1 s2.0 S2214157X21008674 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Case Studies in Thermal Engineering 29 (2022) 101704

Contents lists available at ScienceDirect

Case Studies in Thermal Engineering


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

Heat transfer due to turbulent annular impinging jet with a bullet


extension at the end of the inner blockage rod
Farhana Afroz, Muhammad A.R. Sharif *
The University of Alabama, Aerospace Engineering and Mechanics Department Tuscaloosa, AL, 35487-0280, USA

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

Keywords: The flow and thermal behavior of a turbulent annular jet impinging on a flat plane are numeri­
Jet impingement cally investigated to examine the effect of extending the inner rod of the annulus beyond the exit
Impinging jet heat transfer plane of the jet. The Fluent commercial code is used for conducting the computations employing
Annular jet the k-ε (Realizable) turbulence model. Three different shapes of the extended annular rod end,
Bullet extension of the blockage rod viz., cylindrical, conical, and semi-ellipsoidal, are considered. The length of the extended part is
Realizable k-ε model
taken as 0.5, 1, and 2 times the outer jet diameter. The extension of the annular rod end modifies
the flow structure and eliminates the recirculation region at the exit of the annular jet, which
noticeably affects the impingement plane heat transfer. Depending on the shape and length of the
bullet, the bullet extensions affect the heat transfer differently. Compared to the case without the
extension, the conical extension diminishes the heat transfer, the ellipsoidal extension moderately
increases the heat transfer, while the cylindrical extension enhances the heat transfer for a larger
bullet length (≥1). Also, the location and the magnitude of the peak Nusselt number at the hot
impingement surface vary as the shape and length of the extension change.

1. Introduction
Impinging jets achieve efficient high-intensity localized heat transfer from the impingement plane. There are numerous industrial
applications of the impinging jet such as heating, cooling, drying, de-icing, etc. The most popular jet nozzle shapes used are the circular
(round) jets or the rectangular (slot) jets. Abundant studies have been performed over the years on the jet impingement heat transfer
with the circular and slot jet formations for various flow and geometric configurations and applications. Another jet formation, viz., the
annular jet, is also used in many applications. The annular formation consists of a circular coaxial rod inserted inside a circular pipe.
The annular space within the outer pipe wall and the rod insert is the flow passage. The heat transfer due to a circular impinging jet can
be enhanced using an annular jet [1–3]. Better uniformity of the heat transfer over the target hot surface is also achieved with the
annular jet as opposed to the localized heat transfer with a round jet [4]. This feature is desirable in some jet impingement heat transfer
applications.
The flow structure of the free annular jet is very complex. A graphical sketch of the free annular jet configuration, following Ko and
Chan [5,6] is shown in Fig. 1a. According to them, there exist three different flow regions downstream of the exit plane of the jet. The
zone between the jet exit and the tip of the potential core is termed the initial merging zone. In this zone, an internal recirculation vortex
develops between the potential core and the central axis near the end of the blockage rod. Two different mixing regions; the inner and
outer mixing regions, can be identified in the initial merging zone. Two trains of counter-rotating vortices, separated by the annular jet

* Corresponding author.
E-mail address: msharif@eng.ua.edu (M.A.R. Sharif).

https://doi.org/10.1016/j.csite.2021.101704
Received 11 September 2021; Received in revised form 23 November 2021; Accepted 8 December 2021
Available online 9 December 2021
2214-157X/Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
F. Afroz and M.A.R. Sharif Case Studies in Thermal Engineering 29 (2022) 101704

potential core, are formed. One vortex train forms in the region within the jet potential core inner edge and the geometric axis. The
other vortex train lies in the outer mixing region, between the jet outer edge and the outer edge of the potential core. Within the
potential core the mean velocity is equal to the jet exit velocity. The initial merging zone ends after a certain distance at the downstream
position where the potential core vanishes. The inner and outer mixing region merges and reattaches at a point on the central axis. The
intermediate merging zone is denoted by the zone between the reattachment point and the jet potential core tip (a distance of about 4
times the outer jet diameter). The downstream zone from the intermediate merging zone is called the fully merged zone. The flow
features in this zone are akin to that of a single round jet flow.
For the basic annular free jet, where the inner blockage rod end is made level with the exit plane of the jet, a standing vortex
occupies the region immediately downstream of the exit of the jet (sudden expansion corner) due to the Coanda effect. An array of
toroidal vortices is generated due to the standing vortex, which is swept downstream axially as reported by Ko and Chan [6]. In
addition to the case of the simple free annular jet, Ko and Chan [6] also investigated the case with the bulleted end of the inner rod
extending beyond the jet exit into the jet flow. They modified the end of the blockage rod by adding conical and ellipsoidal bullets
when the bullet length almost covered the range of the initial merging zone (Fig. 1b). Thus, the inner mixing region is mostly covered
by the boundary layer of a uniform flow (of the annular potential core) over a surface with a favorable pressure gradient. The gradual
change of the bullet profile (as opposed to the sudden change for a basic annular jet with no bullet) makes the vortices to be shed from
near the bullet end. In the downstream intermediate merging zone, the absence of the uniform flow potential core implicates that the
wake is no longer immersed in a uniform stream. Thus, the mass and energy transfer rates across the wake are modified due to the
spreading of the surrounding jet stream, causing the wake and the associated vortices to degenerate rapidly [6]. Thus, in the case of
bullet extension of the inner rod for an annular free jet, although the internal recirculating region and the associated vortices are
cleared, wake vortices form in the boundary layer on the bullet surface, which are separated and blown downstream (vortex shedding).
These wake vortices affect the flow downstream up to a distance of about 4Do and within a radial region of about 0.1Do around the
domain axis. The frequency of this vortex shedding depends on the shape and length of the bullet [6].
For an annular jet, impinging perpendicularly on a plane surface, the jet flow is drastically affected by the presence of the
impingement surface when the impinging flow is turned around with subsequent development of a wall jet along the impingement
plane. This phenomenon adjusts the heat transfer (enrichment or reduction compared to that for a jet without the bullet extension)

Fig. 1. Illustration of the various flow regions and features for an annular free jet, (a) without bullet and (b) with bullet extension at the inner rod end; following Ko
and Chan [6].

2
F. Afroz and M.A.R. Sharif Case Studies in Thermal Engineering 29 (2022) 101704

from the impingement plane accordingly. A recent work by the present authors identified three different flow patterns for the annular
turbulent impinging jet, viz., “flow pattern-1, flow pattern-2, and flow pattern-3, based on the non-dimensional jet-to-target separation
distance, H” [12]. Flow pattern-1 of the annular turbulent imping jet is comparable to flow structures of a single round impinging jet
and it occurs when H is large (>3.5). The flow pattern-2 for intermediate separation distance (1.3 < H < 3.5) has a doughnut-shaped
recirculation vortex “generated at the annular nozzle exit and a reverse stagnation flow occurs near the domain axis and the
impingement surface” [12]. For a small separation distance (H < 1.3), the flow pattern-3 has a flow that deviates further to the outer
side of the annulus. A reverse stagnation flow at the impingement plane is also developed near the axis region. Depending on the
impingement plane placement in any of the three zones (initial, intermediate, or fully merging), the heat transfer features are affected
accordingly because of the development of the corresponding flow patterns mentioned above. As explained earlier, the addition of a
bullet extension at the end of the annular rod insert also modifies the flow and heat transfer characteristics for the annular impinging
jet.
Numerous analytical, experimental, and numerical studies on heat transfer from a plane due to annular impinging jets (without the
bullet extension of the inner blockage rod) have been performed over the last few decades. Make and Yabe [7] performed an

Fig. 2. Schematic diagram of the problem configuration and the bullet shapes.

3
F. Afroz and M.A.R. Sharif Case Studies in Thermal Engineering 29 (2022) 101704

experimental study on heat transfer from an annular impinging jet and reported the existence of reversed stagnation flow. Ichimiya [8]
experimentally investigated heat transfer from a plane surface, confined by a parallel surface, due to an annular impinging jet. The
distribution of the local temperature on the surface was measured. Chattopadhyay [9] numerically investigated turbulent heat transfer
from a plane surface due to an annular impinging jet and found that the heat transfer was 20% less compared to that for the circular jet.
Musika et al. [1] studied annular impinging jet heat transfer from a plane surface, both experimentally and numerically. He found heat
transfer enhancement for low jet exit-to-target distance of separation compared to the fully circular impinging jet. For a higher
jet-to-target distance of separation, the heat transfer is comparable to the round jet case. An experimental investigation of the heat
transfer and flow process due to an annular impinging jet was conducted by Terekhov et al. [3] using the PIV method. They observed
enhancement of heat transfer when compared to that for a full round jet. Pal et al. [10,11] conducted a numerical investigation on heat
transfer from a plane surface due to an annular impinging jet. Compared to the full round jet, the heat transfer was found to be
diminished in their case.
From the literature review above, it is noted that, even though the heat transfer for the simple annular impinging jet is investigated
by many researchers, the case where the jet structure is modified by adding a bullet at the end of the inner blockage rod insert has not
been investigated yet. Hence, it is presumed to be worthwhile to explore the effect of the bulleted extension of the blockage rod, on the
jet impingement heat transfer. With this motivation in mind, the heat transfer process caused by a turbulent annular impinging jet with
and without a bulleted end of the rod insert is numerically investigated in this study. Three types of bullet shapes, viz., cylindrical,
conical, and ellipsoidal, are considered where the base diameter of the bullet is equal to the diameter of the inner blockage rod (Di). The
different bullet shapes are considered to realize which shape is more effective for heat transfer improvement in a specific configuration.
The objective is to explore if the bullet extension would enhance the heat transfer and thereby achieve a passive control method for the
heat transfer augmentation under an annular impinging jet. The originality of the present article is the study of the heat transfer
process for the annular impinging jet with a bullet at the end of the inner rod insert.

2. Problem schematic and numerical considerations


2.1. Depiction of the problem
The schematic depiction of the axisymmetric problem geometry and model setup is shown in Fig. 2 on an axial-radial plane. The
inner and outer diameters of the annular jet are represented as Di and Do, respectively. The jet axis is aligned along the x-direction while
the y-axis represents the radial direction. The domain extends as 10Do × 10Do in the axial and radial directions. The bullet shapes
considered here are also sketched at the bottom of Fig. 2. The domain boundary at the top is the outlet while the boundary at the
bottom is the annular jet symmetry axis. The right boundary represents the impingement plane which is kept at a hot constant
temperature Th = 315 K. The whole of the left boundary, excluding the jet inlet and the inner blockage rod part, is the entrainment
boundary on which a constant ambient pressure condition is imposed. At the inlet of the jet, a uniform axial velocity VA and uniform
temperature Tc = 300 K is specified. The ambient temperature is also specified as 300 K. The ratio of the inner and the outer diameter,
Di/Do, is designated as the annular jet blockage ratio (BR). The reference length of the problem is the hydraulic diameter Dh = Do - Di.
The Reynolds number for the problem, Re, is expressed as a function of the average jet velocity, VA, and the hydraulic diameter Dh. The
jet to target distance of separation, h, is made non-dimensional as H = h/Do .

2.2. Simulated cases


The specific Reynolds numbers used in the present study are taken as 5,000 and 25,000, and the value of the blockage ratio, BR, is
given as 0.6; while a constant separation distance H = 2 is employed. This value of H puts the impingement surface in the intermediate
merging zone (1.3 < H < 3.5), which has many practical applications. The value of H > 3.5 puts the impingement surface in the fully
merged zone when the flow behavior is similar to that for a full round jet; whereas H < 1.3 is in the initial merging zone which is too
small to accommodate a bullet and sufficient free space for the jet development beyond the bullet end. The non-dimensional total
length L (= l/DO, l being the dimensional length of the bullet) of the bullet regardless of the shape is varied as 0.5, 1, and 1.5. Table 1
lists a total of 20 cases considered for the simulation and computation of the corresponding flow and thermal fields.

2.3. Development of the numerical model


The numerical model used here is the same as in earlier published work [12] by the present authors. The governing equations, the
details of the numerical method, code validation, mesh refinement study, and turbulence model evaluation is extracted from Ref. [12]
and are briefly described here.

2.3.1. The conservation equations


The equations governing the flow and thermal field with constant fluid properties are expressed in a cylindrical coordinate system
as;

Table 1
Simulated cases; blockage ratio BR = 0.6, separation distance H = 2.

Reynolds number No bullet Cylindrical Conical Ellipsoidal

5,000 L=0 L = 0.5, 1.0, 1.5 L = 0.5, 1.0, 1.5 L = 0.5, 1.0, 1.5
25,000 L=0 L = 0.5, 1.0, 1.5 L = 0.5, 1.0, 1.5 L = 0.5, 1.0, 1.5

4
F. Afroz and M.A.R. Sharif Case Studies in Thermal Engineering 29 (2022) 101704

Conservation of mass
∂u 1 ∂(yv)
+ =0 (1)
∂x y ∂y
Conservation of momentum in the axial and radial directions
( )
∂u2 1 ∂(vu) 1 ∂p ∂(τxx ) 1 ∂ τxy
+ =− + + (2)
∂x y ∂y ρ ∂x ∂x y ∂y
( ) ( )
∂(uv) 1 ∂(yv2 ) 1 ∂p ∂ τxy 1 ∂ yτyy
+ =− + + (3)
∂x y ∂y ρ ∂y ∂x y ∂y
Conservation of energy
[ ( )]
∂(uT) 1 ∂(yvT) 1 ∂(qx ) 1 ∂ yqy
+ = + (4)
∂x y ∂y ρcp ∂x y ∂y

where u and v symbolize the components of the mean velocity in the axial (x) and radial (y) direction, p indicates the mean pressure,
and T represents the mean temperature. In Eqs. (1)–(4), the constant fluid density is represented by ρ, while cp indicates constant
pressure-specific heat. In Eqs. (2) and (3), the stress terms may be written using the indicial notation as
( )
∂ui ∂uj 2
τij = (ν + νt ) + − kδij
∂xj ∂xi 3

where the fluid kinematic viscosity is indicated as ν while νt is the turbulent kinematic viscosity (computed by a chosen turbulence
model), the kinetic energy of turbulence is indicated as k, and δij represents the Kronecker delta. In Eq. (4), the heat fluxes due to
turbulence is expressed as
( )
νt ∂T
qi = − α + . (5)
Prt ∂xi

where the fluid thermal diffusivity is represented as α = λ/ρcp , the fluid thermal conductivity is represented by λ while the turbulent
Prandtl number is represented by Prt .
The distribution of the local Nusselt number on the hot impingement plane is obtained from the expression, Nu = λ(Tq̇D h
h − Tc )
where the
heat flux at the hot plane surface is represented as q̇. For a circular impingement area, the average Nusselt number is obtained as:
∫∫ ∫R ∫2π ∫R
1 1 2
Nu = Nu dA = 2 Nu (ydθdy) = 2 (Nu)(y)dy (6)
A πR R
0 0 0

Fig. 3. Sample mesh.

5
F. Afroz and M.A.R. Sharif Case Studies in Thermal Engineering 29 (2022) 101704

2.3.2. Numerical method


The Ansys Fluent commercial code is used to conduct the present computations. A quadrilateral structured mesh was generated by
the Ansys Meshing code. Fig. 3 presents a sample mesh. The mesh was densely assembled towards the impingement plane in the axial
direction and clustered towards the outer jet wall in the radial direction. Between the outer wall of the jet pipe and the axis of the jet,
the mesh was uniformly spaced in the radial direction. A system of linear algebraic equations, obtained by integrating the governing
equations over the finite volume mesh cells, was sequentially and iteratively solved. The coupling of pressure-velocity was achieved
through the SIMPLE algorithm. The second-order upstream scheme was emplyed to discretize the convection terms while the central
differencing was employed for the diffusion terms. The iteration process was continued until the normalized residuals for each gov­
erning equation dropped less than 10− 6. Additionally, the total surface heat flux on the impingement plane was monitored till a steady
value was reached with iteration.

2.3.3. Model validation and mesh refinement study


A few models for the turbulence simulation, available in the Fluent code, were evaluated by comparing the predicted Nusselt
number distribution against the data (experimental) of Musika et al. [1]. The selected models for the performance evaluation are the
Realizable and RNG k-ε model, the Reynolds stress model, and the Transition SST k-ω model, which are often used by various re­
searchers and in industries for the computation of the jet impingement heat transfer. The Fluent manual and elsewhere contains a
comprehensive description of these turbulence models. The predicted distribution of the local Nusselt number on a radial line on the
impingement plane is compared against experimental data [1] in Fig. 4 for the case with Reynolds number (Re) of 20,000, blockage
ratio (BR) of 0.44, and the separation distance (H) of 2 and 6.
The experimental data is available only for the range r/Do ≤ 3, r being the radial distance from the axis. It is noted from Fig. 4 that
the agreement of the predicted and the experimental data is not very satisfactory at any radial locations for all of the evaluated
turbulence models. For r/Do < 1, the Transition SST model and the RNG k-ε model under-predicted the local Nusselt number while the
other models over-predict the Nusselt number. It is noticed that the Realizable k-ε model produces, on the whole, overall good
agreement against the experimental data. To reinforce this conclusion, the experimental investigation for annular jet impingement
conducted by Yang et al. [13] was also computed using the Realizable k-ε model for the case with Re of 7,000, BR of 0.5, and H of 0.5
and 4.1. The distribution of the experimental and predicted radial pressure coefficient, CP, are compared and presented in Fig. 5, where
a very good agreement is observed. Based on these observations, the Realizable k-ε model is chosen for all calculations in this study.
A systematic grid refinement analysis was also performed with the Realizable the k-ε model to determine the optimum mesh
distribution for the grid-independent solutions. Initially, a coarse mesh of 110 × 108 distribution is considered which was successively
refined. The gradual convergence of the distribution of the predicted local Nusselt number with mesh refinement, as plotted in Fig. 6a,
is monitored. It is perceived that beyond the 180 × 127 mesh, the difference in the successive distribution of the local Nusselt number is
insignificant. Thus, the 180 × 127 mesh distribution is employed for all computations here. It is also recommended that, for the
Realizable k-ε model, the y + values from the wall to the first inside node should be around 1. The plot of the y + distribution for the
various mesh distributions is presented in Fig. 6b, from which it is noted that for the 180 × 127 mesh distribution, the y + values mostly

Fig. 4. Distribution of the local Nusselt number on the hot surface.

6
F. Afroz and M.A.R. Sharif Case Studies in Thermal Engineering 29 (2022) 101704

Fig. 5. Comparison of the distribution of the coefficient of pressure (Cp) on the hot surface against the experimental data of Yang et al. [13].

Fig. 6. Evolution of (a) the distribution of the local Nusselt number, and (b) the distribution of the wall y+ value on the impingement plane with the refinement of
the mesh.

fall below 2, thereby the requirement of very fine mesh near the wall is satisfied.

3. Results and discussion


Representative streamline plots in the flow field (axial-radial plane) for the turbulent annular impinging jet (at Re = 5,000, BR =
0.6, L = 0.5, and H = 2) is presented in Fig. 7 as a sample case. For the standard case without the bullet, the existence of a vortex bubble
(toroidal vortex in full 3D) is visible between the jet axis and the principal jet-stream directly near the exit of the jet. The inner blockage
rod of the basic annular jet generates a sudden expansion corner at the exit of the jet, which is responsible for this recirculation region.
For the cylindrical bullet, a larger and elongated recirculation bubble is formed within the jet mainstream and the axis, extending from
the bullet end to the target surface, which generates a reverse flow region (with negative radial velocity) on the target surface close to
the axis. It also pushes the main jet-stream away from the axis, radially into the domain, and moves the impingement location of the jet
radially out. For the conical and ellipsoidal bullet ends, there is no such development of the vortex bubble. A very weak recirculation
bubble around the junction of the target plane and the jet axis (not visible in the figures), is present for the no-bullet, conical, and
ellipsoidal bullet cases, where a very weak reversed flow region exists. The conical and ellipsoidal bullet does not modify the flow
structure very much other than the elimination of the recirculation region downstream of the bullet end.
The heat transfer effects on the target plane due to the flow modifications by the bullet addition are manifested in the distribution of
the local Nusselt number on the surface, which are presented in Fig. 8. Starting with a small value at the jet axis (r = 0), the Nusselt
number increases to a maximum at the impingement location and then monotonically decreases in the radial direction. The low Nusselt

7
F. Afroz and M.A.R. Sharif Case Studies in Thermal Engineering 29 (2022) 101704

Fig. 7. Streamline plots on the radial-axial plane for with different shapped bullets at Re = 5,000, BR = 0.6, H = 2, L = 0.5.

Fig. 8. Comparison of the distribution of the local Nusselt number along the impingement plane with bullet length, L = 0.5, BR = 0.6, and H = 2.

number value at the axis is attributed to the presence of the reverse radial flow in the corner between the axis and the impingement
plane resulting in a thick boundary layer for the temperature. All bullets reduce the Nusselt number at the axis and increase the peak
Nusselt number value compared to the no-bullet case. The large reverse flow (with stronger recirculation bubble) on the target plane
with the cylindrical bullet makes the Nusselt number at the axis minimum and moves the principal jet-stream impingement location
(radial position of the maximum -Nusselt number) radially more outward. The ellipsoidal bullet shows more enhancement of the heat
transfer compared to the other bullets.
The distribution of the local Nusselt number on the impingement plane for all cases considered (Re = 5,000 and 25,000; L = 0.5, 1,
and 1.5; and cylindrical, conical, and ellipsoidal bullets) are shown in Fig. 9 for a quick comparison of the effect of bullet length on the
thermal transport (Nusselt number). It is seen that, at the lower Re (5,000), the bullet length has an insignificant effect on the heat
transfer for the cylindrical and conical bullets, while the length of the ellipsoidal bullet has a somewhat significant effect on the
distribution of the local Nusselt number. In this case (ellipsoidal bullet), the Nusselt number for the higher bullet length (L = 1.5) is
markedly lower than that for the shorter bullet lengths (L = 0.5 and 1) up to r/Do of about 0.8. At the higher Reynolds number (25,000)
the effect of bullet length on the distribution of the local Nusselt is more significant in the sense that the Nusselt number variation at a
fixed radial position (r/Do) is much higher and much more distinguishable among the different bullet lengths, compared to those for Re
= 5,000, especially around the impingement location.
The minor secondary peak in the distribution of the local Nusselt number for some combinations of the flow parameters at Re =
25,000 in Fig. 9 is a relic of the laminar to turbulence transition. The laminar flow in the jet potential core switches to turbulent flow at
the interface between the potential core and the ambient fluid. If the impingement plane is positioned within the potential core closer
to the jet exit, the secondary peak in the local Nusselt number distribution is produced which is more prominent at a higher Reynolds
number. When the impingement plane is positioned further downstream out of the jet potential core, there is no such transition, and no

8
F. Afroz and M.A.R. Sharif Case Studies in Thermal Engineering 29 (2022) 101704

Fig. 9. Distribution of the local Nusselt number on the hot impingement plane for different shaped bullets with different lengths (L); BR = 0.6 and H = 2.

secondary peak is formed. The potential core has a length of about 4Do. In the present case with a low jet-to-target distance of 2Do (H =
2), the impingement plane is positioned within the jet potential core. Hence the secondary peak of the distribution of the local Nusselt
number is produced.
The overall heat transfer situation is characterized by the average Nusselt number on the impingement plane (the area under
distribution curve for the local Nusselt number, divided by the area of the surface). The average of the Nusselt number (NuAVG) is
plotted in Fig. 10a for all cases considered. It is seen that the average Nusselt number, in general, has an increasing trend as the bullet
length increase, which, however, is not noteworthy for the conical and ellipsoidal bullets at lower Reynolds numbers. For the cy­
lindrical bullet, though, the NuAVG increasing rate with increasing bullet length is notable at both lower and higher Re. To assess
whether the bullet addition enhances the heat transfer or diminishes it, the average surface Nusselt number for the bulleted case is
divided by that for the no-bullet case (NuA). The ratio NuAVG/NuA is plotted in Fig. 10b against varying bullet length L. Data points
above the black solid line indicates heat transfer enhancement while those below indicate reduction. It is observed that the ellipsoidal
bullet consistently enhances the heat transfer for all bullet lengths, albeit to a lesser extent at the lower Re. The conical bullet
consistently diminishes the heat transfer for all bullet lengths, whereas the cylindrical bullet enhances the heat transfer for longer
bullet lengths (L ≥ 1). Without the bullet extension at the end of the inner rod insert, a recirculation bubble emerges at the corner of the
inner rod and the axis immediately downstream of the jet exit due to the sudden expansion effect. With a bullet extension, the
recirculation bubble minimizes, eventually disappearing with increasing bullet length. The absence of the recirculation bubble affects
the heat transfer differently (enhancement or reduction compared to the no-bullet case) depending on the bullet length, the bullet
profile, and the Reynolds number.
Another measure of the heat transport characteristic is the peak Nusselt number, NuP. The peak values of the Nusselt number and
their radial location on the impingement plane, rNuP /Do , for all cases are extracted and charted in Fig. 11a and b. For the higher Re
(25,000), the peak Nu values are significantly higher for the case with the bullet than the case without the bullet, whereas at the lower
Re (5,000), the peak Nu values are slightly higher with the cylindrical bullet and about the same with the conical and ellipsoidal bullets
compared to the case without the bullet. The peak Nusselt number position indicates the jet impingement point on the impingement
plane. The peak Nu location shifts more and more towards the axis (radially inwards) with increasing length for the cylindrical bullet.
This is ascribed to the fact that the recirculation bubble developed for the cylindrical bullet (see Fig. 7) gets thinner and smaller as the
bubble length increases since the distance from the bullet end to the impingement plane is reduced leaving a smaller region for bubble
development. On the other hand, the peak Nu radial location is almost invariant with increasing bullet length for the conical bullet,
whereas the peak Nu location slightly shifts outward radially as the ellipsoidal bullet length increases. Due to the absence of the
recirculation bubble at the corner near the jet exit for the conical and ellipsoidal bullets, the main jet stream flow is not modified
noticeably and the impingement location does not change perceptibly with increasing bullet length.
The differences in the results for the different shaped bullets are demonstrated in the plots in Figs. 7, 10 and 11. Fig. 7 shows the
difference in the flow streamlines which clearly displays the development of the recirculation bubble along the jet axis. For the case
without the bullet, there is a strong recirculation bubble generated immediately downstream of the jet annular exit due to the sudden
expansion (Coanda) effect. With the addition of a conical or ellipsoid bullet extension at the end of the inner blockage rod, the

9
F. Afroz and M.A.R. Sharif Case Studies in Thermal Engineering 29 (2022) 101704

Fig. 10. (a) Average Nusselt number (NuAVG) on the impingement plane versus bullet length (L); (b) ratio of the Nusselt number average with a bullet to that without
bullet (NuAVG/NuA); BR = 0.6 and H = 2.0.

expansion is more gradual and the recirculation bubble at the jet exit is eliminated. In the case of a cylindrical bullet, the sudden
expansion corner at the end of the bullet still exits, forming a larger and elongated recirculation zone covering the entire region from
the end of the bullet to the impingement surface. The presence or absence of this recirculation bubble significantly affects the heat
transfer characteristics of this problem depending on the shape and length of the bullet.
The difference among the heat transfer characteristics for the non-bulleted and bulleted cases are also exhibited in the local Nusselt
number plots in Fig. 8 and the average Nusselt number plots in Fig. 10. The peak value of the local Nusselt number and the location of
the peak value on the impingement surface is different among all cases as shown in Fig. 11. The peak Nusselt number value for the
bulleted cases is always larger than the non-bulleted cases. On the other hand, the local Nuaaselt number values, very close to the
domain axis, are always lesser than that for the non-bulleted case. This is due to the formation of a weak miniature clockwise recir­
culation at the corner between the axis and the impingement surface (not captured in the streamline plots in Fig. 7). Thus, there is a
flow reversal (inward) in that region causing thickening of the thermal boundary layer and a corresponding reduction of the local
Nusselt number values (temperature gradient). These differences in the local Nusselt number distribution are also manifested in the
differences of the average Nusselt number at the impingement surface among the non-bulleted case and the various bulleted cases as
shown in Fig. 10.
The results presented here are at best qualitatively accurate, while the quantitative accuracy is uncertain. The predicted modest
enhancement of the heat transfer (1%–2%) is within the uncertainty level (~5%–10%) of the turbulence model. The quantitative
accuracy may be achieved using higher-order turbulence models such as LES or DNS, which is beyond the scope of the present study.

4. Conclusion
The flow and heat transfer characteristics of an annular impinging jet with three different shaped bullet addition at the end of the

10
F. Afroz and M.A.R. Sharif Case Studies in Thermal Engineering 29 (2022) 101704

Fig. 11. (a) Peak Nusselt number (NuP) on the impingement plane versus bullet length (L); (b) Peak Nusselt number location on the impingement plane (rNup/Do)
versus bullet length (L); BR = 0.6 and H = 2.

inner rod insert have been analyzed numerically. This configuration of the annular jet geometry with the bullet extension of the inner
rod inserts and the corresponding jet impingement heat transfer has not been investigated yet. The bullet end of the annular jet exit
alters the flow structures and eliminates the recirculation region downstream of the annular jet exit. The principal intention of this
investigation was to determine the effect of the shape and length of the bullet extension on the resulting heat transfer. It is found that
the addition of the bullet does not affect the heat transfer significantly. The ellipsoidal bullet slightly improves the average heat
transfer for any bullet length especially at a higher Reynolds number (1%–2% compared to the non-bulleted case). The cylindrical
bullet slightly enhances the heat transfer when the bullet length is greater than 1, whereas the conical end bullet diminishes the
average heat transfer regardless of the bullet length. The peak Nusselt number values slightly increase as the bullet length is increased.
The main goal of this study was to determine whether the bullet extension can improve the heat transfer. Thus, a passive mech­
anism can be developed for the heat transfer enhancement for an annular impinging jet. The results implicate a modest heat transfer
improvement, depending on the bullet length and shape. The reason for this enhancement can be attributed to the elimination of the
internal recirculating zones created immediately downstream of the jet exit without the bullet due to the sudden expansion (Coanda)
effect. When the bullet is inserted at the end of the inner blockage rod, the sudden expansion is reduced to a more gradual expansion
and wake vortices are formed in the boundary layer over the bullet surface. These vortices are detached or shed from the bullet surface
and swept downstream by the jet stream, eventually hitting the impingement surface thereby modifying the thermal boundary layer
thickness on the impingement surface. As a result, the local Nusselt number increases resulting in an overall increase of the average
Nusselt number.

Author statement
The authors of the manuscript hereby declare that they do not have any financial or personal relationships that may be perceived as
influencing their work.

Declaration of competing interest


The authors do hereby declare that there are no conflicts of interest with any individual or organization regarding the material
presented in the submitted manuscript.

11
F. Afroz and M.A.R. Sharif Case Studies in Thermal Engineering 29 (2022) 101704

Nomenclature

BR blockage ratio, Di/Do


Dh hydraulic diameter, Do − Di
Di inner pipe diameter
Do outer pipe diameter
h jet-to-target dimensional separation distance
H jet-to-target non-dimensional separation distance, h/Do
NuAVG average Nusselt number on the impingement plane
NuA NuAVG for the no-bullet case
l dimensional bullet length
L non-dimensional bullet length, l/Do
NuP peak Nusselt number
r radial distance along the impingement plane
rNuP radial distance of the peak Nusselt number
Re Reynolds number of the annular jet
Tc cold inlet temperature
Th impingement plate temperature
VA annular jet exit velocity

References
[1] W. Musika, M. Wae-Hayee, P. Vessakosol, B. Niyomwas, C. Nuntadusit, Investigation of Flow and Heat Trans. Character. Annular Impinging Jet 931 (2014)
1223–1227.
[2] S. Kalinina, V. Terekhov, K. Sharov, Special features of flow in an annular jet impinging on a barrier, Fluid Dynam. 50 (5) (2015) 665–671.
[3] V. Terekhov, S. Kalinina, K. Sharov, An experimental investigation of flow structure and heat transfer in an impinging annular jet, Int. Commun. Heat Mass Tran.
79 (2016) 89–97.
[4] H. Zhen, C. Leung, C. Cheung, Heat transfer characteristics of an impinging premixed annular flame jet, Appl. Therm. Eng. 36 (2012) 386–392.
[5] N. Ko, W. Chan, Similarity in the initial region of annular jets: three configurations, J. Fluid Mech. 84 (4) (1978) 641–656.
[6] N. Ko, W. Chan, The inner regions of annular jets, J. Fluid Mech. 93 (3) (1979) 549–584.
[7] H. Maki, A. Yabe, Heat transfer by the annular impinging jet, Exp. Heat Trans. Int. J. 2 (1) (1989) 1–12.
[8] K. Ichimiya, Heat transfer characteristics of an annular turbulent impinging jet with a confined wall measured by thermosensitive liquid crystal, Heat Mass Tran.
39 (7) (2003) 545–551.
[9] H. Chattopadhyay, Impinging heat transfer due to a turbulent annular jet, Int. J. Transport Phenom. 9 (4) (2007) 287–296.
[10] T.K. Pal, H. Chattopadhyay, D.K. Mandal, S. Bhattacharyya, Numerical investigation of heat transfer under impinging annular jets, MATTER: Int. J. Sci. Technol.
2 (1) (2015) 70–77.
[11] T.K. Pal, H. Chattopadhyay, D.K. Mandal, Flow and heat transfer due to impinging annular jet, Int. J. Fluid Mech. Res. 46 (3) (2019).
[12] F. Afroz, M.A. Sharif, Numerical study of turbulent annular impinging jet flow and heat transfer from a flat surface, Appl. Therm. Eng. 138 (2018) 154–172.
[13] H. Yang, T. Kim, T. Lu, K. Ichimiya, Flow structure, wall pressure and heat transfer characteristics of impinging annular jet with/without steady swirling, Int. J.
Heat Mass Tran. 53 (19) (2010) 4092–4100.

12

You might also like