Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/226909444

An experimental study on the vortical structures and behaviour of jets issuing


from inclined coaxial nozzles

Article  in  Experiments in Fluids · October 2011


DOI: 10.1007/s00348-011-1120-4

CITATIONS READS

7 40

2 authors, including:

T. H. New
Nanyang Technological University
146 PUBLICATIONS   882 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Jet in cross-flow phenomenon View project

Two-phase gas-particle flows in industrial air-slide conveyors View project

All content following this page was uploaded by T. H. New on 25 December 2014.

The user has requested enhancement of the downloaded file.


An experimental study on the vortical
structures and behaviour of jets issuing from
inclined coaxial nozzles

T. H. New* and E. Tsioli

* School of Mechanical and Aerospace Engineering


Nanyang Technological University
50 Nanyang Avenue, Singapore 639798

School of Engineering
University of Liverpool
Liverpool L69 3GH
United Kingdom

Abstract
An experimental study on inclined coaxial jets using laser-induced fluorescence and
particle image velocimetry is presented here. The Reynolds numbers of the inner primary
jet and outer secondary jet were Re=2500 and between Re=500 to 2000 (based on gap
size) respectively, which corresponded to secondary-to-primary jet velocity-ratios (VR) of
VR=0.5 to 2.0. The secondary-to-primary jet area-ratio was 2.25 and 45 and 60 incline-
angles were studied. Flow visualizations show that relatively independent inclined primary
and secondary jet vortex roll-ups were formed at VR=0.5. At VR=1.0, regular pairings and
mergings between primary and secondary jet vortex roll-ups led to large-scale
entrainment of secondary jet and ambient fluids into the primary jet column and conferred
a “serpentile”-shaped outline upon it. While the “serpentile”-shaped outline continued to
exist at VR=2.0, it was a result of stronger secondary jet inner vortex roll-ups which
“pinched” the primary jet column regularly. These flow behaviors are observed to intensify
with an increase in the incline-angle used. Velocity measurements demonstrate that
inclined coaxial nozzles promoted vectoring of the primary jet momentum towards the
longer nozzle lengths when velocity-ratio and/or incline-angle were increased. Lastly,
peak velocity and higher turbulence intensity levels due to augmented vortical interactions
are also detected along shorter nozzle lengths.

Keywords: coaxial jets; indeterminate-origin nozzles; laser-induced fluorescence; particle


image velocimetry

1
Abbreviations

d Annular gap size, d=(D2-D1)/2-tw

D1 Primary nozzle diameter

D2 Secondary nozzle diameter

H Nozzle mean height

tw Nozzle wall thickness

urms Root-mean-square streamwise velocity

U1 Mean primary jet exit velocity

U2 Mean secondary jet exit velocity

x Streamwise distance from nozzle mean height

y Cross-stream distance from nozzle center

 Kinematic viscosity of water

 Nozzle azimuthal location

Re1 Primary jet Reynolds number, Re1=U1D1/

Re2 Secondary jet Reynolds number, Re2=U2d/

St1 Inner shear region vortex formation frequency,

St1=fD1/U1

St2 Outer shear region vortex formation frequency,

St2=fd/U2

AR Area-ratio, AR=(D2/D1)2

VR Velocity-ratio, VR=U2/U1

HWA Hot-wire anemometry

LDA Laser Doppler anemometry

PIV Particle image velocimetry

px Pixel

2
1. Introduction
Single-stream jets and discrete vortex rings discharging from inclined circular nozzles

have been studied previously by Wlezien and Kibens (1986, 1988), Webster and

Longmire (1997, 1998), Lim (1998) and Troolin and Longmire (2010). Such nozzles are

capable of conferring redistributive effects upon the energy levels within the jet shear

layers, as well as resulting in asymmetric jet spreads and vectoring. In vortex rings, the

redistributive effects not only cause them to vector away from the nozzle centreline, but

produce destabilizing effects upon the vortex ring filaments as well. These

characteristics were observed not only in low-speed jet flows but also in supersonic jet

flows as well, which attests to their robustness as passive control devices. These flow

attributes were subsequently deduced to be due to the formation of inclined ring-vortices

and their self-excitations, the extents of which depend on the incline-angle used.

Generally similar flow behaviour has also been observed when inclined elliptic nozzles

were used by New (2009). These studies indicate that imposing simple inclined exits

upon nozzle jets can be a simple but yet effective flow control technique for even more

complex nozzle configurations such as coaxial nozzles.

Jets issuing from conventional non-inclined coaxial nozzles have been well-studied in

numerous earlier investigations. One recurring research interest was focused on how

initial flow conditions affected flow instabilities resulting from the interactions between the

two jet streams, which in turn governed the resulting flow behaviour in the near-field

region. Work carried out by Champagne and Wygnanski (1971), Ko and Kwan (1976),

Kwan and Ko (1977), Ribeiro and Whitelaw (1980), Lu (1983), Ko and Au (1985), Au and

Ko (1987), Wicker and Eaton (1994), Buresti et al. (1994), Tang and Ko (1994), Warda et

al. (2001), Talamelli and Gavarini (2006) and Burattini and Talamelli (2007), just to name

a few, had systematically showed that the area-ratio, velocity-ratio, initial turbulence

intensity levels, shear layer characteristics, nozzle wall thicknesses and presence of

external excitations or swirl played important roles in determining coaxial jet behaviour.

Extensive hot-wire anemometry (HWA) and laser Doppler anemometry (LDA)

measurements in these studies revealed that jet potential core lengths, presence of self-

3
similarity, modes of interaction between the two jet streams and the resultant vortex

structures and behaviour were highly sensitive towards these factors.

Take for instance, early investigation by Champagne and Wygnanski (1971) and Ribeiro

and Whitelaw (1980) showed that velocity ratio significantly influenced the velocity

statistics and associated mixing behaviour arising from interactions between the two jet

streams. Ko and Kwan (1976) and Kwan and Ko (1977) also proposed that coaxial jets

could be looked upon as two single jets superimposed upon each other. Follow-on work

by Ko and Au (1985) and Au and Ko (1987) further demonstrated that the coaxial jet flow

field could be differentiated into well-defined zones – initial, intermediate and fully-merged

merging zones. Substantial dependency of the flow behaviour upon the initial velocity

profile, turbulence intensity levels, velocity-ratio in coaxial jets were also confirmed in

measurements carried out by Buresti et al. (1994), Rehab et al. (1997), Warda et al.

(1999) and Ahmed and Sharma (2000). In their experimental study on coaxial jets

however, Buresti et al. (1994) looked at the turbulent exit profile and observed that the

initial conditions do not affect significantly the spreading of mixing layers. They focused

on high-order analysis of the velocity fluctuations and determined the skewness and

kurtosis of the velocity, providing additional information on the wall effect and the length

of the potential core.

Higher-order velocity statistics had also been determined by Park and Chen (1989) and

Sadr and Klewicki (2003) to shed light upon the movement of the vortex structures and

associated mixing behaviour. Recently, a stability analysis study carried out by Talamelli

and Gavarini (2006) further revealed that it might be possible to control coaxial jet mixing

behaviour by manipulating the velocity-ratio such that an absolute instability mode would

be triggered. The effects of external perturbations upon coaxial jets had also been

investigated by Wicker and Eaton (1994), when they examined excited and unexcited

turbulent jets. Large-scale structures of the secondary jet were found to play an

important role, as they would form independently and ultimately control the primary jet

flow. Primary jet flow excitation was found not to have a significant effect on the

evolution of the outer layer structures in that study. Axial excitation of the annular flow

4
produced large-scale structures on the outer shear layer, similar to a single jet and these

structures provided strong coupling between the two layers.

Note that while the majority of the above studies provided excellent information on coaxial

jets through statistical data such as turbulence intensities, skewness, flow stresses and

velocity spectra, they were less intuitive when it came to appreciating the overall coaxial

jet vortex flow behaviour. This was addressed to some extent when experimental flow

visualization studies were performed by Dahm et al. (1992), Wicker and Eaton (1994),

Tang and Ko (1994), Rehab et al. (1997), Villermaux and Rehab (2000) and Kiwata et al.

(2001). In particular, Dahm et al. (1992) showed good agreements with earlier

measurements when the resultant vortex structures and behaviour observed through

laser-induced fluorescence were highly sensitive towards the jet Reynolds numbers,

velocity-ratio, characteristics of the jet shear layers and wake defects. Interestingly,

additional effects in coaxial jets such as pairing events between the vortex structures of

the two jets as well as “locking-in” effect were also identified in this study, thus supporting

the importance of the interactions between the two jets in the coaxial jet development.

An interesting phenomenon was also observed by Rehab et al. (1997) using similar

visualization technique. In particular, when the ratio of the outer jet to the inner jet

exceeded unity and reached some critical value, the inner jet potential core would be

truncated to form an oscillating recirculation bubble. In addition, the mixing mechanisms

associated with such coaxial jets were further elaborated by a later study by Villermaux

and Rehab (2000). Along the numerical simulation front, Balarac and Métais (2005)

carried out a numerical study to investigate the role of nozzle lip thickness and found that

smaller thicknesses tend to lead to faster transition to turbulence. Furthermore, the

thickness of the nozzle lip also affected the shape and extent of the recirculation bubble

formed, when the velocity-ratio was beyond the critical value. The effects of forcing

upon coaxial jets were also studied by Balarac et al. (2007) recently, where the presence

of excitations generally improved coaxial jet mixing characteristics and supported earlier

findings obtained by Wicker and Eaton (1994).

5
While it is clear that inclined and coaxial nozzles can be useful in applications desiring

significant jet mixing and control improvements through passive means, fundamental flow

behavioural understanding on the combination of these two features together remains

lacking. Hence, the question of how inclined exits may be applied to other more

complex nozzle configurations beyond single-stream jet nozzles provides the main

motivation for the present study. Circular coaxial nozzle jet flows which feature two

independent fluid streams represent a natural extension to single-stream circular nozzle

jet flows. With relatively well documented flow behaviour, it also offers a good starting

point for the question posed above. Therefore, it will be of interest to investigate if flow

features observed in single-stream inclined jets extend to dual-stream inclined coaxial

jets. In particular, the study is interested in understanding whether and how the

production of inclined ring vortices, possible vectoring of the jet and the “serpentile”-

shaped jet column as observed in earlier inclined jets studies are replicated in coaxial jet

configurations. In addition, the not so straight-forward behaviour of the vortex structures

from the primary and secondary jets in coaxial nozzles towards velocity-ratio as observed

by Dahm et al. (1992) invokes further curiosity in inclined coaxial nozzles. Towards

answering these questions, an experimental study has been carried out on inclined

coaxial jets here to inspect the resultant near-field flow structures and behaviour.

Laser-induced fluorescence (LIF) visualizations were used to examine the dynamics of

the vortex structures and behaviour, as the incline angles and velocity-ratio were varied

independently. Particle image velocimetry (PIV) was used as the measurement

technique of choice, such that the resultant flow field information could be obtained in a

relatively straight forward and global manner. Note that at this point in time, it was not

the aim of the present study to obtain detailed velocity statistics similar to earlier studies

from PIV measurements and carry out direct comparisons with historical data. Instead,

the focus was to provide a first-hand look at the most interesting and intuitive flow

changes associated with a unique coaxial jet configuration when important flow

parameters were varied systematically.

6
2. Experimental setup

(a) Test tank and apparatus

The experiments were carried out in a recirculating horizontal Plexiglas water tank with

internal dimensions of 400mm (W) x 400mm (H) x 1000mm (L) as shown in Fig. 1. Two

centrifugal pumps were used to channel water from a reservoir into the coaxial jet

apparatus, along which the individual flow rates are controlled by valves and monitored

using electromagnetic flow meters separately. The coaxial jet apparatus was

constructed such that water for the primary inner jet would pass through a diffuser,

honeycombs and fine screens before reaching the contraction chamber as shown in Fig.

2(a). Water for the secondary annular jet underwent relatively similar flow-conditioning,

except that no diffuser was used due to the jet apparatus design. Eight equally-spaced

inlet ports were also used to allow uniform entrance of secondary jet water into the jet

apparatus, as shown in the figure. Both the primary and secondary jet apparatus were

aligned concentrically for the study here. After both jet streams exhausted from the test

nozzles, excessive water would be diverted from the water tank back to the reservoir to

repeat the flow cycle. Further details on the experimental facility can be found in Tsioli

and New (2009).

(b) Nozzle designs

Three sets of coaxial nozzles were used in the present study, where they consisted of

one conventional non-inclined and two inclined nozzles as shown in Figs. 2(b) to

2(d). For the inclined nozzles, incline-angles of 45o and 60o were used. The diameter of

the primary jet nozzle was D1=20mm throughout, while that of the secondary jet nozzle

depended on the exact area-ratio used. An area-ratio (AR) of AR=(D2/D1)2=2.25 was

studied here, which corresponded to a secondary jet nozzle diameter of D2=30mm. The

mean heights of all nozzles were kept similar at H=50mm or H/D1=2.5 for the sake of

consistency. Due to the nozzle fabrication processes, wall thicknesses were maintained

at 1mm throughout for both primary and secondary nozzles. As the axial extent of the

nozzles varied circumferentially, it was more intuitive to indicate the azimuthal locations of

the nozzle exits by azimuthal angle  as shown in Fig. 3. Within the present study, =0

coincided with the azimuthal location where the nozzle lengths were the longest and

7
increased in the anti-clockwise direction when viewed from the exhaust end of the coaxial

nozzles. In that case, the azimuthal angle would be = at the azimuthal location where

the nozzle lengths were the shortest. This convention was selected to be consistent with

that adopted by Webster and Longmire (1997) for consistency. During the present study

where streamwise LIF visualizations and PIV measurements of the coaxial jet flows were

taken, the laser sheet would be aligned either along =0 plane or =/2 plane, with the

camera and lens orientated perpendicular to the plane accordingly.

(c) Jet exit conditions

The different nozzle configurations were examined under three different velocity-ratios

(VR), where VR=U2/U1=0.5, 1 and 2. For the study, the flow velocity of the primary jet was

kept constant, while that of the secondary jet was varied according to the experimental

requirements. The primary jet mean velocity was U1=0.14m/s, which gave a Reynolds

number of approximately Re1=U1D1/=2500, where  is the kinematic viscosity of

water. The secondary jet velocity was then changed according to the required velocity-

ratios and its Reynolds number, defined here as Re2=U2d/, ranged from Re2=500 to

2000, where d was the dimension of the annular gap between the secondary and primary

nozzles. Turbulence intensity levels for the non-inclined coaxial nozzles were

determined from PIV measurements along x/D1=0.3 location. At that location, the

primary jet centerline initial turbulence level was estimated to be approximately 4.5%.

As for the secondary jet, its initial turbulence levels varied approximately between 8 to

12% at all velocity-ratios used here.

The initial turbulence intensity levels here were relatively different as compared to earlier

studies. Two main reasons accounted for the discrepancies: Firstly, relatively low flow

velocities and Reynolds numbers were used here. In contrast, most of the earlier

studies typically made use of higher flow velocities and Reynolds numbers. Take for

instance, in their investigation, Ko and Au (1985) reported initial turbulence levels of 0.5%

and 1.2% respectively for the primary and secondary jets. Buresti et al. (1994) also

made use of coaxial jets with initial turbulence levels of no more than 3% of the mean

primary jet velocity. In addition, in the study carried out by Tang and Ko (1994), the

8
initial turbulence intensity levels were observed to be approximately 0.45% and 0.55% for

the primary and secondary jets respectively. Note that these studies were air-based

studies. In contrast, the initial turbulence intensity levels observed here were closer to

earlier water-based studies by Sadr and Klewicki (2003), where 9% and 3% turbulence

levels for the primary and secondary jets respectively were ascertained. Secondly, the

significantly different experimental setup and apparatus across different investigations

made it difficult to achieve similar initial flow conditions, as could be seen readily in the

above studies even for air-based investigations. While it would not be entirely

appropriate to compare directly with earlier investigations, comparisons between the

different test cases within the present study, as well as against generally accepted flow

phenomena, remained viable.

(d) Flow visualization and DPIV techniques

For flow visualizations, LIF technique using fluorescein disodium salt was used. The

absorption wavelength of this fluorescent dye was approximately 494nm for maximum

absorption, while its emission wavelength was about 520nm. The fluorescent dye was

premixed and released into the secondary jet fluid through gravity-feed immediately after

the electromagnetic flow meter but before it entered the coaxial jet apparatus to allow

uniform dye distribution within the jet flow. A 5W, 532nm wavelength Laser Quantum

Elite diode-pumped solid-state continuous-wave laser with beam-steering optics and a

rotating-mirror were used to produce thin laser sheets (i.e. approximately 2mm thick) in

the streamwise direction. Despite the discrepancy between the laser and dye absorption

wavelengths, the emission level of the excited dye remained adequate for flow

visualization purposes. It would appear as green when illuminated by the continuous-

wave laser sheet along the nozzle centerline in the streamwise direction. Note that it

appeared as green not because of a direct consequence of using a green wavelength

laser, since the same laser caused rhodamine to fluoresce orange in Tsioli and New

(2009). All flow visualizations were captured using a 3CCD colour video camera with a

17X TV zoom lens recording to a digital video recorder. Recorded digital videos were

then transferred to a workstation for playback analysis and image extraction.

9
A Dantec Dynamics 2D DC-PIV system was used to provide global velocity

measurements of the resultant jet flows. It consisted of a New-Wave Research 50mJ

double-pulsed Nd:YAG laser, 1600px by 1200px 8-bit grayscale CCD camera and

controlled by a workstation with synchronizer and image-grabber cards. The PIV system

was operated at its maximum acquisition rate of 15Hz with the CCD camera in double-

frame mode. The imaged area captured by the CCD camera measured approximately

163.3mm by 120.4mm which resulted in an imaging resolution of approximately

0.1mm/px. 20microns polyamid seeding particles of 1.03g/cm 3 density were used to

seed the water tank and jet water entirely before commencing the PIV experiments. The

seeding particles were non-fluorescent and had a refractive index of approximately 1.5.

Experiments were carried out according to the guidelines suggested by Keane and Adrian

(1992) and Raffel et al. (1998), with an estimated uncertainty level of ±1% for the velocity

vector components. Captured image-pairs were transferred after the experiments to the

workstation for post-processing using Dantec FlowManagerTM software, where a two-pass

multigrid cross-correlation interrogation scheme was used. The initial and final

interrogation window sizes were 64px by 64px and 16px by 16px respectively.

The average particle number in the final 16px by 16px interrogation window was

estimated to be between 5 to 6 particles. Due to the dense seeding used, up to 8

particles could usually be found in some areas, typically within the jet flow regions.

Global and local validation criteria were used to reject spurious vectors from the raw

velocity vector maps, before a 3-point by 3-point smoothing filter was used to provide the

final velocity vector maps. Turbulence intensity levels determined from these velocity

fields were estimated to be accurate within 2.4%. Note that procedures used to

estimate experimental uncertainties here followed similar practises adopted by New

(2009) and New and Tsovolos (2009) under comparable experimental conditions, which

in turned made use of methodologies from Moffat (1988).

10
3. Results and discussions

(a) Vortical structures and behaviour

Figure 4 shows instantaneous LIF flow images captured for the two inclined coaxial jets

along =0 plane when the velocity-ratio was increased from VR=0.5 to 2.0. Generally, it

can be observed that inclined ring-vortices were formed regularly along the shear layers

of primary and secondary jets, similar to single-stream inclined jets observed previously.

Furthermore, flow behaviour associated with incline-angles of 45 and 60 can be

discerned to be generally similar at the same velocity-ratio, though a larger incline-angle

tends to accentuate the flow behaviour. While the interactions between the primary and

secondary jets became more intense along both shorter and longer nozzle lengths with

an increase in the incline-angle, a comparatively larger mixing region still occurred along

the shorter nozzle lengths. This is likely due to the turning of the inclined vortex roll-ups

for both primary and secondary jets expected at relatively large incline-angles, which had

been observed by Wlezien and Kibens (1986) and New (2009) in single-stream inclined-

jets of circular and elliptic geometries. During the turning process, vortex roll-up sections

along the shorter nozzle lengths would accelerate downstream such that the overall

inclination of the vortex roll-ups would decrease. At the flow conditions used here, this

did not produce significant qualitative differences in the flow behaviour. For the two

incline-angles investigated under the flow conditions used here, the main flow behavioral

changes came from the velocity-ratio, followed by the incline-angle.

Nonetheless, it was observed that variations in the incline-angle brought about discernible

changes upon the vortex roll-up frequencies along the inner (i.e. primary-secondary jet

interface) and outer (i.e. secondary jet-ambient interface) shear regions, particularly along

the shorter nozzle lengths. These frequencies were determined from playback of flow

visualization video recordings as no measurements were taken for the purpose of

determining vortex formation frequency. While not as authoritative as frequencies

determined using HWA or LDA measurements, they would provide some insights into the

dependency of vortex formation frequency upon the incline-angle and velocity-ratio.

Note that all vortex formation frequencies were determined at a location close to the

11
nozzle exits, before vortex-pairing events occurred. Based on estimated uncertainty in

the flow velocities provided earlier, a typical uncertainty of approximately one video image

frame in discerning the vortex formation, image frame frequency of the PAL video format

used and the fastest vortex formation frequency observed, the uncertainty level of the

Strouhal numbers was estimated to be about 1.9%. As a benchmark, vortex formation

frequencies along the inner and outer shear regions of the non-inclined coaxial jets were

determined to have Strouhal numbers (St1=fD1/U1 and St2=fd/U2 respectively) of St1=1.79

and 2.13 and St2=0.37 and 0.29 at VR=0.5 and 1.0 correspondingly. Vortex formation

occurred too rapidly at VR=2.0 to be determined reliably, though it would presumably be

higher.

In contrast, for the 45 inclined nozzles, inner shear region vortex formation frequency

along the shorter nozzle lengths exhibited a non-linear relationship with velocity-ratio

changes, where the Strouhal numbers were St1=1.56, 1.70 and 1.55 for VR=0.5, 1.0 and

2.0 respectively. In contrast, when the incline-angle increased to 60, the formation

frequency decreased to St1=1.35, 1.39 and 1.43 correspondingly. Note that the

frequency increased consistently with velocity-ratio for the 60 inclined nozzles only.

The reduction in the vortex formation frequency for the 60 inclined nozzles was due to

the heightened levels of vortex-pairings shortly after the vortex roll-ups formed and

associated with a meandering of the primary jet, as will be described in greater detail

later. Note that the frequency of the jet meandering could not be discerned accurately

from the flow visualization videos and thus, its relationship to the vortex formation

frequency could not be established here.

As for vortex formation frequency along the shorter nozzle length outer shear region, it

decreased with increasing velocity-ratio for the 45 inclined nozzles, where the Strouhal

numbers were St2=0.26, 0.19 and 0.16 at VR=0.5, 1.0 and 2.0 respectively. For the 60

inclined nozzles, similar trend was observed where St2=0.21, 0.20 and 0.14 at

corresponding velocity-ratios. The comparative lower vortex formation frequency at

larger incline-angle could also be attributed to increased vortex-pairing events, similar to

the observations made for the inner shear region. Finally, along the longer nozzle

12
lengths, regardless of the incline-angle used, the vortex formation frequency increased

with velocity-ratio used. However, their variations were less sensitive towards the exact

incline-angle used. For instance, for the 45 inclined nozzles, the inner and outer shear

region had vortex formation frequencies of St1=1.19 and 1.58, and St2=0.13 and 0.31

respectively at VR=0.5 and 1.0. (Note that vortex formations could not be discerned

clearly along the longer nozzle lengths at VR=2 due to the highly turbulent flow field.

Hence, no attempts were made to approximate their frequencies due to the significantly

large uncertainties.) On the other hand, for the 60 inclined nozzles, the corresponding

frequencies were St1=1.03 and 1.58, and St2=0.11 and 0.19 respectively at VR=0.5 and

1.0. Comparing the preceding results, it is clear that vortex formation frequencies along

the shorter nozzle lengths were more sensitive towards changes in the incline-angle.

Before going into the details of the resultant vortex behaviour associated with these

inclined coaxial nozzles, it should be noted that three flow shear regions exist in coaxial

jets as shown in Fig. 3(c) - one along the primary jet shear layer and two along the

secondary jet inner and outer shear layers respectively. Consider the orientation of the

figures where jet flows were issued horizontally from left to right and focusing on the

region along the top-half of the coaxial jets, vortex roll-ups along the primary jet shear

layer would have anti-clockwise rotational sense. On the other hand, those along the

secondary jet inner and outer shear layers would have clockwise and anti-clockwise

rotational senses respectively. Clearly, the rotational senses of these differently located

vortex roll-ups would be opposite along the bottom-half of the coaxial jets. Returning to

Fig. 4, regardless of the incline-angle and velocity-ratio used, it can be discerned that

rolling-up of the primary and secondary jet shear layers occurred earlier along the shorter

nozzle lengths. In contrast, vortex roll-ups along the longer nozzle lengths would begin

to manifest only later downstream, especially when the incline-angle was large. The

vortex formation along the longer nozzle lengths was essentially displaced downstream in

the streamwise direction, which is similar to that observed in single-stream inclined jets.

At VR=0.5 for the inclined nozzles as shown in Figs. 4(a)(i) and 4(b)(i), shear layer

instabilities along the primary-secondary jet interface were formed earlier than those

13
along the secondary jet outer shear layer, once the jets exhausted from the nozzles.

This was partly due to the higher primary jet velocity and Reynolds number as compared

to the secondary jet, which produced earlier onset of wave instabilities along the primary-

secondary jet interface. Furthermore, both primary jet and secondary jet inner shear

layers coexisted within that interface, which would serve to enhance the onset of wave

instabilities. In comparison, only one shear layer region existed between the secondary

jet and the surrounding fluid. Flow observations also suggest that secondary jet shear

layer instabilities were caused by perturbations due to the primary jet shear layer

instabilities. The secondary jet shear layer instabilities eventually rolled-up into discrete

vortices which interacted and paired with their adjacent neighbours shortly thereafter.

These paired vortices do not appear to interact significantly with the primary shear layer

vortex roll-ups until much further downstream. Note also that there are no clear

observations of secondary jet inner shear layer vortex roll-ups due to the higher primary

jet velocity relative to that of the secondary jet. This effectively resulted in relatively

stronger primary jet shear layer vortex roll-ups which tended to dominate over those

associated with the secondary jet inner shear layer.

Increasing the velocity-ratio to VR=1.0 (see Figs. 4(a)(ii) and 4(b)(ii)) would in turn

produce correspondingly stronger coupling between the formation of the primary jet and

secondary jet inner shear layer vortex roll-ups. In this case where the primary and

secondary jet velocities are similar, the secondary jet inner shear layer vortex roll-ups

become more physically apparent as compared to the case at VR=0.5, though the results

strongly suggest that the primary jet shear layers vortex roll-ups were driving the flow

events. Mutual interactions between the three “trains” of vortex roll-ups can also be

observed to become more vigorous, particularly along the shorter nozzle lengths. As a

result, the primary jet became increasingly asymmetric towards the longer nozzle length

region at the same time. More intriguingly, regular flow events occurring along the

shorter nozzle lengths caused the primary jet column to take on a “serpentile”-shaped

outline.

14
To illustrate, Fig. 5 shows a typical LIF flow sequence taken for 45 inclined nozzles at

VR=1 with the general shape of the primary jet indicated in the last three flow images.

Corresponding result for the 60 inclined nozzles is not shown here for the sake of

brevity, due to relatively similar behaviour. The flow images show that secondary jet

outer shear layer vortex roll-ups tend to pair up together along the shorter nozzle lengths,

before proceeding to cross the primary-secondary jet interface and penetrate significantly

into the primary jet. During this process, they interacted further with the primary jet and

secondary jet inner shear layer vortex roll-ups to form a region of merged vorticity with

accompanying rapid reduction in its coherency. Despite the complex vortical

interactions, it is clear that the secondary jet, as well as ambient fluid entrained by the

secondary vortex roll-ups earlier, were entrained into the primary jet. Since the pairing

and merging of the vortex roll-ups was a regular process, the relatively large-scale

entrainment behaviour would occur periodically during the flow visualization experiments.

While relatively similar vortex-pairing and merging processes occurred along the longer

nozzle lengths, the entrainment of the merged vortices into the primary jet was

comparatively weaker and not in-phase with that along the shorter nozzle lengths.

Therefore, regular pairings of secondary jet vortex roll-ups along the shorter nozzle length

and their penetration into the primary jet resulted in a primary jet column which

meandered or flapped about its centerline along the incline-plane which produced the

above-mentioned “serpentile”-shaped jet column. It is worth pointing out that such a jet

column shape has been observed earlier in single-stream inclined jets by Webster and

Longmire (1997). Hence, it appears that single- and dual-stream inclined jets share

some common fundamental flow attributes.

When the velocity-ratio was further increased to VR=2.0 as shown in Figs. 4(a)(iii) and

4(b)(iii), significantly higher levels of flow activities along the shorter nozzle lengths over

those along the longer nozzle lengths were much more apparent. At this velocity-ratio,

the secondary jet velocity was faster than that of the primary jet and as a result,

secondary jet inner shear layer dominated over the primary jet shear layer. As such, the

rotational sense of the dominant vortex structures along the primary-secondary jet

15
interface at the shorter nozzle lengths was clockwise in the figure and opposite to the

case at VR=0.5. Interestingly, the primary jet column retained the “serpentile”-shaped

outline seen at VR=1.0, even though the flow behaviour responsible for it differed from

that seen earlier. Instead of merged primary jet and secondary jet outer shear layer

vortex roll-ups penetrating into the primary jet column at VR=1.0, the stronger clockwise-

rotating secondary jet inner shear layer vortex roll-ups were mainly responsible at

VR=2.0.

These clockwise-rotating vortex roll-ups promoted regular ejection of secondary jet fluid

to be entrained into the primary jet column and produced similar undulation of the primary

jet column shape. Furthermore, out-of-phase but similar behaviour along the longer

nozzle lengths also accentuated the “serpentile”-shaped primary jet column outline (in this

case, the secondary jet inner shear layer vortex roll-ups were rotating anti-clockwise),

though not to the extent observed along the shorter nozzle lengths. In general,

“serpentile-“shaped primary jet columns at VR=1.0 and 2.0 are deduced to be a result of

similar flow events along the shorter and longer nozzle lengths, with those associated

with the latter displaced further downstream to produce alternating flow behaviour. To

illustrate and appreciate the differences in the jet flow behaviour with changes in the

velocity-ratio due to different modes of vortical interactions between the primary and

secondary jets, schematics of the interpreted flow fields are shown in Figs. 6 and 7.

Figure 8 shows corresponding instantaneous LIF images of the same inclined coaxial

nozzles shown earlier in Fig. 4 but along the =/2 plane. Since this plane was non-

inclined, the resultant vortex structures and subsequent developments were expectedly

symmetric about the nozzle centerline along this view. No meandering of either the

primary or secondary jet was observed here as well. Similar to Fig. 4 shown earlier, at a

low velocity-ratio of VR=0.5, primary and secondary jet shear layer vortex roll-ups remain

distinguishable in the near-field region, before they paired up further downstream. As

the velocity-ratio increased to VR=1.0 and beyond, pairing and merging events became

more intense and occurred much nearer to the nozzle exits. Engulfment of these paired

and merged vortices into the primary jet was also regular and brought about enhanced

16
mixing between the primary, secondary and ambient fluids. Thus, the increase in the

velocity-ratio served to promote greater interactions and rapid growth of their mutual

mixing regions along =/2 plane as well. The overall flow behaviour along this plane in

fact resembled closely to conventional non-inclined coaxial nozzles. This can be seen in

Fig. 9 where time-sequenced flow images for 45 inclined nozzles at VR=1 are shown.

However, it is clear from Figs. 4 and 8 that the flow behaviour along the incline-plane (i.e.

=0 plane) was most intriguing and insightful.

Compared to earlier studies by Wlezien and Kibens (1986) and Webster and Longmire

(1997), the flow large-scale behaviour observed here was remarkably similar, despite the

fact that coaxial jets with different velocity-ratios were used in the present study. These

include inclined ring-vortices produced by both the primary and secondary jets, earlier

pairing and enhanced vortical interactions along the shorter nozzle lengths, as well as the

meandering primary jet column. Comparing between this and earlier studies, it is

postulated that the turning of the inclined ring-vortices in the immediate near-field region

brought about increased vortex-pairings, which in turn encouraged further pairing and

merging events between the primary and secondary jet structures. In addition,

increasing the incline-angle also led to corresponding increases in the spreading of jet

momentum along the shorter nozzle lengths, as indicated by the visually enhanced jet

spreads. The near-field interactions between ring-vortices produced by the primary and

secondary jets were also close to those observed by Dahm et al. (1992) and essentially

comprised of rigorous levels of vortex-pairing and merging. Of significant interest here is

the meandering primary jet as it may be possible to utilize the alternating engulfment

events along the shorter and longer nozzle lengths as some form of self-excitation to

confer general or even directional jet-mixing enhancements.

(b) Vectoring of primary jet momentum

To ascertain the flow mean behaviour observed in the visualization results presented

earlier, mean velocity vectors for the two different inclined coaxial nozzles along =0 and

/2 planes in the streamwise direction were taken and shown in Figs. 10 and 11. To

17
appreciate the developments of the cross-stream velocity distributions, only one out every

six vectors along the streamwise direction is shown. Reference vectors corresponding

to u/U=1 were included in both figures. It can be seen that higher levels of flow

interactions and vortex-pairings along the shorter nozzle lengths at VR=1.0 and 2.0

produced not only oscillatory motions in the primary jet column, but also led to vectoring

of the primary jet towards the longer nozzle lengths. To highlight the persistent

vectoring, streamlines emanating from the center of the primary jet for each configuration

and orientation were determined from the mean velocity fields in Figs. 10 and 11 and

overlaid upon them. These streamlines essentially showed the mean trajectories of fluid

particles once they exhausted from the primary jet center. While by no means an

exhaustive way to illustrate the full extent of the vectoring, they nonetheless shed some

light on the relationships between the vectoring and the jet flow configurations. From

Figs. 10(a) and 11(a), the extent of primary jet vectoring can be observed to increase with

velocity-ratio, as well as the incline-angle used. Hence, the mean velocity fields

obtained along =0 plane are consistent with the behaviour captured by the flow

visualization images presented earlier. Correspondingly then, no vectoring of the

primary jet column is observed for all jet flows along =/2 plane as shown in Figs. 10(b)

and 11(b), and agrees well with the symmetric vortex formation and flow behaviour

observed in the LIF flow images along the same plane earlier.

The availability of Figs. 10 and 11 also allows better appreciation of the velocity

distributions in the jet flows due to the differences in the primary and secondary jet

velocities. For instance, Fig. 10(a) demonstrates that peak flow velocities for the 45

inclined nozzles do not occur along the nozzle centerline due to the vectoring of the

primary jet. At VR=0.5, they tend to exist closer to the longer nozzle lengths. On the

other hand, as the velocity-ratio increases to VR=1.0 and 2.0, peak flow velocity locations

will gradually shift towards the shorter nozzle lengths. This is due to relatively enhanced

interactions between the vortex structures along the shorter nozzle lengths as seen

earlier in Fig. 5, where vortex-pairings and their subsequent ingestions into the primary jet

occurred regularly. In contrast, such interactions were more restrained along the longer

nozzle lengths. Increasing the incline-angle to 60 produced relatively similar trends,

18
except that peak flow velocities remain closer to the longer nozzle lengths at VR=1.0.

To confirm the preceding postulation, streamwise turbulence intensity levels were

determined from PIV measurements and shown in Fig. 12. They show comparatively

higher turbulent intensities along the shorter nozzle lengths at VR=1 and 2.0, which agree

well with the earlier flow visualization results depicting the onset of heightened vortex-

pairing and merging behaviour. Upon closer inspection, it is interesting to note that for

both inclined coaxial nozzles, turbulence intensity levels start out higher along the longer

nozzle lengths at VR=0.5. But as the velocity-ratio increases from VR=0.5 to 1.0,

turbulence levels along the shorter nozzle lengths increases drastically to reach

comparable levels. And as the velocity-ratio increases further to VR=2.0, turbulence

levels along the longer nozzle lengths begins to decrease below those along the shorter

nozzle lengths. Clearly, these observations indicate increasingly comparatively

favourable jet mixing characteristics towards the shorter nozzle lengths when the velocity-

ratio increases.

As can be deduced from the results presented here, the use of inclined exits in coaxial

nozzle flows offers another control parameter in the manipulation of the resultant jet

behaviour. It is clear that interesting flow behaviour observed previously in single

inclined jets, particularly jet vectoring, continues to manifest in inclined coaxial jets. As

such, they may prove to be useful for diversion of the primary jet along preferential

azimuthal directions, by controlling the velocity-ratio, incline-angle and possibly area-ratio,

3. Conclusions
The present study investigated jet flows exhausting from 45 and 60 inclined coaxial

nozzles at various velocity-ratios using LIF visualization and PIV measurement

techniques. Visualization results show that earlier and higher levels of vortex-pairings

and associated flow interactions occurred between inclined primary and secondary jet

vortex roll-ups along the shorter nozzle lengths, the extent of which was also generally

enhanced by the use of higher velocity-ratio and incline-angle. An interesting

phenomenon where the primary jet underwent oscillatory flapping motion and produced a

19
“serpentile”-shaped outline along the incline-plane, was observed for both inclined coaxial

nozzles at velocity-ratios of VR=1.0 and above. This is similar to that observed by

Webster and Longmire (1997) in single-stream inclined jets and deduced to be a result of

rigorous vortex-pairing and merging behaviour between the primary jet and secondary jet

shear layer vortex roll-ups along the shorter nozzle lengths at VR=1.0. Regular large-

scale ingestion of secondary and ambient fluid into the primary jet was responsible for its

undulating outline.

At VR=2.0 however, the “serpentile”-shaped outline was caused by stronger secondary

jet inner shear layer vortex roll-ups, which “pinched” and penetrated into the primary jet

asymmetrically along the shorter and longer nozzle lengths. In contrast, vortex

structures and behaviour observed along the plane orthogonal to the incline-plane are

highly symmetric and resemble closely to those observed in previous studies on non-

inclined coaxial nozzles. Lastly, mean velocity fields indicate that both inclined coaxial

nozzles caused the primary jet to vector towards the longer nozzle lengths. The extent

of the vectoring was observed to increase with both velocity-ratio and incline-angle, which

signifies that enhanced levels of vortex-pairings and interactions along the shorter nozzle

lengths are responsible.

Acknowledgments
The authors would like to acknowledge the support for the study by UK Engineering and

Physical Science Research Council under project grant EP/F003102/1.

References
Ahmed MR, Sharma SD (2000) Effect of velocity ratio on the turbulent mixing of confined,

co-axial jets. Exp Therm Fluid Sci 22:19-33

Au H, Ko NWM (1987) Coaxial jets of different mean velocity ratios, Part 2. J Sound

Vibrations 100:211-232

20
Balarac G and Métais O (2005) The near field of coaxial jets: A numerical study. Phys

Fluids 17:065102

Balarac G, Métais O, Lesieur M (2007) Mixing enhancement in coaxial jets through inflow

forcing: A numerical study. Phys Fluids 19: 075102

Burattini P, Talamelli A (2007) Acoustic control of a coaxial jet. J Turbul 8:N47

Buresti G, Talamelli A, Petagna P (1994) Experimental characterization of the velocity

field of a coaxial jet configuration. Exp Therm Fluid Sci 9:135-146

Champagne FH, Wygnanski IJ (1971) An experimental investigation of coaxial turbulent

jets. Int J Heat Mass Transfer 14:1445-1464

Dahm WJA, Frieler CE, Tryggvason G (1992) Vortex structure and dynamics in the near-

field of a coaxial jet. J Fluid Mech 241:371-402

Keane RD, Adrian RJ (1992) Theory of cross-correlation analysis of PIV images. Appl.

Sci. Res. 49:191-215

Kiwata T, Okajima A, Kimura S (2001) Flow visualization of vortex structure of an excited

coaxial jet. J Visual 4:99-107

Ko NWM, Au H (1985) Coaxial jets of different mean velocity ratios. J Sound Vibrat

100:211-232

Ko NWM, Kwan ASH (1976) The initial region of subsonic coaxial jets. J. Fluid Mech

73:305-332

Kwan ASH, Ko NWM (1977) The initial region of subsonic coaxial jets, Part 2. J. Fluid

Mech 82:273-287

Lim TT (1998) On the breakdown of vortex rings from inclined nozzles. Phys Fluids

10:1666-1671

Lu HY (1983) Effect of excitation on coaxial jet noise. AIAA J 21:214-220

Luff JD, Drouillard T, Rompage AM, Linne MA, Hertzberg JR (1999) Experimental

uncertainties associated with particle image velocimetry PIV based vorticity algorithms.

Exp. Fluids 26:36-54

Moffat RJ (1988) Describing the uncertainties in experimental results. Exp. Therm. Fluid

Sci. 1:3-17

New TH, Tsovolos D (2009) Influence of nozzle sharpness on the flow fields of V-notched

nozzle jets. Phys Fluids 21:084107

21
New TH (2009) An experimental study on jets issuing from elliptic inclined nozzles. Exp

Fluids 46:1139-1157

Park CJ, Chen L-D (1989) Experimental investigation of confined turbulent jets. I: single-

phase data. AIAA J 27:1506-1510.

Raffel M, Willert C, Kompenhans J (1998) Particle image velocimetry: A practical guide.

Springer, Berlin

Rehab H, Villermaux E, Hopfinger EJ (1997) Flow regimes of large-velocity-ratio coaxial

jets. J Fluid Mech 345:357-381

Ribeiro MM, Whitelaw JH (1979) Coaxial jets with and without swirl. J. Fluid Mech

96:769-795

Sadr R, Klewicki JC (2003) An Experimental Investigation of the Near Field Flow

Development in Coaxial Jets. Phys Fluids 15:1233-1246

Saiki Y, Suzuki Y, Kasagi N (2010) Active control of swirling coaxial jet mixing with

manipulation of large-scale vortical structures. Flow Turbul Combust DOI:

10.1007/s10494-010-9274-3

Talamelli A, Gavarini I (2006) Linear instability characteristics of incompressible coaxial

jets. Flow Turbul Combust 76:221-240

Tang SK, Ko NWM (1994) Experimental investigation of the structure interaction in an

excited coaxial jet. Exp Therm Fluid Sci 8:214-229

Timmerman BH, Skeen AJ, Bryanston-Cross PJ, Graves MJ (2009) Large-scale time-

resolved digital particle image velocimetry (TR-DPIV) for measurement of high subsonic

hot coaxial jet exhaust of a gas turbine engine. Meas Sci Technol 20:074002

Troolin DR, Longmire EK (2010) Volumetric velocity measurements of vortex rings from

inclined exits. Exp Fluids 48:409-420

Tsioli E, New TH (2009) Near-field vortex structures of inclined coaxial jets. Proceedings

of Sixth International Symposium on Turbulence and Shear Flow Phenomena, Seoul,

Korea, 499-504

Villermaux E, Rehab H (2000) Mixing in Coaxial Jets. J Fluid Mech 425:161-185

Warda HA, Kassab SZ, Elshorbagy KA, Elsaadawy EA (2001) Influence of the magnitude

of the two initial velocities on the flow field of a coaxial turbulent jet. Flow Meas Instrum

12:29-35

22
Webster DR, Longmire EK (1997) Vortex dynamics in jets from inclined nozzles. Phys

Fluids 9:655-666

Webster DR, Longmire EK (1998) Vortex rings from cylinders with inclined exits. Phys

Fluids 10:400-416

Wicker RB, Eaton JK (1994) Near field of a coaxial jet with and without axial excitation.

AIAA J 32:542-546

Wlezien RW, Kibens V (1986) Passive control of jets with indeterminate-origins. AIAA J

24:1263-1270

Wlezien RW, Kibens V (1988) Influence of nozzle asymmetry on supersonic jets. AIAA J

26:27-33

Table

Nozzle type Nozzle region Shear region VR=0.5 VR=1.0 VR=2.0

Inner (St1) 1.79 2.13 Too


Uniform nozzle
Non-inclined turbulent to
length Outer (St2) 0.37 0.29 determine

Shorter nozzle Inner (St1) 1.56 1.70 1.55


length Outer (St2) 0.26 0.19 0.16
45 inclined
Inner (St1) 1.19 1.58 Too
Longer nozzle
turbulent to
length Outer (St2) 0.13 0.31 determine

Shorter nozzle Inner (St1) 1.35 1.39 1.43


length Outer (St2) 0.21 0.20 0.14
60 inclined
Inner (St1) 1.03 1.58 Too
Longer nozzle
turbulent to
length Outer (St2) 0.11 0.19 determine

Table 1 Summary of vortex formation frequencies for the non-inclined and inclined coaxial

nozzles

23
List of captions
Fig. 1 Schematics of the experimental recirculating water-tank facility used in the present

study

Fig. 2 Design schematics of the (a) coaxial jet apparatus, (b) reference, (c) 45 inclined

and (d) 60 inclined nozzles used in the present study

Fig. 3 (a) 3D-view and (b) view from downstream of the inclined nozzles. Primary and

secondary jet shear layers associated with coaxial jets are illustrated in (c)

Fig. 4 Streamwise LIF flow images for the (a) 45 inclined and (b) 60 inclined jets at

VR=0.5, 1.0 and 2.0 along =0 plane

Fig. 5 Time-sequenced LIF images for 45 inclined nozzles along =0 plane at VR=1

depicting a typical large-scale ingestion process and resultant meandering primary jet

Fig. 6 Interpretations of the observed flow behaviour for the inclined nozzles at VR=1.0

Fig. 7 Interpretations of the observed flow behaviour for the inclined nozzles at VR=2.0

Fig. 8 Streamwise LIF flow images for the (a) 45 inclined and (b) 60 inclined jets at

VR=0.5, 1.0 and 2.0 along =/2 plane

Fig. 9 Time-sequenced LIF images for 45 inclined nozzles at VR=1 along =/2 plane

Fig. 10 Velocity distributions for the 45 inclined jets at VR=0.5, 1.0 and 2.0 along (a) =0

and (b) =/2 plane

Fig. 11 Velocity distributions for the 60 inclined jets at VR=0.5, 1.0 and 2.0 along (a) =0

and (b) =/2 plane

Fig. 12 Streamwise turbulence intensity distributions for the 45 and 60 inclined jets

along =0 plane

24
Fig. 1

Fig.2

25
Fig. 3

Fig. 4

26
Fig. 5

27
Fig. 6

28
Fig. 7

Fig. 8

Fig. 9

29
Fig. 10

Fig. 11

30
Fig. 12

31

View publication stats

You might also like