Você está na página 1de 25

See

discussions, stats, and author profiles for this publication at:


https://www.researchgate.net/publication/305278980

Round impinging jets with relatively large


stand-off distance

Article in Physics of Fluids July 2016


DOI: 10.1063/1.4955167

CITATIONS READS

0 13

4 authors, including:

Mehrdad Shademan Ram Balachandar


University of Toronto University of Windsor
25 PUBLICATIONS 53 CITATIONS 201 PUBLICATIONS 1,405 CITATIONS

SEE PROFILE SEE PROFILE

Ronald Barron
University of Windsor
133 PUBLICATIONS 726 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate, Available from: Mehrdad Shademan
letting you access and read them immediately. Retrieved on: 30 September 2016
Round impinging jets with relatively large stand-off distance
Mehrdad Shademan, Ram Balachandar, Vesselina Roussinova, and Ron Barron

Citation: Physics of Fluids 28, 075107 (2016); doi: 10.1063/1.4955167


View online: http://dx.doi.org/10.1063/1.4955167
View Table of Contents: http://scitation.aip.org/content/aip/journal/pof2/28/7?ver=pdfcov
Published by the AIP Publishing

Articles you may be interested in


Large-eddy simulation of cavitating nozzle flow and primary jet break-up
Phys. Fluids 27, 086101 (2015); 10.1063/1.4928701

Large eddy simulation of flow development and noise generation of free and swirling jets
Phys. Fluids 25, 126103 (2013); 10.1063/1.4833215

Role of coherent structures in supersonic impinging jetsa)


Phys. Fluids 25, 076101 (2013); 10.1063/1.4811401

Effects of passive control rings positioned in the shear layer and potential core of a turbulent round
jet
Phys. Fluids 24, 115103 (2012); 10.1063/1.4767535

On large streamwise structures in a wall jet flowing over a circular cylinder


Phys. Fluids 16, 2158 (2004); 10.1063/1.1703531

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
PHYSICS OF FLUIDS 28, 075107 (2016)

Round impinging jets with relatively large


stand-off distance
Mehrdad Shademan,1 Ram Balachandar,2,3,a) Vesselina Roussinova,2
and Ron Barron2,4
1
Institute for Aerospace Studies, University of Toronto, Toronto, M3H 5T6 Ontario, Canada
2
Department of Mechanical, Automotive & Materials Engineering, University of Windsor,
Windsor, N9B 3P4 Ontario, Canada
3
Department of Civil and Environmental Engineering, University of Windsor, Windsor,
N9B 3P4 Ontario, Canada
4
Department of Mathematics and Statistics, University of Windsor, Windsor,
N9B 3P4 Ontario, Canada
(Received 4 February 2016; accepted 22 June 2016; published online 13 July 2016)

Large eddy simulation and particle image velocimetry measurements have been
performed to evaluate the characteristics of a turbulent impinging jet with large
nozzle height-to-diameter ratio (H/D = 20). The Reynolds number considered is
approximately 28 000 based on the jet exit velocity and nozzle diameter. Mean
normalized centerline velocity in both the free jet and impingement regions and pres-
sure distribution over the plate obtained from simulations and experiments show good
agreement. The ring-like vortices generated due to the Kelvin-Helmholtz instabilities
at the exit of the nozzle merge, break down and transform into large scale structures
while traveling towards the impingement plate. A Strouhal number of 0.63 was found
for the vortices generated at the exit of the nozzle. However, this parameter is reduced
along the centerline towards the impingement zone. A characteristic frequency was
also determined for the large scale structures impinging on the plate. The expansion,
growth, tilt, and three-dimensionality of the impinging structures cause dislocation of
the impinging flow from the centerline, which is significantly larger when compared
with flows having small H/D ratios. Contrary to the behavior of impinging jets with
small stand-off distance, due to the loss of coherence, the large scale structures do
not result in significant secondary vortices in the wall jet region and consequently
less fluctuations were observed for wall shear stress. Published by AIP Publishing.
[http://dx.doi.org/10.1063/1.4955167]

I. INTRODUCTION
Impinging jets have many practical applications in cooling, heating, metal cutting, and indus-
trial cleaning. Such jets are characterized by the presence of a free jet region, a stagnating flow,
and a wall jet flow. The characteristics of round impinging jets strongly depend upon parameters
such as the Reynolds number, distance between the nozzle and the plate, nozzle geometry, and the
turbulence intensity introduced at the inlet to the domain.1
The potential core of the free jet is surrounded by a shear region that grows with axial distance
from the nozzle. In this region the development of Kelvin-Helmholtz instabilities results in the
formation of ring vortices. With increasing downstream distance, the ring vortices undergo transi-
tion and change into large eddies. According to the definition of Yule,2 an eddy may be described
as a vorticity containing region of fluid which can be identified as a moving coherent structure in
the flow. Therefore, following the suggestion of Yule,2 the term eddy is used to denote the ring

a) Author to whom correspondence should be addressed. Electronic mail: rambala@uwindsor.ca. Tel: (+1) 519-253-3000 ext.
3572.

1070-6631/2016/28(7)/075107/23/$30.00 28, 075107-1 Published by AIP Publishing.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-2 Shademan et al. Phys. Fluids 28, 075107 (2016)

vortices after transition. These eddies are significant features of the turbulent region of the jet and
may remain coherent for a substantial distance. However, features like three-dimensionality and
irregularity of the vorticity field restrict us in denoting them as a ring vortex. A coherent structure,
as explained by Hussain,3 is a part of the flow with a connected turbulent fluid mass with instan-
taneously phase-correlated vorticity over its spatial extent. Ring vortices, which change into large
eddies after transition, influence the flow field and cause pressure fluctuations on the plate.4 This
phenomenon causes an unsteady behavior in the radial distributions of the wall shear stress5 and
wall pressure due to the separation and reattachment of the flow. This unsteadiness also influences
the rate of the heat transfer from the plate.6
The flow parameters undergo significant changes while the jet travels towards the impingement
zone. The jet exit velocity remains constant inside the core part of the free jet region.7 Once the
core reaches its maximum penetration, a sharp decay in the jet centerline velocity occurs which is
caused by the radial expansion of the jet. In the impinging zone, the flow loses its axial velocity and
changes direction due to the presence of the plate. A wall jet is formed along the radial direction on
the impingement plate which is completely under the influence of the unsteady structures impinging
on the plate.
Impinging jets have been classified based on the nozzle height-to-diameter (H/D) ratio, where
H is nozzle height (distance from the impingement plate) and D is the nozzle diameter. Based on the
arrangement suggested by Beltaos and Rajaratnam,8,9 type I includes impinging jets with H/D < 6,
type II is for 6 < H/D < 8, and type III refers to impinging jets with H/D > 8.
In type I jets (H/D < 6), the core of the jet continues to be in a developing state as it reaches
the plate. Therefore, distinction of the core part and impingement region is difficult.10 Based on
the initial conditions at the nozzle exit, the ring vortices generated due to Kelvin-Helmholtz insta-
bilities may convert to eddies before reaching the plate and gain a three-dimensional behavior.11
This phenomenon significantly changes the flow field in the wall region and consequently affects
wall parameters. There are a number of experimental and numerical studies on impinging jets
with H/D < 6.4,6,12 On the numerical front, most studies discuss the flow features and difficulties
associated with simulating impinging jets. Reynolds Averaged Navier-Stokes (RANS) modeling of
impinging jets was a part of the 2nd ERCOFTAC-IAHR Workshop on Refined Flow Modelling.13
Craft et al.14 published their research using different turbulence models to analyze the heat transfer
in the impingement region of the jet. They observed that the results were not in good agreement
with experimental data15 and attributed this to the weakness associated with the eddy viscosity
stress-strain relationship in the turbulence models used. Large Eddy Simulation (LES) seems to
be a more appropriate choice for the unsteady analysis of impinging jets. This method demands
reasonably fine meshes at higher Reynolds numbers, but it is more flexible compared with direct
numerical simulation. Many of the jet studies reported in the literature deal with simple plane
jets at low Reynolds numbers. Examples include LES studies on plane impinging jets by Voke
and Gao16 at Re = 6500 and by Beaubert and Viazzo17 at Re = 3000 and 7500. Hadziabdic and
Hanjalic4 used LES to analyze a circular impinging jet at Re = 20 000 and H/D = 2. Their analysis
showed that due to the small distance between the nozzle and the plate, the generated vortices are
short-lived and undergo a faster stretching breakdown than in a free jet due to the radial deflection.
They also noticed that because of jet flapping, the stagnation point meanders in time around the jet
geometrical center. They concluded that the second peak in the Nusselt number observed along the
plate was due to the reattachment of the recirculation bubble and associated turbulence production.
Bogey et al.18 performed LES to evaluate the effects of moderate Reynolds numbers ranging from
2.5 104 to 2 105 on subsonic round impinging jets with highly perturbed nozzle-exit boundary
layers at Mach number M = 0.9. They observed that as the Reynolds number increases, the mixing
layer develops more slowly, with smaller integral length scales and lower level of velocity fluctu-
ations. They also noticed that large scale components existing in turbulent shear layer which were
characterized by Strouhal number of 0.013 remain dominant. Uddin et al.19 used LES to model the
impinging jets at two Reynolds numbers of 13 000 and 23 000 for H/D = 2 in order to extract the
reason for the second peak observed in the radial distribution of the Nusselt number profile. They
noticed that the flow acceleration in the developing region of the boundary layer is closely related to
the observed phenomenon in the flow.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-3 Shademan et al. Phys. Fluids 28, 075107 (2016)

There are also a number of unsteady analyses of impinging jets with small nozzle-to-diameter
ratios (H/D < 6). These analyses include investigations on the unsteady behavior of the wall shear
stress, pressure distribution, and separation and reattachment of flow along the wall jet zone.
Hall and Ewing6 carried out instantaneous and time-averaged measurements of the fluctuating
wall pressure for incompressible turbulent impinging jets with H/D of 2 and 4, at Re = 23 300.
They noticed that the pressure fluctuations in the stagnation region are more three-dimensional,
similar to the findings of Kataoka et al.20,21 for a jet with H/D = 6. El Hassan et al.5 carried
out time-resolved particle image velocimetry and polarographic measurements simultaneously to
investigate the influence of the large-scale vortical structures on the wall shear stress in a round
impinging jet at a Reynolds number of 1260 with H/D = 2. They found that the instantaneous wall
shear stress is strongly dependent on the vortex dynamics. As the literature shows, a wide range of
mean and unsteady studies have been carried out for different features of impinging jets with small
height-to-diameter ratios (type I).
In type II jets (6 < H/D < 8), the core of the jet is at its maximum penetration when it reaches
the plate. The distinction of the free jet part from the impingement zone is still a challenge in
this case. The RANS study of Shademan et al.22 on the effect of nozzle stand-off distance on the
flow parameters showed that for an impinging jet with H/D = 6, no decay occurs for the centerline
velocity up to a station close to the plate which confirms that the core of the jet develops to a
level between 5 < x/D < 6. A similar behavior was observed by Giralt et al.23 when they conducted
experiments for impinging jets with H/D ratios ranging from 3 to 25 at Re = 34 000 80 000. No
decay in jet centerline velocity was observed up to x/D = 6.67 for the case with H/D = 25. Basi-
cally, the main flow feature in this type of impinging jet is the initiation of jet velocity centerline
decay as soon as the core section of the jet ends.
In type III jets (H/D > 8), the potential core of the jet is clearly distinguishable, the ambient
flow fully penetrates the jet, the tip of the potential core is quite far from the impingement plate,
and the three distinct regions described earlier occur in this type of jet. The analysis of this type
of impinging jet is important due to their engineering application in jets issuing from hydraulic
structures, impinging on erodible surfaces and for industrial spraying.10 Due to the large dis-
tance between the nozzle and the plate, large eddies having a wide range of sizes and trajectories
are formed, the jet undergoes a transition and starts approaching a fully developed condition.2
The mechanism of transition to a fully developed condition has been analyzed by Yule,2 who
showed that the transition process starts when the ring vortices generated near the nozzle lose their
two-dimensional form while moving downstream, followed by vortex pairing and coalescence. The
resulting vortices become more three-dimensional and generate more entrainment of the ambient
flow. As these vortices travel downstream, entanglement occurs between the vortices which have
different phases. This process causes the vortices to change into large eddies and consequently lose
their interconnected shape. Ball et al.24 showed that the transition region refers to stations between
7 < x/D < 70. They also named the region near to intermediate field (NIF) for distances between
0 < x/D < 30 where upstream conditions can significantly influence mass and momentum transfer
in the flow. They concluded that the far field region located at x/D > 70 can be considered as the
fully developed or self-similar part of the flow.
As the literature shows, different numerical and experimental simulations have been conducted
to analyze the flow field with H/D > 8. On the experimental side, detailed unsteady analysis of flow
structures is limited. On the numerical side, most of the unsteady studies are either RANS based or
only cover small stand-off distances. The objective of the current study is to answer a number of
unresolved questions for impinging jets with large stand-off distance (H/D > 8). These include how
the pressure fluctuation and wall shear distribution vary with time; how much the jet axis meanders
around the center in the impingement region; how the flow structure changes in moving from the
nozzle towards the plate; and how the separation and reattachment occurs in the impinging zone.
In the first phase of the current study, particle image velocimetry (PIV) experiments were
carried out on an impinging jet with H/D = 20 at a Reynolds number of 28 000. Different flow
parameters including mean velocity and fluctuating velocity components were measured in the
entire domain. In the second phase, Large Eddy Simulation (LES) was performed for the same flow
conditions to capture the unsteady flow parameters. The PIV results were used as the benchmark

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-4 Shademan et al. Phys. Fluids 28, 075107 (2016)

to validate the LES analysis. The LES results were analyzed to address the questions raised in the
previous paragraph.

II. EXPERIMENTAL SETUP (PIV MEASUREMENT)


Impinging jet experiments were conducted in a water jet facility 2 m long, 1 m wide, and 0.7 m
deep, where a turbulent jet was generated using a round nozzle with an exit diameter D = 0.01 m
as shown in Fig. 1. The nozzle and the flow conditioning system have been described in detail in
previous free jet studies.7,25 A stable flow was produced by maintaining a constant head of 2 m in
the overhead tank. Water flow was controlled using a valve and monitored during the experiment
with a flowmeter. A sharp crested weir located at the downstream end of the tank helped maintain
the depth of the water in the tank at 0.6 m while being re-circulated in the system.
To create an impinging jet, an additional acrylic plate was installed inside the jet facility at
H/D = 20 as shown in Fig. 1. The nozzle exit was mounted flush with the inner tank wall. The jet
exit Reynolds number (Re = UjD/) based on the velocity at the nozzle exit (Uj) and the nozzle
diameter (D) was 28 000.
The velocity measurements were carried out using a two-dimensional PIV system. To illumi-
nate the flow a laser sheet was generated using a 50 mJ/pulse Nd:YAG laser. An arrangement of
spherical and cylindrical lenses produced a 1 mm thick laser sheet at the measurement plane. The
laser sheet was aligned along the (xz) plane and it was ensured that it passed through the centerline
axis of the jet. The centerline of the jet is defined as a straight line emanating from the axis of
the nozzle, perpendicular to the nozzle exit plane, and positive along the flow direction. Since it
was desired to capture flow near the nozzle at x/D = 0 (Zone 1), in the free jet region at x/D = 7
(Zone 2) and near the impinging plate at x/D = 13 (Zone 3), PIV measurements were conducted
separately in each of these regions to adequately resolve their flow characteristics. The resulting
horizontal (xz) fields-of-view (FOV) at the jet centerline were approximately 50 50 mm2 for
Zone 1 and 71.5 71.5 mm2 for Zones 2 and 3. The spatial resolution for Zones 2 and 3 was
slightly higher due to the need to cover more of the jet since the jet expands as it travels away from
the nozzle. Each of the flow zones required a specific time delay between the laser pulses, with
the near jet nozzle data acquired at t = 50 s, the free jet region acquired at t = 150 s and the
impinging plate data at t = 360 s. The latter two zones required a larger t in order to resolve
the slower velocity field due to the jet decay. The tap water used in the jet facility was filtered

FIG. 1. Impinging jet experiment setup.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-5 Shademan et al. Phys. Fluids 28, 075107 (2016)

through a 1 m filter and the flow was seeded with silver coated hollow glass spheres of average
diameter 12 m and specific gravity 1.13. These particles have lower Stokes settling velocity and
their particle response time is suitable to follow the flow.7,25
A TSI Powerview Plus 4MP CCD camera was synchronized with the laser pulse to capture
illuminated seeding particles at the 50 50 mm2 region of the jet. The camera has a resolution
of 2048 2048 pixels which results in each pixel representing 0.024 mm of the jet. In order to
provide adequate statistics, 2000 pairs of images were captured in Zones 1 and 2, and 3000 pairs of
images were captured in Zone 3. Two pairs of successive images were captured with a frequency of
1.0 Hz. All raw images were processed using a recursive, two-frame cross correlation methodology
using TSI Insight 3G software package. The interrogation process for every image pair was done
in two steps. The first step utilized 64 64 pixels interrogation spots for all zones. The tentative
displacement data from the initial step was used in the second pass interrogation to locally offset
the individual 32 32 pixels interrogation windows. The offsetting was carried out to enhance the
signal-to-noise ratio by diminishing the count of random pairing of images from various particles.
In all cases, a 50% overlap of the first frame interrogation windows was employed to satisfy the
Nyquist criterion. This provides velocity data on a 16 16 pixel grid, which results in velocity
vectors every 0.39 mm for Zone 1 and 0.56 mm for Zones 2 and 3 over the entire FOV. A Gaussian
curve fitting method was used to evaluate the correlation peak to sub-pixel accuracy. In order to
reduce the peak locking effect, the ratio of the particle image diameter to the pixel was kept to be
approximately two. Subsequent to the accomplishment of the image correlation, deficient veloc-
ity vectors were rejected using a variable threshold technique based on a cellular neural network
method similar to that described by Shinneeb et al.26 This method utilizes the data associated with
the local velocity gradient in the domain to appropriately select the threshold. Less than 2% of
the vectors were rejected and those were replaced by the Gaussian-weighted mean of the adjacent
vectors.
The overall uncertainty associated with the velocity measurements is a summation of bias and
precision components. Many sources can contribute to the bias error in the PIV measurements such
as inability of the particle to follow the fluid, inaccuracy in PIV timing setup, and errors associated
with determining the particle displacements. In the present measurements, the seeding particles
were 12 m silver coated hollow glass spheres with a density of 1130 kg/m3 and estimated Stokes
settling velocity of 0.01 mm/s. The value of the Stokes velocity is small and the ratio of particle
response time to the turbulence time scale of a free jet at Reynolds number of 30 000 is 2 103,
so the error associated with particle inability to follow the flow is assumed insignificant. Following
recommendations of Prasad et al.,27 in the present PIV measurements, the mean particle image
diameter was approximately 3-4 pixels, yielding a bias error of about 0.1 pixel. The time delay
between the PIV images for a given experiment was chosen to yield an average displacement of
about 8-10 pixels in the centerline of the jet which resulted in bias error of 1.4% in the centerline
velocity. All statistical fields of velocity were obtained from an ensemble of 2000 instantaneous
vector fields (at the jet central plane) and 3000 instantaneous fields (at the impinging plate) for
each position. The uncertainties in the mean velocities at 95% confidence level were estimated to be
3% and 1.4% of the local velocity close to and away from the impingement plate, respectively.
Based on the evaluation of a global bias error and of the statistical error related to the data scattering
around the mean value, the uncertainties in the Reynolds stresses are estimated to be 6% in the
mid-region of the jet and 10% near the impingement plate.

III. NUMERICAL METHOD


In the experiments, the fluid first passed through a well-designed nozzle prior to entering the
jet tank. To maintain consistency with the experimental setup, the flow through the nozzle used in
these studies has been modelled. In the second step, the jet flow in the tank was simulated. The flow
parameters extracted from the nozzle exit were set as the inlet conditions for the jet simulation. The
Secs. III A and III B describe the modelling of the flow through the nozzle and the tank.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-6 Shademan et al. Phys. Fluids 28, 075107 (2016)

A. Nozzle flow modelling


The Finite Volume Method (FVM) is used to discretize the incompressible Navier-Stokes
equations governing the fluid flow. The second-order upwind scheme is used for discretizing
the convective terms in the momentum equations and the diffusion terms are discretized using a
second-order central scheme. The standard scheme28 for the pressure interpolation is used. For the
pressurevelocity coupling, the SIMPLE algorithm developed by Patankar and Spalding29 has been
applied. Time-marching is performed using a fully implicit second-order scheme with a time step of
1 105. The corresponding CFL number reached to 1 in places with high velocity but in most of the
regions in the domain it was less than 1. ANSYS Fluent30 has been used to perform the simulations.
A hybrid mesh combining a structured mesh in the wall region and an unstructured mesh in the
rest of the nozzle was used. A total of 3.3 106 cells was created for the entire nozzle geometry. A
mass flow rate of 0.35 kg/s was set at the inlet to the nozzle, resulting in a velocity of 2.86 m/s at the
exit of the nozzle. A pressure outlet condition was applied for the outlet and a no-slip no-penetration
condition was set for the walls.
A finite volume RANS simulation using the Reynolds Stress Model (RSM) was performed for
the flow through the nozzle. The RSM was selected because, as a non-isotropic model, it has the
capability of predicting the turbulence quantities in different directions, i.e., the Reynolds stresses
uu, vv, ww, uv, uw, and vw, where uu, vv, and ww are normal stresses in x, y, and z direc-
tions, and the rest are shear stresses in x-y, x-z, and y-z planes, respectively. Profiles of turbulence
components and the mean velocity, shown in Fig. 2, were extracted from the exit of the nozzle and
introduced as inlet conditions for the next stage of the simulation.

B. Jet modelling
The dimensions of the computational domain for the tank were chosen to simulate the exper-
imental setup described above. To simplify the meshing and reduce the computational cost, the
domain was taken to be a large circular cylinder (radius of 0.1 m) with its axis perpendicular to the
nozzle outlet plane (Fig. 3(a)). In order to reduce the lateral extent of the computational domain,
the water was allowed to escape to the ambient through the outer cylindrical boundary of the
computational domain by applying a constant pressure at the side boundary (pressure outlet condi-
tion). The impingement plate was considered to be a no-slip boundary. The jet flow was modelled
using LES. The governing equations used in LES are the filter based unsteady incompressible
Navier-Stokes equations. The filtered equations are as follows:31
ui
= 0, (1)
xi
ui 1 p ij 2ui
+ uiuj = + .

(2)
t xj xi xj xjxj

FIG. 2. Normalized streamwise (a) velocity, (b) turbulence intensity used at the inlet.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-7 Shademan et al. Phys. Fluids 28, 075107 (2016)

FIG. 3. (a) Virtual tank dimensions, (b) cross section of the mesh inside the tank, (c) cross section of the mesh on the plate.

Here, xi (i = 1, 2, 3) indicates the spatial directions, p is the pressure, is the density, is the
kinematic viscosity, ui is the resolved velocity field, and ij is the subgrid-scale (SGS). Reynolds
stresses defined by
ij = uiuj ui uj, (3)
where the overbar indicates a resolved quantity.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-8 Shademan et al. Phys. Fluids 28, 075107 (2016)

TABLE I. Grid data (virtual tank).

Number of cells in each direction

Mesh Axial Radial Circumferential Total number of cells

#1 300 150 120 6.3 106


#2 400 150 120 7.6 106
#3 400 150 160 10.6 106

Three different mesh sizes (Table I) were chosen for evaluation, each constructed in such a
way that the mesh requirements for LES are satisfied. The three meshes are of a hybrid type with
triangular prisms in the region r/D < 0.5, where r is the radial distance from the jet centerline. For
the rest of the domain, hexahedral elements are used. Details of the computational domain and the
mesh generated for one of the LES cases are shown in Figs. 3(a)-3(c).
The flow near a wall can be either resolved or modelled in LES, depending on the mesh
resolution close to the wall. Since, in the current simulations, the analysis of the impingement
zone is of primary interest, more emphasis has been placed on the resolution of the mesh close
to the target plate. Chapman32 determined that the resolution needed for the outer region of a
boundary layer is proportional to Re0.4, while for the wall layer the number of grid points required
increases to Re1.8. The impingement plate wetted wall surface area increases with increase in
the radial distance from the axis, resulting in a larger domain and consequently a need for more
cells. In the area of interest (r/D < 4.0), the mesh resolution should satisfy the generally accepted
LES criteria for wall-attached flows suggested by Piomelli and Chasnov,33 which requires that
r+ < 100, (r)+ < 20, and h+ < 2 (h is the distance in the normal direction to the wall, is
circumferential coordinate). Here r+ = r(u/), (r)+ = (r)(u/), and h+ = h(u/), where
u is the friction velocity and is the kinematic viscosity. Having a mesh with h+ < 1 resolves the
laminar sub-layer. Therefore the laminar stress-strain relationship (u+ = h+) can be implemented to
+
determine u, where u+ is the non-dimensional velocity equal to u/u, and h is the non-dimensional
normal distance from the wall equal to hu/, where u = w/ and w is the wall shear stress.
The number of grid points in each direction was systematically increased to evaluate whether
the mesh complied with the above-mentioned criteria. Figure 4(a) shows that the value of h+ is

FIG. 4. Mesh requirement analysis for different meshes, (a) mesh resolution in the axial direction, (b) dimensionless spacing
in the axial direction, (c) dimensionless spacing in the circumferential direction, and (d) dimensionless spacing in the radial
direction (: mesh #1, : mesh #2, : mesh #3).

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-9 Shademan et al. Phys. Fluids 28, 075107 (2016)

less than one for all three meshes for the near-wall cells. Figure 4(b) confirms that increasing the
number of grid points in the axial direction produces smaller h+ values close to the wall and
meets the requirement of h+ < 2. Figures 4(c) and 4(d) show the effect of the mesh density in
the near-wall region. To check for the radial and circumferential direction requirements (r+ < 100,
(r)+ < 20),33 the values of r+ and (r)+ for the three meshes are plotted against the radial
distance from the axis. Figures 4(c) and 4(d) show that in the region of interest (r/D < 4) the values
of (r)+ and r+ for all meshes are less than 5 and 100, respectively. However, some increase in
this value can be observed in regions far from the axis (r/D > 4), which is inevitable due to the
expansion of the mesh in the radial direction.
Figures 4(c) and 4(d) also show that an increase in grid points in the circumferential direction
does not have a significant effect on the results close to the wall. Figures 4(a)-4(d) show that even
though all three meshes meet the basic LES requirements, mesh #3 is preferable since it provides
values well within the criterion.
The mesh resolution quality can also be evaluated by comparing the mesh cell size =
(r r h)1/3 to the Kolmogorov length scale = (3/)1/4. Here is the molecular viscosity
and is the dissipation rate. The value of can be estimated from a separate RANS simulation
of the impinging jet flow using, for example, the Realizable k turbulence model. For isotropic
turbulence, a cell size of 12 or less is required in order to resolve the major contributions to the
dissipation.34 Therefore, the cell size for each mesh in the current study was evaluated to determine
whether the value of / was less than 12 in the regions of interest.
Figures 5(a) and 5(b) compare the mesh cell size with the Kolmogorov length scale at two
radial distances from the axis, r/D = 0 and 1, respectively. It can be seen that at r/D = 0 meshes #2
and #3 meet the required criteria and mesh #1 is very close to this requirement. At locations far from
the axis the cell size increases and satisfying the above mentioned criterion becomes more difficult.
It can be seen that, at r/D = 1, mesh #3 is very close to this requirement while mesh #1 does not
adhere to this condition. According to Fig. 5, the / ratio decreases as the plate is approached,
which confirms that mesh #3 is in the safe side for analyzing the impingement region and the wall
jet part.
Based on the above analysis, mesh #3 was selected for the current simulation. Further details on
the evaluation of mesh #3 are presented in Sec. IV A by comparing the mean value results with the
available experimental data. The average time used for each simulation for each mesh was about 2
months using a cluster with 28 (2.2 GHz AMD Opteron) CPUs.
The SGS stresses (Equation (3)) can be decomposed into three parts if the subgrid scale
velocity ui = ui ui is used,31
ij = uiuj ui uj = Lij + Cij + Rij, (4)

where Lij = ui uj ui uj are the Leonard stresses, Cij = uiuj + ujui are the cross terms, and Rij = uiuj
are the SGS Reynolds stresses. The interactions between resolved scales which results in subgrid-
scale contribution are represented by Leonard stresses. The cross term represents the interactions

FIG. 5. Comparison of mesh cell size with the Kolmogorov length scale at (a) r/D = 0, (b) r/D = 1 (: mesh #1, : mesh #2,
: mesh #3).

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-10 Shademan et al. Phys. Fluids 28, 075107 (2016)

between resolved and unresolved scales and the SGS Reynolds stresses are for interactions between
small and unresolved scales.31
The subgrid-scale stresses resulting from the filtering operation are unknown and require
modeling. The majority of the subgrid-scale models in use today are eddy viscosity models. These
models assume proportionality between the anisotropic part of the SGS stress tensor ij (1/3)ijkk
and the resolved scale strain rate tensor Sij as
1
ij ijkk = 2tSij, (5)
3
where t is the subgrid-scale turbulent viscosity and Sij is defined by
1 ui uj
( )
Sij = + . (6)
2 xj xi
One major drawback of the eddy viscosity subgrid-scale stress models used in LES is their
inability to correctly represent the turbulent field in rotating or sheared flows near solid walls with
a single universal model constant. The dynamic SGS model resolves this issue by dynamically
computing the model coefficient as the calculation progresses rather than being specified as a
priori input. The model is based on an algebraic identity between the subgrid-scale stresses at
two different filtered levels and the resolved turbulent stresses. The subgrid-scale stresses obtained
using the dynamic model vanish in laminar flow and at a solid boundary and have the correct
asymptotic behavior in the near-wall region of a turbulent boundary layer. In the current simulations
the dynamic Smagorinsky method presented by Germano et al.35 is used for the modelling of the
subgrid-scale stresses. The data collection for averaging the results has been started long after the
start of the simulation, when the flow has reached a fully established condition.

IV. RESULTS
A. Time averaged results
Validation of the numerical model was carried out by comparing the predicted mean and turbu-
lent characteristics with the current PIV results and other available experimental results.10,23,25,36
To evaluate the accuracy of the inlet profiles, mean and turbulent profiles in the jet, obtained from
the current LES analysis using mesh #3, were compared with the experimental results of Tandalam
et al.25 close to the nozzle exit, at x/D = 1, noting that the same nozzle was used in both experi-
ments. As shown in Fig. 6, there is good agreement between the results of the LES and the PIV data.
Considering the significant influence of the inlet condition on the development of the flow in the
domain, the accuracy observed in the mean and turbulent profiles in the vicinity of the nozzle exit
provides confidence in the validity of the numerical model.

FIG. 6. Comparison of (a) mean axial velocity, (b) turbulent axial velocity, at x/D = 1 (: LES, : experiments25).

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-11 Shademan et al. Phys. Fluids 28, 075107 (2016)

FIG. 7. Streamwise mean centerline velocity (CFD : LES with H/D = 20,  : RANS22 with H/D = 18.5;
experiments : PIV with H/D = 20, : free jet,36 x: H/D = 18.5,10 : H/D = 15.5,23 : H/D = 2223).

Figure 7 shows the normalized mean centerline velocity from the current simulations using
mesh #3 in comparison with the experimental results for 0 < x/D < 25. Results from the present
simulations follow the expected trend observed for different H/D values. The agreement between
the results obtained from LES and current PIV and other experiments supports the mesh require-
ment analysis presented in Section III B (Figs. 4 and 5) and the selection of mesh #3 for subsequent
simulations. Up to about x/D = 15, the flow is not influenced by the impingement wall and essen-
tially follows the behavior of a free jet. The computational model predictions agree well with the
experimental data as the flow approaches the plate. Up to approximately x/D = 4, the core of the jet
continues to develop and no decay in the centerline velocity can be observed. For x/D > 4, the free
jet region starts to develop and a large decay in centerline velocity occurs up to about x/D = 15 as
the ambient fluid is entrained into the jet. For x/D > 15, the flow senses the presence of the plate
and a sharper decay in the centerline velocity can be seen due to the transfer of momentum from the
axial to the radial direction.
Further validations were carried out by comparing other key flow parameters obtained from the
numerical simulation with the experimental results. Figure 8(a) compares the mean static pressure
distribution on the plate. The static pressure values are normalized by the pressure at the stagnation
point (Ps). The radial direction is normalized by the jet half width (r1/2), which is the radial posi-
tion where P = 0.5Ps. Higher pressure is observed in regions close to the impingement zone, and
the pressure decreases in the radial direction. The numerical prediction obtained from the current
simulation is in good agreement with the experiments37 and a slight improvement over the RANS
results.22
Figure 8(b) displays the mean wall shear stress distribution on the plate. In this figure, the radial
direction is normalized by the impingement distance, while the shear stress is normalized by U2j.
The quantity plotted along the vertical axis is chosen to be consistent with other studies. A similar
trend is observed between the data obtained from the simulations and the experimental data.37,8 The
difference can likely be attributed to different Reynolds numbers and variations in the experimental
setups and numerical models.
Further simulations were carried out at Re = 14 000 and 56 000 to investigate the dependence
of wall parameters including pressure distribution and wall shear stress on the variation of the
Reynolds number. The results showed a very marginal change in the distribution of these parameters
relative to the Reynolds numbers. These results are consistent with the results obtained by Tandalam
et al.25 who investigated the flow at three different Reynolds numbers of 10 000, 30 000, and 55 000.
They found that the effect of Reynolds number on the number, size, and circulation of vortices is
mainly in the nozzle near-field range of x/D < 5.
In Fig. 9 the turbulence intensities (urms, vrms, and wrms) normalized with Uj are plotted to
investigate the turbulence anisotropy along the centerline. Figure 9 shows that all three intensities

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-12 Shademan et al. Phys. Fluids 28, 075107 (2016)

FIG. 8. (a) Mean static pressure along the wall, (b) mean wall shear stress (CFD : LES with H/D = 20 (Ps = 440 Pa),
 : RANS22 with H/D = 18.5 (Ps = 410 Pa); experiments: H/D = 18,37 : H/D = 218).
start from a minimum value at x/D = 0 and reach a peak value at 6 < x/D < 8, indicating that the
maximum rate of mixing occurs in this region. Streamwise turbulence intensity has larger values
compared to the other components, suggesting that LES more accurately captures the physics of
impinging jet flows than models based on the assumption of isotropic turbulence.
The results also show how the variation of the turbulence intensities is influenced by the pres-
ence of the impingement plate. Near the plate (x/D > 19), the vrms and wrms (turbulence components
in y and z directions) exceed urms due to the strong anisotropy caused by the turning flow along the
plate. In proximity of the plate, the large decay in streamwise turbulent velocity fluctuation occurs
due to the change in the flow direction, where vrms and wrms components show a sudden increase. A
good match between numerical and experimental data is observed in the region of 10 < x/D < 18.

FIG. 9. Turbulence intensities obtained from LES (: urms/Uj, : vrms/Uj, : wrms/Uj) and PIV experiments ( :
urms/Uj, -+: vrms/Uj).

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-13 Shademan et al. Phys. Fluids 28, 075107 (2016)

FIG. 10. (a) Mean radial velocity profiles (V/Uj), (b) turbulent velocity profiles (vrms/Uj) in wall region (PIV - : r/D = 0,
: r/D = 1, : r/D = 2, x: r/D = 3; LES - : r/D = 0, : r/D = 1, : r/D = 2, : r/D = 3).

Following impingement, the flow spreads out in the radial direction. Since the fluid flows along
the wall, a boundary layer is expected to be formed. The mean and fluctuating radial velocities
normalized with the jet exit velocity are plotted as a function of (H x)/D in Fig. 10, where (H
x) is the normal distance from the plate. Figure 10 illustrates the comparison between the results
obtained from the LES and experiments at different r/D stations along the plate. These results show
that both mean and fluctuating components of the velocity follow the behavior of a developing wall
jet. As discussed above, the flow changes direction from axial to radial when it gets close to the
plate. This phenomenon can be observed in the mean radial velocity values in Fig. 10(a), as they
start to increase from zero at r/D = 0 to higher values at r/D = 1, 2, and 3. The space between
the profiles decreases along the wall jet, indicating that the flow is tending to a fully developed
condition. For the turbulence values, although quantitative agreement between the experiments and
CFD is not good in the proximity of the plate, the trends of the profiles in this region are quite
similar. This issue may be expected due to the higher levels of uncertainty associated with the
measurements in the wall proximity, as well as the difficulties associated with accurate prediction in
this region from the CFD model. In the region where the wall effect is negligible, good agreement
between the CFD and experiments can be observed.
Figures 11(a), 11(c), and 11(e) demonstrate the contours of mean velocity magnitude super-
imposed with the sectional streamlines, streamwise turbulent velocity fluctuations, and shear stress
distribution in the entire domain, respectively. Figures 11(b), 11(d), and 11(f) show the contours of
the same parameters in the proximity of the plate. These figures confirm that there is a symmetrical
behavior for the flow parameters with respect to the jet axis. Sectional flow streamlines show the
presence of a stagnation streamline which is located along the jet axis. As can be seen in these fig-
ures, turbulent velocity fluctuations and shear stresses also show symmetrical behavior with respect
to the axis. The marginal asymmetry observed is due to the averaging time during the simulation
period. An unsteady analysis will reveal the behavior of these parameters over time and elucidate
their deviation from the mean values.

B. Unsteady results
In this section, the evolution of the ring vortices generated in the domain due to the entrainment
of the ambient flow into the jet and their effect on different flow regions will be discussed. This
includes behavior of structures in the free jet region, impingement zone, and wall jet portion of the
flow.
Different methods can be employed to visualize the three-dimensional flow structures. The
literature shows that intuitive methods such as vorticity contours, pressure minima, and streamline

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-14 Shademan et al. Phys. Fluids 28, 075107 (2016)

FIG. 11. Contours of ((a) and(b)) mean velocity magnitude superimposed with sectional streamlines, ((c) and (d)) streamwise
turbulent velocity fluctuations, and ((e) and (f)) shear stress in the whole domain and close to the plate.

plots have largely been used for this purpose. In the current study the 2 criterion,38 which can be
used to identify the core of the vortices, has been implemented.
The 2 criterion was derived by taking the gradient of the Navier-Stokes equations and de-
composing the acceleration gradient term into symmetric and antisymmetric parts, expressed as
SijSij + ijij, where Sij and ij are the symmetric and antisymmetric parts of the velocity gradient
tensor, respectively. The Hessian of the pressure can then be connected to the vortical motions in
the flow. According to the theory of multivariable calculus, the SijSij + ijij tensor has three real
eigenvalues (1 2 3). The point of local pressure minimum requires two eigenvalues of this
tensor to be negative. 2, which corresponds to the second largest eigenvalue of this tensor, is the
representative of the local pressure minima region. The iso-surface of 2 can be used to visualize
instantaneous structures in the flow.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-15 Shademan et al. Phys. Fluids 28, 075107 (2016)

1. Free jet region


Figure 12(a) shows the iso-surface of 2 colored with velocity magnitude contours (0 < x/D <
3) and Fig. 12(b) shows the iso-surface of static pressure colored with vorticity magnitude contours
(0 < x/D < 20). For the current case (H/D = 20), from x/D = 0 15 the jet behaves very similar to
a free jet, in which the core of the jet undergoes a decay before reaching the plate.
The vortices generated due to the Kelvin-Helmholtz instabilities form a street of rolled up
vortex rings as illustrated in Fig. 12(a). These vortices, while traveling towards the plate, interact,
pair, and coalesce with neighboring vortices with increased azimuthal instability, thereby reducing
their circumferential coherence. The distance between these vortices increases and the ring-like
shape is lost when the flow approaches the impingement plate (Fig. 12(b)).
The expansion and stretching of the ring-like vortices caused by the azimuthal instabilities
are demonstrated in Figs. 13(a)-13(i), which show vorticity magnitude contours in the y-z plane at

FIG. 12. Iso-surfaces of (a) 2 criterion colored with velocity magnitude, (b) static pressure colored with vorticity magnitude.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-16 Shademan et al. Phys. Fluids 28, 075107 (2016)

FIG. 13. Deformation of shear layer in y-z plane visualized by the instantaneous vorticity magnitude contours at (a) x/D = 1,
(b) x/D = 1.25, (c) x/D = 1.5, (d) x/D = 1.75, (e) x/D = 2, (f) x/D = 4, (g) x/D = 7, (h) x/D = 15, (i) x/D = 20.

different axial stations. The color scale has been kept the same for all figures except in Fig. 13(h)
because of the reduced vorticity magnitude at this axial location (x/D = 15). In regions very close
to the nozzle (x/D = 11.5), the ring vortices have a diameter close to the nozzle exit diameter with
the highest vorticity located around the ring at about r/D = 0.5. As the vortices move in the axial
direction, they get deformed with vortical patches which are nested around the ring by the time they
reach x/D = 2. These patches of instability serve as a mechanism for transforming into the large
scale eddies downstream. Figure 13 shows that the number of patches reduces as the fluid flows
towards the plate. This may be attributed to the pairing and growth of the structures.
At a particular distance from the nozzle, the ring-like vortices start to breakdown due to
tilting and three-dimensional effects which cause the vortices to lose their axial symmetry. This
results in the broken down structures striking the plate at different time instances. The asymmetric
impingement of the large scale structures results in jet flapping and meandering around the jet axis.

2. Stagnation zone and wall jet region


To analyze the frequency of the ring vortices generated, the Strouhal number St = f1D/Uj was
calculated, where f1 is the frequency of generation of the ring vortices, D is the nozzle diameter,
and Uj is the jet exit velocity. One should note that the Strouhal number is strongly dependent on
Reynolds number, initial velocity profile, and most importantly the distance between the nozzle and
the plate.4,39 In the current study, to determine the frequency of the shear layer instabilities, the
energy-density spectra, obtained from a time series of the instantaneous static pressure at x/D = 2,

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-17 Shademan et al. Phys. Fluids 28, 075107 (2016)

TABLE II. Experimental and numerical values of the Strouhal number.

Experiments x/D Reynolds number Strouhal number

Yule,2 free jet 0.4 21 000 0.60


Yule,2 free jet 56 21 000 0.330.4
Han and Goldstein,40 free jet 1.0 8 000 0.65
Han and Goldstein,40 free jet 1.0 120 000 0.60
Tsubokura et al.,39 H/D = 10 38 6 000 0.37
Hadziabdic and Hanjalic,4 H/D = 2 0.1 21 000 0.64
Current LES study, H/D = 20 2.0 28 000 0.63

r/D = 0.5, was examined. Applying a FFT on the time history of the pressure signal yields a domi-
nant frequency of 180 Hz which corresponds to a Strouhal number of 0.63. For comparison, Table II
presents Strouhal numbers from previous studies and from the current simulation. As can be seen,
the St predicted in the current simulation is in the range of previous experimental results at similar
Reynolds numbers.
As previously mentioned, due to the mixing and the level of the turbulence, the ring vortices
lose their form, become entangled and evolve into large scale structures as they move towards
the plate. As the ring vortices approach the impingement zone, they break down and lose their
shape. The decrease in the number of ring vortices can be interpreted as a reduction in their fre-
quency. To investigate the existence of a characteristic frequency (or period T2) for the structures
striking the plate, the pressure history at x/D = 18 and r/D = 5 was recorded. The time history
of the static pressure monitored at this point is presented in Fig. 14. FFT was carried out to
determine any possible peak frequency for this time history. Based on the frequency determined
for the large scale structures impinging the plate, the characteristic period is T2 = 0.18 s. This
characteristic period is much higher than the period determined for the vortices close to the nozzle
(T1 = 1/f1 = 1/180 = 0.0055 s). Using this characteristic period, the movement of the structures in
the wall proximity was analyzed and the behavior of the wall shear stresses was monitored as time
continues.
To analyze the flow structures in the impingement region, velocity magnitude contours in the
y-z plane and the static pressure contours in x-y and x-z planes superimposed with the sectional
flow streamlines at different time instances over the characteristic period (T2) are plotted in Fig. 15.
The y-z plane parallel to the plate is located at (H x)/D = 0.05, which is slightly above the plate.
Five equal time intervals with t = T2/5 were chosen for evaluation. In this comprehensive figure,
(a), (b), (c), etc. denote the flow field at different time instants. The index 1 refers to the y-z plane
(e.g., Figure 15(a1)), index 2 refers to the x-y plane (e.g., Figure 15(a2)), and index 3 refers to the
x-z plane (e.g., Figure 15(a3)).
In Figs. 15(a1)-15(f1), the red patches indicate the high velocity regions. In each of these
figures the red patches are located in different regions of the flow field. It appears that the location of
these patches depend on the shape of the large scale structures impinging on the plate. These figures
confirm the asymmetrical impact of the large scale structures on the plate.

FIG. 14. History of static pressure at x/D = 18 and r/D = 5.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-18 Shademan et al. Phys. Fluids 28, 075107 (2016)

The asymmetrical impingement of the large scale structures also influences the location of the
stagnation region, which can be observed by the behavior of sectional streamlines in the y-z plane.
The dotted circle in the middle of Figs. 15(a1)-15(f1) indicates the location of the nozzle above
the plate. As seen in these figures, different stagnation patterns exist in the impingement zone at

FIG. 15. Instantaneous velocity magnitude (y-z plane) and static pressure contours (x-y and x-z planes) with sectional
streamlines over the characteristic period (T2), (a) t/T2 = 0, (b) t/T2 = 1/5, (c) t/T2 = 2/5, (d) t/T2 = 3/5, (e) t/T2 = 4/5, and
(f) t/T2 = 5/5 (the small circle in the y-z plane shows the projection of the nozzle on the plate).

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-19 Shademan et al. Phys. Fluids 28, 075107 (2016)

FIG. 15. (Continued.)

different times. One single patch of low velocity magnitude flow (in blue color with streamlines
emanating from that location) representing the stagnation zone can be seen at the beginning of the
characteristic period. As time goes on the shape of the stagnation zone changes and becomes more
distributed on the plate. Different disperse patches of blue color (with streamlines starting from that
region) representing the stagnation zones can be observed on the plate. These zones are located
significantly off the axis of symmetry relative to the beginning of the period.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-20 Shademan et al. Phys. Fluids 28, 075107 (2016)

The stagnation regions reach a maximum distance of r/D = 2 from the jet axis during the char-
acteristic period. Note that, this dispersed form of stagnation regions is a result of the impingement
of the broken down, disconnected, three-dimensional structures. At the end of the period, the stag-
nation region gets close to the jet axis and takes a shape close to its original form at the beginning of
the period, and no other stagnation zone can be observed on the plate.
Figure 15 also shows the flow streamlines in the x-y and x-z planes at different time intervals
of the characteristic period (example, Figs. 15(a2) and 15(a3) and 15(b2) and 15(b3), etc.). The
pressure contours superimposed with flow streamlines show the asymmetric and three-dimensional
movement of the flow structures approaching the plate. The fluid structures have different shapes
and sizes and are deflected in the radial direction while getting close to the plate. This behavior
influences the flow in the impingement zone as can be observed in the x-y and x-z planes. One of
these influences is the change in the pressure contours noticed in the impingement zone. The pres-
ence of an adverse pressure gradient in different directions allows for the generation of secondary
vortices in the impingement zone.
Due to the large distance between the nozzle and the plate (H/D = 20), the maximum disloca-
tion of the impingement point from the axis of symmetry is quite different from impinging jets with
small nozzle-to-plate ratio (H/D < 4). Hadziabdic and Hanjalic4 noticed a maximum dislocation
of r/D = 0.1 and a change of the shape of the impingement point to two lines for H/D = 2. They
suggested that the origin of the impingement point oscillation arises from the instability and tilt of
the structures generated from the shear layer of the jet. In the current study, due to the large distance
between the nozzle and the plate, the oscillation of the jet centerline is larger. At the beginning
of the characteristic period, the stagnation region (dark blue regions in Fig. 15(a1)) has the form
of a point with a maximum dislocation of r/D = 0.5 from the jet axis. As time goes on different
stagnation regions can be observed on the plate (at t/T2 = 3/4 and 4/5). The maximum distance of
these stagnation regions is about r/D = 2.
After impingement the flow changes direction from axial to radial and a thin boundary layer
is formed on the plate (referred to as the wall jet region). The sectional streamlines show an
asymmetric behavior caused by the large scale structures. The structures occurring in the wall jet
are remnants of the ring-like vortices generated from the nozzle shear layer which survive the
impingement.
According to the analyses of Dairay et al.41 for an impinging jet with H/D = 1 and Hadzi-
abdic and Hanjalic4 for H/D = 2, several secondary vortices occur in the wall-jet region which
are generated due to the influence of deflected nozzle vortices. This behavior causes unsteadiness
and fluctuations in the wall pressure and shear stress. Based on the sectional flow streamlines and
pressure contours presented in Figs. 15(a)-15(f), the deflected large scale structures do not result
in significant secondary flows in the wall region. This is due to the fact that in impinging jets
with large stand-off distance, the large scale structures lose their coherence while approaching the
impingement plate, unlike the cases with small stand-off distance where the structures almost keep
their interconnected shape. This behavior illustrates another major difference between impinging
jets with large and small stand-off distances.

3. Wall shear stress


Figures 16(a) and 16(b) show the instantaneous sectional streamlines in the x-y plane (not
scaled) and the wall shear stresses in the y direction, respectively. Similarly, the rest of the graphs
in Fig. 16 show the distributions at other time instants over the characteristic period. In these figures
the mean wall shear stresses over the characteristic period (T2) are compared with the instantaneous
results. The wall shear stress () is normalized with the maximum value of mean shear stress
(meanmax) along the y direction. Note that these figures are presented at the same instantaneous
time steps as presented in Fig. 15. Sudden changes from high positive values to low negative values
occur in this region. The negative wall shear stress values are a result of the separation of flow in
the impingement zone and generation of secondary vortices (see Figs. 16(a), 16(c), 16(e), 16(g),
16(i), and 16(k)). Another important phenomenon observed in these figures is the maximum ratio of
the instantaneous wall shear stress to the maximum mean wall shear stress during the characteristic

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-21 Shademan et al. Phys. Fluids 28, 075107 (2016)

FIG. 16. Instantaneous sectional streamlines (black) and mean wall shear stress (red) in the x-y plane over the characteristic
period, ((a) and (b)) t/T2 = 0, ((c) and (d)) t/T2 = 1/5, ((e) and (f)) t/T2 = 2/5, ((g) and (h)) t/T2 = 3/5, ((i) and (j)) t/T2 = 4/5,
and ((k) and (l)) t/T2 = 5/5.

time period. A maximum ratio of +3 was observed at t/T2 = 5/5 and a minimum ratio of 1.8 was
seen at t/T2 = 4/5. By increasing the r/D ratio, the rate of the fluctuations decreases and the behavior
of the instantaneous and mean wall shear stresses gets close to each other.
Figures 16(a)-16(f) show that, unlike that seen in the mean profile, there are multiple peaks
with negative and positive values in the instantaneous wall shear stress profiles. These fluctuations
occur mostly within the r/D < 5 stations with a number of negative values, while for regions
with r/D > 5 significant fluctuations were not observed. This is due to the deflection of the large
scale structures from the centerline on approaching the plate. This is in contrast to the behavior
of impinging jets with small stand-off distance. El Hassan et al.5 showed that for impinging jets
with H/D = 2, due to the coherence of the vortex rings, the vortices survive the impingement pro-
cess and generate secondary vortices in the wall region. This phenomenon contributes to increased
fluctuations in the instantaneous wall shear stress.

V. CONCLUSIONS
Analysis based on LES was carried out to characterize a turbulent round impinging jet with
a large stand-off distance. Modelling this type of long jet is computationally more expensive than
short impinging jets normally discussed in the literature.
Three meshes with different number of cells (6.3 106, 7.6 106, 10.6 106) were tested to
ensure the mesh requirements in LES were satisfied. The decay of centerline velocity from the three
different meshes was compared with the available experimental data. Based on the mesh require-
ment analysis and the quality of the results, the 10.6 106 cell mesh was selected for subsequent
computations. These evaluations demonstrated that the LES framework showed good capability

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-22 Shademan et al. Phys. Fluids 28, 075107 (2016)

in capturing the mean value fields in all three sub-regions of the domain including the free jet,
impinging zone, and wall jet regions.
Turbulence generation at the inlet of the computational domain plays an important role in large
eddy simulations, significantly affecting the downstream flow field. To introduce a correct shear
flow and a proper level of turbulence at the exit of the nozzle, flow inside the nozzle was modelled
separately. Good agreement between the experimental and computational results close to the nozzle
and inside the tank suggests that the flow inside the nozzle has been accurately modelled.
The dynamics of the roll-up vortices created by the instabilities in the initial shear layer close
to the nozzle influences the entire flow field including the free jet region, impinging zone, and wall
jet region. It was found that these roll-up vortices have a Strouhal number of 0.63 near the nozzle.
However, this parameter is reduced along the centerline closer to the impingement zone. While
moving towards the plate, these vortices merge, breakdown, or change into large scale structures.
Up to four diameters from the nozzle exit, these vortices retain their ring-like shape. In the range
of 4 < x/D < 7, a transition occurs and the vortices start to change into large scale structures. For
x/D > 7, there is no sign of the ring-like vortices as they are transitioned to form large scale struc-
tures. Due to the asymmetric behavior of the structures, these eddies are tilted from the streamwise
direction. The expansion, growth, tilt, and unsteady behavior of the impinging structures cause
dislocation of the impinging flow from the centerline, which is significantly larger when compared
to cases with small H/D ratios. Unlike in the cases with small H/D ratios, for large stand-off cases,
the large scale structures do not present a self-organizing tendency towards a ring-like shape after
impingement.
In the current study a pressure probe (numerical monitoring) was set at a location close to the
plate to determine the frequency of the structures striking the plate. Based on the pressure history, a
FFT operation was carried out on the signal and the dominant frequency was captured. The behavior
of the structures in the impinging zone and in the wall region was analyzed using this characteristic
frequency.
Different patterns of velocity distribution were observed in the impingement region over one
cycle. Initially, the stagnation region is in the form of an organized circular shape. As time passes,
different patches of low velocity magnitude contours representing different stagnation regions were
observed. Due to rotation, tilting, and unsteadiness of the large scale structures approaching the
plate and also considering the large distance between the nozzle and the plate, a large dislocation of
the jet stagnation regions was observed compared to cases with small stand-off distances.
Due to the loss of coherence, the large scale structures do not generate secondary vortices in
the wall region far from the centerline (r/D > 5). As a result, the wall shear stress demonstrates a
smooth trend in this part of the flow. This phenomenon is in contrast to the behavior of imping-
ing jets with small H/D ratio, where ring-like vortices retain their coherence and cause secondary
vortices which are associated with large fluctuations in the wall shear stress.

ACKNOWLEDGMENTS
This research was supported by NSERC through the Vanier Canada Graduate Scholarship pro-
gram. The facilities of the Shared Hierarchical Academic Research Computing Network (SHARC-
NET) were used to conduct the simulations.
1 R. Manceau, R. Perrin, M. Hadiabdic, and S. Benhamadouche, Investigation of the interaction of a turbulent impinging
jet and a heated, rotating disk, Phys. Fluids 26, 035112 (2014).
2 A. J. Yule, Large-scale structure in the mixing layer of a round jet, J. Fluid Mech. 89, 413432 (1978).
3 A. K. M. F. Hussain, Coherent structures and turbulence, J. Fluid Mech. 173, 303356 (1986).
4 M. Hadziabdic and K. Hanjalic, Vortical structures and heat transfer in a round impinging jet, J. Fluid Mech. 596, 221260

(2008).
5 M. El Hassan, H. H. Assoum, R. Martinuzzi, V. Sobolik, K. Abed-Meraim, and A. Sakout, Experimental investigation of

the wall shear stress in a circular impinging jet, Phys. Fluids 25, 077101 (2013).
6 J. W. Hall and D. Ewing, On the dynamics of the large-scale structures in round impinging jets, J. Fluid Mech. 555,

439458 (2006).
7 J. Tian, V. Roussinova, and R. Balachandar, Characteristics of a jet in the vicinity of a free surface, ASME J. Fluids Eng.

134(3), 31204 (2012).


8 S. Beltaos and N. Rajaratnam, Impinging circular turbulent jets, J. Hydraul. Eng. ASCE 100(10), 13131328 (1974).

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47
075107-23 Shademan et al. Phys. Fluids 28, 075107 (2016)

9 S. Beltaos and N. Rajaratnam, Impingement of axisymmetric developing jets, J. Hydraul. Res. 15(4), 311326 (1977).
10 N. Rajaratnam, D. Z. Zhu, and S. P. Rai, Turbulence measurements in the impinging region of a circular jet, Can. J. Civ.
Eng. 37(5), 782786 (2010).
11 P. Bradshaw, The effect of initial conditions on the development of a free shear layer, J. Fluid Mech. 26(2), 225236

(1966).
12 K. Nishino, M. Samada, K. Kasuya, and K. Torii, Turbulence statistics in the stagnation region of an axisymmetric imping-

ing jet flow, Int. J. Heat Fluid Flow 17(3), 193201 (1996).
13 Second ERCOFTAC-IAHR Workshop on Refined Modelling, Round normally impinging turbulent jet and turbulent

flow through tube bank sub-channel, 16th Meeting of the IAHR Working Group on Refined Flow Modelling (University
of Manchester, Institute of Science and Technology, UK, 1993).
14 T. Craft, L. Graham, and B. E. Launder, Impinging jet studies for turbulence model assessment-II. An examination of the

performance of four turbulence models, Int. J. Heat Mass Transfer 36, 26852697 (1993).
15 D. Cooper, D. C. Jackson, B. E. Launder, and G. X. Liao, Impinging jet studies for turbulence model assessment-I.

Flow-field experiments, Int. J. Heat Mass Transfer 36, 26752684 (1993).


16 P. R. Voke and S. Gao, Numerical study of heat transfer from an impinging jet, Int. J. Heat Mass Transfer 41, 671680

(1998).
17 F. Beaubert and S. Viazzo, Large eddy simulation of plane turbulent impinging jets at moderate Reynolds numbers, Int.

J. Heat Fluid Flow 24, 512519 (2003).


18 C. Bogey, O. Marsden, and C. Bailly, Effects of moderate Reynolds numbers on subsonic round jets with highly disturbed

nozzle-exit boundary layers, Phys. Fluids 24, 105107 (2012).


19 N. Uddin, S. O. Neumann, and B. Weigand, Les simulations of an impinging jet: On the origin of the second peak in the

Nusselt number distribution, Int. J. Heat Mass Transfer 57(1), 356368 (2013).
20 K. Kataoka, Y. Kamiyama, S. Hashimoto, and T. Komai, Mass transfer between a plane surface and an impinging turbulent

jet: The influence of surface-pressure fluctuations, J. Fluid Mech. 119, 91105 (1982).
21 K. Kataoka, M. Suguro, K. Degawa, K. Maruo, and I. Mihata, The effect of surface renewal due to large-scale eddies on

jet impingement heat transfer, Int. J. Heat Mass Transfer 30, 559567 (1987).
22 M. Shademan, R. Balachandar, and R. M. Barron, CFD analysis of the effect of nozzle stand-off distance on turbulent

impinging jets, Can. J. Civ. Eng. 40(7), 603612 (2013).


23 F. Giralt, C. Chia, and O. Trass, Characterization of the impingement region in an axisymmetric turbulent jet, Ind. Eng.

Chem. Fundam. 16(1), 2128 (1977).


24 C. Ball, H. Fellouah, and A. Pollard, The flow field in turbulent round free jets, Prog. Aerosp. Sci. 50, 126 (2012).
25 A. Tandalam, R. Balachandar, and R. M. Barron, Reynolds number effects on the near-exit region of turbulent jets, J.

Hydraul. Eng. 136(9), 633641 (2010).


26 A. M. Shinneeb, J. D. Bugg, and R. Balachandar, Variable threshold outlier identification in PIV data, Meas. Sci. Technol.

15, 17221732 (2004).


27 A. K. Prasad, R. J. Adrian, C. C. Landreth, and P. W. Offutt, Effect of resolution on the speed and accuracy of particle

image velocimetry interrogation, Exp. Fluids 13, 105116 (1992).


28 C. M. Rhie and W. L. Chow, Numerical study of the turbulent flow past an airfoil with trailing edge separation, AIAA J.

21(11), 15251532 (1983).


29 S. V. Patankar and D. B. Spalding, A calculation procedure for heat, mass and momentum transfer in three-dimensional

parabolic flows, Int. J. Heat Mass Transfer 15, 17871806 (1972).


30 ANSYS Academic Research, Release 14.0, Help System. Fluent Users Guide. ANSYS, Inc., USA, 2011.
31 U. Piomelli, Large-eddy simulation: Achievements and challenges, Prog. Aerosp. Sci. 35, 335362 (1999).
32 D. R. Chapman, Computational aerodynamics development and outlook, AIAA J. 17(12), 12931313 (1979).
33 U. Piomelli and J. R. Chasnov, Large-eddy simulations: Theory and applications, in Transition and Turbulence Modelling,

edited by A. Henningson, K. Hallback, L. Alfredsson, and M. Johansson (Kluwer Academic Publishers, Dordrecht, 1996),
pp. 269336.
34 S. B. Pope, Turbulent Flows (Cambridge University Press, Cambridge, UK, 2000).
35 M. Germano, U. Piomelli, P. Moin, and W. H. Cabot, A dynamic subgrid-scale eddy viscosity model, Phys. Fluids A3,

17601765 (1991).
36 A. M. Shinneeb, J. D. Bugg, and R. Balachandar, Quantitative investigation of vortical structures in the near-exit region

of an axisymmetric turbulent jet, J. Turbul. 9(19), 120 (2008).


37 P. Bradshaw and E. M. Love, The normal impingement of a circular air jet over a flat surface, in Aeronautical Research

Council Reports and Memoranda (Aeronautical Research Council, 1959), 3205.


38 J. Jeong and F. Hussain, On the identification of a vortex, J. Fluid Mech. 285, 6994 (1995).
39 M. Tsubokura, T. Kobayashi, N. Taniguchi, and W. P. Jones, A numerical study on the eddy structures of impinging jets

excited at the inlet, Int. J. Heat Fluid Flow 24, 500511 (2003).
40 B. Han and R. J. Goldstein, Instantaneous energy separation in a free jet. Part I. Flow measurement and visualization, Int.

J. Heat Mass Transfer 46, 39753981 (2003).


41 T. Dairay, V. Fortun, E. Lamballais, and L. E. Brizzia, Direct numerical simulation of a turbulent jet impinging on a heated

wall, J. Fluid Mech. 764, 362394 (2015).

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP:
107.173.142.153 On: Mon, 01 Aug 2016 13:20:47

Você também pode gostar