You are on page 1of 14

International Journal of Thermal Sciences 151 (2020) 106257

Contents lists available at ScienceDirect

International Journal of Thermal Sciences


journal homepage: http://www.elsevier.com/locate/ijts

Numerical investigation of heat transfer from a plane surface due to


turbulent annular swirling jet impingement
Farhana Afroz *, Muhammad A.R. Sharif
Aerospace Engineering and Mechanics Department, The University of Alabama, Tuscaloosa, AL 35487-0280, United States

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

Keywords: The heat transfer characteristic of an annular turbulent swirling jet impingement on a heated plane surface is
Annular turbulent jet investigated numerically. The annular jet configuration causes instabilities and fluctuations in the flow.
Impinging jet Depending on the combinations of different parameters, the annular jet impingement may have positive or
Swirling jet
negative effects on the heat transfer from the impingement surface. Swirl, on the other hand, introduces vorticity
Reverse stagnation
and fluid mixing in an impinging jet, which is desirable in some applications. The axisymmetric two-dimensional
flow domain on a radial-axial plane is considered in the numerical model. Both non-swirling and swirling jet
impingement is studied. The numerical computations are performed using the ANSYS Fluent CFD code. The
Realizable k-ε turbulence model with enhanced wall treatment is used in the computation. The computational
process is validated against other published data on similar flow configuration for non-swirling annular
impinging jets. The flow and geometric parameters are the jet exit Reynolds number, Re (5000 to 25,000), the
swirl strength Sw (0–0.77), the jet exit to the impingement surface distance, H (0.5–4.0), and the moderate
blockage ratio of the annular jet, BR (0.4–0.6). The thermal-hydraulic field in the domain is computed for various
combinations of these parameters. The effects of these parameters on the Nusselt number distribution on the
impingement surface are analyzed. For short separation distance (H ¼ 0.5), the swirling motion positively affects
the overall heat transfer, and the average Nu is increased as high as 8% for certain combinations of Re, Sw, and
BR; compared to the non-swirling annular jet impingement. For higher separation distances, the average and
peak Nusselt number is initially reduced and then increased with increasing swirl strength.

Reynolds number or smaller the jet to target separation distance, the


1. Introduction larger is the heat transfer. Furthermore, the placement of the impinge­
ment surface within or outside of the jet potential core makes a signif­
Impinging jets are widely used in many industrial applications like icant difference in the heat transfer. Placement of the impingement
heating, cooling, and drying processes. The jets are usually issued from a surface within the jet potential core produces a secondary peak of the
circular nozzle or rectangular slots and impinged normally on the target Nusselt number distribution at a radial location of about twice the jet
surface. Numerous studies [1–5], on the performance of the slot and diameter from the jet axis. This secondary peak of the Nusselt number, in
round impinging jet heat transfer in various configurations, have been addition to the primary peak around the jet axis, occurs due to the
conducted over the years due to their importance in the practical in­ transition from the laminar flow within the potential core to the tur­
dustrial application. These studies have revealed the key flow and heat bulent flow at the edge of the potential core where the moving jet fluid
transfer features of the slot and round impinging jets. Slot or round interacts with the quiescent ambient fluid.
impinging jet can achieve a high localized heat transfer at the In order to enhance the heat transfer and distribute it more uniformly
impingement location (stagnation point) on the surface. The peak value over the impingement surface (which is desirable in some applications),
of the heat transfer occurs around the intersection of the jet axis with the various jet configurations, as well as active or passive flow control
impingement surface, which quickly diminishes in the outward radial methods, have been proposed and investigated in the past. Active con­
direction in the wall jet region. The jet impingement heat transfer trols include imposing jet pulsations [6] or adding acoustic excitations
principally depends on the jet exit Reynolds number and the jet exit to to alter the impinging jet flow characteristics [7]. The passive controls
the impingement surface separation distance. In general, the higher the are achieved by changing the inlet nozzle shape [8–10]. Passive control

* Corresponding author.
E-mail address: fafroz@crimson.ua.edu (F. Afroz).

https://doi.org/10.1016/j.ijthermalsci.2019.106257
Received 22 February 2018; Received in revised form 30 December 2019; Accepted 31 December 2019
Available online 9 January 2020
1290-0729/© 2020 Elsevier Masson SAS. All rights reserved.
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

Nomenclature QA annular jet discharge


r radial direction
A area of the impingement surface Re annular jet Reynolds number
BR blockage ratio Sw Swirl number
Cp pressure coefficient Tc cold jet temperature at the inlet
Dh hydraulic diameter, Do Di Th hot impinging surface temperature
Di inner jet diameter T.K.E Turbulent Kinetic Energy
Do outer jet diameter vz axial velocity
H jet-to-target separation distance vr radial velocity
Hc1 first critical distance vθ azimuthal velocity
Hc2 second critical distance VA average annular jet exit velocity
k fluid thermal conductivity ρ fluid density
Nu Nusselt number μ fluid viscosity
Pr Prandtl number

configurations also include the twin slot jet [11] or annular jet between an outer truncated cone (frustum) and an inverted inner
impingement [12]. The annular jet configuration can be obtained easily truncated cone. Chan and Ko [13] identified three different zones in a
by inserting a coaxial rod inside a circular tube. free annular jet through experimental investigations. The zone from the
The flow and geometric parameters for the annular jet impingement jet exit up to the tip of the annular potential core in the streamwise
flow are; (i) the blockage ratio, BR ¼ Di =Do where Di is the diameter of direction is designated as the initial merging zone. The flow region in the
the inner rod insert and Do is the outer diameter of the jet pipe , (ii) the cross-stream direction in this initial merging zone is characterized as; (i)
hydraulic diameter, defined as Dh ¼ ðDo Di Þ, (iii) the jet exit Reynolds the outer mixing region bounded by the jet outer edge and the outer
number, Re ¼ VA Dh =ν where VA is the average jet inlet velocity, and (iv) edge of the annular potential core and (ii) the inner mixing region be­
the non-dimensional jet exit to impingement surface separation dis­ tween the inner edge of the annular potential core and the jet axis. Two
tance, H ¼ h=Do , where h is the dimensional separation distance. counter-rotating vortex trains separated by the annular jet potential core
Many experimental and numerical studies have been conducted on occupy the region between the jet axis and the jet outer edge in the inner
the annular impinging jet flow and associated heat transfer process in and the outer mixing regions. Further downstream from the potential
the last several decades. Chan and Ko [13], and Ko and Chan [14,15] core tip, the counter-rotating vortex trains continue to exist and keep the
published articles characterizing the structure of free annular jets. At two mixing regions separate, but after some distance, the boundary
around the same time, Maki and Yabe [16] published the results of between the two mixing regions merges at the reattachment point
experimental research on annular impinging jet heat transfer where they forming one vortex train between the axis and the edge of the jet. The
identified three different flow patterns and flow regions. Experimental streamwise zone from the annular potential core tip to slightly down­
studies were conducted by Ichimiya [17] to investigate the heat transfer stream of the reattachment point is designated as the intermediate
characteristics of the turbulent annular impinging jet using a merging zone. The zone beyond the intermediate mixing zone is desig­
thermo-sensitive liquid crystal sheet. He observed that higher heat nated as the fully merged zone where the annular jet merges together and
transfer could be achieved for smaller H and larger BR. Chattopadhaya the flow pattern behaves like a single circular jet flow. In the case of
[18,19] conducted numerical studies of annular impinging jet heat annular jet impinging on a surface, the placement of the impingement
transfer from flat surfaces in both the laminar and turbulent jet flow surface plays a crucial role in the establishment of the flow pattern and
regime. He concluded that for the same mass and momentum efflux at the subsequent heat transfer process. Three distinct flow patterns,
the nozzle exit, the average Nusselt number of annular jets are about demarcated by sudden changes, are evolved when the impingement
20% lower compared to the circular jets. However, heat transfer was surface is placed normal to the jet axis in each of the zones described
enhanced in the near-axis region of the annular jet when the jet to target above.
separation distance was reduced. Through numerical and experimental The jet exit to target plate separation distances (H) at which the
investigations, Travnicek et al. [20] and Travnicek and Tesar [21] sudden changes in the jet flow pattern happens are designated as the
discovered the phenomena of bistability and hysteresis in annular critical distances. When the impingement surface is placed at a large
impinging jets. Zhen et al. [22] conducted an experimental investigation distance in the fully merged zone, the flow behaves like that of a circular
on pre-mixed annular jet flame impingement heat transfer from a flat impinging jet with a positive axial velocity up to the impingement sur­
surface. They observed higher and more uniform heat transfer at lower face and positive radial velocity in the wall jet region. As the separation
values of H in the turbulent flow regime, compared to the circular jet. distance is gradually reduced by moving the impingement surface closer
Musika et al. [23] experimentally and numerically investigated annular to the jet exit, the first sudden change in the flow pattern occurs at a
jet impingement heat transfer at Re ¼ 20,000 and BR ¼ 0.44 for a range critical distance designated as Hc1, when a flow reversal (negative axial
of H. They found that for same mass flow rate, the heat transfer for an and radial velocity) occurs around the jet axis near the impingement
annular impinging jet is higher compared to the round jet impingement. surface and the stagnation point shifts off of the impingement surface
Kalinina et al. [24] and Terekhov et al. [25] recently conducted exper­ into the flow on the jet axis. This phenomenon is described as the
imental measurement of the flow and thermal field for annular reversed stagnation flow. The sudden change in the flow pattern is
impinging jet and found heat transfer enhancement compared to the detected by the onset of the negative axial (as well as radial) velocity.
round jet case. When the separation distance is further decreased gradually, another
Even though the annular jet geometry and the associated sudden change occurs at a distance, designated as Hc2 when the vortex
manufacturing is simple, its flow physics is very complex. For round jet, between the jet axis and the annular jet potential core, in the inner
the potential core is a conical region extending from the jet exit up to mixing zone, comes in contact with the impingement surface. The sud­
several jet diameter (~4–5) along the axis with the cone base at the jet den change in the flow pattern is detected by the sudden change of the
exit and the tip on the jet axis. In the case of the annular jet, the potential pressure coefficient distribution on the impingement surface. The flow
core has the shape of an annular cone having the shape of the region pattern produced when the impingement surface is placed in the fully

2
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

Fig. 1. (a) Schematic illustration of the annular swirling impinging jet, (b) sample mesh; for H ¼ 2.0.

merged zone (H > Hc1), is designated as the flow pattern-1 (FP1). When and Reynolds number. Recently, Sharif [30] numerically investigated
the impingement surface is placed in the intermediate merging zone non-swirling and swirling impinging jet heat transfer from a flat surface
(Hc2 < H < Hc1), the flow pattern is designated as the flow pattern-2 for a range of the flow parameters (Re ¼ 23,000, 0 � Sw � 2, and H ¼ 2,
(FP2). When the impingement surface is placed in the initial merging 6, and 10). The results indicated a negative effect on heat transfer when
zone (H < Hc2), the flow pattern is designated as the flow pattern-3 (FP3) swirl is present. For low-to mid-range swirl strength when Sw � 0.77,
[12,15]. the average Nusselt number drops mildly as Sw is increased from zero
As much as the annular jet flow is complex, another level of com­ (non-swirling case). When Sw � 1, the average Nusselt number increases
plicacy is imposed on the flow physics and resulting heat transfer pro­ mildly with increasing Sw but remains less than that for the non-swirling
cess if the swirling motion is imparted on the inlet annular jet. Swirling case. Also when Sw � 0.77, the average Nusselt number increases mildly
jet is used for various applications. The jet spreading rate is altered when with increasing jet-to-target separation distance. On the other hand,
swirl is introduced at the jet thereby altering the heat transfer process for when Sw � 1.00, the average Nusselt number moderately drops as H
the flow. The jet decay, ambient fluid entrainment, and growth of the jet, increases.
etc., are influenced by the imposed swirl strength. The swirl strength or Relatively very few studies on annular swirling jet impingement flow
swirl number, Sw, is another important parameter (in addition to Re, BR, process and associated heat transfer from a flat surface have been pub­
and H) for the annular swirling jet flow configuration, which is the ratio lished in the literature. The effect of swirl on the annular jet impinge­
of the azimuthal momentum flux to the axial momentum flux at the jet ment flow within the initial merging zone was visualized by Particle
exit (detailed definition for Sw is given in section 3). Image Velocimetry (PIV) by Vanierschot and Bulck [31]. They reported
Heat transfer, due to swirling round jet impingement, from a flat that the flow in the initial merging zone (immediately downstream of
surface has been investigated experimentally and numerically by several the annular jet exit) is very complex. The central rod acts as a bluff body
researchers in the past. Huang and El-Genk [26] performed flow visu­ which results in a central recirculation zone (CRZ) behind it. With the
alization and experimental studies on non-swirling and swirling increase of swirl strength, a vortex breakdown bubble appears at an
impinging jet heat transfer. They observed that the swirling round axial downstream location, which moves axially upstream and eventu­
impinging jet produces more uniform heat transfer compared to the ally merges with CRZ. Thus, a very complex flowfield is generated at
non-swirling round impinging jet at larger separation distances. Lee higher swirl strengths. Numerical investigation of a swirling annular
et al. [27] experimentally studied round swirling jet impingement heat two-phase jet (liquid jet injected in a gaseous environment) by Siamas
transfer from a flat surface at Re ¼ 23,000 with H ¼ 2–10 and Sw ¼ et al. [32] also showed the existence of a vortex bubble downstream of
0–0.77. They found that, while swirl enhances the heat transfer at low the central recirculation zone. The preceding two articles reported on
swirl strength (Sw � 0.21) and low separation distance (H � 6), it the flow physics of the annular jet without any heat transfer studies.
negatively affects the heat transfer at higher Sw and H, compared to the Hallqvist and Fuchs [33] conducted a numerical investigation of annular
non-swirling round jet case. Swirl also achieves more uniformity (with swirling turbulent impinging jet heat transfer for the combination of Re
some reduction) in the heat transfer over the impingement surface ¼ 23,000, BR ¼ 0.4, and H ¼ 2. The swirl strength was varied as Sw ¼ 0,
compared to the non-swirling jet. Nanan et al. [28] conducted experi­ 0.5 and 1.0. Due to the swirling motion, the axial jet experienced
mental studies on swirling jet impingement heat transfer for a range of spreading and the heat transfer was affected negatively within the
Reynolds number (Re ¼ 4000–16,000) and separation distance (H ¼ stagnation region. Yang et al. [34,35] experimentally investigated
2–8). They used twisted tape with different twist ratios in the jet pipe to non-swirling and swirling annular jet impingement heat transfer char­
impart the swirling motion. They reported that, for a specific swirl acteristics. The Reynolds number (based on the outer diameter, Do, and
strength, the impinging jet at lower separation distances (H ¼ 2, 4) re­ the average jet velocity) and the blockage ratio in both of these works
sults in higher heat transfer rate compared to non-swirling jets but for were taken as 7000 and 0.5, respectively. The swirl strength in Ref. [34]
larger separation distances (H ¼ 6, 8), the scenario is reversed. The was 0.92 and the separation distance H was varied from 0.3 to 8.1
experimental and numerical study of Salman et al. [29], to investigate whereas the swirl strength in Ref. [35] was 0.45 and H was varied from
the effect of the swirling strength, separation distance, and Reynolds 0.5 to 4.1. Thus, the impingement surface was placed in the initial, in­
number on the heat transfer, demonstrated that more uniform distri­ termediate, full merging zones for the annular jet configuration. They
bution of heat transfer can be achieved, depending on the swirl strength observed that at short and intermediate separation distances, the

3
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

swirling motion introduced in the exit flow of the annular impinging jet � �
∂ðvz vθ Þ 1 ∂ðrvr vθ Þ vr vθ ∂ðτzθ Þ 1 ∂ðr2 τrθ Þ
caused strongly non-uniform distributions of wall pressure and heat Azimuthal θ direction: þ ¼ þ 2
transfer on the impingement plate. In sharp contrast, for large separation ∂z r ∂r r ∂z r ∂r
distances, the wall pressure and heat transfer of the swirling annular (4)
impinging jet were distributed more uniformly than those of the refer­ Energy equation
ence jet without swirling. Furthermore, at large separation distances, � � �
whilst the reference non-swirling jet exhibited simple round impinging ∂ðvz TÞ 1 ∂ðrvr TÞ 1 ∂ qz
þ ¼ þ
1 ∂ðrqr Þ
(5)
jet-like features, no such features were found for a swirling annular jet. ∂z r ∂r ρcp ∂z r ∂r
It is evident from the above literature review that, the majority of the
research on annular impinging jet heat transfer has been conducted for where vz, vr, and vθ are the mean axial (z), radial (r), and azimuthal (vθ)
non-swirling and swirling round jet impingement as well as for non- velocity components, p is the mean pressure, and T is the mean tem­
swirling annular jet impingement. Published research on the swirling perature. Here, constant fluid density and specific heat at constant
annular jet impingement heat transfer is not abundant. The combined pressure are represented by ρ and cp . The stress terms in Eqs. (2) – (3) are
effect of swirl and annular configuration on impinging jet heat transfer written in terms of the velocity gradients in indicial notation as
would result in complex and interesting convective heat transfer �
∂ui ∂uj

2
behavior, which remains relatively under-explored. It might have a τij ¼ ðν þ νt Þ þ
∂xj ∂xi 3
kδij
positive or negative effect on the heat transfer depending on the com­
binations of the different parameters. More systematic and detailed where ​ ν is fluid kinematic viscosity, νt is the turbulent viscosity
research efforts, covering a good range of the flow and geometric pa­ calculated by a suitable turbulence model, k is the kinetic energy of
rameters, are needed to explore and reveal the combined effect of turbulence, and δij is the Kronecker delta. The turbulent heat fluxes in
annular configuration and swirl on the impinging jet heat transfer. With Eq. (5) are expressed in terms of the temperature gradient as
that motivation, the swirling annular jet impingement heat transfer from � �
a flat surface is numerically investigated in the present study for a range qi ¼ αþ t
ν ∂T
of the flow and geometric parameters (Re, BR, H, and Sw). The flow Prt ∂xi
structure and heat transfer characteristics of the turbulent swirling
annular impinging jet is computed for various combinations of these where α ¼ k=ρcp is the fluid thermal diffusivity, k is the fluid thermal
parameters. The resulting thermal-hydraulic field is critically analyzed conductivity and Prt is the turbulent Prandtl number.
and the results are presented. The Reynolds number at the annular jet exit is expressed as
VA Dh VA ðDo Di Þ VA Do ð1 BRÞ
2. Problem description Re ¼ ¼ ¼ (6)
ν ν ν
The discharge for the annular jet can be expressed as
The schematic diagram for the axisymmetric swirling annular jet
� �
impingement flow configuration is shown in Fig. 1(a) in the axial-radial VA π D o 2 D i 2 VA πDo 2 1 BR2 Reð1 þ BRÞπDo ν
plane where Z is the axial direction and r is the radial direction. The QA ¼ ¼ ¼ (7)
4 4 4
height and length of the computational domain are 10Do � 10Do, based
The average velocity at the jet inlet is obtained dividing the
on the literature corroborations that this length is adequate to eliminate
discharge by the annular cross-sectional area as
the end effect [36]. The top boundary is the outlet and the bottom
boundary is the axis of symmetry. The left boundary is set as open Re ν
VA ¼ (8)
entrainment surface (excluding the jet inlet part) and the right boundary ð1 BRÞDo
is set as hot isothermal impingement surface maintained at 315 K.
On the hot impingement surface, the local Nusselt number is defined
Constant pressure condition is applied on the outlet and open entrain­
as
ment surface. At the jet inlet boundary, a uniform velocity, VA, is
specified based on the jet exit Reynolds number, while the cold jet inlet qðD
_ o Di Þ
Nu ¼ ; (9)
temperature is maintained at 300 K. kðTh Tc Þ

3. Numerical procedure where q_ is the surface heat flux. The average Nusselt number for a cir­
cular impingement surface of Radius R is obtained as:
3.1. Governing mathematical equations ZZ Z R Z 2π Z R
1 1 2
Nu ¼ Nu dA ¼ 2 Nu ðrdθdrÞ ¼ 2 ðNuÞðrÞdr (10)
The governing equations in this current study are the conservation of A πR 0 0 R 0
mass, momentum, and energy. The equations are expressed in terms of The swirl number, Sw, at the jet exit is defined as the ratio of the
the axisymmetric cylindrical coordinate system with constant fluid azimuthal momentum (Gθ) to the axial momentum (Gz),
properties as: R
Continuity equation Gθ ρvz vθ ðrdrÞ
(11)
jet ​ exit
Sw ¼ ¼R
Gz jet ​ exit
ρvz vz ðrdrÞ
∂vz 1 ∂ðrvr Þ
þ ¼0 (1)
∂z r ∂r In practical applications, the jet exit swirl is imposed by using swirl
generator vane or inserting a twisted tape inside the jet pipe. Thus, if the
Momentum equation
swirl velocity information at the jet exit is not accessible for a specific
��
∂v2 1 ∂ðvr vz Þ 1 ∂p ∂ðτzz Þ 1 ∂ðτrz Þ swirl guide vane or twisted tape, then numerically it is very challenging
Axial z direction: z þ ¼ þ þ (2)
∂z r ∂r ρ ∂z ∂z r ∂r to impose swirl velocity at the jet inlet. A simple substitute method is to
� impose a rigid body rotation as w ¼ rω at the jet inlet, where ω is a
∂ðvz vr Þ 1 ∂ rv2r 1 ∂p ∂ðτrz Þ 1 ∂ðrτrr Þ constant angular speed. Thus, Eq. (11) with uniform jet inlet axial ve­
Radial ðrÞ direction: þ ¼ þ þ (3)
∂z r ∂r ρ ∂r ∂z r ∂r locity VA , can be written as

4
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

Fig. 3. Comparison of pressure coefficient (Cp) along the hot impingement


surface against experimental data for a non-swirling annular jet.

Fig. 2. Comparison of local Nusselt number along the hot impingement surface
against experimental data of Musika et al. [11] for a non-swirling annular jet.

R Do =2 ,0 1
ω r2 dr1 ωðDo DiÞ
(12)
Di =2 @Do
Sw ¼ R ¼ or ω ¼ 3 SwVA Di A
VA Do =2 rdr 3 VA
Di =2

3.2. Computational details

All the computations are performed using the commercial software


ANSYS Fluent 18.1. Integration over the finite volume meshes produces
a set of linear equations, which are solved sequentially. The coupling
between pressure and velocity is attained by the SIMPLE method. The
convergence criteria for residual error is set to 10 6 and the total heat
flux at the impingement surface is monitored until it becomes invariant
with iterations. A sample mesh of the whole domain for H ¼ 2 of the
present numerical model is shown in Fig. 1(b). The mesh is clustered
toward the right impingement wall in the Z direction, and toward the
Fig. 4. Mesh refinement study of (a) local Nusselt number distribution along
edge of the jet in the r direction with a bias factor. The mesh between the
the hot impingement surface and (b) distribution of wall yþ along the hot
axis and the jet outer wall along r direction is uniformly spaced.
impingement surface for a non-swirling annular jet.

3.3. Turbulence model validation


Nusselt number distribution that matches very well with the experi­
mental data for any of the radial locations. The local Nusselt number
In order to choose an appropriate turbulence model among the
distribution for r/D < 1 is under-predicted by the RNG k-ε model and
various turbulence model options given in the ANSYS Fluent code, the
Transition SST model. Among all of the turbulence models, the Realiz­
performance of four selected models is evaluated against the experi­
able k-ε model provides overall good agreement with experimental data
mental data [23,34] for the computation of the annular impinging jet
for the local Nusselt number distribution. Moreover, the pressure coef­
heat transfer. The selected models included the RNG k-ε model, the
ficient, CP, along the radial direction of impingement surface at Re ¼
Realizable k-ε model, the Transition SST model, and the Reynolds stress
7000 with BR ¼ 0.5 for two different separation distance (H) of 0.5 and
model, which are frequently used in academia and industry. The detail
4.1, predicted using the Realizable k-ε model, are validated with the
information of these models are available in ANSYS fluent user manual
experimental data [34] in Fig. 3 and shows a very good agreement.
and elsewhere and is not included here. The comparison of the local
Hence, the Realizable k-ε model is chosen over the other three turbu­
Nusselt number distribution along the radial direction on the hot
lence models and is used for all of the further computations.
impingement surface at Re ¼ 20,000, BR ¼ 0.44, and H ¼ 2 and 6 with
the experimental data [23] is shown in Fig. 2. The experimental data is
reported for r/Do � 3, only. None of the turbulence models predicted the

5
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

Fig. 5. (a) Streamlines for different separation distances, and corresponding (b) axial velocity distribution along the jet axis (inset is a blown up view near the
impingement surface), (c) pressure coefficient distribution in the radial direction on the impingement surface, and (d) the Nusselt number distribution in the radial
direction on the impingement surface (non-swirling annular jet at Re ¼ 5000 and BR ¼ 0.6).

3.4. Grid refinement study strength, Sw, is varied as 0, 0.21, 0.44, and 0.77. Thus, a total of 36
combinations of BR, Sw, and H are considered for the computations.
Using the Realizable k-ε model, a systematic grid refinement study is Additional computations for 24 more combinations of the parameters, at
also conducted by first choosing a coarse mesh (110 � 108) and then two other Reynolds number of Re ¼ 15,000 and 25,000 and at a fixed BR
successively refining it. The gradual convergence of the predicted local of 0.5 with the same combinations of Sw and H for Re ¼ 5,000, are also
Nusselt number distribution in the radial direction with mesh refine­ computed to see the effect of Reynolds number on the heat transfer
ment is shown in Fig. 4. It may be noticed that no significant difference process.
in the Nusselt number distribution is produced beyond the 180 � 127
mesh. Hence, this mesh distribution is used for all simulations in the
present study. The yþ distance to the first internal node from the wall is 4.1. Non-swirling annular impinging jets (Sw ¼ 0)
required to be around 1.0 for the Realizable k-ε model. The distribution
of the yþ distance to the first internal node along the hot impingement For a non-swirling annular impinging jet, the jet-to-target separation
surface is plotted for different mesh sizes in Fig. 4 (b), which shows that distance strongly influences the flow behavior and pressure coefficient
for mesh resolution of 180 � 127, yþ value falls mostly below 2, which on the impingement surface. As mentioned before, three different flow
satisfies the very fine mesh refinement condition near the impingement and heat transfer characteristics evolve depending on whether the
surface. impingement surface is placed in the initial merging zone, the inter­
mediate merging zone, or the fully merged zone. These flow patterns are
4. Results and discussions recognized by the sudden changes in the flow behavior as the separation
distance is gradually varied. In order to illustrate this phenomenon, a
The flow and thermal field computations are performed for various representative case of the non-swirling case with Re ¼ 5000 and BR ¼
combinations of the flow and geometric parameters. The annular pipe 0.6 is computed where the value of the separation distance, H, is varied
has the outer diameter of Do ¼ 0.03 m and the blockage ratio, BR, is as 0.5, 2, and 4. The effect of varying the separation distance on the flow
varied as 0.4, 0.5, and 0.6. The Reynolds number is taken as 5000 for all field (streamline plots), pressure coefficient and local Nusselt number
computations. The separation distance, H, is varied as 0.5, 2, and 4. distribution along the radial direction on the impingement surface, and
These three separation distances place the impingement surface in the the axial velocity profile along the jet axis from the jet exit to the
initial merging zone, the intermediate merging zone, and the fully impingement surface is illustrated in Fig. 5(a)-(d).
merged zone, respectively, for a non-swirling annular jet. The swirl The common feature present in all three streamline plots (at H ¼ 0.5,
2, and 4) is the presence of the clockwise recirculating flow region

6
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

Fig. 6. Streamlines, contours of the swirl velocity components, and isotherms on the axial-radial plane for different swirl strength; Re ¼ 5000, BR ¼ 0.5 and H ¼ 0.5.

immediately downstream of the jet exit between the jet axis and the the impingement surface at a downstream radial location in the wall jet
main jet stream. This is due to the presence of the inner blockage rod, region. The axial velocity profile along the jet axis goes through a
which creates a sudden expansion corner configuration at the jet exit. negative, positive, and then another negative region along the jet axis at
The recirculation bubble remains attached to the impingement surface this intermediate separation distance. For the higher separation distance
at a lower value of H (0.5) but is detached from the impingement surface (H ¼ 4), the annular jet merges into a single jet at a downstream axial
as the separation distance increases (H ¼ 2 and 4). Because of the location and impinges on the surface like a single round jet. The axial
recirculation bubble, there is a flow reversal (or negative axial velocity) velocity along the axis is negative in the recirculation zone near the jet
along the jet axis in that part of the axis at all separation distances. At the exit and gets positive after the merging location and remains positive all
intermediate separation distance (H ¼ 2) the primary recirculation the way up to the impingement surface without any flow reversal. These
bubble is detached from the impingement surface, but another weak features of the distribution of the axial velocity profile along the jet axis
clockwise recirculation bubble is formed at the corner between the jet at each of the three separation distances, at Re ¼ 5000 and BR ¼ 0.6, are
axis and impingement surface. This causes axial and radial flow reversal captured in the velocity profile plots in Fig. 5(b) where xa is the non-
in that corner, the stagnation point moves off the impingement surface dimensional distance (normalized by Do) along the axis from the jet
inside on the jet axis. Another stagnation point (or a circle) is formed on exit. Notice the close-up view of the axial velocity profile near the

7
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

Fig. 7. Streamlines, contours of the swirl velocity components, and isotherms on the axial-radial plane for different swirl strength; Re ¼ 5000, BR ¼ 0.5 and H ¼ 2.

impingement surface at xa/H around 1 in the inset in Fig. 5(b) where the akin to that for the regular round jet impingement case with large jet-to-
negative axial velocity along the axis for H ¼ 0.5 is clearly visible. This is target separation distance. The peak of the CP and Nu lie on the jet axis
the implication of the flow reversal at the intermediate separation (r/Do ¼ 0) which then monotonically drops in the radial direction in the
distance. wall jet region. At the lower H values (0.5 and 2) the Cp and Nu starts
The distribution of the pressure coefficient and the local Nusselt with a lower value at the jet axis and reaches a peak at around r/Do ¼ 0.7
number on the impingement surface in the radial direction is plotted in from which the monotonically drop in the radial direction. This is due to
Fig. 5(c) and (d), respectively, for the three separations distances. It is the presence of the reversed stagnation flow near the impingement
noticed that, at the larger separation distance (H ¼ 4), the Cp and Nu surface around the jet axis at the lower jet-to-target distance. The sudden
distribution along the radial direction on the impingement surface is changes in the flow characteristics among the three separation distances

8
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

Fig. 8. Streamlines, contours of the swirl velocity components, and isotherms on the axial-radial plane for different swirling strength; Re ¼ 5000, BR ¼ 0.5 and H
¼ 4.

are also evident in the pressure coefficient and local Nusselt number more robust, especially at high swirling strengths. A small recirculation
distribution plots in Fig. 5(c) and (d). bubble form at the lower right corner (stagnation region) when Sw ¼
0.21 and the recirculation bubble gets bigger as swirling strength in­
4.2. Swirling annular impinging jet hydrodynamics creases. This recirculation region pushes the annular jet outside of
annular jet slot and impingement location moves to a downstream
Representative plots of the streamlines, swirl velocity contours, and location. This stronger swirling motion causes the isotherms to be much
isotherms are presented in Figs. 6–8 for H ¼ 0.5, 2.0 and 4.0, respec­ thicker in the location around the jet axis. The vortex breakdown of
tively. The domain extent in the radial (r) direction is truncated beyond swirling jets might be the reason behind the formation of recirculation
which there is no critical flow feature present. bubble.
It is discernible from these plots that the swirl significantly modifies In Fig. 8, for larger H ¼ 4.0, the flow dynamics of non-swirling
the flow patterns (streamlines) and isotherms by various degree, annular jet impingement is similar to the conventional non-swirling
compared to the non-swirling cases. The main jet stream of the annular round jet impingement. Low swirling strength does not have much ef­
jet impinges on the surface, then turns, and forms a wall jet along the fect on the flow dynamics and heat transfer, as seen in the numerical
surface. As the swirl strength is increased, various levels (increasing investigation of Sharif [30]. However, at higher swirling strength (Sw �
size) of recirculation flows emerge between the jet axis and main jet 0.44), recirculation bubble forms and drives the mainstream slightly
stream. The isotherms are modified accordingly, which affects the heat downstream. As a result, the thermal boundary layer becomes thicker
transfer from the surface. For lower separation distance (H ¼ 0.5), the around the jet axis.
effect of swirl is not as significant compared to that for higher H (2 and
4). As seen from the streamlines in Fig. 6, with the increase of swirl 4.3. Effect of swirl on the heat transfer
strength, the recirculation bubble on hot impingement surface grows
slightly bigger in size and spreads both to the downstream and upstream The characteristics of the swirling annular impinging jet heat
position. The magnitude of swirl velocity in the region, immediately transfer are evaluated in terms of the local Nusselt number distribution
after it leaves the jet exit, gets higher with the increase of swirl strength. on the impingement surface. The predicted local Nusselt number dis­
Finally, the isotherms show that higher swirling strength causes a tribution along the radial direction on the impingement surface is
slightly thicker thermal boundary layer near the jet axis. extracted for all cases considered and are plotted in Fig. 9. The general
In Fig. 7, for H ¼ 2, the effect of imposed swirl on flow dynamics are trend of the distribution is that the local Nu starts with a low value at the

9
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

Fig. 9. Distribution of local Nusselt number (Nu) along radial direction on the hot impingement for various combination of H, BR, and Sw for Re ¼ 5000.

jet axis where the thermal boundary layer is thicker and then rises to a separation distances (H ¼ 2 and 4) the average Nu initially decreases
maximum at the impingement location of the main jet stream where the slightly with increasing Sw, up until Sw ¼ 0.44 at BR ¼ 0.4 and 0.5, and
thermal boundary layer is the thinnest (isotherms are packed close up until Sw 0.21 at BR ¼ 0.6. Beyond these ranges of Sw, the average Nu
together) and then gradually and monotonically drops in the radially increases with increasing Sw.
outward direction. The effect of swirl is only noticed up to the radial The swirl adds another velocity component to the flow in addition to
location of r/Do ~2 beyond which the profiles for all Sw merge together the axial and radial velocity components thereby increasing the
for any specific combination of Re, BR, and H. The effect of swirl on the magnitude of the convective velocity, which in turn affect the heat
local Nusselt number distribution is minimal at lower separation dis­ transfer from the impingement surface. The swirl velocity is mostly
tance (H ¼ 0.5) where the Nusselt number slightly increases with concentrated around the jet axis and is quickly diffused (weakened) in
increasing swirl at any radial location. At higher separation distances (H the radial direction in the wall jet region (see the swirl velocity contours
¼ 2 and 4), the local Nu and its peak value significantly decrease and the in Figs. 6–8). Thus, at lower separation distance and at higher swirl
location of the peak Nu shifts radially outward with increasing Sw. strength, the heat transfer (average Nu) increases with increasing Sw
The effect of the swirl strength on the overall heat transfer is due to increased convective velocity magnitude within the wall jet
comprehensible in terms of the average Nusselt number on the boundary layer. At higher separation distance, however, the swirl does
impingement surface. The average Nusselt number, at a fixed Re of 5000 not penetrate strongly into the wall jet boundary layer and negatively
for all combinations of H and Sw, is calculated using Eq (10); and is affects the heat transfer yielding decreasing average Nu trend with
plotted in Fig. 10(a) as a function of swirl strength. The variation trend increasing swirl strength.
of the average Nu is not consistent and differs for the different separation The average Nusselt number is also affected by the blockage ratio.
distances. At low H (0.5), the average Nu almost linearly increases with For example, at H ¼ 0.5 and Sw ¼ 0.77, the average Nusselt number
increasing swirl strength, the increase is about 6% at BR ¼ 0.4 to 8% at increases by 6.45%, 7.78%, and 7.96% compared to non-swirling
BR ¼ 0.6, compared to the non-swirling case. However, at higher annular jet when the BR is varied as 0.4, 0.5, and 0.6, respectively. A

10
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

Fig. 10. Variation of average Nusselt number and peak Nusselt number for various combination of H, BR, and Sw for Re ¼ 5000.

similar increasing trend of the average Nusselt number with increasing swirling annular jet.
blockage ratio is also noticed at higher separation distances (H ¼ 2 and
4) at lower Sw (0.21 and 0.44). 4.4. Effect of jet exit Reynolds number on the heat transfer
The peak Nusselt number at the impingement surface is another
indication of the intensity of the heat transfer at the main jet stream In order to investigate the effect of jet exit Reynolds number on
impingement location. The peak Nusselt number variation as a function annular turbulent swirling impinging jet heat transfer, additional com­
of the swirling strength at Re ¼ 5000 is shown in Fig. 10(b) for the putations are performed for Re ¼ 15,000 and 25,000 at a fixed BR of 0.5,
different separation distances. Like the average Nusselt number, the while H is varied as 0.5, 2.0 and 4.0 and Sw is varied as 0, 0.21, 0.44 and
peak Nusselt number almost linearly increases with increasing Sw at H 0.77. The streamlines, isotherms and swirling velocity contours (not
¼ 0.5. On the other hand, at H ¼ 2 and 4, the peak Nusselt number has a shown) have a similar variation trend to those for Re ¼ 5000 presented
decreasing trend with increasing Sw. The peak Nusselt number increases in Fig. 7. The local Nusselt number distribution along the radial direc­
with increasing BR at any specific swirl strength. For H ¼ 0.5 and Sw ¼ tion on the impingement surface for Re ¼ 15,000 and 25,000 at various
0.44, while BR varies as 0.4, 0.5 and 0.6, the peak Nusselt number in­ combinations of separation distance and swirling strength are presented
creases by 1.61%, 2.89% and 4.24%, respectively, compared to the non- in Fig. 11. The local Nusselt number at any specific radial location at any

11
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

Fig. 11. Local Nusselt number distribution along the radial direction on the impingement surface for various combinations of H and Sw at BR ¼ 0.5.

specific H significantly increases as the Reynolds number is increased attributed to the three different flow patterns associated with the cor­
from 5000 to 25,000. However, near the jet axis, the local Nusselt responding separation distances. The peak Nusselt numbers for all
number does not vary significantly with the variation of the Reynolds combinations of Re, H, and Sw (at BR ¼ 0.5) are extracted and plotted
numbers due to the presence of the low-pressure zone here, especially at also in Fig. 12. The variation trend for the peak Nu is similar to that for
a high swirling rate. The effect of swirl on the local NuNu distribution is the average Nu.
also conceivable in these plots, which is significant especially at higher
separation distance. 5. Conclusion
The average Nusselt number at a fixed BR of 0.5 for all combinations
of H and Sw plotted in Fig. 12 as a function of Sw. In general, the average Numerical investigation of the turbulent annular swirling jet
Nusselt number increases with increasing Reynolds number at a impingement heat transfer is performed in this study. The Realizable k-ε
particular Sw. The swirl strength, however, influences the average Nu turbulence model is chosen for the computations after the performance
differently. At low separation distance (H ¼ 0.5), the average Nu in­ evaluation of a few different turbulence models against previously
creases slightly with increasing Sw while at higher separation distance published experimental data. The flow Reynolds number is varied from
(H ¼ 4), the average Nu decreases slightly with increasing Sw. At the 5000 to 25,000, the annular nozzle diameter ratio, BR, is varied from 0.4
intermediate separation distance (H ¼ 2), the average Nu initially de­ to 0.6, the jet-to-target separation distance, H, is varied from 0.5 to 4,
creases and then increases with increasing Sw. This behavior can be and the swirl strength, Sw, is varied from 0 to 0.77. Flow and thermal

12
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

Fig. 12. Effect of swirl on the average and peak Nusselt number for various combination of H, Sw for BR ¼ 0.5 at different Reynolds number.

field computations are performed for many combinations of the geo­ 2. At low jet-to-target separation distance (H ¼ 0.5), the average Nus­
metric and flow parameters to investigate the effects of swirl and other selt number increases almost linearly with increasing swirl strength,
parameters on the heat transfer from the impingement surface. The while at higher separation distance (H ¼ 2 and 4), the average
streamlines, swirl velocity contour, isotherms, and the local Nusselt Nusselt number initially decreases mildly (at low Sw) and then in­
number distribution on the impingement surface, and the average creases moderately (at higher Sw) with increasing Sw.
Nusselt number on the impingement surface, for various combinations 3. Certain combinations of BR, Sw, and Re can improve the average
of H, BR, Re, and Sw, is presented. Summary of the results are as follows: Nusselt by about 6–8% compared to the non-swirling jet, while other
combinations (lower Sw and higher H) reduces the heat transfer.
1. The swirl affects the heat transfer from the impingement surface both 4. The local Nusselt number distribution on the impingement surface
positively and negatively depending on the combinations of the flow starts with a low value at the jet axis and quickly rises to a maximum
parameters (Re, H, BR, and Sw). Shorter jet-to-target separation value at around r/Do ~0.7 to 1.2, from where it quickly decreases
distances cause reverse stagnation flow, which significantly reduces monotonically to a constant value for r/Do � 3. The location of the
the heat transfer near the jet axis for the non-swirling case. The local Nusselt number maximum shifts to downstream radial location
imposed swirl moderately modifies (enhancement or reduction) the as swirling strength increases. For intermediate distance (H ¼ 2.0)
heat transfer. and high swirling strength (Sw � 0.44), the flowfield near the jet axis

13
F. Afroz and M.A.R. Sharif International Journal of Thermal Sciences 151 (2020) 106257

is characterized by a recirculation bubble on the impingement sur­ [15] N. Ko, W. Chan, The inner regions of annular jets, J. Fluid Mech. 93 (1979)
549–584.
face. The low-pressure zone due to this recirculation bubble causes
[16] H. Maki, A. Yabe, Heat transfer by the annular impinging jet, Exp. Heat Transf. Int.
the Nusselt number to be very low on impingement surface close to J. 2 (1989) 1–12.
the jet axis. [17] K. Ichimiya, Heat transfer characteristics of an annular turbulent impinging jet
5. The Reynolds number has a strong positive effect on the heat transfer with a confined wall measured by thermosensitive liquid crystal, Heat Mass Transf.
39 (2003) 545–551.
and the average Nusselt number increases almost linearly with [18] H. Chattopadhyay, Numerical investigations of heat transfer from impinging
increasing Reynold number. annular jet, Int. J. Heat Mass Transf. 47 (2004) 3197–3201.
6. The blockage ratio has a moderate effect on heat transfer. At any [19] H. Chattopadhyay, Impinging heat transfer due to a turbulent annular jet, Int. J.
Transp. Phenom. 9 (2007) 287.
specific combination of the Reynolds number and swirl strength, the [20] Z. Tr�
avnıć�ek, K. Peszy�nski, J. Ho�sek, S. Wawrzyniak, Aerodynamic and mass
average Nusselt number at the impingement surface decreases with transfer characteristics of an annular bistable impinging jet with a fluidic flip–flop
increasing blockage ratio at a specific separation distance. control, Int. J. Heat Mass Transf. 46 (2003) 1265–1278.
[21] Z. Tr�
avní�cek, V. Tesa�r, Hysteresis in annular impinging jets, Exp. Therm. Fluid Sci.
44 (2013) 565–570.
Appendix A. Supplementary data [22] H. Zhen, C. Leung, C. Cheung, Heat transfer characteristics of an impinging
premixed annular flame jet, Appl. Therm. Eng. 36 (2012) 386–392.
[23] W. Musika, M. Wae-Hayee, P. Vessakosol, B. Niyomwas, C. Nuntadusit, Investig.
Supplementary data to this article can be found online at https://doi. Flow Heat Transf. Charact. Annul. Impinging Jet 931 (2014) 1223–1227.
org/10.1016/j.ijthermalsci.2019.106257. [24] S. Kalinina, V. Terekhov, K. Sharov, Special features of flow in an annular jet
impinging on a barrier, Fluid Dyn. 50 (2015) 665–671.
[25] V. Terekhov, S. Kalinina, K. Sharov, An experimental investigation of flow
References
structure and heat transfer in an impinging annular jet, Int. Commun. Heat Mass
Transf. 79 (2016) 89–97.
[1] J. Baughn, S. Shimizu, Heat transfer measurements from a surface with uniform [26] L. Huang, M. El-Genk, Heat transfer and flow visualization experiments of swirling,
heat flux and an impinging jet, J. Heat Transf. 111 (4) (1989) 1096–1098. multi-channel, and conventional impinging jets, Int. J. Heat Mass Transf. 41 (1998)
[2] K. Jambunathan, E. Lai, M. Moss, B. Button, A review of heat transfer data for 583–600.
single circular jet impingement, Int. J. Heat Fluid Flow 13 (2) (1992) 106–115. [27] D.H. Lee, S.Y. Won, Y.T. Kim, Y.S. Chung, Turbulent heat transfer from a flat
[3] B. Webb, C. Ma, Single-phase liquid jet impingement heat transfer, Adv. Heat Tran. surface to a swirling round impinging jet, Int. J. Heat Mass Transf. 45 (2002)
26 (1995) 105–217. 223–227.
[4] R. Viskanta, Heat transfer to impinging isothermal gas and flame jets, Exp. Therm. [28] K. Nanan, K. Wongcharee, C. Nuntadusit, S. Eiamsa-Ard, Forced convective heat
Fluid Sci. 6 (2) (1993) 111–134. transfer by swirling impinging jets issuing from nozzles equipped with twisted
[5] H. Martin, Heat and mass transfer between impinging gas jets and solid surfaces, tapes, Int. Commun. Heat Mass Transf. 39 (2012) 844–852.
Adv. Heat Tran. 13 (1977) 1–60. [29] S.D. Salman, A.A.H. Kadhum, M.S. Takriff, A.B. Mohamad, Experimental and
[6] P. Xu, B. Yu, S. Qiu, H.J. Poh, A.S. Mujumdar, Turbulent impinging jet heat transfer numerical investigations of heat transfer characteristics for impinging swirl flow,
enhancement due to intermittent pulsation, Int. J. Therm. Sci. 49 (7) (2010) Adv. Mech. Eng. 6 (2014) 631081.
1247–1252. [30] M. Sharif, Numerical investigation of round turbulent swirling jet impingement
[7] T. Liu, J. Sullivan, Heat transfer and flow structures in an excited circular heat transfer from a hot surface, Comput. Therm. Sci.: Int. J. 8 (2016).
impinging jet, Int. J. Heat Mass Transf. 39 (17) (1996) 3695–3706. [31] M. Vanierschot, E. Van Den Bulck, Influence of swirl on the initial merging zone of
[8] L.A. Brignoni, S.V. Garimella, Effects of nozzle-inlet chamfering on pressure drop a turbulent annular jet, Phys. Fluids 20 (2008) 105104.
and heat transfer in confined air jet impingement, Int. J. Heat Mass Transf. 43 (7) [32] G.A. Siamas, X. Jiang, L.C. Wrobel, Numerical investigation of a perturbed swirling
(2000) 1133–1139. annular two-phase jet, Int. J. Heat Fluid Flow 30 (2009) 481–493.
[9] N. Gao, H. Sun, D. Ewing, Heat transfer to impinging round jets with triangular [33] T. H€allqvist, L. Fuchs, Numerical study of swirling and non-swirling annular
tabs, Int. J. Heat Mass Transf. 46 (14) (2003) 2557–2569. impinging jets with heat transfer, AIAA Paper 5 (2005) 5153.
[10] D. Colucci, R. Viskanta, Effect of nozzle geometry on local convective heat transfer [34] H. Yang, T. Kim, T. Lu, K. Ichimiya, Flow structure, wall pressure and heat transfer
to a confined impinging air jet, Exp. Therm. Fluid Sci. 13 (1) (1996) 71–80. characteristics of impinging annular jet with/without steady swirling, Int. J. Heat
[11] F. Afroz, M. Sharif, Numerical study of heat transfer from an isothermally heated Mass Transf. 53 (2010) 4092–4100.
flat surface due to turbulent twin oblique confined slot-jet impingement, Int. J. [35] H. Yang, T. Kim, T. Lu, Characteristics of annular impinging jets with/without
Therm. Sci. 74 (2013) 1–13. swirling flow by short guide vanes, Sci. China Technol. Sci. 54 (2011) 749–757.
[12] F. Afroz, M. Sharif, Numerical study of turbulent annular impinging jet flow and [36] H. Laschefski, T. Cziesla, G. Biswas, N. Mitra, Numerical investigation of heat
heat transfer from a flat surface, Appl. Therm. Eng. 138 (2018) 154–172. transfer by rows of rectangular impinging jets, Numer. Heat Transf. Part A Appl. 30
[13] W. Chan, N. Ko, Coherent Structures in the outer mixing region of annular jets, (1996) 87–101.
J. Fluid Mech. 89 (1978) 515–533.
[14] N. Ko, W. Chan, Similarity in the initial region of annular jets: three configurations,
J. Fluid Mech. 84 (1978) 641–656.

14

You might also like