You are on page 1of 19

Numerical Heat Transfer, Part A: Applications

An International Journal of Computation and Methodology

ISSN: (Print) (Online) Journal homepage: www.tandfonline.com/journals/unht20

Numerical analysis of concentric jet impingement


heat transfer to the concave surface of the cone

Anilkumar Vajubhai Gorasiya

To cite this article: Anilkumar Vajubhai Gorasiya (28 Dec 2023): Numerical analysis of
concentric jet impingement heat transfer to the concave surface of the cone, Numerical Heat
Transfer, Part A: Applications, DOI: 10.1080/10407782.2023.2296575

To link to this article: https://doi.org/10.1080/10407782.2023.2296575

Published online: 28 Dec 2023.

Submit your article to this journal

Article views: 34

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=unht20
NUMERICAL HEAT TRANSFER, PART A: APPLICATIONS
https://doi.org/10.1080/10407782.2023.2296575

Numerical analysis of concentric jet impingement heat transfer


to the concave surface of the cone
Anilkumar Vajubhai Gorasiyaa,b
a
Department of Mechanical Engineering, Indian Institute of Technology Bombay, Mumbai, Maharashtra, India;
b
Department of Mechanical Engineering, Faculty of Technology and Engineering, The Maharaja Sayajirao
University of Baroda, Vadodara, Gujarat, India

ABSTRACT ARTICLE HISTORY


This work includes numerical investigation of concentric jet impingement Received 21 September 2023
heat transfer to the concave surface of 30 apex angle conical surface. Revised 13 December 2023
Heat transfer results are represented in terms of nondimensional heat Accepted 13 December 2023
transfer coefficient Nusselt number along the slant edge of the cone.
KEYWORDS
Turbulence intensity, static pressure distribution along slant edge of the Concentric jet
cone, and velocity contours in the domain are also presented for a detail impingement; conical
understanding of the flow behavior. Different geometrical parameters like surface; Nusselt number
diameter of pipe (10 mm, 14 mm, and 20 mm) and L/D (4.25, 6.25, and
10.25) are considered for heat transfer analysis. Jet Reynolds number is var-
ied between 25000 and 82000 for each diameter and L/D case. It is
observed that Nusselt number value is minimum close to the apex of the
cone due to stagnation region, then increases with increase in the distance
from the apex of the cone and becomes maximum due to maxima in the
turbulence intensity and decreases with increases in the distance. Nusselt
number value decreases with increase in the L/D. Nusselt number increases
with increase in the jet Reynolds number. There is an insignificant influ-
ence of the diameter of pipe on Nusselt number for identical jet Reynolds
number and L/D.

1. Introduction
Jet impingement heat transfer is a fast-cooling technique utilized for highly localized heat flux
locations in electronic chip, metal billet, turbine blade, rocket nozzle, and anti-icing of aircraft at
higher altitude. Present work is a possible application of heating inside the surface of the aircraft
engine nose to prevent ice formation on its external surface when aircraft runs at higher altitude.
Several people have reported results on jet impingement cooling of the different surfaces experi-
mentally and numerically.
Casanova and Granados-Ortiz [1] analyzed impingement heat transfer characteristics to the
flat plate by making dimples, bumps, and both on the surface numerically. Heat transfer at the
stagnation region is increased with dimpled geometry. Flat plate with bump enhances heat trans-
fer over a whole region. Combination of dimple and bump is not giving better heat transfer
results compared to the flat plate without any modifications. Nabadavis and Mishra [2] numeric-
ally analyzed jet impingement heat transfer characteristics for a flat plate. They have used jet
Reynolds number, L/D, and inclination of twin jet as parameters. It was found that Nusselt num-
ber increases with increase in the jet Reynolds number and there was optimal value for L/D. Heat
transfer distribution over the surface improves with outward inclination of the jet and reaches

CONTACT Anilkumar Vajubhai Gorasiya avgorasiya@gmail.com


ß 2023 Taylor & Francis Group, LLC
2 A. V. GORASIYA

Nomenclature
D diameter of pipe (mm) TI turbulence intensity
h heat transfer coefficient (W/m2 k) X distance along the streamwise direction.
k thermal conductivity of fluid (W/mK) Ø apex angle of surface ( )
L distance between jet exit and the tar- l dynamic viscosity of fluid (Ns/m2)
get (mm) q density of fluid (kg/m3)
Nu Nusselt number
O offset of jet axis from the cone axis (in Subscripts
terms of diameter of pipe) Air air
q heat flux (W/m2) j jet
Re jet Reynolds number w wall
T temperature ( C) x along stream wise direction

maximum, after that heat transfer decreases with increase in the angle. Kannan and Senthilkumar
[3] numerically studied jet impingement heat transfer to plate with and without grooved surface.
They observed that K-x Shear Stress Transport (SST) turbulence model predicts accurate
impingement heat transfer compared to other models. There is an adverse influence of the
grooves on the heat transfer. Pakhomov and Terekhov [4] studied fluid flow and heat transfer
characteristics of intermittent turbulent impinging circular jet. Different parameters were used for
analysis like jet Reynolds number, pulse frequency, ratio of on-time to total cycle time, and L/D.
The Nusselt number increases with increase in the pulse frequency, jet Reynolds number, and L/
D for L/D  4. Sagot et al. [5] investigated experimentally and numerically jet impingement heat
transfer to a flat plate at constant temperature. Jet Reynolds number and L/D were taken as
parameters. They have identified that K-x SST turbulence model predicts well with experimental
results. Nusselt number increases with jet Reynolds number and decreases along the plate from
staganation point due to the development of the boundary layer.
Lu et al. [6] carried out numerical work for impingement heat transfer with temperature of
coolant and wall as the parameter. They have used K-x SST turbulence model for predicting heat
transfer. They have observed that Nusselt number decreases with increase in the temperature
ratio (Wall temperature to coolant temperature). Lam et al. [7] studied jet impingement heat
transfer on flat plate numerically with array of jets. Jet Reynolds number, velocity ratio and non-
dimensional channel height were taken as parameters. They observed that Nusselt number
increases with increase in jet Reynolds number and velocity ratio and with decrease in the chan-
nel height-to-length ratio. Shah [8] investigated jet impingement heat transfer with multiple jets
and moving plates numerically. The average Nusselt number is unaffected by the plate movement
for velocity ratio up to 0.75 then increases drastically with increase in the velocity ratio for single
jet. Whereas average Nusselt number increases gradually with velocity ratio for multiple jets.
Afroz and Sharif [9] performed the numerical analysis of turbulent twin oblique jet impingement
on isothermally heated plate. The jet Reynolds number, L/D, jet-to-jet separation, and inclination
of jet were taken as parameters. They tested various turbulence model for predicting Reynolds
stress and observed that K-x SST turbulence model predicts accurately with respect to experi-
mental results. They observed that average Nusselt number decreases with decrease in the
impingement angle and increases with jet Reynolds number. Wen et al. [10] did numerical work
on jet impingement heat transfer on flat plate with multiple jets. They took different shape of the
geometry for the nozzle like: circular, semicircular, rectangle, cross, square, isosceles triangle, and
flower for heat transfer analysis. They used K-x SST turbulence model for predicting impinge-
ment heat transfer. They observed that cross section of the geometry plays a significant role in
the jet momentum, velocity profile from the geometry, and accordingly interaction of the neigh-
boring jets. Pawar and Patel [11] conducted a numerical analysis of slot jet impingement heat
transfer on a flat plate. The jet Reynolds number, impingement angle, and plate velocity were
NUMERICAL HEAT TRANSFER, PART A: APPLICATIONS 3

taken as parameters. They used K-x SST turbulence model for modeling. The average Nusslet
number increases with increase in the jet Reynolds number and decreases with decrease in the
impingement angle. Influence of plate velocity is significant at lower impingement angle. Behera
and Rathore [12] carried out numerical work of impingement heat transfer over a moving plate
with fixed jet Reynolds number and L/D. They observed significance influence of plate motion on
local and average Nusselt numbers.
Modak et al. [13] studied the influence of nano-fluids (aluminum oxide 0.15% and 0.6% vol-
ume ratio) and aqueous high-alcohol surfactant (100–400 ppm) on jet impingement heat transfer.
There is enhancement in heat transfer by 140%, 207%, and 117% with 0.15% aluminum oxide,
0.6% aluminum oxide, and 150 ppm aqueous high-alcohol surfactant, respectively. Chang et al.
[14] studied the effect of nanoparticle (Sio2) on jet impingement heat transfer numerically. They
observed that there was 67.9% enhancement in heat transfer with 3% nanoparticle inclusion.
Rehman et al. [15] carried out numerical work for slot jet impingement on a curved plate with
jet Reynolds number and L/D as parameters. They observed that with increase in the jet
Reynolds number and L/D Nusselt number increases. Hemmami and Hammami [16] performed
numerical work for impingement cooling of rotating milling tool. Heat transfer performance was
analyzed based on the jet Reynolds number with single and multiple jets. They observed that heat
transfer is enhanced with jet Reynolds number and with multiple jets compared to single jet.
Bhansali and Ekkad [17] investigated jet impingement heat transfer over rotating cylinder numer-
ically with different shape fins like triangular, cylindrical, square, and dome-shaped over the sur-
face. They observed that there was enhancement in heat transfer by 220% at low rotation speed
for triangular shape fin and 170% for cylindrical fin at high rotation speed. Satish and
Venkatasubbaiah [18] analyzed numerically turbulent multiple jet impingement heat transfer on a
heated square block in a channel. They used different parameters like jet Reynolds number,
height of the block, number of jets, distance between the center of the jets, and distance of the jet
from the block for heat transfer optimization. They observed that heat transfer enhances with
increase in the number of jets due to interaction between jets compared to single jet and found
that four number is the optimal for maximum heat transfer. With increase in the center distance
between jets heat transfer from the block is enhanced due to proximity of the jet to the corner of
the block.
Yang et al. [19] reported numerical results for local and average Nusselt number values for
slot jet impingement onto the semicircular concave surface with different L/S (where S is the slot
width) and curvature ratio as parameter. Maximum Nusselt number at 0.5 L/S due to increased
acceleration between nozzle exit and target surface and at L/S greater or equal to 4.0 there is
insignificant change in stagnation Nusselt number. Maeda and Ohmori [20] reported Nusselt
number data for V-shaped wedge with slot jet impinging on the internal side of the surface for
10000 jet Reynolds number. Angle of the wedge surface was changed from 30 to 180 . With
decrease in the wedge angle maxima shift away from the geometric impingement point and it is
approximately 75% of the stagnation Nusselt number for wedge angle equal to 180 . This maxima
shift further toward outlet with increase in the L/S. Cornaro et al. [21] reported experimental
Nusselt number values for jet impingement on the convex cylindrical test section. There is sec-
ondary peak in the Nusselt number for L/D equal to 1.0 and 2.0 for 16000 jet Reynolds number
due to transition of flow from laminar to turbulent region whereas it is absent for 6000 jet
Reynolds number. Apart from this, there is insignificant influence of L/D on Nusselt number
results for both jet Reynolds number values. Singh et al. [22] performed numerical work for
impinging jet over circular cylinder with jet Reynolds number and L/D as parameters. They
found that stagnation point Nusselt number increases with decrease in L/D. Rahman et al. [23]
investigated impingement heat transfer characteristics over hemispherical plate with jet Reynolds
number, L/D, and thickness of plate with different materials having different thermal conductivity
as parameters. They observed that there is more uniform temperature distribution with higher
4 A. V. GORASIYA

thermal conductivity material. Ying et al. [24] reported data of curvature effect on heat transfer
for jet impinging onto a concave surface with curvature ratio from 0.005 to 0.03 that was
obtained by changing either target surface diameter or nozzle diameter. An increase in the aver-
age Nusselt number with increase in the curvature ratio (change in diameter of target surface)
was reported which was attributed to increase in the mass flow rate with increase in the pipe
diameter and formation of vortices in the streamwise direction. However, with a change in curva-
ture due to change in the pipe diameter, average Nusselt number remains more or less constant.
Bu et al. [25] reported jet impingement heat transfer of concave surface with the piccolo tube
with jet Reynolds number, L/D and jet angle as parameters. They concluded that Nusselt number
at stagnation region is enhanced with increase in jet Reynolds number and the maximum value
of Nusselt number was observed at L/D around 4–5.75. The heat transfer at the stagnation zone
enhances with increase in the jet impingement angle. Joshi and Sahu [26] analyzed influence of
nozzle geometry by taking circular and elliptical cross section for flat and concave surface. They
observed that Nusselt number increases at the stagnation region with increase in the aspect ratio
(ratio of major axis length to minor axis length) of the geometry. Gorasiya and Vedula [27] per-
formed experimental work of the jet impingement heat transfer to the concave surface of the
cone. They conducted experiment for concentric and offset jet impingement for 30 and 70 con-
ical surface with different parameters like jet Reynolds number, L/D and diameter of the pipe for
analysis. From that, it was concluded that average Nusselt number increases with jet Reynolds
number and decreases with increases in the L/D. And there is insignificant influence of the diam-
eter of the pipe on local Nusselt number for identical experimental conditions. Hadipour and
Zargarabadi [28] analyzed jet impingement heat transfer numerically and experimentally on con-
cave surface at small distance from the target. They have varied jet Reynolds number, L/D, and
nozzle diameter as the parameter. They found that K-x SST turbulence model predicts closely
velocity profile and Nusselt number for impingement heat transfer.
From the literature, it is observed that there are several works available for impingement heat
transfer for flat plate and curved surface. However, numerical work for jet impingement to the
concave surface of the cone is not available. Our earlier work was experimental investigation of
jet impingement heat transfer on the concave surface of the cone. It is very difficult to get vel-
ocity contours, turbulence intensity, and pressure variation along the slant edge of the cone
through experimental work. And to very several L/D and diameter of the tube is not feasible
through experimental work. Hence present work is carried out to get all these details and analysis
based on these parameters. Experimental data of Nusselt number are taken from the article of
Gorasiya and Vedula [27] for comparison of the numerical results.

2. Validation of numerical solver


Since there is no numerical work available for jet impingement to the concave surface of cone,
initial numerical simulations for a jet impinging normally on a flat plate were carried out and
compared with the data reported in the literature.

2.1. Problem statement mesh detail


Two-dimensional axisymmetric simulation was carried out due to the symmetrical nature of the
problem. Figure 1 shows the geometrical and boundary condition detail for the simulation carried
out in the ANSYS software. At the inlet of pipe velocity inlet condition was given, for the plate
(heater wall) uniform heat flux (1,400 W/m2) condition was given, for the wall of the pipe insu-
lated and no-slip wall and for other side pressure outlet condition was given. Length of the pipe
was taken as 60D to make the flow fully developed at the exit of the pipe. L is the distance
between nozzle exit and the plate. X is the distance along the plate from the center of the jet. Air
NUMERICAL HEAT TRANSFER, PART A: APPLICATIONS 5

Figure 1. Geometry and boundary condition detail of the computational domain.

Figure 2. Mesh of domain.

was taken as the working fluid and flow regime was taken as turbulent. SIMPLE algorithm was
used for solving pressure velocity coupled equation.
Figure 2 shows some portion of the domain with mesh, mesh was kept fine near the stagna-
tion region and where the jet exited from the pipe and interacted with the ambient fluid.
Minimum length of element was selected based on Y-plus requirement of K-x SST model. Size
6 A. V. GORASIYA

of the elements increases gradually from the wall and from the interaction region and ratio of
successive element length is 1.015. This increment was in the direction of normal to the surface
and normal to the axis of the jet and was given on either side of the pipe wall. There are 66000
quadrilateral cells in the domain and these cells are orthogonal hence there is no skewness of the
element. Mass, momentum, and energy equations were solved and convergence criteria for mass,
momentum, and energy equation were kept 10−6 (with lower residual more accurate solution can
be obtained).

2.2. Governing equations and methodology of data reduction


The averaged governing equations of continuity (1), momentum (2), energy (3), and for turbulent
flow separate equations for k (turbulent kinetic energy) (7) and x (specific dissipation rate) (8) at
steady and incompressible case were solved in the ANSYS software by Finite Volume Method
(FVM) method. These equations are as follows:
Continuity equation:
r  ðViÞ ¼ 0 (1)
Momentum equation:
" !#
@qVj Vi @Vi @Vj 2 @Vi @ðqVi0 Vj0 Þ
¼ −rp þ r  l þ − dij þ (2)
@xj @xj @xi 3 @xi @xj

Energy equation:
 
@Vi ðqe þ pÞ @ @T
¼ Keff (3)
@xi @xi @xi
where e is the total energy given by Eq. (4)
p V2
e¼h− þ (4)
q 2
where Keff is effective thermal conductivity in turbulent flow, summation of thermal conductivity
of fluid (K), and turbulence thermal conductivity which is due to heat transfer among the fluid
particles due to eddies present in the turbulent flow. Kt is calculated from the eddy viscosity and
turbulent Prandtl number as shown in Eq. (6).
Keff ¼ K þ Kt (5)
Cp lt
Kt ¼ (6)
Prt
Governing equations for k and x
 
@ðVi kÞ @ @k
¼ f þ Kk − wk (7)
@xi @xi k @xi
 
@ðVi xÞ @ @x
¼ f þ Kx − w x þ Ux (8)
@xi @xi x @xi
where f corresponds to effective diffusivity, K and w represent generation and dissipation of tur-
bulent quantity, and U is a cross-diffusion term.
Constant heat flux boundary condition was given to the target surface in the numerical simu-
lation and from the converged solution temperature data was used to reduce the Nusselt number
as shown by Eqs. (9) and (10).
NUMERICAL HEAT TRANSFER, PART A: APPLICATIONS 7

qw
hx ¼ (9)
Twx − Tj
hx D
Nux ¼ (10)
Kair
where Twx is the wall temperature obtained from numerical simulation, Tj is the jet temperature
at inlet, and D is the pipe diameter.

2.3. Comparison with literature results


Comparison of present local Nusselt number data for L/D ¼ 2.0 and 23000 jet Reynolds number was
made with Kannan and Senthilkumar [3] and Sagot et al. [5] as shown in Figure 3. Difference
between the present result and the Sagot result is 5% near the stagnation region and 12% away from
the stagnation where values are bit low. Near the stagnation region difference is high and away from
the stagnation region difference is around 10%–12% between the present result and Kannan and
Senthilkumar’s result. This might be due to distribution of grid elements through the domain of ana-
lysis. Hence present numerical solver can be used for further analysis. Figure 4 shows variation of tur-
bulence intensity and static pressure along the streamwise direction of the present numerical work.
Turbulence intensity increases and at around 1.4D there is maxima in the turbulence intensity and
then onwards decreases. Near the jet impingement, there is maximum pressure due to stagnation
region and then there is drastic reduction in the pressure which is due to increase in the velocity.
Maxima in the turbulence intensity corresponds to peak in the local Nusselt number.

3. Numerical work on internal cooling of conical surfaces


Computations for local Nusselt number along the slant edge of the cone with a jet impinging on
the inner surface are presented in this section. This section comprises of problem statement,
mesh detail, and grid-independent test.

3.1. Problem statement


The geometry was kept identical to that of experimental work of Gorasiya and Vedula’s [27] art-
icle. Truncated cone of SS-304 with copper ring brazed at the end was used for experimental
work. Copper ring at the ends was brazed to SS-truncated cone for electrical power supply. High-
velocity jet is impinged concentrically to the concave surface of the cone. Temperature of the jet
and surface were measured with the help of thermocouple and thermal camera. Nusselt number
was reduced from these temperature and heat flux measurement. Figure 5 shows the geometrical
and numerical details of the conical test section. Two-dimensional axisymmetric assumption was

Figure 3. Comparison of present work results of local Nusselt number for L/D ¼ 2.0 and Re ¼ 23000 with literature results.
8 A. V. GORASIYA

Figure 4. Turbulence intensity and static pressure along the slant edge of the cone for present work.

Figure 5. Numerical domain for conical test section.

Figure 6. Mesh of the domain.

assumed to be valid due to the symmetrical nature of the flow which resulted reduction in the
computation time. A uniform heat flux boundary condition was applied for the slant edge of the
cone, near the apex insulated wall condition is given. At pipe inlet velocity inlet condition was
given, for pipe wall insulation, and no-slip condition was given. Pressure outlet condition was
given for the flow outlet at bigger diameter of the cone. Reynolds number of the jet was varied
between 25000 and 82000. Distance between apex and jet exit “L” is changed to vary L/D.

3.2. Mesh of the domain


Figure 6 shows the mesh of the domain. Triangular domain was divided into multiple subdo-
mains to make the structured mesh and to reduce the skewness of elements due to which chances
of physically unrealistic results were reduced. Mesh near the wall was kept fine to capture the
NUMERICAL HEAT TRANSFER, PART A: APPLICATIONS 9

Figure 7. The local Nusselt number along the slant edge of the cone for Ø ¼ 30 , D ¼ 20 mm, O/D ¼ 0.0, L/D ¼ 4.25, Re ¼ 82000.

Figure 8. Velocity contours: (a) L/D ¼ 4.25, Re ¼ 29000; (b) L/D ¼ 10.25, Re ¼ 82000.

flow physics and away from the wall mesh size increased gradually (in the ratio of 1.05) toward
the center of the domain. There were 117,487 elements in the domain. Convergence criteria for
continuity, momentum, and energy equation were kept 10−6. Turbulence model and solver were
the same as those used for validation of flat plate simulation.

3.3. Grid-independent study


Three grids were used for grid independence test. Grid number one contains 93,424 elements,
grid number two contains 117,487 elements and grid number three contains 140,428 elements.
10 A. V. GORASIYA

Figure 7 shows the local Nusselt number results along the slant edge of the cone for all grids.
From that, it is observed that there is insignificant change in the result from grid two to three
hence grid two can be used for further analysis.

4. Results and discussion


In this section, numerical results of velocity contours, Turbulence intensity, static pressure, and
local Nusselt number are presented. The Nusselt number results are compared with the experi-
mental results of Gorasiya and Vedula [27].

4.1. Velocity contours


Figure 8a shows the velocity contours throughout the domain at L/D equal to 4.25 for 20 mm diam-
eter pipe. In the region which is very close to the apex, there is very little fluid motion. The incoming
fluid from the jet is unable to reach the apex of the cone and therefore fluid is nearly stagnant. The jet
entrains the surrounding fluid due to velocity gradient between incoming and outgoing fluid as it goes
toward the apex of the cone due to which jet area increases. After traveling some distance jet turns
back and moves toward the outlet of the test section. Fluid experiences acceleration in this region. At
some point downstream of the jet there occurs sever bending of the jet due to high pressure gradient
which promotes high heat transfer at the bending. From the bending point onwards the jet flows
toward the exit and expands in the cross-stream direction due to increase in the cross-stream area.
Figure 8b shows the velocity contours for 82000 jet Reynolds number and 20 mm diameter pipe at
L/D equal to 10.25. The momentum of the jet decreases as it moves toward the apex of the cone due
to more interaction with the outgoing fluid as L/D increases. In addition, the spread of the incoming
jet increases with increase in the L/D due to more interaction between incoming and outgoing fluids.
Hence length of the very slow-moving fluid domain increases near the apex of the cone. The bending
of the jet occurs away from the apex as compared to lower L/D case due to the low momentum of
the fluid. The remaining characteristics are identical to the lower L/D case.

4.2. Local Nusselt number results


Figure 9a shows the comparison of numerical results with experimental results of local Nusselt
number along the slant edge of the cone. Experimental result is taken from the Gorasiya and
Vedula [27]. The trend of the numerical results and experimental measurements can be observed
to be matching with each other. Maximum difference between experimental and numerical results
is 15% at X/D around 5 and at other places difference is within 5%.
Figure 9b shows the plot of static pressure along the slant edge of the cone for 82000 jet
Reynolds number. The static pressure is maximum near the apex of the cone due to stagnation
region which leads to conversion of kinetic energy into pressure energy. As the velocity of the
fluid is very low, the local Nusselt number is also small at these locations. Then onwards static
pressure decreases (local Nusselt number increases) and reaches minima where the kinetic energy
is maximum corresponds to maxima in turbulence intensity as seen from the subsequent plot.
This location has maxima in the Nusselt number due to transition of the flow from laminar to
turbulent. Pressure at some locations goes below the atmospheric pressure due to very high kin-
etic energy. Then static pressure increases slightly and goes slightly above atmosphere pressure.
The local Nusselt number decreases sharply due to rapid deceleration of the fluid in the region 3
< X/D < 7. Then onwards there is not much change in the static pressure with stream-wise dis-
tance along the slant edge of the cone and pressure remains close to the atmospheric pressure.
Figure 9c shows the plot of local Nusselt number with the turbulence intensity for 10 mm
diameter pipe at 29000 jet Reynolds number. Near the apex of the cone, there is low turbulence
NUMERICAL HEAT TRANSFER, PART A: APPLICATIONS 11

Figure 9. Local Nusselt number plot with turbulence intensity and static pressure along the slant edge of the cone. (a)
Comparison of numerical and experimental result for Ø ¼ 30 , D ¼ 10 mm, O/D ¼ 0.0, L/D ¼ 6.25, Re ¼ 63000. (b) Static pressure
and Nusselt number for Ø ¼ 30 , D ¼ 10 mm, O/D ¼ 0.0, L/D ¼ 4.25, Re ¼ 82000. (c) Turbulence intensity and Nusselt number
for Ø ¼ 30 , D ¼ 10 mm, O/D ¼ 0.0, L/D ¼ 4.25, Re ¼ 29000.
12 A. V. GORASIYA

Figure 10. Local Nusselt number plot with turbulence intensity and static pressure along the slant edge of the cone for 20 mm
diameter pipe. (a) Turbulence intensity and Nusselt number for Ø ¼ 30 , D ¼ 20 mm, O/D ¼ 0.0, L/D ¼ 4.25, Re ¼ 82000. (b)
Static pressure and Nusselt number for Ø ¼ 30 , D ¼ 20 mm, O/D ¼ 0.0, L/D ¼ 4.25, Re ¼ 82000. (c) Turbulence intensity and
Nusselt number for Ø ¼ 30 , D ¼ 20 mm, O/D ¼ 0.0, L/D ¼ 4.25, Re ¼ 29000. (d) Static pressure and Nusselt number for Ø ¼ 30 ,
D ¼ 20 mm, O/D ¼ 0.0, L/D ¼ 4.25, Re ¼ 29000.
NUMERICAL HEAT TRANSFER, PART A: APPLICATIONS 13

Figure 11. Local Nusselt number along the slant edge of the cone for three different diameter pipes at L/D is equal to 4.25. (a)
Ø ¼ 30 , O/D ¼ 0.0, L/D ¼ 4.25, Re ¼ 29000. (b) Ø ¼ 30 , O/D ¼ 0.0, L/D ¼ 4.25, Re ¼ 44000. (c) Ø ¼ 30 , O/D ¼ 0.0, L/D ¼ 4.25,
Re ¼ 63000. (d) Ø ¼ 30 , O/D ¼ 0.0, L/D ¼ 4.25, Re ¼ 82000.
14 A. V. GORASIYA

Figure 12. Local Nusselt number along the streamwise direction for three different diameter pipes at L/D is equal to 6.25. (a)
Ø ¼ 30 , O/D ¼ 0.0, L/D ¼ 6.25, Re ¼ 29000. (b) Ø ¼ 30 , O/D ¼ 0.0, L/D ¼ 6.25, Re ¼ 44000. (c) Ø ¼ 30 , O/D ¼ 0.0, L/D ¼ 6.25,
Re ¼ 63000. (d) Ø ¼ 30 , O/D ¼ 0.0, L/D ¼ 6.25, Re ¼ 82000.
NUMERICAL HEAT TRANSFER, PART A: APPLICATIONS 15

Figure 13. Local Nusselt number along the streamwise direction for three different diameter pipes at L/D is equal to 10.25. (a)
Ø ¼ 30 , O/D ¼ 0.0, L/D ¼ 10.25, Re ¼ 29000. (b) Ø ¼ 30 , O/D ¼ 0.0, L/D ¼ 10.25, Re ¼ 44000. (c) Ø ¼ 30 , O/D ¼ 0.0, L/
D ¼ 10.25, Re ¼ 63000. (d) Ø ¼ 30 , O/D ¼ 0.0, L/D ¼ 10.25, Re ¼ 82000.
16 A. V. GORASIYA

intensity due to a very small value of velocity that is attributed to lower heat transfer. Then
onwards turbulence intensity increases leading to enhancement in local Nusselt number. At some
downstream distance, there is maxima in the turbulence intensity due to shear interaction
between incoming and outgoing jet and acceleration of the flow from the stagnation region which
leads to the maxima in the local Nusselt number. Location of the maxima in turbulence intensity
and maxima in local Nusselt number is more or less identical. Then onwards local Nusselt num-
ber and turbulence intensity decreases sharply from the maxima up to X/D equal to 7. Then
onwards turbulence intensity decreases gradually and reaches close to zero value near X/D equal
to 10. For X/D > 10 magnitude of the turbulence intensity remains close to zero due to low vel-
ocity of the fluid and local Nusselt number value is also low in this region of the cone.
Figure 10a shows the turbulence intensity plot for 20 mm diameter pipe at 82000 jet Reynolds
number. Characteristics of the curve are more or less identical to the 10 mm diameter pipe but
there seems little bit shift of turbulence intensity maxima compared to Nusselt number maxima.
Static pressure plot has the identical characteristic to the low diameter pipe result as shown in
Figure 10b. There is decrease in the turbulence intensity and static pressure with decrease in the
jet Reynolds number as seen in Figures 10c, d, respectively.
Figure 11a shows the local Nusselt number plot along the slant edge of the cone for three dif-
ferent diameters of the pipe at L/D equal to 4.25 for 29000 jet Reynolds number. Near the apex
of the cone Nusselt number is minimum due to inability of the jet to reach the apex and low vel-
ocity of the fluid as discussed earlier. A peak in local Nusselt number is observed around the
region where the jet turns and interacts with the outgoing fluid and detailed reasoning was dis-
cussed in the earlier section. Numerical data show peak at X/D equal to 3.7 from the apex of the
cone. Then Nusselt number decreases with increases in the stream-wise distance due to deceler-
ation of the flow due to expansion of the jet in the cross-stream direction as area increases. The
local Nusselt number value is more or less equal for different diameter of the pipe. Hence there
is insignificant influence of the diameter of the pipe on Nusselt number for identical geometrical
and flow conditions. These observations are consistent with the experimental results.
Figures 11b–11d show the plots of the local Nusselt number for 44000, 63000, and 82000 jet
Reynolds numbers, respectively. Characteristics of these curves are identical to 29000 jet Reynolds
number curves. Here also diameter influence is insignificant on local Nusselt number for identical
geometrical and flow conditions. The local Nusselt number increases with increases in the jet
Reynolds number due to increase in the jet momentum.
Then influence of the change in the L/D has been studied for all three diameters of pipe and
at all jet Reynolds numbers. L/D has been varied from 4.25 to 6.25 and then 10.25 and plots of
local Nusselt number along the streamwise direction are shown in Figures 12 and 13. The general
trend of the curve and characteristics of data is similar to L/D equal to 4.25 case. The local
Nusselt number decreases with increase in the L/D. As L/D increases jet is positioned further
away from the apex hence jet has to travel more distance to reach the apex of the cone. During
this path, incoming jet interacts more with the outgoing jet hence its momentum reduces and jet
spreads in the cross-stream direction. The location of the maxima in the Nusselt number also
shifts away from the apex with increase in the L/D. Stagnant region is extended further toward
the exit of the cone with increase in the L/D. That leads to shift of turbulence intensity maxima
location toward the exit of the cone hence maxima in local Nusselt number is also shifted toward
the exit of the cone. These observations are consistent with the experimental results.

5. Conclusion
The numerical analysis was conducted for the internal cooling of a conical surface using concen-
tric jet impingement for 30 apex angle cone. Various parameters like, jet Reynolds number, L/D
NUMERICAL HEAT TRANSFER, PART A: APPLICATIONS 17

and diameter of pipe were analyzed with respect to local Nusselt number. The following conclu-
sions were drawn from the numerical work:
The characteristics of numerical and experimental results were matching with each other. The
difference between experimental and numerical results was 5%–15%. The jet was unable to reach
the apex of the cone and near apex there was a slowly moving fluid. The fluid as a jet moves
toward the apex and at some downstream location, it starts bending and flows toward the exit.
Then fluid decelerates due to increase in area and the heat transfer decreases. The maxima in
local Nusselt number were found downstream of the geometric impingement location due to
inability of fluid to penetrate up to the apex of cone. Maxima in the local turbulence intensity
was observed due to the interaction between the incoming and outgoing streams and this location
also records maxima in the local Nusselt number. Then local Nusselt number decreases at further
downstream locations due to spreading of the fluid into the progressively increasing area of the
cone. The local Nusselt number was observed to increase with the jet Reynolds number for all
cases. The local Nusselt number decreases at the impingement location with increase in the L/D
due to reduction in the momentum at the impingement location due to interaction between
incoming and outgoing jets. Influence of diameter on local Nusselt number for identical L/D and
jet Reynolds number is insignificant.

Disclosure statement
The author reports no conflicts of interest, and there is no financial disclosure for this work.

Funding
There is no funding agency involved in this work.

References
[1] J. Ortega-Casanova and F. J. Granados-Ortiz, “Numerical simulation of the heat transfer from a heated
plate with surface variations to an impinging jet,” Int. J. Heat Mass Transf., vol. 76, pp. 128–143, 2014.
DOI: 10.1016/j.ijheatmasstransfer.2014.04.022.
[2] A. Nabadavis and D. P. Mishra, “Numerical investigation of jet impingement heat transfer on a flat plate,”
ASI Carbon Sci. Technol., vol. 4, pp. 1–12, 2016.
[3] B. T. Kannan and S. Senthilkumar, “Steady state jet impingement heat transfer from axisymmetric plates
with and without grooves,” Procedia Eng., vol. 127, pp. 25–32, 2015. DOI: 10.1016/j.proeng.2015.11.320.
[4] M. A. Pakhomov and V. I. Terekhov, “Numerical study of fluid flow and heat transfer characteristics in an
intermittent turbulent impinging round jet,” Int. J. Therm. Sci., vol. 87, pp. 85–93, 2015. DOI: 10.1016/j.
ijthermalsci.2014.08.007.
[5] B. Sagot, G. Antonini, A. Christgen and F. Buron, “Jet impingement heat transfer on a flat plate at a con-
stant wall temperature,” Int. J. Numer. Methods Heat Fluid Flow, vol. 47, no. 12, pp. 1610–1619, 2008. DOI:
10.1016/j.ijthermalsci.2007.10.020.
[6] S. Lu, Q. Deng, P. M. Ligrani, H. Jiang and Q. Zhang, “Effects of coolant and wall temperature variations
on impingement jet array thermal performance,” Numer. Heat Transf. Part A Appl., vol. 79, no. 1, pp. 68–
82, 2021. DOI: 10.1080/10407782.2020.1814593.
[7] L. Kumar Prasanth Anand and K. A. Prakash, “A numerical investigation and design optimization of
impingement cooling system with an array of air jets,” Int. J. Heat Mass Transf., vol. 108, pp. 880–900,
2017. DOI: 10.1016/j.ijheatmasstransfer.2016.12.017.
[8] S. Shah, “Numerical analysis of heat transfer between multiple jets and flat moving surface,” Int. J. Heat
Mass Transf., vol. 171, pp. 121088, 2021. DOI: 10.1016/j.ijheatmasstransfer.2021.121088.
[9] F. Afroz and M. A. R. Sharif, “Numerical study of heat transfer from an isothermally heated flat surface
due to turbulent twin oblique confined slot-jet impingement,” Int. J. Therm. Sci., vol. 74, pp. 1–13, 2013.
DOI: 10.1016/j.ijthermalsci.2013.07.004.
18 A. V. GORASIYA

[10] Z.-X. Wen, Y.-L. He, X.-W. Cao and C. Yan, “Numerical study of impinging jets heat transfer with differ-
ent nozzle geometries and arrangements for a ground fast cooling simulation device,” Int. J. Heat Mass
Transf., vol. 95, pp. 321–335, 2016. DOI: 10.1016/j.ijheatmasstransfer.2015.12.022.
[11] S. Pawar and D. K. Patel, “Study of conjugate heat transfer from the impingement of an inclined free slot
jet onto the moving hot surface,” Int. Commun. Heat Mass Transf., vol. 111, pp. 104429, 2020. DOI: 10.
1016/j.icheatmasstransfer.2019.104429.
[12] V. M. Behera and S. K. Rathore, “Numerical investigation on conjugate behavior of heat transfer and its
augmentation in a turbulent wall-jet flow on a heated plate in motion,” Numer. Heat Transf. Part A Appl.,
vol. 1, pp. 1–19, 2023. DOI: 10.1080/10407782.2023.2200216.
[13] M. Modak, A. K. Sharma and S. K. Sahu, “An experimental investigation on heat transfer enhancement in
circular jet impingement on hot surfaces by using Al2O3/water nano-fluids and aqueous high-alcohol sur-
factant solution,” Exp. Heat Transf., vol. 31, no. 4, pp. 275–296, 2018. DOI: 10.1080/08916152.2017.
1381655.
[14] S. Chang, J. Lv, P. Wang and M. Bai, “Influence of particle deposition on heat transfer characteristics for
nanofluid free impinging jet,” Numer. Heat Transf. Part A Appl., vol. 84, no. 11, pp. 1297–1322, 2023. DOI:
10.1080/10407782.2023.2175085.
[15] M. M. Rahman, C. F. Hernandez and J. C. Lallave, “Free liquid jet impingement from a slot nozzle to a
curved plate,” Numer. Heat Transf. Part A Appl., vol. 57, no. 11, pp. 799–821, 2010. DOI: 10.1080/
10407781003800706.
[16] Z. Hemmami and A. Hammami, “Improving a rotating steel milling tool cooling by an air jet impingement
flow,” Numer. Heat Transf. Part A Appl., vol. 83, no. 5, pp. 467–477, 2023. DOI: 10.1080/10407782.2022.
2091896.
[17] P. S. Bhansali and S. V. Ekkad, “Numerical study of heat transfer enhancement over a rotating surface,”
Numer. Heat Transf. Part A Appl., vol. 1, pp. 1–20, 2023. DOI: 10.1080/10407782.2023.2209696.
[18] N. Satish and K. Venkatasubbaiah, “Numerical investigations of turbulent multiple jet impingement on a
heated square block in a confined channel,” Therm. Sci. Eng. Prog., vol. 14, pp. 100415, 2019. DOI: 10.
1016/j.tsep.2019.100415.
[19] Y. Yang, T. Wei and Y. Wang, “Numerical study of turbulent slot jet impingement cooling on a semi-circu-
lar concave surface,” Int. J. Heat Mass Transf., vol. 54, no. 1–3, pp. 482–489, 2011. DOI: 10.1016/j.ijheat-
masstransfer.2010.09.021.
[20] M. Maeda and K. Ohmori, “Heat transfer of jet impinging on V-shape inverse wedge,” J. Chem. Eng.
Japan, vol. 13, no. 3, pp. 194–198, 1980. DOI: 10.1252/jcej.13.194.
[21] C. Cornaro, A. S. Fleischer, M. Rounds and R. J. Goldstein, “Jet impingement cooling of a convex semi-
cylindrical surface,” Int. J. Therm. Sci., vol. 40, no. 10, pp. 890–898, 2001. DOI: 10.1016/S1290-
0729(01)01275-3.
[22] D. Singh, B. Premachandran and S. Kohli, “Numerical simulation of the jet impingement cooling of a cir-
cular cylinder,” Numer. Heat Transf. Part A Appl., vol. 64, no. 2, pp. 153–185, 2013. DOI: 10.1080/
10407782.2013.772869.
[23] M. M. Rahman, J. C. Lallave and C. F. Hernandez, “Convective heat transfer from a thick hemispherical
plate during free liquid jet impingement,” Numer. Heat Transf. Part A Appl., vol. 54, no. 6, pp. 581–602,
2008. DOI: 10.1080/10407780802289327.
[24] Y. Zhou, G. Lin, X. Bu, L. Bai and D. Wen, “Experimental study of curvature effects on jet impingement
heat transfer on concave surfaces,” Chinese J. Aeronaut., vol. 30, no. 2, pp. 586–594, 2017. DOI: 10.1016/j.
cja.2016.12.032.
[25] X. Bu, L. Peng, G. Lin, L. Bai and D. Wen, “Jet impingement heat transfer on a concave surface in a wing
leading edge: experimental study and correlation development,” Exp. Therm. Fluid Sci., vol. 78, pp. 199–
207, 2016. DOI: 10.1016/j.expthermflusci.2016.06.006.
[26] J. Joshi and S. K. Sahu, “Heat transfer characteristics of flat and concave surfaces by circular and elliptical
jet impingement,” Exp. Heat Transf., vol. 1, pp. 1–26, 2021. DOI: 10.1080/08916152.2021.1995082.
[27] A. V. Gorasiya and R. P. Vedula, “Heat transfer characteristics of jet impingement onto the concave surface
of a cone,” Exp. Heat Transf., vol. 35, pp. 1–25, 2022. DOI: 10.1080/08916152.2022.2126029.
[28] A. Hadipour and M. R. Zargarabadi, “Heat transfer and flow characteristics of impinging jet on a concave
surface at small nozzle to surface distances,” Appl. Therm. Eng., vol. 138, pp. 534–541, 2018. DOI: 10.1016/
j.applthermaleng.2018.04.086.

You might also like