gives the
minimum cold stream temperature. In Fig. 10, the cooling, heating, and energy separation
efficiencies are plotted against the cold mass fraction (𝜉𝜉) for different values of truncation
lengths (𝑙𝑙ℎ ). These plots show that the efficiency for the cold stream is found to be best for a
𝑙𝑙ℎ value of 3 mm followed by 𝑙𝑙ℎ equal to 8 mm. It is maximum when 𝜉𝜉 is in the range of 0.6 –
0.7. The hot stream efficiency is best when 𝑙𝑙ℎ is between 6 and 8 mm and its maximum occurs
at a 𝜉𝜉 value of 0.5. The energy separation efficiency is also found to be best for a 𝑙𝑙ℎ value of 8
mm, only slightly higher than the efficiency at 𝑙𝑙ℎ equal to 6 mm with their maximums occurring
at 𝜉𝜉 equal to 0.5.

16
(a) (b)

(c)

FIGURE 10 Comparison of (a) the cooling efficiency, (b) the heating efficiency, and (c)
the energy separation efficiency of 110° CVT for different truncated lengths of the hot
conical valve as a function of cold mass fraction.

8.2. EFFECT OF TUBE CURVATURE

The effect of tube curvature is numerically investigated using curvature angles 𝜃𝜃 =


0°, 110°, 180° and 270° in the counterflow vortex tube. The cold stream temperature has been
found to decrease in the 180° and 270° CVTs when compared to the values obtained for the
110° tube. The total temperature values of hot and cold streams are provided in Table 6 for
different curvature angles of the vortex tube with different cold mass fractions. The cold stream
temperature difference (Δ𝑇𝑇𝑐𝑐 ), the hot stream temperature difference (Δ𝑇𝑇ℎ ) and the stream
temperature separation (Δ𝑇𝑇ℎ𝑐𝑐 ) values are plotted for different cold mass fractions (𝜉𝜉) in Figs.
11(a) to 11(c) respectively, whereas the variation of cooling, heating and energy separation
isentropic efficiencies with 𝜉𝜉 are shown in Figs. 12(a) to 12(c) respectively. It can be seen from
Figs. 12(a) and 12(b) that the maximum cold stream and hot stream isentropic efficiencies
respectively are found to be almost identical for both 180° and 270° CVTs but the performance

17
of 180° CVT is slightly better. However, the smaller end-to-end distance in 270o CVT can
justify its use without much loss in efficiency in applications where it is considered a critical
parameter such as in thermo-electric power generation. The energy separation efficiency has
been found to improve in 270° CVT as shown in Fig. 12(c). In Fig. 13, the total temperature
variation in SVT and different CVTs is shown at a 𝜉𝜉 value of 0.3. The minimum temperature
at the cold outlet is found to be 284.32 K and 284.34 K for 180° and 270° CVTs respectively
whereas for the SVT and the 110° CVT these values are 285.11 K and 286.40 K respectively.

The stagnation point is the location where the axial velocity is zero and is considered as the
endpoint of secondary circulation which is an important parameter in the analysis as described
by Sadeghiseraji et al. [24] and Pourmahmoud et al. [32]. The length of the secondary
circulation loop (SC-loop), which is the distance from the cold outlet to the stagnation point, is
often used to compare the performance of different CVTs.

TABLE 6 Total Temperature distribution at hot and cold exit of the vortex tube of
different curvature angles

𝜃𝜃 = 0° 𝜃𝜃 = 110° 𝜃𝜃 = 180° 𝜃𝜃 = 270°


𝜉𝜉 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐
0.1 302.13 287.42 301.49 289.48 301.54 287.16 301.56 287.56
0.2 303.62 286.15 303.19 287.62 303.23 285.37 303.26 286.49
0.3 305.18 285.11 304.82 286.40 305.53 284.32 305.47 284.34
0.4 307.01 285.64 306.64 286.67 307.71 284.81 307.67 285.11
0.5 309.23 286.51 309.14 287.07 309.57 285.56 309.87 285.75
0.6 312.01 287.90 311.64 288.26 311.98 287.42 312.08 287.76
0.7 314.65 289.68 314.21 289.59 314.28 289.45 314.28 289.37
0.85 317.30 292.28 316.78 292.99 317.32 292.95 317.69 292.85

18
(a) (b)

(c)

FIGURE 11 A comparison of (a) the cold temperature difference, (b) the hot temperature
difference, and (c) the hot and cold stream temperature difference of CVT with different
curvature angles varying the cold mass fraction.

19
(a) (b)

(c)

FIGURE 12 A comparison of (a) the cooling efficiency, (b) the heating efficiency, and (c)
the energy separation efficiency of CVT with different curvature angles varying the cold
mass fraction.

20
FIGURE 13 Total temperature contours in SVT and CVTs with different curvature angles
at a cold mass fraction of 𝜉𝜉 = 0.3.

In Fig. 14, the velocity streamlines are superimposed over the static temperature contours. The
free and forced vortices combine and meet at the stagnation point. The number of vortices and
the flow circulation patterns formed in different vortex tubes vary with tube curvature and
different flow patterns may be observed. The length of the SC-loop also increases with the
curvature angle, however, the number of vortices became constant beyond a certain curvature
angle and this number is found to be 9 for all the curvature angles considered in the present
work. These observations were also confirmed earlier by Bovand et. al. [29] who also reported
9 vortices in the multi-circulation region when the curvature angle was varied from 60o to 150o.
The energy separation is also seen to improve as the length of the SC-loop increases.

21
FIGURE 14 A comparison of velocity streamline of CVTs with different curvature angles
at a cold mass fraction of 𝜉𝜉 = 0.3.

8.3. EFFECT OF INLET SWIRL

The vortex tube performance also depends on the initial swirl in the flow at inlet. For a given
inlet pressure, the initial swirl may be varied by varying the number of inlet nozzles. To study
this effect, the 110o CVT is tested with 2, 3, 4 and 6 equally angularly spaced nozzles placed
around the circumference of the vortex tube at the inlet section. The performance of these CVTs
is found to be best with 3 inlet nozzles with a cold stream temperature of 285 K. The
performance deteriorates when this number is increased or decreased. There is however a
consensus among the researchers that no conclusive argument can be made in this respect as
the performance based on this parameter would depend on a number of other geometric and

22
operational parameters. For a constant nozzle diameter of 4 𝑚𝑚𝑚𝑚, the cold outlet temperature
increases due to strong swirling flow inside the tube. The total temperature values of the hot
and the cold streams are provided in Table 7 for different nozzle counts as a function of the
cold mass fraction (𝜉𝜉). The Δ𝑇𝑇𝑐𝑐 , Δ𝑇𝑇ℎ , Δ𝑇𝑇ℎ𝑐𝑐 values for different nozzle counts are plotted versus
𝜉𝜉 and the variation is shown in Fig. 15. The change in temperature differences is quite
significant as the nozzle count is increased. All three temperature differences increased when
the nozzle count is increased from 2 to 3 but they decreased when the count is increased beyond
3.

In Fig. 16 the effect of increasing the number of inlet nozzles is shown by plotting the
temperature contours for the 110o CVT at a 𝜉𝜉 value of 0.3. It can be seen from this figure that
as the number of nozzles decreased, the free vortex region at the periphery of the vortex tube
widened causing the hot temperature region to expand. The number of inlet nozzles is an
important parameter and hence must be chosen carefully. It affects the initial swirl, higher the
initial swirl, higher will be the energy separation which would gradually deteriorate under the
influence of the secondary circulation flow. The reason for the decline in the energy separation
process is turbulent mixing. Though the multi-circulation region is still present but the increase
in the mass flow rate resulted in the expansion of the free vortex region.

TABLE 7 Total temperatures at hot and cold exits of the 𝟏𝟏𝟏𝟏𝟏𝟏° CVT with different inlet
nozzle counts

𝑁𝑁 = 2 𝑁𝑁 = 3 𝑁𝑁 = 4 𝑁𝑁 = 6
𝜉𝜉 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐 𝑇𝑇ℎ 𝑇𝑇𝑐𝑐
0.1 301.49 289.48 302.19 288.78 302.42 289.39 302.11 290.21
0.2 303.19 287.62 305.09 285.57 303.54 287.44 302.76 289.48
0.3 304.82 286.40 306.82 285.00 305.01 286.38 303.34 288.89
0.4 306.64 286.67 308.01 285.01 306.18 286.95 303.96 288.70
0.5 309.14 287.07 310.29 285.72 307.72 287.56 304.85 289.72
0.6 311.64 288.26 313.04 286.21 309.87 288.80 306.02 290.85
0.7 314.21 289.59 316.91 287.75 313.10 291.19 308.01 293.61
0.85 316.78 292.99 320.79 289.29 315.24 294.78 312.81 296.52

23
(a) (b)

(c)

FIGURE 15 A comparison of (a) the cold temperature difference, (b) the hot temperature
difference, and (c) hot and cold stream temperature difference of 110° CVT for different
nozzle counts and varying the cold mass fraction.

24
FIGURE 16 A comparison of total temperature contours of 110° CVT with different
nozzle counts at cold mass fraction of 𝜉𝜉 = 0.3.

The thermal performance of each of these tubes is examined on the basis of isentropic
efficiencies of the cold stream, the hot stream and energy separation for 2, 3, 4, and 6 nozzles
at different 𝜉𝜉 values and the results are presented in Fig. 17(a) to 17(c) respectively. It may be
observed from these figures that best hot and cold stream efficiencies are obtained with three
inlet nozzles, but the energy separation efficiency is found to be better with two nozzles.

25
(a) (b)

(c)

FIGURE 17 A comparison of (a) the cooling efficiency, (b) the heating efficiency, and (c)
the energy separation efficiency of 110° CVT for different nozzle counts and varying the
cold mass fraction.

8.4. OPTIMIZED CURVED VORTEX TUBE

Based on the results presented in sections 8.1 to 8.3, the performance of the CVT can be
optimized by choosing the proper tube curvature, correct number of inlet nozzles and the right
profile for conical flow control valve. Simulations are carried out using the optimized values
of the above parameters (lh=8 mm, θ =180o, N=3) for the given inlet and boundary conditions.
The optimized vortex tube resulted in a maximum cold stream temperature (𝑇𝑇𝑐𝑐 ) = 281. 62 K
at a cold mass fraction of 0.3. This temperature is the lowest cold stream temperature obtained
in the present analysis, it provided an improvement in Δ𝑇𝑇𝑐𝑐 value of about 23.43 percent over
the SVT and about 17.22 percent over the 180o CVT with two inlet nozzles and a non-truncated
conical valve. The temperature differences Δ𝑇𝑇𝑐𝑐 , Δ𝑇𝑇ℎ , Δ𝑇𝑇ℎ𝑐𝑐 are plotted as a function of 𝜉𝜉 for the
optimized and straight vortex tubes in Fig. 18. The results clearly show that the performance
of the optimized tube is better than SVT under all conditions and for all applications of its

26
intended use whether it be cooling, heating or power generation. In Fig. 19, the cold stream,
the hot stream and the energy separation isentropic efficiencies of the optimized tube are
compared with the SVT, where both cooling and heating efficiencies have improved in the
optimized tube. However, the energy separation efficiency is still found to be higher in the SVT
than for the optimized tube.

FIGURE 18 A comparison of the cold temperature difference, the hot temperature


difference, and hot and cold stream temperature difference of optimized 180° CVT (𝑁𝑁 = 3,
𝑙𝑙ℎ = 8 𝑚𝑚𝑚𝑚) with straight vortex tube varying the cold mass fraction.

FIGURE 19 A comparison of cooling, heating, and energy separation efficiencies of


optimized 180° CVT (𝑁𝑁 = 3, 𝑙𝑙ℎ = 8 𝑚𝑚𝑚𝑚) with straight vortex tube varying the cold mass
fraction.

27
In Fig. 20 the total temperature contours of the optimized CVT (lh=8 mm, θ =180o, N=3) are
compared with the SVT (lh=0 mm, θ =0, N=2) at 𝜉𝜉 = 0.3 whereas in Fig. 21 the variation of
the static temperature is shown with velocity streamlines superimposed on them for the
optimized and straight vortex tubes. The maximum temperature separation is found in the
optimized CVT (lh=8 mm, θ =180o, N=3) at 𝜉𝜉 = 0.85 which is 37.29 percent higher than the
SVT.

FIGURE 20 A comparison of total temperature contours of optimized 180° CVT with


nozzle count of 𝑁𝑁 = 3 and truncation length 𝑙𝑙ℎ = 8 𝑚𝑚𝑚𝑚 to the straight vortex tube at cold
mass fraction of 𝜉𝜉 = 0.3.

28
FIGURE 21 A comparison of velocity streamline of optimized 180° CVT with nozzle
count of 𝑁𝑁 = 3 and truncation length 𝑙𝑙ℎ = 8 𝑚𝑚𝑚𝑚 to the straight votex tube at cold mass
fraction of 𝜉𝜉 = 0.3.

The total temperature contours and the velocity streamlines superimposed on static temperature
contours of the optimized tube are compared with the SVT (lh=0 mm, θ =0, N=2) in Figs. 22
and 23 respectively.

29
FIGURE 22 A comparison of total temperature contours of optimized 180° CVT with
nozzle count of 𝑁𝑁 = 3 and truncation length 𝑙𝑙ℎ = 8 𝑚𝑚𝑚𝑚 to the straight vortex tube at cold
mass fraction of 𝜉𝜉 = 0.85.

30
FIGURE 23 A comparison of velocity streamline of optimized 180° CVT with nozzle
count of 𝑁𝑁 = 3 and truncation length 𝑙𝑙ℎ = 8 𝑚𝑚𝑚𝑚 to the straight votex tube at cold mass
fraction of 𝜉𝜉 = 0.85.

9. CONCLUSIONS

In this paper, the effects of truncating the conical flow control valve, variation of tube curvature
and varying the inlet swirl by increasing the number of inlet nozzles on the thermal
performance of the vortex tube have been studied through numerical simulation. All have been
found helpful in one way or the other to achieve better thermal performance from the curved
vortex tubes over straight tubes. The combined effect of all these parameters resulted in the
lowering the cold stream temperature and increasing the hot stream temperature, thereby
creating larger temperature separation for the optimized curved vortex tube. The standard 𝑘𝑘 −
𝜀𝜀 model provided accurate computational results for the straight and 110° curved vortex tubes.
The energy separation is studied with the help of vortex generation and secondary circulation.
The computational results of this study are validated with the published numerical and
experimental results of straight and curved vortex tubes. The major findings of this study may
be summarized as follows:

31
1. The maximum value of the cold stream temperature difference is mostly found to occur
at a cold mass fraction of 0.3 whereas the maximum temperature separation in majority
of simulations take place at a cold mass fraction value of 0.7 or higher.

2. If the vortex tube is to be used as a cooling device, then its performance can be improved
by truncating the hot conical valve. The optimum truncated length 𝑙𝑙ℎ is found to be 8
mm. The maximum Δ𝑇𝑇𝑐𝑐 at this length is 16.54 K which is 21.62% more than the
nontruncated conical valve.

3. When the vortex tube is to be used as a power generation device, then maximum Δ𝑇𝑇ℎ𝑐𝑐
is equal to 30.02 K, obtained for a truncated length (𝑙𝑙ℎ ) of 8 mm at a cold mass fraction
of 0.85. It provided an improvement of approximately 26.19 percent in temperature
separation over non-truncated conical valve.

4. The curvature angles of 180° and 270° produced similar results for curved vortex tubes
for which the maximum Δ𝑇𝑇𝑐𝑐 obtained is 15.7 K, about 5.3 percent higher than the
straight vortex tube.

5. The temperature separation is found to be higher for curved vortex tubes (180o& 270o)
when compared with straight tubes. However, for curvature angles less than 150o, the
temperature separation in curved vortex tubes is found to be less than in straight tubes.

6. The number of inlet nozzles is also an important parameter that should be carefully
selected for designing the curved vortex tube. The results show that maximum Δ𝑇𝑇𝑐𝑐
value of 15 K is obtained with three inlet nozzles at a cold mass fraction of 0.3. The use
of three inlet nozzles is found to provide an improvement of approximately 10.3 percent
in the value of Δ𝑇𝑇𝑐𝑐 over curved vortex tube with two nozzles.

7. Also, the maximum value of Δ𝑇𝑇ℎ𝑐𝑐 = 31.5 K is found to occur with three inlet nozzles
at a cold mass fraction of 0.85. The temperature separation decreased when the number
of nozzles is either increased or decreased.

8. The maximum Δ𝑇𝑇𝑐𝑐 for the optimized vortex tube (𝜃𝜃 = 180°, 𝑁𝑁 = 3, 𝑙𝑙ℎ = 8 𝑚𝑚𝑚𝑚) is
found to be 18.38 K which is 23.43 percent higher than the straight vortex tube.

32
9. The maximum Δ𝑇𝑇ℎ𝑐𝑐 for the optimized vortex tube (𝜃𝜃 = 180°, 𝑁𝑁 = 3, 𝑙𝑙ℎ = 8 𝑚𝑚𝑚𝑚) is
found to be 34.35 K which was 37.29 percent higher than the straight vortex tube at a
𝜉𝜉 value of 0.85.

10. The hot and cold stream isentropic efficiencies are found to be better for curved vortex
tubes than straight tubes. However, the energy separation efficiency of the straight tube
is higher than its curved counterparts.

REFERENCES

[1] Arsalan, M., J.Ahmad, T., & S. Saeed, A., "Energy Harvesting for Downhole
Applications in Openhole Multilaterals," Society of Petroleum Engineers, 2018,
November 12.

[2] Arsalan, M., Ahmad, T. J., & Noui Mehidi, M. N., "Review of Downhole Wireless
Communication Techniques.," Society of Petroleum Engineers, 2014, November 10.

[3] Huo, A.Q.; He, Y.Y.; Wang, Y.L.; Tang, N.; Cheng, W.B., "Study of Downward
Communication Receiver in Rotary Steerable Drilling Tool Based on Turbine
Generator," in In Proceedings of the 29th Chinese Control Conference, Beijing, China,
29–31 July 2010.

[4] Jun Zheng, Bin Dou, Zilong Li, Tianyu Wu, Hong Tian and Guodong Cui, "Design and
Analysis of a While-Drilling Energy-Harvesting Device Based on Piezoelectric Effect,"
Energies, no. 1266, p. 14, 2021.

[5] Muhammad Arsalan, Talha Jamal Ahmad, and Abubaker S. Saeed, "Thermal Energy
Harvesting from Oil and Gas Flow," in Offshore Technology Conference , Asia, OTC-
30456-MS, 2020.

[6] Ahmad, T. J., Arsalan, M., Black, M. J., & Noui-Mehidi, M. N., "Piezoelectric Based
Flow Power Harvesting for Downhole Environment," Society of Petroleum Engineers,
2015, September 15.

[7] Sharma, T. K., Rao, A. P., Murthy, M. K., "Numerical Analysis of a Vortex Tube: A
Review," Arch Computat Methods Eng, vol. 24, no. 2017, pp. 251-280, 2016.

[8] G. J. Ranque, "Experiences sur la détente giratoire avec production simultanes d’un
echappement d’air chaud et d’un echappement d’air froid” (Experiments on vorticial
expansion with the simultaneous production of an outflow of hot air and an outflow of
cold air)," Journal de Physique et le Radium, pp. 112S-114S, 1933.

[9] R. Hilsch, "The use of the expansion of gases in a centrifugal field as cooling process,"
Review of Scientific Instruments., vol. 18, no. 2, p. 108–113, 1947.

33
[10] C. D. Fulton, "Ranque’s tube," Journal of the Americal Society of Refrigeration
Engineers, vol. 58, pp. 473-479, 1950.

[11] J. C. Maxwell, Theory of Heat, London, UK: Longman, 1891, pp. 338-339.

[12] Polihronov, J.; al., et, "The maximum coefficient of performance (COP) of vortex
tubes," Canadian Journal of Physics, vol. 93, no. 11, p. 1279–1282, 2015.

[13] T. Dutta, K. P. Sinhamahapatra, S. S. Bandyopadhyay, "CFD Analysis of Energy


Separation in Ranque-Hilsch Vortex Tube at Cryogenic Temperature," Journal of
Fluids, vol. 2013, no. Article ID 562027, p. 14 pages, 2013.

[14] Kassner R., Knoernschild E., "Friction Laws and Energy Transfer in Circular Flow,"
Headquarters Air Material Command Wright Patterson Air Force Base, Dayton, OH,
1948.

[15] S. GW, "THe vortex tube - internal flow data and heat transfer theory," J ASRE, pp.
985-990, 1951.

[16] Deissler R.G, Perlmutter M., "Analysis of the flow and energy separation in a turbulent
vortex," Int J Heat Mass Trans., vol. 1, pp. 173-191, 1960.

[17] A. J. Reynolds, "Energy flow in a vortex tube," Z. Angew. Math. Phys., vol. 12, pp.
343-356, 1961.

[18] H. Bruun, "Experimental investigation of the energy separation in vortex tube," Journal
of Mechanical Engineering Science, vol. 11, no. 6, pp. 567-582, 1969.

[19] O. Hinze, Turbulence, McGraw-Hill Series in Mechanical Engineering, New York:


McGraw-Hill, 1975.

[20] C. M. Gao, K. J. Bosschaart, J. C. H. Zeegers, A. T. A. M. de Waele, "Experimental


study on a simple Ranque-Hilsch vortex tube," Cryogenics, vol. 45, no. 2005, pp. 173-
183, 2005.

[21] Ahlborn B, Groves S., "Secondary flow in a vortex tube," Fluid Dynamics Research,
vol. 21, no. 2, pp. 73-86, 1997.

[22] Sachin U. Nimbalkar, Michael R. Muller, "An experimental investigation of the optimu
geometry for the cold end orifice of a vortex tube," Applied Thermal Engineering, vol.
29, no. 2009, pp. 509-514, 2009.

[23] M. H. Saidi, M. S. Valipour, "Experimental modeling of vortex tube refrigerator,"


Applied Thermal Engineering, vol. 23, pp. 1971-1980, 2003.

[24] Sadeghiseraji, J., Moradicheghamahi, J. & Sedaghatkish, A., "Investigation of a vortex


tube using three different RANS-based turbulence models," Journal of Thermal
Analysis and Calorimetry, 2020.

34
[25] Seyed Ehsan Rafiee, M.M. Sadeghiazad, "Three-dimensional and experimental
investigation on the effect of cone length of throttle valve on thermal performance of a
vortex tube using ke turbulence model," Applied Thermal Engineering, vol. 66, pp. 65-
74, 2014.

[26] M. Yilmaz, M. Kaya, S. Karagoz, S. Erdogan, "A review on design criteria for vortex
tubes," Heat Mass Transfer, vol. 45, pp. 613-632, 2009.

[27] Mohammad Sadegh Valipour, Nima Niazi, "Experimental modeling of a curved


Ranque-Hilsch vortex tube refrigerator," International Journal of Refrigeration, vol.
34, no. 4, pp. 1109-1116, 2011.

[28] Masoud Bovand, Mohammad Sadegh Valipour, Kevser Dincer, Ali Tamayol,
"Numerical analysis of the curvature effects on Ranque–Hilsch vortex tube
refrigerators," Applied Thermal Engineering, vol. 65, no. 1-2, pp. 176-183, 2014.

[29] Masoud Bovand, Mohammad Sadegh Valipour, Smith Eiamsa-ard, Ali Tamayol,
"Numerical analysis for curved vortex tube optimization," International
Communications in Heat and Mass Transfer, vol. 50, pp. 98-107, 2014.

[30] Celik IB, Ghia U, Roache PJ, "Procedure for Estimation and Reporting of Uncertainty
Due to Discretization in CFD Applications," J Fluids Eng., vol. 130, no. 7, p. 078001 (4
pages), 2008.

[31] H.M. Skye, G.F. Nellis, S.A. Klein, "Comparision of CFD analysis to empirical data in
a commercial vortex tube," International Journal of Refrigeration, vol. 29, no. 1, pp.
71-80, 2006.

[32] Pourmahmoud N., Rahimi M., Rafiee S. E., Hassanzadeh A., "A Numerical Simulation
of The Effect Of Inlet Gas Temperature On The Energy Separation In A Vortex Tube,"
Journal of Engineering Science and Technology, vol. 9, no. 1, pp. 81-96, 2014.

35
NOMENCLATURE

𝐿𝐿: length of vortex tube [𝑚𝑚𝑚𝑚]

lh: truncation length of the cone [mm]


𝐷𝐷: inside diameter of vortex tube [𝑚𝑚𝑚𝑚]
𝑅𝑅: ideal gas constant [𝐽𝐽 𝐾𝐾 −1 𝑚𝑚𝑚𝑚𝑙𝑙−1 ]
𝑚𝑚̇: mass flow rate [𝑘𝑘𝑘𝑘 𝑠𝑠 −1 ]
𝑃𝑃: pressure [𝑃𝑃𝑃𝑃]
𝑇𝑇: temperature [𝐾𝐾]
𝑣𝑣: velocity [𝑚𝑚 𝑠𝑠 −1 ]
𝑐𝑐𝑝𝑝 : specific heat at constant pressure [𝐽𝐽 𝑘𝑘𝑔𝑔−1 𝐾𝐾 −1 ]
𝑘𝑘: turbulent kinetic energy [𝑚𝑚2 𝑠𝑠 −2 ]

GREEK SYMBOLS

𝜌𝜌: density [𝑘𝑘𝑘𝑘 𝑚𝑚−3 ]


Δ𝑇𝑇: temperature difference [𝐾𝐾]
𝜇𝜇: dynamic viscosity [𝑘𝑘𝑘𝑘 𝑚𝑚−1 𝑠𝑠 −1 ]
𝜈𝜈: kinematic viscosity [𝑚𝑚2 𝑠𝑠 −1 ]
𝜀𝜀: turbulent dissipation rate [𝑚𝑚2 𝑠𝑠 −3 ]
𝛿𝛿𝑖𝑖𝑖𝑖 : Kronecker delta
𝜂𝜂: efficiency
𝜉𝜉: Cold mass fraction
𝛾𝛾: specific heat ratio

SUBSCRIPTS

𝑐𝑐: cold
ℎ: hot
𝑖𝑖: inlet
𝑛𝑛: nozzle

hc: temperature/energy separation

𝑎𝑎: atmospheric

36
ABBREVIATIONS

CFD: Computational Fluid Dynamics

RANS: Reynolds Averaged Navier Stokes

SVT: Straight vortex tube

CVT: Curved vortex tube

37

You might also like