You are on page 1of 16

International Journal of Thermal Sciences 154 (2020) 106389

Contents lists available at ScienceDirect

International Journal of Thermal Sciences


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

Experimental and numerical study of inclined free surface liquid


jet impingement
Kuldeep Baghel *, Arunkumar Sridharan, Janani Srree Murallidharan
Department of Mechanical Engineering, Indian Institute of Technology Bombay, Mumbai, 400076, India

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

Keywords: Experimental and computational investigation is carried out to study heat transfer characteristics and to explain
Inclined liquid jet the underlying physics for the case of inclined free surface liquid jet impinging on a uniformly heated flat plate.
Free surface Experimentally, four Reynolds number 17541, 26,311, 35,081, and 438,521 were considered while in the nu­
Heat transfer
merical study, Reynolds number of 17,541 and 26,311 were considered. In all the experiments, jet to surface
Jet impingement
Oblique jet
spacing was fixed (H=d ¼ 4) and inclination angles of 45o , 30o , 30o and 0o (orthogonal jet) were considered.
SST-SAS model Local Nusselt number contours are presented for the above range of Reynolds number and inclination angle.
Also, profiles of Nusselt number are analyzed and compared for different Reynolds numbers and inclination
angles (θ). The numerical study was performed using SST-SAS model and VOF method was used to capture the
liquid-air interface. The aim of the numerical study is to explain the underlying physics of heat transfer char­
acteristics through flow dynamics. Streamlines, pressure, velocity, and turbulence intensity are calculated and
plotted for a few cases. It is observed that with increasing inclination angle, the location of peak Nusselt number
is shifted towards the compression side or uphill side, while the magnitude remains nearly the same. The point of
maximum Nusselt number was found to be independent of Reynolds number.

the geometric orientation of heated regions. Also, the effect of increasing


inclination is known to affect heat transfer [4–7] significantly.
1. Introduction Measurements of temperature have been the primary concern in jet
impingement studies. Liquid crystal technique was used in various
Jet impingement is known for high heat removal from a localized impinging air jet studies to capture the thermal profile of the heater
area of higher heat load due to the direct impact of cooling fluid on the surface [4,8]. Yan and Saniei [4] presented experimental results of local
heated surface. Jet impingement technique is being utilized in many heat transfer for oblique air jet impingement on a heated plate using
industrial applications for cleaning and heat removal [1–3]. Due to its liquid crystal technology to capture the temperature profile. It was
high heat removal capacity, jet impingement is widely investigated. shown that the location of peak heat transfer is shifted towards the
Depending on the complexity of the location where the heat is to be uphill direction of the wall from the geometric center [4,9,10]. The shift
removed, many orientations of jet impingement are possible. A simple was more prominent for higher inclination. Bhagwat and Sridharan [5]
possible configuration is an orthogonal jet impingement on a flat sur­ numerically investigated inclined circular air-jet on a heated surface
face, where the surface shape could be ribbed, rough, curved, elliptic, using v2 f turbulence modeling. As found in experimental literature, the
etc. The jet could be confined or free surface jet, depending on the point of peak Nusselt number shifted from the geometric center towards
environment it is discharging into. Air jet impingement is widely studied the uphill direction of the impingement plate. Goldstein and Franchett
in the literature, while studies related to water jet impingement are [8] have given a correlation for local Nusselt number in terms of angle of
limited considering the complexity of problem and difficulty associated impingement and distance. Choo et al. [11] investigated the effect of the
with it, such as splattering of liquid jet and bending of thin foil used as a lower jet to surface spacing (H=d � 1), and pumping power on heat
heater among others. Jet impingement studies include the effect of pa­ transfer is characteristics for oblique air jet. They showed that increased
rameters like Reynolds number (Re), jet to surface spacing (H= d), the heat transfer with the increasing inclination for H=d � 1 is due to high
effect of geometry at the nozzle exit. Few applications require the pumping power. For H=d � 1, a decrease in Nusselt number with
application of inclined jet due to the complexity and constraint due to

* Corresponding author.
E-mail address: kuldeepbaghel.11@gmail.com (K. Baghel).

https://doi.org/10.1016/j.ijthermalsci.2020.106389
Received 30 July 2019; Received in revised form 10 March 2020; Accepted 17 March 2020
Available online 6 April 2020
1290-0729/© 2020 Elsevier Masson SAS. All rights reserved.
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

Nomenclature Nus stagnation Nusselt number


Pr Prandtl number
Greek symbol q’’ heat flux W=m2
α volume fraction Re Reynolds number
μ kinematic viscosity ðm2 =sÞ T temperature (K)
ρ density kg=m3 angle∘ uj jet exit velocity (m=s)
Uref reference velocity or jet exit velocity m=s
List of symbol
V Voltage drop across the plate V
A area of the heated surface m2
cp specific heat at constant pressure j=ðkg ⋅KÞ v2 variance of the normal component of the turbulent velocity
d nozzle exit diameter m m2 s2
f elliptic relaxation function kg=ðm3 ⋅sÞ W width of heated surface m
g gravitational acceleration m=s2 x=d non dimensional longitudinal axis
H nozzle to plate spacing m y=d non dimensional transverse axis
h convective heat transfer coefficient W=ðm2 KÞ yþ non dimensional distance from wall
I current A Abbreviations
k thermal conductivity W=m⋅K exp: experimental
L length of the heated surface m num: numerical
l length of the nozzle pipe m vel: velocity
Nu Nusselt number

increasing inclination is due to loss of momentum of wall jet. Interest­ Investigation of these parameters was mostly performed using air-jet
ingly, with fixed pumping power, the Nusselt number was found to be impingement. Literature is scarce for inclined free surface jet impinge­
insensitive of the inclination angle. O’Donovan and Murray [12] pre­ ment involving liquid such as water. In air-jet, surrounding air is known
sented fluctuations in heat transfer on the impingement plate and ve­ to affect hydrodynamics of jet, and hence heat transfer from the surface
locity fluctuations along wall jet for inclined air jet impingement on a is affected due to vortices formed by strong interactions between
heated flat plate for Re ¼ 10000. The structure of vortices was affected ambient air and jet fluid [12]. For free surface liquid jet, the effect of air
by the inclination angle and nozzle to surface distance. on jet might be different due to the difference in dynamic viscosity of
Beltaos [13] performed an analytical and experimental investigation liquid and air [17]. In the case of orthogonal free surface liquid jet
to study planar oblique jet impingement. He suggested a theoretical impingement, geometric center and location of peak heat transfer co­
method to present shear stress in the symmetry plane. The pressure was efficient coincide with each other. However, in the case of an inclined
predicted by a semi-empirical method. Also, flow characteristics were jet, the point of peak heat transfer is known to be shifted towards the
predicted by a method based on boundary layer approximations and compression side for both air-jet as well as liquid jet impingement
similarity of radial velocity profiles. Beitelmal et al. [7] showed that the studies [4,6,9,10,16]. Nusselt number distribution also changes in the
location of the stagnation point was insensitive to Reynolds number for transverse direction, analysis of such results and relevant physics are
fixed inclination angle, and peak Nusselt number was found to decrease scarce in the literature. Few experimental studies focused on a free
with decreasing inclination angle. Benmouhoub and Mataoui [14] surface inclined liquid jet [6,16], but there is a scarcity of research
numerically studied the problem of an inclined planar jet impinging on a concerning flow dynamics of inclined free surface jet impingement on a
moving wall numerically. The aim was to control the stagnation point on flat surface.
the moving wall through the inclination angle. Moreover, none of the numerical studies performed are for the free
Sparrow and Lovell [15] have used the naphthalene sublimation surface liquid jet with inclined nozzle configurations. Hence, a numer­
technique, which estimates the mass transfer coefficient. The mass ical investigation could be useful to understand hydrodynamics and its
transfer coefficient was utilized to calculate the heat transfer coefficient relation to heat transfer. The complex asymmetrical nature of three-
using the analogy between two processes. They showed that maximum dimensional flow along with the interaction of interface with the sur­
and average heat transfer both were nearly independent of the jet roundings are the associated difficulties in numerical modeling for the
inclination angle. Ma et al. [6] studied inclined, free surface high Prandtl free surface liquid jet.
number liquid (transformer oil) jet impingement on a vertical wall in the The present study report results in both the directions (longitudinal
free-surface jet configuration. Jet inclination angle was found to affect and transverse) and explains the physics relevant to heat transfer dis­
the peak value of heat transfer. They developed a correlation which tribution. Both experimental and numerical findings are presented in the
predicts location and magnitude of peak heat transfer. It was shown that paper. IR thermal camera is used to capture thermal images. Experi­
increasing the jet inclination angle increases symmetry as the jet mental Nusselt number contours are reported for Reynolds number of
approached orthogonal arrangement. Stevens and Webb [16] studied 17,541, 26,311, 35,081 and 43,852 and inclination angle of 45� , 30� ,
heat transfer characteristics of liquid jet impingement on the heater 15� and 0� (orthogonal case). A numerical study is performed for Rey­
surface in the free surface configuration. As found in air-jet studies, the nolds number of 17,541, 26,311 for all the four inclination angles. In the
location of peak heat transfer shifted towards the uphill side. Addi­ numerical study, the inclination angle was kept the same as was used in
tionally, the maximum shift in the location of peak heat transfer from the experimental work. First, general heat transfer characteristics are pre­
geometric center was limited to x=d ¼ 0:5, which was different from sented, followed by numerical results explaining the nature of heat
earlier submerged jet studies or air-jet studies. transfer characteristics. Presentation of numerical results are limited
The literature review presented above suggests that focus in inclined and only used to explain flow phenomena concerning heat transfer since
jet studies were to investigate the effect of inclination [4–7], peak Nusselt number values are overpredicted by CFD study. However, the
Nusselt number [4,9,10]and its location, also developing correlations to nature of the curves can be well presented by the numerical investiga­
predict location and magnitude of peak Nusselt number in terms of tion. CFD study is performed using SST SAS turbulence model,
inclination angle [8], jet to plate spacing [11] and Reynolds number. whereas, to capture the interface, Volume of Fluid (VOF) model is used.

2
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

2. Experimental facility circular pipe diameter.

Experimental facility to study inclined free surface water jet


2.1. Flow set up and heater design
impingement over the horizontal flat surface was designed and fabri­
cated at IIT Bombay, in steam power laboratory. The schematic of the
Schematic of flow setup and design of heater assembly is shown in
setup is shown in Fig. 1(a and b). A centrifugal pump (component 2 in
Fig. 1 (b). Four different holes, which support acrylic pipe firmly at
Fig. 1(a), Rating: 1 HP) is used to pump the water liquid from the
different angles (0� , 15� , 30� , 45� ) from vertical (normal upward di­
overhead tank (component 1 in Fig. 1(a)) to the circular pipe from which
rection from heater plate) were drilled on an acrylic block (6 cm width
the jet emerges. The liquid flow rate was controlled using a control valve
and 100 cm length and 20 mm thickness). Pipes were then inserted in
and measured using a rotameter. Rotameter was calibrated by catch and
these holes to provide the specific impingement angle arrangements.
time method using a stopwatch. Water was then passed through the
Acrylic block (component 6 in Fig. 1 (b)) is firmly clamped above the
circular acrylic pipe (6 mm diameter) and having a sufficiently long
heater assembly on the top of the chamber (component 7 in Fig. 1(a)
length (l=d � 60) to produce fully developed flow at pipe exit before
using clamps. Acrylic pipes at different angles through mounting as­
impingement. The heater assembly (component 6a in Fig. 1(a) is
sembly ensure repeatability since this setup eliminates the requirement
mounted in the chamber (component 7 in Fig. 1(a) which is open at the
of manually setting the angle and hence reduces the error in angle
top. Pipe exit positions and distance from the impingement surface could
measurements. The maximum error in angle measurement is 2� . A
be manually adjusted over the heater. The chamber was mounted on a
flexible pipe connects the outlet of rotameter to one of the acrylic pipes
stand with one side facing the jet, so that liquid impinges on the flat
from which the jet emerges. Spirit level was used to check the surface
surface (Fig. 1(b)). An outlet was provided at the bottom of the chamber
level of this block as well as the heater surface to ensure horizontal
to discharge water from this chamber through a connecting pipe. Dis­
placement.
charged water collected in a tank (component 11 in Fig. 1(a) was
Fig. 2 shows a schematic of the heater assembly. The heater foil was
brought to the laboratory tank (component 1 in Fig. 1(a) using an
made of 0.15 mm thick stainless steel foil (SS304, part 1 in Fig. 2) which
additional centrifugal pump (component 12 in Fig. 1(a), Rating: 1 HP)
has a width of 6 cm and length 10 cm. A ten millimeter foil strip
thereby completing the loop. Water passing through the bypass valve
(component 1 in Fig. 2) was soldered with the copper bus bars
(component 3 in Fig. 1(a) was collected in a small container, and the
(component 2a and 2b in Fig. 2) from both the sides to avoid contact
temperature was measured using the thermometer for each set of ex­
resistance. This heater foil and copper bus-bar assembly were fitted on
periments. In the present case, the temperature of water at the nozzle
the acrylic sheet (component 5 in Fig. 2), in which a rectangular cut
exit was found between 28o C to 29o C . This temperature (T∞ ) was then
section of 50 mm and 100 mm length is made to have a clear view of
considered as free stream temperature to calculate Nusselt number.
heater plate surface which faces IR thermal camera. RTV anabond sili­
Properties of water used to calculate Reynolds number were taken at
con is used to avoid any leakage. This assembly (component 6a in Fig. 1
28:5� C for both experimental analysis and numerical study. Reynolds
(b)) was placed inside the chamber (component 7 in Fig. 1 (b)). The
number was calculated using the expression Re ¼ ρvd= μ, where d is
backside of the foil was painted by high emissivity black paint to provide

Fig. 1. (a) Line diagram of experimental setup: 1 Laboratory water tank, 2 centrifugal pump, 3 bypass valve, 4 rotameter, 5, 5a, 5b circular pipes, 6 nozzle stand, 6a
SS304 foil (heater), 7 chamber, 8 discharge pipe, 9 thermal camera, 10 data collection PC, 11 discharge tank, 12 centrifugal pump (b) three dimensional view of flow
and heater foil arrangement.

3
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

experimental and numerical data. For a few numerical cases, co­


ordinates were transformed to present data in notations or axis, as
shown in Fig. 3(a and b). For both experimental and numerical work,
coordinates (x ¼ 0, y ¼ 0z ¼ 0) denotes geometric center. Jet to plate
spacing (H) is the distance between the jet exit to the impingement
surface in the nozzle axis direction, as shown in Fig. 3(a). The uphill side
is the region of negative x-direction at the impingement surface; simi­
larly, the downhill side is the positive x-axis (x � 0). Two primary di­
mensions are used in the context: longitudinal and transverse direction.
The longitudinal direction is x axis direction, whereas transverse di­
rection is the y axis direction. Both directions are parallel to the
impingement surface, as shown in Fig. 3(b).

2.1.1. Experimental calculations


Experiments were performed for jet impingement on a uniformly
heated SS 304 foil. The heat transfer coefficient was then calculated
using Eq. (1) through Eq. (4). Heat flux (q00 ) can be given by;
Fig. 2. Schematic of arrangement of heater assembly.
q〞 ¼ h△T (1)
uniform emissivity distribution for thermal images. A thermal camera
Power was the product of current passing through the SS foil and the
(Model: FLIR A600) was placed away from the backside of the plate to
voltage drop across the heater surface and is calculated by Eq. (2).
capture thermal images during water jet impingement. The copper bus
bars were connected to AC power supply (0 10 V, 0 500 A, 5 kVA P ¼ VI (2)
capacity transformer) for heating the SS304 foil by passing electric
Heat flux (q’’ ) is then calculated by dividing this power by area of the
current. This was done using connection holes 6a and 6b (Fig. 2) which
SS304 foil. △T (Eq. (3)) is the difference between the local temperature
connects copper bus bar to AC transformer. The current was indicated on
at the impingement surface and water temperature at jet exit.
the transformer display, and it was confirmed with tong type ammeter.
Voltage measurement was done using a multimeter probe, which was △T ¼ T T∞ (3)
connected to the heater plate for measuring the actual voltage drop
The Nusselt number given by
across the heater surface. Water jet impinges on the top side of the
heating plate, and thermal images were captured by FLIR thermal hd q’’d
Nu ¼ ¼ (4)
camera from the backside. The power supply to SS304 foil was set such kf kf △T
that the temperature difference (Ts T∞ ), where Ts is the lowest tem­
perature, was in the range of 9 C to 10 C for each set of experiment.
� �
2.1.2. Uncertainty and validation of setup
Fig. 3(a and b) shows notations and axes used to present the Uncertainties in Reynolds number and Nusselt number were evalu­
ated by the method proposed by Coleman [18]. Maximum uncertainty in
Reynolds number was found to be 2:10%, whereas maximum uncer­
tainty in Nusselt number has 6:8%.
To validate the experimental setup, stagnation Nusselt number of
present experimental work Nuo is compared with the correlations pro­
posed in literature for orthogonal free surface liquid jet impingement
[16,19–21]. Different correlations and their specifications are shown in
Table 1. From Fig. 4(a and b), it is clear that, stagnation Nusselt number
matches well with the literature data. Along with this, local Nusselt
number profiles with available data in literature [16,22] also compared
for vertical orthogonal free surface liquid jet impingement on the heated
surface. Similar nature and trends of curves confirm the suitability of
setup. Comparison of Nusselt number profile and stagnation Nusselt
number gives us enough confidence to extended this study further to
investigate inclined jet impingement on a flat surface. Also, it is shown
in the literature that jet to surface spacing has negligible influence on
stagnation Nusselt number [19,23] and Nusselt number profile for free

Table 1
Correlations in literature for stagnation Nusselt number for free surface liquid jet
impingement.
Correlation Reynolds Pr H/d
number

1 7700–23500 Pr � 5 H=d � 7
Nuo ¼ 0:938Re0:488 Pr3 [19]
1 3000–34000 Pr > 3 NA
Nuo ¼ 0:797Re0:5 Pr3 [20]
Nuo ¼ 0:7212Re0:5 Pr0:37 [21] NA 3 � Pr � 10 NA
Nuo ¼ 1:51Re0:44 Pr0:4 ðH=dÞ 0:11 4000–52000 NA H=d � 35
Fig. 3. Notations to present experimental and numerical data (a) Isometric [16]
view (b) top view of heater surface.

4
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

Fig. 4. Comparison of present experimental Nusselt number with experimental data in literature (a) Stagnation Nusselt number (Nuo ) (b) Nu= Nuo [16,24].

surface liquid jet impingement for H=d � 1 [17]. It was also verified in experimental data predicted by a different model (k ε, k ω, and v2 f)
the earlier numerical work [17], that only small jet to plate spacing, (H= for axisymmetric liquid jet impingement on a horizontal surface [17]. It
d < 1 cases) shown dependency on H=d, and stagnation Nusselt number was concluded that v2 f turbulence model predicted hydrodynamic and
was nearly same for H=d ¼ 1,1.5 and 2. Given this, jet to plate spacing is heat transfer characteristics better compared to other models with
fixed to H=d ¼ 4 for the present experimental work. respect to experimental data available in the literature. In the present
work, SST-SAS model is chosen as a primary model based on the com­
2.2. Numerical method parison of different models. VOF model is used to capture the interface
between water and air. Detailed governing equations for v2 f model, VOF
Schematic of the physical and computational domain is shown in method, wall treatment, choosing first layer thickness based on appro­
Fig. 5(a and b). Due to the nature of flow and orientation of the surface priate yþ value, and solution methodology were presented in detail in
and circular pipe, the flow would be symmetric. Thus, half of the ge­ work by the same authors [17].
ometry would be sufficient (with symmetry boundary condition) for the SST-SAS turbulence model is the addition to SST model by intro­
study of inclined circular free surface liquid jet impinging on a flat plate, ducing source term (Eq. (5)) in ω equation of SST model [25].
as shown in Fig. 5 (b). Appropriate boundary conditions are also indi­ � � �2 � � �
cated in Fig. 5 (b) at respective boundaries. The boundary on plane xy is L 2k 1 ∂ω ∂ω 1 ∂k ∂k
QSST SAS ¼ ρ � max ζ2 κ C � max 2 ; 2 ;0
specified as the symmetry boundary. Pressure outlet boundary condition LvK σφ ω ∂xi ∂xi k ∂xi ∂xi
was used at outer faces of the computational domain. Constant heat flux (5)
of 250 kW=m2 was used at the bottom surface along with no-slip wall Model constants are given as: ζ2 ¼ 3:51, σφ ¼ 2=3, C ¼ 2. Where, L
boundary condition. Fig. 5 (d) shows a possible distribution of jet and is the length scale, given by Eq. (6)
liquid film on the surface at symmetry plane. Here, the direction of
pffiffi . 1=4 �
gravity is specified as normal downward from the bottom wall. In ex­ L¼ k cμ � ω cμ ¼ 0:09 (6)
periments, circular pipes which were used to produce jet were of suffi­
cient length (l � 60d) to produce fully developed flow at pipe exit for all LvK is Von Karman length scale (Eq. (7)):
the cases. In the simulation, flow in a circular pipe (L ¼ 40d) has been pffiffiffiffiffiffiffiffiffiffiffi
� �
κS 1 ∂Ui ∂Uj
simulated separately. It was assumed that no additional effort had been LvK ¼ rffiffiffiffiffiffiffiffiffiffiffiffiffiffi; S ¼ 2Sij Sij ; Sij ¼ þ κ ¼ 0:41 (7)
2 ∂xj ∂xi
made to induce any turbulence in the pipe, except turbulence generated ∂2 Ui ∂2 Ui
∂x2k ∂x2j
due to flow within the pipe, hence minimal value 0:05% of turbulence
intensity is used at pipe inlet and resulting fully developed velocity and Closer to the wall, LvK is given by Eq. (8)
turbulence profiles are used at inlet of present computational domain.
∂U=∂y
Velocity at pipe inlet has been specified based on Reynolds number LvK ¼ κ � (8)
calculated using circular diameter and properties of liquid water at 28:5� ∂2 U ∂y2
C (Table 2), and resulting conditions (velocity and turbulence kinetic
energy) were superimposed at present domain inlet boundary. For the 2.3. Solution method
case of Re ¼ 26311 these profiles are shown in Fig. 5 (e). Dimensions of
the computational domain were chosen based on nozzle exit diameter. ANSYS Fluent CFD tool is used for the numerical study of the free
Extension of domain represented by primary dimensions p= d, q= d, w= d surface liquid jet impingement on the flat surface. A transient, pressure-
and H=d respectively, which are 3, 3, 2.33 and 3 respectively, where based, 3D numerical solver is used. Gravity is specified normal to heater
d (2r) was kept constant (6 mm). Inclination angle (θ) is varied appro­ surface in the downward direction (g ¼ 9:81 m=s2 ). For pressure-
priately for parametric variations. velocity coupling, SIMPLE algorithm is adopted. For discretization, the
v2 f turbulence model was chosen in our published work based on the second-order upwind scheme is chosen for turbulent kinetic energy,
comparison of Nusselt number prediction and velocity profile with volume fraction, and turbulent dissipation rate. Bounded central

5
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

Fig. 5. (a) Schematic of physical phenomena and (b) computational domain and boundary conditions (c) dimensions of computational domain (d) inclination angle
and gravity (e) velocity and turbulence kinetic energy profiles at nozzle exit (Re ¼ 26311). Note: not to scale.

two different domains are tested to investigate domain dependency. The


Table 2
details of different domain sizes are shown in Table 3. Pressure outlet
Properties of air and water used in the computational study.
boundaries are sufficiently far away from the jet impact region in
Fluid ρðkg =m3 Þ μðkg =m ⋅sÞ cp ðJ =kg ⋅KÞ kðW =m ⋅KÞ Pr Domain II as compared to Domain I. Highest Reynolds number (Re ¼
Water 995.62 0.000803 4070.8 0.62 5.3
26311), and largest inclination angle (θ ¼ 45) are chosen to test the
Air 1.1707 1:86 � 10 5 1006.4 0.026230 0.71402 domain dependency and grid dependency to get the confidence in nu­
merical work.
Fig. 6 (a) shows Nusselt number distribution along the symmetry line
differencing for momentum. The transient formulation uses bounded for both the domains. The region in uphill side at which both the do­
second-order implicit scheme. Convergence has been monitored by mains show different results is encircled for clarity. This is the region
observing the residuals. Residuals is chosen as 10 6 for momentum, closest to pressure outlet boundary for Domain I. Our interest was to
continuity, k, ω, and energy equations. It was observed that these re­ understand flow physics near the impact region, and as clear at all other
sidual criteria had been satisfied as the solution approaches a steady locations, Nusselt number curves for both the cases asymptotically
state. Initially, the time step was chosen as 10 6 sec. This reduced to merged. The reason why the domain does not affect the results is the
10 3 sec as numerical simulation approached near steady-state condi­ dynamic viscosity difference between air and water. In this case, water
tion. Pressure and temperature at the stagnation point were observed
during the iterations. When these variables approached nearly constant Table 3
value, solution at that state was considered as the steady-state solution. Dimensions used in domain dependency test.
p=d q=d h=d w=d
2.3.1. Domain and grid independence study
The size of the computational domain is shown to have a strong in­ Domain I 3 3 2.33 3
fluence on air jet impingement studies [26–28]. In the present study, Domain II 6 5.6 4.12 6

6
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

Fig. 6. (a) Domain dependency test (b) grid dependency test.

interface can affect air significantly, but the effect of airflow near the
vicinity of the interface in water liquid film is not significant, which
could affect hydrodynamics and heat transfer aspect of free surface jet
impingement. Results of domain dependency give us the confidence to
continue further study with Domain I to study free-surface liquid jet
impingement for inclined jet cases.
Different element sizes were chosen to perform a grid independence
test for numerical simulations. Mesh was refined near the wall and in the
region of liquid jet flow. The near-wall mesh was created based on an
appropriate yþ criteria, as discussed in the earlier work [17]. Three
different grids of element size of 80 μm (Grid I) and 160 μm (Grid II) and
320 μm (Grid III) were tested for grid dependency. These element sizes
specified were corresponding to near the wall and in the higher gradient
region of the computational domain. Nusselt number profiles along the
line of symmetry are shown in Fig. 6 for all three grid cases. Nusselt
number profiles are nearly the same for all the grids. Nusselt number
profile predicted by grid II and grid III collapse on each other. Nusselt
number profiles predicted by the grid I show the relatively lower value
of Nusselt number, especially near the peak Nusselt number locations
and in the downhill side. However, the difference in Nusselt number is
Fig. 7. Comparison of local Nusselt number obtained by two different turbu­
less than 4% at any location compared to the largest element size tested. lence model and experiments.
The reason for these differences is due to the nature of SAS model, which
is independent of grid resolution; this means that this model can capture
turbulence structures corresponding to smaller-scale if sufficient grid
hence, further numerical investigation is performed using the SST-SAS
resolution is provided (refined grids). In the absence of grid resolution,
model.
this model behaves as the SST model, which can give a better prediction
of average quantities based on URANS formulation. In present case, grid
3. Results and discussion
with 80 μm element size takes large time for computations (nearly 15�
24 hours). Given this, the grid size of 160 μm (Grid II) is chosen for
A total of sixteen (16) cases were studied in the present experimental
further simulations.
work. Details of all the cases are shown in Table 4. For each inclination,
Results of two different turbulence models are compared to choose
the best-suited model for the present study. Fig. 7 compares Nusselt
number along the line of symmetry employing v2 f, SST-SAS model,
along with the experimental Nusselt number. Results are presented for
Table 4
the inclination angle of 0 and 45� . Significantly different values of
Parametric variations in experimental study.
Nusselt number are observed near the impact for both the turbulence
models. Both models over-predicted peak Nusselt number as compared Cases θ d (mm), H=d Re

to experimental cases. Error for v2 f model for orthogonal and inclined 1 4 0� 6, 4 17,541, 26,311, 35,081. 43,852
cases are 83% and 136% respectively, for SAS model these values are 5 8 15� 6, 4 17,541, 26,311, 35,081. 43,852
54% and 62% respectively. SST-SAS model predicts the trend of Nusselt 9 12 30� 6, 4 17,541, 26,311, 35,081. 43,852
number and values closer to experimental cases. The objective of the 13 16 45� 6, 4 17,541, 26,311, 35,081. 43,852
present numerical work is to understand the physics of heat transfer;

7
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

four different Reynolds numbers were studied, as shown in the table. For the downhill side 3 � x=d � 4:5.
all the cases, jet to plate spacing (H=d ¼ 4) was kept constant. The focus Fig. 8 (c) shows Nusselt number distribution along the lateral di­
of the parametric study was to investigate the effect of inclination angle rection (y=d) at different longitudinal locations (x=d ¼ 1, 0, 1). As
and Reynolds number on heat transfer distribution at the impingement expected, the Nusselt number curves are nearly symmetric at all longi­
surface. tudinal locations (x=d ¼ 1; 0; 1). Nusselt number curve towards the
uphill side (x=d ¼ 1) is sharp and has higher magnitude of Nusselt
3.1. Nusselt number distribution at the impingement surface number near the impact zone (y=d ¼ 0) as compared to Nusselt number
curve at x=d ¼ 0 and x=d ¼ 1. All the Nusselt number curves at different
Fig. 8(a–c) shows local Nusselt number obtained experimentally. x=d locations 1; 0; 1 merge away from the impact y=d � 2 and y=d �
Fig. 8 (b) shows Nusselt number contours for the case of 30o inclination 2.
(Re ¼ 26311, H=d ¼ 4). Origin (0, 0) in the graph represents geometric Fig. 9(a–d) shows experimentally obtained Nusselt number distri­
center. It is seen that nature of Nusselt number distribution is asym­ bution at the impingement surface for different inclination cases for
metric which is in contrast to orthogonal jet impingement configuration, Reynolds number of 17,541. Nusselt number contours are nearly sym­
where, Nusselt number distribution is generally symmetric. For inclined metric for orthogonal jet configuration (θ ¼ 0� ). As inclination angle
jet, the region covered by of contour lines becomes nearly semicircular increases, asymmetrical Nusselt number distribution is observed. For
near the impact. The region of higher Nusselt number shifts towards the highest inclination case θ ¼ 45� , contours become nearly semicircular
uphill direction of the heater surface. Local Nusselt number profiles near the point of impact (innermost contour line) as clear from the 9 (d).
along different locations passing through the plate are shown in Fig. 8 The magnitude of peak Nusselt number first decreases slightly with
(a) and (c). Fig. 8 (a) shows Nusselt number profiles along the longitu­ inclination (θ ¼ 0� to θ ¼ 15� ), then it decreases with further increase in
dinal direction (x=d) at different lateral locations (y= d ¼ 0, 1, 2). For y= inclination angle (θ ¼ 15� to θ ¼ 45� ). Fig. 10(a–d) shows local Nusselt
d ¼ 0, location of peak Nusselt number is shifted towards the uphill side number contours for highest Reynolds number case studied (Re ¼
of the heater surface. Nusselt number decreases along both the di­ 43852) and different inclination angles as indicated in figure. Near the
rections (uphill and downhill), but decrease is more sharp towards the impact, similar nature and dependency are observed, as was observed
uphill side as compared to the downhill side. Towards the downhill di­ for lower Reynolds number. In this case, contours are not smooth, as was
rection, Nusselt number nearly attains a constant value in the region observed for lower Reynolds number owing to instability at such higher
3:5 � x=d � 6. There is a slight increase beyond this location (x= d � 6). Reynolds number (Vmean ¼ 5:9 m/s). Interesting feature for this case is
Nusselt number curves at other locations (y=d ¼ 1; 2) show that location downstream heat transfer (3 � x=d � 10), Nusselt number is highly
of peak Nusselt number shifts towards the geometric center (x= d ¼ 0) sensitive to inclination angle for Re ¼ 43852, whereas for Re ¼ 17541
from uphill side, and as we go along the traverse direction (y= d). All the Nusselt number is nearly independent of inclination angle in downhill
Nusselt number curves, (y=d ¼ 0; 1; 2) asymptotically merge towards region.

Fig. 8. (a) Local Nusselt number profiles at different transverse locations (y=d) (b) Nusselt number contour at the impingement surface (c) Nusselt number profile at
different longitudinal location (x=d).

8
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

Fig. 9. Experimental Nusselt number contours for different inclination angles (Re ¼ 17541).

Fig. 10. Experimental Nusselt number contours for different inclination angles (Re ¼ 43852).

3.2. Heat transfer characteristics along longitudinal direction plateau. Then it increases gradually attains a local maximum value
(x=d � 5 for Re ¼ 43852) and then decreases in the further downhill
Fig. 11(a and b) show experimentally obtained local Nusselt number direction. The location corresponding to the second maximum value of
and its dependency on the Reynolds number for inclination angle of Nusselt number shifts towards the stagnation point with increasing
0� and 45� respectively along the line of symmetry. For orthogonal case, Reynolds number. As shown symbolically by an inclined dotted line (B)
peak Nusselt number coincides with the geometric center. Nusselt in Fig. 11 (a). A similar trend is observed for the location corresponding
number decreases sharply following the stagnation point, attains to the local minimum value (line A). Interestingly, the distance between

9
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

Fig. 11. Experimental Nusselt number profiles along the symmetry line (y=d ¼ 0) for different Reynolds number (a) θ ¼ 0� (b) θ ¼ 45� .

both the points is nearly constant for all the Reynolds numbers studied. However, the difference in Nusselt number reduces further downstream
Nusselt number is found to be highly sensitive to Reynolds number away (x=d � 8).
from the impact mainly in the transition region as compared to near the For inclined case (11 (b)), sharp changes in Nusselt number curves
stagnation point. For example, at the stagnation point, the difference are observed near the point of impact for all the Reynolds number cases.
between Nusselt number for lowest and highest Reynolds number case is Nusselt number decreases sharply in the uphill side; whereas, the
around 55%; whereas, the difference is around 177% at x= d ¼ 5. decrease is found to be relatively gradual in downhill direction till

Fig. 12. Experimental Nusselt number profile along the symmetry line (y=d ¼ 0) for different inclination angle (a)Re ¼ 17541 (b) Re ¼ 26311 (c) Re ¼ 35081 (d)
Re ¼ 43852.

10
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

x=d � 1:8. Further downhill, higher Reynolds number (Re ¼ 35081,


43,852) shows local minimum values near x=d � 2, followed by increase
and then decrease in further downstream. Intermediate Reynolds num­
ber case (26,311) attains a nearly constant Nusselt number in downhill
direction after x=d ¼ 4. Low Reynolds number case (Re ¼ 17541) does
not show a significant change in the local Nusselt number curve in the
downhill side. Similar to the orthogonal jet case, similar inclined lines
are drawn to distinguish local minimum and maximum (line A and line B
in Fig. 11 (b)). The distance between both the lines is relatively higher,
as was observed for the orthogonal jet case. Similar to the orthogonal jet
case, heat transfer for the inclined case is also found to sensitive to
Reynolds number as compared to the orthogonal case in the downhill
direction, which can be clearly seen from Fig. 11 (b). Here, the differ­
ence between the local Nusselt number is found to be significantly
higher at x=d ¼ 5 as compared to orthogonal jet case. The location of
peak Nusselt number is found to be weakly dependent on Reynolds
number, and it shifts slightly towards the uphill side with increasing
Reynolds number.
Fig. 13. Average Nusselt number at the impingement surface for different
3.3. Nusselt number dependency on inclination angle inclination angle and Reynolds numbers.

Fig. 12(a–d) show the experimental Nusselt number along the line of is used in origin to calculate the average Nusselt number.
symmetry on the impingement surface for different inclination angles Z
1
for different Reynolds number. With increasing inclination, the location Nu ¼ NudA (9)
corresponding to the peak Nusselt number is shifted to uphill direction A
A
for all the Reynolds number cases. For Re ¼ 17541 (Fig. 12 (a)), peak
Nusselt number first decreases slightly as inclination increases from where, A is the area considered to calculate average Nusselt number,
0� to 15� , with further increase in inclination angle, peak Nusselt which is the area between 3d to 3d in y direction and 3d to 9d in x
number increases and highest value to peak Nusselt number observed direction. Fig. 13 shows the average Nusselt number at the impingement
for largest inclination angle studied (45). A similar trend of peak Nusselt surface for different cases studied. For higher Reynolds number cases
number is observed for higher Reynolds number, as shown in Fig. 12 (Re � 26511) highest inclination case shows a higher average Nusselt
(b–d). However, difference between peak Nusselt number for orthogonal number as compared to other inclination angle cases. The difference
case (θ ¼ 0� ) and inclined case (θ ¼ 45� ) increases with increase in between average Nusselt number for 45� case and orthogonal (0) in­
Reynolds number. For all the Reynolds numbers studied, Nusselt num­ creases with increasing Reynolds number. However, the maximum dif­
ber decreases sharply towards the uphill side with increasing inclina­ ference is below 12% for both cases. The general trend of the average
tion. Higher value of Nusselt number is observed for orthogonal case in Nusselt number for other cases could not be confirmed since the dif­
uphill side away from the impact near x=d ¼ 3 for Reynolds number of ference in Nusselt number is lower. The main reason for the higher value
35,081 and 43,852 (Fig. 12(c and d)). For lower Reynolds number, all of average Nusselt number for the case of a higher inclination angle is
the curves asymptotically merge towards the uphill side near x= d ¼ the contribution due to the heat transfer on the downhill side, especially
3, as seen from Fig. 12(a and b). For all the Reynolds number cases, in the region beyond x=d ¼ 2. It must be noted, only for a small region
Nusselt number near the geometric center (0 � x=d � 1:2) decreases near x=d ¼ 0, the values of Nusselt number are higher for the 45 incli­
with an increasing inclination angle. In downhill side (x= d � 3) Nusselt nation case as compared to the orthogonal case.
number for lower Reynolds number is found to be nearly independent of The main contribution for higher average Nusselt number for higher
inclination angle, whereas for higher Reynolds number (Re � 26311) inclination angle is downhill heat transfer, since near the impact only a
inclination significantly affect heat transfer and Nusselt increases with small region showed higher Nusselt number for inclined case (θ ¼ 45 )

increasing inclination angle. Vertical straight lines are drawn in Fig. 12 as compared to orthogonal case (θ ¼ 0� ).
(b–d) to distinguish these characteristics. For Re ¼ 43852, Nusselt
number is nearly 16% higher at x=d ¼ 5 for θ ¼ 45 as compared to 3.5. Comparison of experimental and numerical results
orthogonal case, whereas the difference is 48% in further downhill at x=
d ¼ 9. For orthogonal jet case, momentum distribution is symmetric in To understand the physics behind the nature of Nusselt number
all directions following the impact. For inclined jet, net momentum is distribution, a numerical study is carried out. Using the turbulence
higher in the downhill direction. For higher Reynolds number case, free model in numerical study imposes a condition to satisfy yþ criteria to
stream higher momentum, especially in downhill, could enhance tur­ resolve boundary layer flow. High Reynolds number cases (Re � 26311)
bulence fluctuations and enhance the heat transfer. require the first layer of grid to be smaller than 5 μm size near the wall.
This order of grid size results in a high aspect ratio of computation cells,
3.4. Average Nusselt number which are avoided. Given this, the upper limit of the Reynolds number
was 26,311 for the CFD study. However, there is no restriction for the
As seen from the aforementioned discussion that if the inclination lower Reynolds number; hence computations were performed for Rey­
angle increases, the Nusselt number curves become sharp near the nolds number of 10,000, 17,541, and 26,311 considering 6 mm circular
impact. Higher Nusselt number in a narrow region for inclined jet near pipe diameter. In the numerical study, two Reynolds numbers and four
the impact can result in a lower overall Nusselt number. To calculate inclination angle values exactly match with the experimental condition,
average Nusselt number, an integral numerical approach is adopted over which is a total of eight cases are identical to experiments.
the impingement surface. Since local Nusselt number was available in Fig. 14(a and b) compares experimental and numerical Nusselt
matrix form corresponding to different locations over the impingement number profiles. Nusselt number profiles along the symmetry line for
surface. OriginPro 2018 (software) inbuilt function is used to calculate four different inclination angles and two different Reynolds numbers.
the average Nusselt number. The following form of expression (Eq. (9)) The general trend of Nusselt number profiles obtained numerically

11
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

matches well with the experimental cases for all inclination angles. The phenomena.
values of peak Nusselt number in the experimental and numerical study
first decrease initially with an increasing inclination angle, and then it 3.6. Hydrodynamics of jet impingement
increases with a further increase in inclination angle.
Fig. 14(c and d) compares local Nusselt number profiles for incli­ As stated, present numerical work aims to explain the flow phe­
nation angle of 30� considering Reynolds number of 17,541 and 26,311 nomena concerning Nusselt number distribution at the impingement
respectively. Experimental and numerical Nusselt number profiles are surface. First, velocity streamlines and its relation to pressure and heat
shown at different traverse locations (y=d) and compared. Nusselt transfer are discussed. Then, velocity variation in the liquid film is
number profiles are found to be similar in the experimental and nu­ presented, followed by turbulence intensity distribution. The coupled
merical study along with different locations at the impingement surface. effect of mean velvety and turbulence on the peak value of Nusselt
Peak value of Nusselt number in experiential and numerical study differs number and Nusselt number profiles, especially along the symmetry line
by maximum of 62% for highest inclination angle (θ ¼ 45) and highest is discussed.
Reynolds number (Re ¼ 26311) as shown in Fig. 14 (b). Overall com­ Fig. 15(a) and (b) present streamlines in computational domain for
parison of experimental and numeral results reveals that Nusselt number Re ¼ 17541 (H=d ¼ 4) considering orthogonal case (θ ¼ 0� ), and in­
trends and profiles along with dependency on inclination can be well clined case (θ ¼ 45� ) respectively. Solid black lines in Fig. 15 (b) present
explained by numerical results, especially near the point of impact. location of dividing streamline. When jet comes close to the wall, the
There is enough confidence to use numerical results to understand the point of stagnation depends on the orientation of the surface and nozzle.
hydrodynamics of inclined jet and underlying physics of heat transfer For inclined case, jet streams towards the uphill side are affected by the

Fig. 14. Comparison of experimental and numerical Nusselt number along the symmetry line for different inclination angle (a) Re ¼ 17541 (b) Re ¼ 26511.
Comparison of local Nusselt number at different transverse locations (Z=d) for inclined jet (θ ¼ 30) (c) Re ¼ 17541 (d) Re ¼ 26511.

12
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

Fig. 15. Streamlines for Re ¼ 17,541, (a) θ ¼ 0 (b) (a) θ ¼ 45 (c) symbolic representation of velocity near jet exit (d) pressure variation along the symmetry line for
� �

different inclination angels.

plate earlier, and hence stagnation point shifts towards the uphill side of velocity magnitude starts to decrease radially; for this case, the contour
the impingement surface. In Fig. 15(a and b), G.C represents the geo­ is symmetric. For inclined case, as shown in Fig. 16 (b), there is a shift in
metric center, whereas an arrow in Fig. 15 (b) represents the stagnation momentum towards the downhill side due to the orientation of the
point. For orthogonal case, dividing streamline corresponds to the cen­ nozzle. This results in the higher magnitude of the velocity and hence
tral axis of the circular nozzle, whereas, for the inclined case, dividing higher momentum in the downhill liquid film near the impact as is clear
streamlines shifts towards the wall of the circular pipe. The symbolic from the contours. It should be noted that momentum along the trans­
representation of velocities at the nozzle exit, corresponding to dividing verse direction also decreases with the increasing inclination angle.
streamlines is shown in Fig. 15 (c). For the orthogonal case, dividing As discussed, velocity near to dividing streamline for the inclined jet
streamlines belong to the maximum velocity of jet exit, whereas, for the has lower velocity streams as compared to the case of the orthogonal jet.
inclined jet, dividing streamline belongs to a relatively lower jet exit Given this, it is hypothesized that higher velocity liquid particles that are
velocity point, as can be seen from Fig. 15 (b). For orthogonal case, flowing with dividing streamlines would affect the boundary layer
stagnation point (zero velocity) coincides with the geometric center, thickness inside the film, and it was expected that orthogonal case would
whereas, for the inclined jet (θ ¼ 45� ), stagnation point shifts towards result in lower boundary layer thickness. In order to understand this,
the uphill direction of the geometric center. velocity contours for orthogonal and inclined cases close to the
Pressure distribution near the impact is dominated by dividing impingement surface (z=d ¼ 0:00833) are presented in Fig. 17(a and b).
streamline and corresponding velocity at jet exit. Fig. 15 (d) shows Velocity along both the directions of the geometric center for the
pressure profiles along the symmetry line for Reynolds number of orthogonal case is found to be higher as compared inclined case. These
17,541 (d ¼ 6 mm, H=d ¼ 4), and different inclination angles. Stagna­ differences in mean velocity affect the velocity gradient at the wall and
tion pressure for the orthogonal case is found to be higher as compared hence, boundary layer thickness.
to the inclined cases. This is due to the fact that, for the orthogonal case, Fig. 17 (c) shows Nusselt number along the symmetry line for
the liquid particle that comes to stagnant condition belongs to the orthogonal and inclined jet (θ ¼ 45� ) case for Reynolds number of
highest velocity at the jet exit; whereas, the corresponding velocity is 17,541. For inclined jet case, relatively lower mean velocity following
lower for the inclined jet cases, as discussed. This causes the peak the stagnation point results in higher boundary layer thickness as
pressure to decreases with increasing inclination angle, as clear from compared to orthogonal jet case. This causes lower convective heat
Fig. 15 (d). transfer following the stagnation region, and hence Nusselt number
In order to understand momentum distribution in liquid film, ve­ drops sharply for the inclined jet case. In the uphill side, the mean ve­
locity magnitude inside the film above the impingement surface (z= d ¼ locity drops sharply, and this is seen to manifest as a sharp drop of
0:08333) is presented for 0� , and 45� inclination cases, as shown in Nusselt number. However, in the downhill side, the Nusselt number
Fig. 16(a and b). For orthogonal case (θ ¼ 0� ), jet strikes the surface, drop is gradual due to the gradual drop of mean velocity (momentum).
starts accelerating radially away from the stagnation region, which is Higher mean velocity near the wall results in a higher Nusselt number
reflected in the velocity contour, as shown in Fig. 16 (c). Afterward, the for the orthogonal case as compared to the inclined jet case along the

13
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

higher inclination still needs to be understood. In order to understand


this, turbulence intensity variation at the symmetry plane is presented.
Fig. 18(a–d) shows turbulence intensity distribution at the symmetry
plane considering different inclination angle for Re ¼ 17541. As the jet
comes closer to the plate, there is a loss of mean kinetic energy. Thus,
there is an increase in turbulence kinetic energy, which is clear from the
turbulence intensity contours. For inclined cases, just after impact, there
is a slightly higher turbulence level observed in the liquid film near the
impact towards the uphill direction in comparison to the downhill side,
as shown by the encircled region in Fig. 18(c and d). Following the
impact, liquid streams in the uphill side change direction significantly
and flow towards the uphill direction. These sudden changes in the di­
rection of a liquid stream create turbulence fluctuation, and hence tur­
bulence intensity is slightly higher in these regions. In the downhill
direction, change in the direction of streamlines are smooth, and hence,
lower turbulence intensity is observed in the respective region.
It was not possible to infer the differences in the values of turbulence
intensity in the encircled regions in Fig. 18(c and d). Hence, for clarity
near the wall turbulence intensity curves are presented in Fig. 18 (e),
through the line passing parallel to the impingement surface (y=d ¼
0:0833) along the symmetry plane for different inclination angles. For
orthogonal case, turbulence intensity profile is found to be symmetric,
increases following the stagnation point attain peak value, decreases
following it. Inclined jet case (θ ¼ 30� , θ ¼ 45� ) dominates in terms of
turbulence fluctuations at the locations where liquid streams turns to­
wards the uphill direction near the impact. In these regions, higher
turbulence intensity is observed for a higher inclination angle, as shown
in the encircled region in Fig. 18 (d). Higher level of turbulence fluc­
tuations cause the sharp peak of Nusselt number for inclined jet cases
Fig. 16. Velocity magnitude contours at symmetry plane for Re ¼ 17541 (a) (θ ¼ 30 and θ ¼ 45 ).
� �

θ ¼ 0 (b) θ ¼ 45 . Velocity magnitude contours at the plane parallel to the wall The following two phenomena are discussed, (I) how the velocity of
� �

passing through y=d ¼ 0:08333 (c) θ ¼ 0� . the dividing streamlines decreases with increasing inclination angle,
which also caused lower pressure near the impact for inclined jet cases
larger portion of the symmetry line, especially after x= d � 0:7, A and (II), turbulence intensity, which is seen to increase with increases
vertical dotted line is drawn to differentiate this in Fig. 17 (c). inclination angles near the impact. The first phenomenon decreases the
This discussion explains the trend of Nusselt number following the heat transfer near the impact, and the second one contributes to
peak value for inclined jet case, but the higher peak value for the case of enhancing the heat transfer from the impingement surface. For

Fig. 17. Velocity magnitude contours at the plane parallel to the wall passing through y=d ¼ 0:008333 (a) θ ¼ 0 (b) θ ¼ 45� (c) Nusselt number profile along the

symmetry line for orthogonal (θ ¼ 0� ) and inclined jet (θ ¼ 45� ) case.

14
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

Fig. 18. Turbulence intensity contours at the symmetry plane for Re ¼ 17541 (a) θ ¼ 0 (b) θ ¼ 15 (c) θ ¼ 30 (d) θ ¼ 45 . (e) Near wall turbulence intensity profile
� � � �

for different inclination angles.

orthogonal jet, the mean velocity near the dividing streamlines domi­ the underlying physics of the phenomena. The values of Nusselt number
nates, which influences overall heat transfer. As inclination increases, are off by nearly 61% compared to experimental work. However, near
turbulence fluctuations dominate and affect the heat transfer. Coupled the impact, the Nusselt number was similar to the orthogonal jet case.
effect of velocity and turbulence intensity is the reason why peak Nusselt This implies for higher Reynolds number and higher inclination angle
number first decreases as inclination angle changes from 0� to 15� and cases, multiple jet impingement cooling should be explored with care­
then it increases further as inclinations changes from 15 to 45� as seen fully chosen jet location and angle. The position of nozzles should be

in experimental work. fixed in such a way that it enhances the heat transfer, especially on the
uphill side of other jet locations. Based on the analysis of experimental
4. Conclusions and numerical findings, the following conclusions have been drawn.

This study presents experimental and numerical investigation of in­ � Local Nusselt number decreases sharply towards the uphill side of the
clined free surface liquid jet impingement on a flat surface. IR ther­ jet, while this decrease was found to be relatively gradual toward the
mography was used for experiments to capture temperature downhill side. This is due to the momentum distribution (velocity) of
distribution. Four different inclinations (0� , 15� , 30� , 45� ) and four liquid film near the wall.
different Reynolds number were used for each inclination in experi­ � Peak Nusselt number first decreases with increasing inclination angle
mental study. It is shown that a higher inclination angle (� 30� ) and (0� –15� ) then it increases with further increasing inclination angle
higher Reynolds number (Re � 26511) results in a significantly higher up to 45� . It is seen that the coupled effect of turbulence and velocity
Nusselt number in the downhill side as compared to orthogonal jet. distribution causes this trend of peak Nusselt number.
Numerical work was done and presented only with the aim of explaining

15
K. Baghel et al. International Journal of Thermal Sciences 154 (2020) 106389

� Location corresponding to peak Nusselt number, peak pressure shifts [12] Tadhg S. O’Donovan, Darina B. Murray, Fluctuating fluid flow and heat transfer of
an obliquely impinging air jet, Int. J. Heat Mass Transf. 51 (25–26) (2008)
towards the uphill side with increasing inclination angle, as the
6169–6179, https://doi.org/10.1016/j.ijheatmasstransfer.2008.04.036. ISSN
dividing streamline shifts towards one side from the nozzle axis. 00179310.
� Peak pressure decrease with increasing inclination angle, and anal­ [13] S. Beltaos, Oblique impingement of circular turbulent jets, J. Hydraul. Res. 14 (1)
ogy is explained by the idea of dividing streamline velocity. (jan 1976) 17–36, https://doi.org/10.1080/00221687609499685. ISSN
00221686, http://www.tandfonline.com/doi/abs/10.1080/00221687609499685.
[14] Dahbia Benmouhoub, Amina Mataoui, Heat transfer control of an impinging
Appendix A. Supplementary data inclined slot jets on a moving wall, Heat Transf. Asian Res. 44 (6) (sep 2015)
568–584, https://doi.org/10.1002/htj.21138. ISSN 15231496.
[15] E.M. Sparrow, B.J. Lovell, Heat transfer characteristics of an obliquely impinging
Supplementary data to this article can be found online at https://doi. circular jet, J. Heat Transf. 102 (2) (may 2009) 202, https://doi.org/10.1115/
org/10.1016/j.ijthermalsci.2020.106389. 1.3244261. ISSN 00221481, http://heattransfer.asmedigitalcollection.asme.or
g/article.aspx?articleid¼1437157.
[16] J. Stevens, B.W. Webb, Local heat transfer coefficients under an axisymmetric,
References single-phase liquid jet, J. Heat Transf. 113 (1) (1991) 71–78, https://doi.org/
10.1115/1.2910554. ISSN 15288943, http://heattransfer.asmedigitalcollection.
[1] D.E. Metzger, R.S. Bunker, Local heat transfer in internally cooled turbine airfoil asme.org/article.aspx?articleid¼1440630.
leading edge regions: Part II-Impingement cooling with film coolant extraction, [17] Kuldeep Baghel, Arunkumar Sridharan, Janani Srree Murallidharan, Numerical
J. Turbomach. 112 (3) (1990) 459–466, https://doi.org/10.1115/1.2927681. ISSN study of free surface jet impingement on orthogonal surface, Int. J. Multiph. Flow
15288900, http://turbomachinery.asmedigitalcollection.asme.org/article.aspx? 113 (2019) 89–106, https://doi.org/10.1016/j.ijmultiphaseflow.2019.01.001.
articleid¼1463529. ISSN 03019322.
[2] Bradley W. Bartilson, Air Jet Impingement on Miniature Pin-Fin Heat Sinks for [18] Hugh W. Coleman, W. Glenn Steele, Experimentation, Validation, and Uncertainty
Cooling Electronic Components, jan 1992. https://patents.google.com/patent/ Analysis for Engineers, fourth ed., 2018, ISBN 9781119417989, https://doi.org/
US5083194 https://patents.google.com/patent/US5083194A/en. 10.1002/9781119417989.
[3] J. J€
org, Silvano Taraborrelli, Garikoitz Sarriegui, Rik W. De Doncker, [19] Feng Gao, Yongchang Chen, Jianbo Cai, Chongfang Ma, Experimental study of free-
Reinhold Kneer, Wilko Rohlfs, Johannes Jorg, Silvano Taraborrelli, surface jet impingement heat transfer with molten salt, Int. J. Heat Mass Transf.
Garikoitz Sarriegui, Rik W. De Doncker, Reinhold Kneer, Wilko Rohlfs, Direct 149 (2020) 119160, https://doi.org/10.1016/j.ijheatmasstransfer.2019.119160.
single impinging jet cooling of a MOSFET power electronic module, IEEE Trans. ISSN 00179310.
Power Electron. 33 (5) (may 2017) 4224–4237, https://doi.org/10.1109/ [20] X. Liu, J.H. Lienhard, J.S. Lombara, Convective heat transfer by impingement of
TPEL.2017.2720963. ISSN 08858993, http://ieeexplore.ieee.org/document circular liquid jets, J. Heat Transf. 113 (3) (1991) 571–582, https://doi.org/
/7962209/. 10.1115/1.2910604. ISSN 15288943.
[4] Xiaojun Yan, Nader Saniei, Heat transfer from an obliquely impinging circular air [21] C.F. Ma, Y.H. Zhao, T. Masuoka, T. Gomi, Analytical study on impingement heat
jet to a flat plate, Int. J. Heat Fluid Flow 18 (6) (1997) 591–599, https://doi.org/ transfer with single-phase free-surface circular liquid jets, J. Therm. Sci. 5 (4)
10.1016/S0142-727X(97)00051-9. ISSN 0142727X. (1996) 271–277, https://doi.org/10.1007/BF02653234. ISSN 10032169.
[5] Abhishek B. Bhagwat, Arunkumar Sridharan, Numerical simulation of oblique air [22] H. Sun, C.F. Ma, W. Nakayama, Local characteristics of convective heat transfer
jet impingement on a heated flat plate, J. Therm. Sci. Eng. Appl. 9 (1–10) (2016), from simulated microelectronic chips to impinging submerged round water jets,
https://doi.org/10.1115/1.4034913. ISSN 1948-5085. J. Electron. Packag. Trans. ASME 115 (1) (1993) 71–77, https://doi.org/10.1115/
[6] C.F. Ma, Q. Zheng, H. Sun, K. Wu, T. Gomi, B.W. Webb, Local characteristics of 1.2909304. ISSN 15289044.
impingement heat transfer with oblique round free-surface jets of large Prandtl [23] B.W. Webb, C.F. Ma, Single-Phase liquid jet impingement heat transfer, Adv. Heat
number liquid, Int. J. Heat Mass Transf. 40 (10) (jul 1997) 2249–2259, https://doi. Transf. 26 (C) (1995) 105–217, https://doi.org/10.1016/S0065-2717(08)70296-X.
org/10.1016/S0017-9310(96)00310-9. ISSN 00179310, https://www.sciencedirec ISSN 00652717, http://www.sciencedirect.com/science/article/pii/S0065271
t.com/science/article/pii/S0017931096003109. 70870296X http://linkinghub.elsevier.com/retrieve/pii/S006527170870296X.
[7] Abdlmonem H. Beitelmal, Michel A. Saad, Chandrakant D. Patel, The effect of [24] H. Sun, C.F. Ma, Y.C. Chen, Prandtl number dependence of impingement heat
inclination on the heat transfer between a flat surface and an impinging two- transfer with circular free-surface liquid jets, Int. J. Heat Mass Transf. 41 (10)
dimensional air jet 21 (2000) 156–163. (1998) 1360–1363, https://doi.org/10.1016/S0017-9310(97)00156-7. ISSN
[8] R.J. Goldstein, M.E. Franchett, Heat transfer from a flat surface to an oblique 00179310.
impinging jet, J. Heat Transf. 110 (February 1988) (1988) 84–90, https://doi.org/ [25] F.R. Menter, Y. Egorov, The scale-adaptive simulation method for unsteady
10.1115/1.3250477. ISSN 00221481, http://heattransfer.asmedigitalcollection. turbulent flow predictions. part 1: theory and model description, Flow, Turbul.
asme.org/data/Journals/JHTRAO/27504/84_1.pdf. Combust. 85 (1) (2010) 113–138, https://doi.org/10.1007/s10494-010-9264-5.
[9] Yahya Erkan Akansu, Mustafa Sarioglu, Kemal Kuvvet, Tahir Yavuz, Flow field and ISSN 13866184.
heat transfer characteristics in an oblique slot jet impinging on a flat plate, Int. [26] T. Cziesla, G. Biswas, H. Chattopadhyay, N.K. Mitra, Large-eddy simulation of flow
Commun. Heat Mass Transf. 35 (7) (2008) 873–880, https://doi.org/10.1016/j. and heat transfer in an impinging slot jet, Int. J. Heat Fluid Flow (2001), https://
icheatmasstransfer.2008.03.005. ISSN 07351933. doi.org/10.1016/S0142-727X(01)00105-9. ISSN 0142727X.
[10] Haydar Eren, Nevin Celik, Cooling of a heated flat plate by an obliquely impinging [27] F. Beaubert, S. Viazzo, Large eddy simulations of plane turbulent impinging jets at
slot jet, Int. Commun. Heat Mass Transf. 33 (3) (2006) 372–380, https://doi.org/ moderate Reynolds numbers, Int. J. Heat Fluid Flow (2003), https://doi.org/
10.1016/j.icheatmasstransfer.2005.10.009. ISSN 07351933. 10.1016/S0142-727X(03)00045-6. ISSN 0142727X.
[11] Kyosung Choo, Tae Yeob Kang, Sung Jin Kim, The effect of inclination on [28] A. Shukla, A. Dewan, Flow and thermal characteristics of jet impingement:
impinging jets at small nozzle-to-plate spacing, Int. J. Heat Mass Transf. 55 (13–14) comprehensive review, Int. J. Heat Technol. 35 (1) (2017) 153–166, https://doi.
(2012) 3327–3334, https://doi.org/10.1016/j.ijheatmasstransfer.2012.02.062. org/10.18280/ijht.350121. ISSN 03928764, http://iieta.org/sites/default/files/
ISSN 00179310. Journals/IJHT/35.1_21.pdf.

16

You might also like