You are on page 1of 21

International Journal of Heat and Mass Transfer 167 (2021) 120832

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Impingement heat transfer on flat and concave surfaces by


piston-driven synthetic jet from planar lobed orifice
Yuan-wei Lyu, Jing-zhou Zhang∗, Jun-wen Tan, Yong Shan
College of Energy and Power Engineering, Key Laboratory of Thermal Management and Energy Utilization of Aircraft, Nanjing University of Aeronautics and
Astronautics, Nanjing 210016, China

a r t i c l e i n f o a b s t r a c t

Article history: The heat transfer characteristics of a single synthetic jet (SJ) issuing from planar lobed orifice are in-
Received 8 September 2020 vestigated experimentally in current study. Two specific planar-lobed orifices (petal-shaped orifice and
Revised 28 October 2020
arch-shaped orifice) and two typical targets (flat target and semi-cylindrical concave target) are taken
Accepted 12 December 2020
into consideration. For each lobed orifice, three configurations are included. In the petal-shaped orifices,
Available online 28 December 2020
the lobe numbers are chosen as N=4, 6 and 8 in turns. While in the arch-shaped orifices, the lobe num-
Keywords: ber is fixed as N=6 but the lobe aspect ratios are altered as AR=1, 2 and 3 respectively. All lobed orifices
Synthetic jet have the same exit area with the baseline round orifice. The experimental tests are performed under a
Heat transfer series of operational frequencies (ranging from 10Hz to 25Hz) and dimensionless jet-to-target distances
Planar lobed orifice (ranging from 2 to 14 for flat target, and from 6 to 14 for concave target). Under the present conditions, SJ
Flat and concave targets impingement heat transfer enhancement is mostly achieved by the planar lobed orifices when compared
Piston-driven actuator
to the baseline round orifice, which depends tightly on the lobe shape and target shape. Among current
petal-shaped orifices, the N=6 petal-shaped orifice is demonstrated to be a most promising orifice con-
figuration. With regard to current arch-shaped orifices, a moderate lobe aspect ratio is demonstrated to
be the superior. When compared to the flat target, SJ impingement heat transfer on the concave target
is seriously reduced, regardless of orifice shapes. Interestingly, the relationship of SJ impingement heat
transfer on the concave target, between the N=6 petal-shaped orifice and AR=2 arch-shaped orifice, is
found contrary to that on the flat target. On the concave target, the N=6 petal-shaped orifice does not
show its advantages on SJ heat transfer enhancement as appeared on the flat target. However, the AR=2
arch-shaped orifice still shows its positive role on SJ heat transfer enhancement obviously in comparison
to the baseline round orifice.
© 2020 Elsevier Ltd. All rights reserved.

1. Introduction with the flow morphology regimes, several heat transfer schemes
were also identified in the synthetic jet impingent (Valiorgue et al.
Recently, the synthetic jet (SJ) has received growing attention [9], Persoons et al. [10]). When the stroke length of SJ is large, both
in the applications of active flow control and highly-efficient ther- the primary vortex rings that form from periodic ejecting at the
mal management because of its unique zero-net-mass-flux prop- orifice and the secondary vortex rings that form from shearing and
erty (ZNMF) [1]. Also, due to its inherent coherence of ‘synthe- paring in the shear layer were contributive attributive for the heat
sized’ formation from the surrounding fluid directly, SJ generally transfer together. However, when the stroke length of SJ is small,
exhibits more sophisticated flow dynamics when compared to the the heat transfer on the target was mainly attributed to the sep-
traditional continuous jet [2–4]. arate propagating of a train of primary vortex rings. Recently, Xu
Referring previous investigations on the detailed flow fields and Wang [11] performed a PIV (Particle Image Velocimetry) mea-
(such as Krishnan and Mohseni [5], McGuinn et al. [6], Greco surement to reveal the combined effects of the stroke length and
et al. [7,8]), different flow morphology regimes could take place in the Reynolds number on synthetic jet vortex rings impinging onto
an impinging synthetic jet, dependent on the impinging distance, a solid wall. It was identified that the effect of stroke length on an
stroke length and Strouhal number of a jet tightly. In accordance impinging synthetic jet is mainly reflected in the vortex ring co-
herence before impacting the wall. The effect of Reynolds number
was more behaved for a small-stroke-length impinging synthetic

Corresponding author. jet. During the last decade, vast efforts had been devoted to illus-
E-mail address: zhangjz@nuaa.edu.cn (J.-z. Zhang). trating the multi-parameter influence on SJ heat transfer in a great

https://doi.org/10.1016/j.ijheatmasstransfer.2020.120832
0017-9310/© 2020 Elsevier Ltd. All rights reserved.
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

[31,32]). From detailed measurement or numerical simulation of


Nomenclature the three-dimensional flow fields, it had been well illustrated that
the chevrons or lobes could induce an array of streamwise vortices
A area (m2 ) in the near-filed of the jet, thus changing the azimuthal coher-
a diameter of semi-circular lobe (m) ence of round jet and promoting the turbulence level of approach-
AR aspect ratio of lobe ing jet toward the target surface. Subsequently, the heat transfer
d equivalent diameter of orifice (m) enhancement of jet impingement was generally obtained by the
d1 inner diameter corresponding to lobe valleys (m) use of these advanced nozzles. In the synthetic jet impingement,
d2 outer diameter corresponding to lobe peaks (m) the orifice shapes were also concerned by many researchers re-
D piston diameter (m) cently. Bhapkar et al. [33] studied the influences of orifice shape
Dc concave target diameter (m) on the thermal and acoustic performances of SJ impingement ex-
f operational frequency (Hz) perimentally, by taking several orifices (e.g. circular, square, rect-
h convective heat transfer coefficient (W/(m2 K)) angular and elliptical) into consideration. At small impinging dis-
H jet-to-surface distance (m) tances, the elliptical orifice was found superior to the other ori-
k thermal conductivity (W/(mK)) fices generally. However, at large impinging distances, both the cir-
ks thermal conductivity of heater sheet (W/(mK)) cular and square orifices were demonstrated better than the other
L piston stroke length (m) orifices. Mangate and Chaudhari [34] experimentally investigated
L0 synthetic jet stroke length (m) two non-circular orifices (diamond and oval). When compared to
Lc longitudinal length of concave target (m) the baseline circular orifice, the diamond orifice and the oval ori-
lc clearance between piston top dead and orifice plate fice could increase the maximum average Nu about 17% and 7%
(m) respectively at 200Hz. Further, they considered the multiple-orifice
N lobe number synthetic jet impingement wherein the orifices were arranged in
Nu Nusselt number a central-satellite mode [35]. It was found that the maximum
Q heat power (W) spatially-averaged heat transfer coefficient could be increased 12%
q heat flux (W/m2 ) by the use of multiple-orifice with respect to the single orifice. The
R radius from jet stagnation point (m) central orifice played a dominate role on heat transfer enhance-
Re characteristic Reynolds number of synthetic jet ment. Liu et al. [36] performed an experimental study to illus-
r radial direction trate the role of round-orifice diffusion on SJ impingement. With
s thickness of orifice plate (m), curvilinear direction respect to the non-diffusion orifice, the orifice diffusion with an
Sr Strouhal number appropriate diffused-angle was found to provide a 30% increase at
T temperature (K) best under small impinging distances. Iwana et al [37] proposed a
U0 characteristic synthetic jet velocity (m/s) combined active-passive device on the jet impingement enhance-
x,y,z directions ment, wherein triangular tabs and synthetic jets were integrated
together. The use of chevron nozzles on SJ heat transfer enhance-
Greek letters
ment was explored by Crispo et al. [38–40] and Lyu et al. [41] more
δ thickness of heater sheet (m)
recently. For the SJ with a large dimensionless stroke length, the
ρ density (kg/m3 )
chevron nozzle was identified to yield a 20% heat transfer increase
Subscripts at small jet-to-surface distances with respect to the round noz-
a relative to ambient zle. Its role on enhancing impingement heat transfer was more
aw relative to adiabatic wall demonstrated for the SJ with a small dimensionless stroke length,
b relative to back surface of insulator in a wider scope of impinging distance. Giachetti et al. [42,43] per-
cond relative to conduction formed experimental investigations on the heat transfer enhance-
eff relative to effective heat transfer ment by the use of multi-perforation synthetic jets under laminar
joule relative to joule heating and turbulence cross-flow situations. From their study, the sensi-
jet relative to synthetic jet tive parameters (such as jet frequency, piston amplitude displace-
l-av circumferentially line-averaged ment and cross-flow velocity) were evaluated. It was concluded
loss relative to heat loss that the heat transfer from the synthetic jet device can be ampli-
ref reference fied by 23%~175% among the concerned situations.
s-av spatially-averaged As the SJ was generated from its self-synthesized process, the
w relative to wall effects of orifice geometry on its impinging heat transfer were
demonstrated more intricate than the traditional continuous jet.
In particular, because of its inherent ‘synthesized’ feature, an ap-
variety (such as driving mode of actuator, operational frequency, parent temperature rise could take place in the SJ during impinge-
impinging distance, multiple-jet interaction, presence of cross-flow, ment process under small jet-to-target distances, wherein the im-
etc. [12–19]). The detailed of advance about SJ impingement was pinged warmer fluid could be partly suctioned into the actuator
comprehensively also reviewed by Krishan et al. [20] and Arshad cavity (Gillespie et al. [44]). To alleviate the temperature rise in
et al. [21] more recently. SJ impingement heat transfer, Bhapkar et al. [45] proposed a tech-
In common sense, the orifice or nozzle shape is a very impor- nologic way by modifying the orifice-cavity shapes. It was iden-
tant factor affecting the flow dynamics and heat transfer charac- tified that a properly-contoured orifice-cavity shape can improve
teristics of jet impingement. It was regarded as a major passive the SJ heat transfer characteristics significantly under small jet-to-
strategy for heat transfer enhancement in the continuous jet situ- target distances. Rylatt and O’Donovan [46,47] proposed another
ations (Carlomagno and Ianiro [22]). Two typical fluidic-excitation technologic way by equipping an additional ducting close to the
nozzles that gained much attention recently were the tabbed or SJ orifice. From experimental tests the role of additional ducting
chevron nozzle (Violato and Scarano [23], Violato et al. [24], Yu was identified in the SJ impingement, with a maximum heat trans-
et al. [25,26], Vinze et al. [27]) and the lobed nozzle (Martin fer increase of 27% in the vicinity of jet stagnation. Bhapkar et al.
and Buchlin [28], Sodjavi et al. [29], Trinh et al. [30], He and Liu [48] performed a PIV measurement on the radial wall jet formed

2
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

by a normal synthetic jet impingement. Under a small jet-to-target sionless jet-to-target distances. From current study, several inter-
distance, a large recirculation zone was clearly observed near the esting findings relating to the effects of orifice shape and target
stagnation region and in the wall jet region, indicating a strong surface shape on SJ heat transfer are brought forward.
confinement effect. Apparently, the temperature-rise phenomenon
in SJ impingement onto a concave target would behave more pro- 2. Experimental procedures
nouncedly when compared to the flat target, because the ‘confined’
effect of impinged fluid between the heated target and actuator 2.1. Test setup
orifice became more serious. Besides, the evolution of an imping-
ing SJ would be also affected by the re-circulated flow inside con- Fig. 1 displays the schematic diagram of the test setup. A
cave cavity. piston-driven actuator was utilized in the current study. Its main
In order to bring out the effects of orifice shape and target geometric parameters were summarized in Table 1, including the
surface shape on SJ impingement heat transfer performance more internal diameter (D), piston stroke (L), orifice plate thickness (s),
deeply, a series of experimental tests are conducted in the present and the cavity clearance (lc ) wherein the piston at its reciprocat-
work, by taking two specific planar-lobed orifices (namely, petal- ing top-dead. On this orifice plate, a central hole was made, with a
shaped orifice and arch-shaped orifice) and two typical targets diameter (d) of 10 mm for the baseline round orifice. The recipro-
(flat target and semi-cylindrical concave target) into considerations, cating movement of the piston was driven by a variable-frequency
with the use of a piston-driven actuator. The experimental tests are electric motor, through a crankshaft-rod mechanism. In the present
performed under a series of operational frequencies and dimen- tests, the operational frequency (f) varied from 10Hz to 25Hz. Re-

Fig. 1. Schematic diagram of experimental setup.

3
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 2. Schematic diagram of lobed orifices.

Table 1 Two specific planar lobed orifices were considered in this study,
Main geometric parameters of piston-driven SJA.
as shown in Fig. 2. One was referred as petal-shaped lobe orifice,
Parameter Value similar to that adopted by Martin and Buchlin [28] as well as He
Internal diameter (D) 52 mm and Liu [31,32] in continuous jet impingement. Another was re-
Piston stroke (L) 64 mm ferred as arch-shaped lobe orifice, similar to that adopted by Sod-
Orifice plate thickness (s) 1 mm javi et al. [29] and Trinh et al. [30] in continuous jet impingement.
Orifice diameter (d) 10 mm For the petal-shaped orifices, three configurations were taken into
Cavity clearance (lc ) 10 mm
consideration by altering the lobe numbers (N), as seen in Fig. 2(a).
Each petal took on a semi-circular shape with a diameter of a. Cor-
responding to the 4-lobe, 6-lobe and 8-lobe petal-shaped orifices,
Table 2
Characteristic velocities, Reynolds numbers and Strouhal numbers under different the diameters of petals were 5.5mm, 4mm and 3.1mm in turns.
excitation frequencies. For the arch-shaped orifices, the lobe number was fixed as 6, but
the aspect ratio of the lobe was changed. As seen in Fig. 2(b), the
Frequency (Hz) 10 15 20 25
arch took on a semi-circular shape with a diameter of a=2mm. By
U0 (m/s) 12.44 18.66 24.88 31.1
changing the inner diameter (d1 ) corresponding to the lobe valleys
Re 7820 11740 15650 19550
Sr 0.008 0.008 0.008 0.008
and the outer diameter (d2 ) corresponding to the lobe peaks, three
different aspect ratios of the lobe (defined as AR = 0.5(d2 − d1 )/a)
were obtained. They were 1, 2 and 3, respectively. All of the shaped
orifices had the same exit area as the baseline round orifice. In the
ferring our previous study [41], this piston-driven SJ had a large other ward, they had the same equivalent orifice diameter (d) as
stroke length (L0 ) of 1.24m, regardless of variation of the oper- the baseline round orifice. When compared to the baseline round
ational frequency. The characteristic velocity (U0 ) was varied lin- orifice, the lobed orifice enlarged the interaction length of the ori-
early along the operational frequency. Table 2 presents the charac- fice with the working fluid while remained the orifice exiting area.
teristic velocity (U0 ), Reynolds number (Re) and Strouhal number Two target surfaces were taken into consideration. The flat tar-
(Sr) of piston-driven synthetic jet in current tests. The variation of get had a square size with a side-length of 200 mm. For the semi-
operational frequency had rare influence on the Strouhal number cylindrical concave surface, as displayed in Fig. 3, the diameter (Dc )
for the piston-driven SJ. and the longitudinal length (Lc ) were designed as 100 mm and 200

4
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

variations at the specified monitoring points on the target surface


were all identified to approach their respectively saturate mean-
levels. Then, the sampling of the IR camera was triggered by using
a synchronizer device linking to the signal of synthetic jet actu-
ator. The sampling time was assigned as 20 seconds for the pur-
pose of eliminating the effect of temperature fluctuation in the un-
steady jet impingement. Thus 600 frames of thermal images were
obtained to determine the time-averaged temperature. From the
above measurement procedure, the time-averaged heat transfer co-
efficient and Nusselt number were determined as
(Q joule − Qloss )/A − qcond q joule − qloss − qcond
h(x, y ) = = (1)
Tw − Tre f Tw − Tre f
h(x, y )d
Nu(x, y ) = (2)
k
where Qjoule is the joule heating power, Qloss is the heat loss from
the target rear-surface to the surroundings. A is the heater-sheet
area. qjoule and qloss are the joule-heating heat flux and heat-loss
heat flux respectively. qcond is the tangential heat-conduction heat
flux in the heater sheet. Tw is the target surface temperature and
Tref is referred as the reference temperature. k is the working-fluid
thermal conductivity at the reference temperature.
In Eq. (1), Qjoule was directly determined from the records of
electric current (I) and voltage (V) imposing on the heater sheet.
Qloss was estimated with the use of an empirical relation relating
to the effective heat transfer coefficient (heff,b ) according to Li et al.
[49], as illustrated in Eq. (3). The tangential heat conduction was
treated according to Greco et al. [50], as illustrated in Eq. (4).
Qloss = he f f,b Ab (Tb − Ta ) he f f,b = 0.11(Tb − Ta ) + 12.5 (3)
where Ab is the surface area of insulator exposed to the environ-
ment. Tb is the average surface temperature on the insulator. Ta is
the ambient temperature. In the current tests, the heat loss was
Fig. 3. Schematic diagram of concave target test model. found to be within 6% of the joule heating power.
 
∂ 2 Tw ∂ 2 Tw
qcond = ks δ + (4)
mm, respectively. The target plate was a composite assembly by ∂ x2 ∂ y2
mounting a thin Inconel-625 sheet (0.025mm-thick) onto a 5mm- where δ is the heater-sheet thickness. ks is the heater-sheet ther-
thick perforated transparent plastic plate wherein a square-shaped mal conductivity.
opening (100 mm × 100 mm) was fabricated. It was fastened on a The selection of Tref in the determination of SJ impingement
traverse platform to make the jet-to-target distance (H) adjustable heat transfer coefficient was a notable issue. More recently, Lyu
continuously. The Inconel-625 sheet was heated by inputting the et al. [51] performed a series of experimental tests to study the
joule heating power (Qjoule ) through two copper bars that clamped temperature-rise phenomenon in a piston-driven SJ. By using three
on its bi-lateral edges. To reduce the heat losses from the rear sur- selections of the reference temperature, namely, the ‘real’ synthetic
face of target (opposite to the SJ impingement) to the environment jet temperature (Tjet ), the adiabatic wall temperature (Taw ) and the
as far as possible, an insulator box was set up. On the top of insu- ambient temperature (Ta ), they illustrated that Tjet -based and Taw -
lator box a rectangular opening (60 mm in y-direction and 100 mm based Nusselt numbers could be twice times of Ta -based Nusselt
in x-direction) was made for installing a same-size infrared glass. number. Generally, the Tjet or Taw might be more properly used as
This infrared window was perfectly transmissive to the IR camera reference temperature in the definition of convective heat transfer
operating at 8-14μm band of the infrared spectrum. Beside, on the coefficient [39,44,45], because the temperature-variation effect of
insulator outer-surface ten K-type thermocouples were fixed to de- SJ on the convective heat transfer coefficient definition was entirely
termine the average temperature on it, for the purpose of estima- or partly dispelled. However, as the temperature-variation effect of
tion of heat losses. SJ was extremely complicated, unfortunately, the accurate determi-
During the tests, the target surfaces that exposed to IR camera nation of Tjet or Taw was indeed a difficult problem in the applica-
were carefully treated in advance, by spraying a uniform thin-coat tions. For this cause, the driving temperature difference T=(Tw -
with a perfect emissivity of 0.96. On each target surface, four K- Ta ) was still a common selection in the SJ impingement situations.
type thermocouples were also used to monitor the heat transfer As the temperature-variation effect of SJ was included in the Ta -
process of SJ impingement. The temperature measurement errors, based convective heat transfer coefficient, this specific definition
for IR camera and K-type thermocouple, were pre-calibrated form reflected the comprehensive performance of ‘heat removal’ by the
respective tests that had been well described in our previous study synthetic jet impingement, in accordance with the inherent ‘syn-
[41]. thesized’ feature of SJ. Furthermore, it was noted that the current
study mainly focused on illustrating the influences of orifice shape
2.2. Measurement and data reduction and target surface shape on SJ impingement heat transfer perfor-
mance, the conclusions relating the influencing roles were not af-
The heat transfer measurement was conducted under a quasi- fected by the definition scheme of Nusselt number. Therefore, in
steady state of SJ impingement process, wherein the temperature current study the reference temperature (Tref ) was selected as Ta .

5
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 4. Local time-averaged Nu contours on flat target at f=15Hz for round orifice.

Table 3
Measured or used parameters and corresponding uncertainties.

Parameter Error source Maximum uncertainty


q joule
qjoule Electric voltage measurement (V), ±2%Electric current measurement (I), ±1%Effective area of heating foil, ±1% q joule −qloss
= ±2.6%
q s
qloss Temperature measurement (Tb ), ±0.5°CTemperature measurement (Ta ), ±0.5°CEmpirical relation (hnb ), ±10% q joule −qloss
= ±1.0%
Tw
Tw Temperature measurement, ±1.0°C Tw −Tre f
= ±5.6%
Tre f
Tref Temperature measurement, ±0.5°C Tw −Tre f
= ±2.8%
k
k Temperature effect, ±2% k
= ±2.0%
d
d Manufacture error, ±1.5% d
= ±1.5%

Table 4
Test arrangements and corresponding descriptions.

No Orifice - Target Description

1 round orifice - flat target f=10-25Hz, H/d=2-14


2 petal-shaped orifice - flat target f=10-25Hz, H/d=2-14, N=4, 6 and 8
3 arch-shaped orifice - flat target f=10-25Hz, H/d=2-14, AR=1, 2 and 3
4 round orifice - concave target f=10-25Hz, H/d=6-14
5 petal-shaped orifice - concave target f=10-25Hz, H/d=6-14, N=6
6 arch-shaped orifice - concave target f=10-25Hz, H/d=2-14, AR=2

6
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 5. Effects of impinging distance on line-averaged Nu for round orifice on flat


target.

In current study, two types of averaged Nusselt numbers were


applied, referring as line-averaged Nusselt number (Nul-av ) and
spatially-averaged Nusselt number (Nus-av ). The former was de-
termined in a circumferential line around the jet stagnation and
the latter was determined in a specified region with a radius of
R=4d. For the concave surface, local Nusselt number distribution
was firstly transferred on a two-dimensional flat plane and then
the averaged Nusselt numbers were calculated.
Regarding Eqs. (1)–(4), the main affecting parameters, as well
as their respective uncertainties involved in the uncertainty anal-
ysis of Nu measurement, were summarized in Table 3. They were
qjoule , qloss , Tw , Tref , k and d. For the joule heating power per unit
area (qjoule ), its uncertainty was determined from instrument ac-
curacy (such as electric current I, electric voltage U) and direct
measurement (such as heater-sheet area A). Considering that the Fig. 6. Effects of operational frequency on line-averaged Nu for round orifice on flat
heat loss was within 6% of the joule heating power, the maximum target.
uncertainty of qjoule was estimated to be ±2.6%. As the heat loss
(qloss ) was estimated approximately by using the empirical rela-
tion, on account that the empirical relation on determining effec- the target surface and the surrounding ambient was mostly be-
tive heat transfer coefficient had an accuracy of 90%, the maximum yond 20°C. The maximum uncertainties induced by independent
uncertainty induced by qloss was then estimated to be ±1.0%. In Tw and Tref were estimated to be ±5.8% and ±2.8%, respectively.
the current tests, the measured temperature difference between The uncertainty in the thermal conductivity of working fluid due

7
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 7. Local time-averaged Nu contours on flat target at f=15Hz and H/d=2 for petal-shaped orifices.

to temperature effect was determined to be ±2.0% according to ranged, as displayed in Table 4. They could be really classified into
the ambient temperature fluctuation during tests. The uncertainty two groups, according to the target shapes. In the present tests,
in the nozzle diameter due to manufacture error was determined the dimensionless jet-to-target distance (H/d) was adjusted from 2
to be ±1.5% from the direct measurement. Therefore, by using the to 14 for the flat-target situations. While for the concave-target sit-
error-transfer methodology [52], the maximum uncertainty of Nus- uations, the dimensionless jet-to-target distance was only adjusted
selt number was estimated from Eq. (5) to be ±7.3% approximately, from 6 to 14, due to the geometric interference of SJA with the
which occurred at a small jet-to-surface distance of H/d=2 and op- concave cavity at small impinging distances.
eration frequency of f=25Hz.
 2  2  2  2  2  2
Nu q joule qloss Tw Tre f d k
=± + + + + + (5)
Nu q joule − qloss q joule − qloss Tw − Tre f Tw − Tre f d k

3. SJ Impingement on flat target

2.3. Test arrangements 3.1. Baseline round orifice

In the current study, both the orifice shape and the target shape Fig. 4 presents local Nu contours on the flat surface at dif-
were concerned. Totally, six orifice-target combinations were ar- ferent dimensionless impinging distances under f=15Hz (corre-

8
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 8. Local time-averaged Nu contours on flat target at f=15Hz and H/d=8 for petal-shaped orifices.

sponding Re=11740). Fig. 5 displays the influence of H/d on line- fied to be f=25Hz. Generally, as the jet-to-target distance in-
averaged Nu distributions for the baseline round orifice. It is seen creases, the suitable operational frequency of SJA is increased
from Fig. 4(a) that a ‘double peaks’ pattern in the local Nu con- accordingly.
tours appears at H/d=2, same as that demonstrated by previous The above heat transfer characteristics are interpreted here, in
researches [19,39,41]. Seen from Fig. 5, the optimal dimensionless the viewings of flow morphology regimes and temperature varia-
jet-to-target distance is identified to be H/d=8 at f=15Hz and re- tion effects together. Referring to Valiorgue et al. [9] and Persoons
placed by H/d=12 at f=25Hz (corresponding Re=19550). Generally, et al. [10], the impingement heat transfer of a large-stroke-length
SJ impingement heat transfer at H/d=2 is the worst among the cur- SJ (a threshold value of L0 /H is suggested to be 2.5 by Valiorgue
rent range of jet-to-target distances. et al. [9] and Persoons et al. [10]) is dominated by the primary
Fig. 6 shows the effect of frequency on line-averaged Nu dis- vortex rings and the secondary vortex rings simultaneously. With
tributions for the baseline round orifice. Under a small imping- regard to the present piston-driven SJ, its stroke length is 1.24m,
ing distance (i.e. H/d=2), f=15Hz is confirmed to be an op- far larger than the threshold value of L0 /H, therefore, it is cer-
timal working frequency, and f=10Hz (corresponding Re=7820) tain that the current SJ is a large-stroke-length SJ. As its charac-
is the next. Under a moderate impinging distance (i.e. H/d=8), teristic velocity is varied along f in a nearly linear mode, the in-
f=15Hz may also be regarded as an appropriate value, how- crease of f promotes synthetic Reynolds number and subsequently
ever, higher operational frequencies are demonstrated to be su- would enhance the coherence intensity of vortical structures as
perior to f=10Hz. When r/d exceeds 2, the line-averaged Nu well as the development capacity of the approaching jet toward
achieved at f=20Hz is even larger than that at f=15Hz. Un- the target. It exhibits a positive aspect for SJ impingement heat
der H/d=14, the appropriate operational frequency is identi- transfer. However, the increase of f would aggravate the degrada-

9
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 9. Comparison of line-averaged Nu between petal-shaped orifices and round orifice on flat target.

tion of SJ quality due to the temperature-rise effect, presenting 3.2. Petal-shaped orifices
a negative aspect for SJ impingement heat transfer on the other-
wise [44,45,51]. Therefore, SJ impingement heat transfer is appar- Fig. 7 displays local Nu contours on the flat target under H/d=2
ently associated with the competition of the above two aspects, and f=15Hz, for the petal-shaped orifices. As shown in Fig. 7(a),
namely SJ capacity and SJ quality. Under a small H/d, the temper- a clear ‘satellite-similar’ pattern of the local Nu distribution is
ature variation effect will become more pronounced. The SJ qual- demonstrated in the 4-lobe petal-shaped orifice SJ impingement.
ity would be severely affected when the operational frequency is Four separated islands that occupy high heat transfer coefficient
too large. On the other hand, the SJ capacity would be severely are distributed around the jet stagnation, in accordance with the
weakened when the operational frequency is too small wherein respective lobe valley of 4-lobe petal-shaped orifice. Another cen-
the corresponding synthetic Reynolds number is abated. In such tral island that occupies a high heat transfer coefficient is appeared
a situation, moderate operational frequencies (also moderate syn- in the jet stagnation. Looking at the petal-shaped orifice configura-
thetic Reynolds numbers) would be appropriate because both SJ tion (as seen in Fig. 2(a)), the lobe valley between adjacent petals
capacity and SJ quality could be well matched. Under a large H/d, seems similar to a tilted tab with respect to the baseline orifice.
as the heated target is far away from the orifice, the warm-air Referring Reeder and Samimy [53], the tab acts as a vortex gen-
suction is effectively alleviated. In such a situation, the SJ qual- erator to perturb vortical motions in the evolution of a jet. Due
ity is rarely affected by the operational frequency so that the SJ to the tab-excited streamwise vortices shedding from the lobe val-
impingement heat transfer is completely dominated by its native ley, the local heat transfer is enhanced, agrees well with the pre-
coherence of vortical structures. Subsequently, as the operational vious findings in a continuous jet impingement study performed
frequency increases, the SJ impingement heat transfer is increased by Gao et al. [54]. For the 6-lobe petal-shaped orifice, as displayed
monotonously. in Fig. 7(b), the local Nu in the vicinity of jet stagnation is obvi-

10
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 10. Local time-averaged Nu contours on flat target at f=15Hz and H/d=2 for arch-shaped orifices.

ously improved in comparison with the 4-lobe petal-shaped ori- downstream region. Also, seen from Fig. 8, it is identified that
fice. Besides, the adjacent satellite-type islands that occupy high the 6-lobe petal-shaped orifice produces the highest heat transfer
heat transfer coefficient are connected together gradually. Around among the current petal configurations.
the jet stagnation, six satellite-type islands that occupy small heat Fig. 9 shows a direct comparison of the line-averaged Nu dis-
transfer coefficient are still observed in accordance with the re- tributions on flat target between the petal-shaped orifices and the
spective lobe peak of 6-lobe petal-shaped orifice. With regard to baseline round orifice, under some cases. It is confirmed that the
the 8-lobe petal-shaped orifice, it is found that the local Nu in the petal-shaped orifice is able to play a positive role in the synthetic
vicinity of jet stagnation is significantly reduced when compared to jet heat transfer, tightly dependent on petal-shaped orifice design.
the other petal configurations. In addition, as seen in Fig. 7(c), the In general, the 6-lobe petal-shaped orifice is a most promising ori-
Nu distribution remains ring pattern with a ‘double peaks’ feature, fice configuration.
similar to that of the round orifice. Looking at the petal-shaped
orifice configuration, with the increase of lobe number, the tilted 3.3. Arch-shaped orifices
length of lobe valley is relevantly reduced. As a consequence, the
vortex pair is conjectured to be increased but the vortex scale re- Figs. 10 and 11 display local Nu contours on the flat target at
duced. When compared to the baseline round orifice, the ‘double f=15Hz for the arch-shaped orifices, under H/d=2 and H/d=8, re-
peaks’ feature in 8-lobe petal-shaped orifice is more distinct. spectively. Under a small jet-to-target distance, the ‘lobe-shaped’
Fig. 8 displays local Nu contours on the flat target under H/d=8 pattern of Nu contours is vigorously displayed for all arch-shaped
and f=15Hz, for the petal-shaped orifices. Under a larger jet-to- orifices, resembling the orifice configuration. The lobed pattern of
target distance, the characteristic pattern in local Nu contours takes Nu distribution is staggered with the lobed geometry of orifice. Be-
on an axisymmetric ring shape, because the lobed jet issuing from tween two adjacent arch-shaped lobes, the shearing vortices are
a petal-shaped orifice develops to an axisymmetric jet in the far coupled together in the synthetic jet development so that a higher

11
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 11. Local time-averaged Nu contours on flat target at f=15Hz and H/d=8 for arch-shaped orifices.

convective heat transfer is resulted. Under H/d=8, the character- and Chaudhari [35]. Besides, as the inner diameter corresponding
istic ‘lobe-shaped’ pattern of Nu contours disappears, owing to to the lobe valleys is reduced, the lobe valley is narrowed down
the same cause mentioned in the petal-shaped orifices. Seen from to affect the vortices-synthesize from this region. Therefore, in
Figs. 10 and 11, it is confirmed that a moderate lobe aspect ra- the lobed orifice design, the lobe-aspect-ratio should be carefully
tio (AR=2) provides the best synthetic jet impingement heat trans- concerned.
fer among the current arch-shaped orifices. Referring Fig. 2(b), as Fig. 12 shows a direct comparison of the line-averaged Nu
the arched-lobe has the same width, thus the increase of lobe as- distributions on flat target between the arch-shaped orifices and
pect ratio leads to a relative reduction in the inner diameter (d1 ) the baseline round orifice, under some cases. Similar to that in
that corresponds to the lobe valleys but a relative increase in the the petal-shaped orifices, the properly-designed arch-shaped ori-
outer diameter (d2 ) that corresponds to the lobe peaks. With re- fice exhibits an obviously positive action on the SJ heat transfer
spect to AR=2, the smaller lobe aspect ratio (such as AR=1) re- enhancement with respect to the baseline round orifice, in partic-
sults in a reduction of the interaction length of the orifice with ular the 6-lobe arch-shaped orifice with a moderate lobe aspect
the working fluid, making the fluidic excitation weaker. However, ratio.
if the lobe aspect ratio is too big (such as AR=3), the circum-
ferential interval between adjacent lobe peaks is enlarged to af- 3.4. Comparison between N=6 petal-shaped orifice and AR=2
fect the vortices integration. In some sense, the lobed-orifice with arch-shaped orifice
a big lobe aspect ratio seems more close to a central-satellite
multi-orifice in the geometric configuration, wherein the lobes act Form the above mentions, it is confirmed that N=6 petal-
as the satellite orifices. Looking at Fig. 10(c), it is identified that shaped orifice and AR=2 arch-shaped orifice remove the most heat
the central part of lobed-orifice plays a dominate role on SJ im- in their respective categories. Therefore, they are chosen here for
pingement, agreeing well with the previous finding of Mangate the purpose of direct comparison.

12
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 12. Comparison of line-averaged Nu between arch-shaped orifices and round orifice on flat target.

Fig. 13 presents direct comparisons of line-averaged Nu dis- f=20Hz and f=25Hz for the lobed orifice is also identified oppo-
tributions on the flat target between N=6 petal-shaped orifice site to the round orifice, either under H/d=2 or H/d=8, by compar-
and AR=2 arch-shaped orifice at two operational frequencies. ing Figs. 14 and 6. These differences are conjectured to be caused
Fig. 14 presents direct comparisons of line-averaged Nu distribu- by dual effects of the shaped orifice on the synthetic jet impinge-
tions between N=6 petal-shaped orifice and AR=2 arch-shaped ment, also referring as SJ capacity and SJ quality. For the lobed ori-
orifice under three dimensionless impinging distances. Although fice, as the streamwise vortical structures are excited in the jet,
the influencing trends of operational frequency and jet-to-target the turbulence mixing would be enhanced (Nastase and Meslem
distance on the lobed-orifice SJ heat transfer are found gener- [55], Lyu et al. [56]). On the other hand, the orifice shape also
ally in agreement with corresponding ones observed in the base- affects the temperature variation in the impinging synthetic jet.
line round orifice, some differences are also demonstrated between From this comprehensive standpoint, the above-mentioned differ-
them. Taking N=6 petal-shaped orifice as an example, it is seen ences could be explained. For an instance, under a smaller jet-to-
from Fig. 13 that the relationship of SJ impingement heat trans- target distance, although the fluidic-excitation effect of lobed ori-
fer between H/d=4 and H/d=12 at f=15Hz, and the relationship fice is stronger than that in a far-field flow, the negative influence
between H/d=10 and H/d=14 at f=25Hz are opposite to the cor- of temperature-variation effect on the SJ quality would also be am-
responding ones in the baseline round SJ impingement (seen in plified. Referring Figs. 13(a) and 5(a), the SJ heat transfer near the
Fig. 5). Also, in the baseline round orifice situation, the stagnation- jet stagnation is greater under H/d=12 than that under H/d=4 at
nearby line-averaged Nu at f=10Hz is slightly higher than that at f=15Hz for the N=6 petal-shaped orifice, but the situation is op-
f=15Hz, under H/d=2 (seen in Fig. 6(a)). But in the lobed-orifice posite for the baseline round orifice. This change is attributed to
situation, as seen in Fig. 14(a), the line-averaged Nu numbers at the cause that the ‘net’ positive output of the shaped orifice on SJ
f=15Hz are distinctly greater than that at f=10Hz over the entire impingement behaves more pronouncedly under H/d=12 than that
target. The relationship of SJ impingement heat transfer between under H/d=4. Similarly, referring Figs. 14 and 6, the lobed-orifice

13
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 13. Effects of impinging distance on line-averaged Nu for two 6-lobe shaped
orifices on flat target.

SJ produces a higher heat transfer at f=25Hz than that at f=20Hz


under H/d=2 and H/d=8, opposite to the baseline round orifice.
With the increase of operational frequency, the vortical excitation
produced by the lobed orifice becomes stronger, which could pro-
vide a more pronounced ‘net’ positive role on the synthetic jet heat
transfer.
Evaluated from the spatially-averaged Nu in the specified region
with a radius of R=4d, on the flat target, the relative increases pro-
duced by N=6 petal-shaped orifice with respect to baseline round
orifice are generally a little larger than the AR=2 arch-shaped ori-
fice, as demonstrated in Fig. 15. At the most, the spatially-averaged
Nu is increased by 26% for the N=6 petal-shaped orifice and 21%
for the AR=2 arch-shaped orifice, at f=15Hz and H/d=10. It is also
noted that the active role of the shaped orifice on heat trans-
fer enhancement behaves more pronouncedly under the situations Fig. 14. Effects of operational frequency on line-averaged Nu for two 6-lobe shaped
wherein the dimensionless jet-to-target distance is departure from orifices on flat target.
the ‘optimal jet-to-target distance’ of the round-orifice synthetic optimal jet-to-target distance for the baseline round-orifice syn-
jet impingement in accordance with its respect operational fre- thetic jet impingement is increased with the operational frequency
quency. For the current piston-driven SJ, the stroke length and the (or Reynolds number). Beside, different from the continuous jet
Strouhal number are fixed. With the increase of the operational impingement, the effect of Reynolds number on SJ impingement
frequency, the characteristic velocity and Reynolds number are in- behaves more complicatedly, associated with the SJ capacity and
creased in a nearly linear mode. As discussed in Section 3.1, the SJ quality. At small and moderate jet-to-target distances, the ‘heat

14
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 15. Relative increases of spatially-averaged Nu by lobed orifices with respect to round orifice on flat target.

Fig. 16. Local time-averaged Nu contours on concave target at f=15Hz for round orifice.

15
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 17. Effects of operational frequency on line-averaged Nu for round orifice on Fig. 18. Comparison of line-averaged Nu between concave target and flat target for
concave target. round orifice.

16
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

removal’ outcome is not monotonously increased with the increase


of SJ Reynolds number. For these causes, the relative-increase of
spatially-averaged Nu produced by the use of shaped orifices is
tightly dependent on the jet-to-target distance and the operational
frequency (or Reynolds number). Generally, the dimensionless jet-
to-target distance that corresponds to the weakest relative-increase
of spatially-averaged Nu is increased. It appears under H/d=4 or 6
at f=10Hz (Re=7820), H/d=6 or 8 at f=15Hz (Re=11730), H/d=8 at
f=20Hz (Re=15640), and H/d=12 at f=25Hz (Re=19550).

4. SJ Impingement on concave surface

4.1. Baseline round orifice

Fig. 16 presents local Nu contours on the concave target at dif-


ferent dimensionless impinging distances under f=15Hz. It is seen
that the local Nu distribution takes on oval contours, with the
major axis located at the curvilinear axis of the concave target.
With respect to the SJ impingement on a flat target, it is also
found that the variation of SJ impingement heat transfer along
with the jet-to-target distance becomes weaker on the concave tar-
get. Fig. 17 presents the line-averaged Nu distributions at different
frequencies and jet-to-target distances, for the baseline round ori-
fice. As the operational frequency increases, the line-averaged Nu is
monotonously increased when the dimensionless jet-to-target dis-
tance is beyond 8.
Fig. 18 compares the line-averaged Nu distributions on con-
cave target and flat target, produced by the baseline round ori-
fice. When compared to the flat target, the SJ impingement heat
transfer is seriously reduced on the concave target. Referring pre-
vious investigations relating to the continuous jet impingement on
concave targets [57–60], the large scale circulation of fluid is iden-
tified to be an important flow feature inside the concave cavity,
affecting the flow stability and subsequently the heat transfer per-
formance. Due to the curvature effect, the concave target provides
a ‘nature’ confinement to the impinging jet, which is a very un-
favorable factor to the SJ impingement because the confinement
will make the exhaust or impinged fluid more difficult to escape
from the entrainment of SJ actuator during the suction stage, even
under a large impinging distance. Furthermore, the shearing and
pairing of vortical structures during synthetic jet development are
also conjectured to be weakened in the concave-surface situations
because of the presence of circulating flow inside the concave cav-
ity. Fig. 19 presents the relative reductions of spatially-averaged Nu
on concave target with respect to the flat target, for the round-
orifice SJ impingement. It is seen that the severe situations gener-

Fig. 20. Comparison of line-averaged Nu between concave target and flat target for
N=6 petal-shaped orifice.

ally occur at smaller jet-to-target distances and larger operational


Fig. 19. Relative reductions of spatially-averaged Nu on concave target with respect frequencies. At the most, the spatially-averaged Nu is reduced by
to flat target for round orifice. 25% on the concave target with respect to flat target. Only when

17
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 22. Relative reductions of spatially-averaged Nu on concave target with respect


to flat target for lobed orifice.

4.2. Lobed orifices

Two lobed orifices are involved in SJ impingement on the con-


cave target, referring as N=6 petal-shaped orifice and AR=2 arch-
shaped orifice, on account of their best heat removal roles in re-
spective categories.
Figs. 20 and 21 provide direct comparisons of line-averaged Nu
distributions between concave target and flat target for the N=6
petal-shaped orifice and AR=2 arch-shaped orifice, respectively. It
is also identified that the concave target weakens the SJ heat trans-
fer compared with the flat target. By comparing two shaped ori-
fices, it is found that the N=6 petal-shaped orifice produces a more
serious reduction of spatially-averaged Nu than the AR=2 arch-
shaped orifice, as displayed in Fig. 22. For the N=6 petal-shaped
Fig. 21. Comparison of line-averaged Nu between concave target and flat target for orifice, in most situations the spatially-averaged Nusselt numbers
AR=2 arch-shaped orifice. are reduced up to 25% on the concave target in comparison with
flat target. While for the AR=2 arch-shaped orifice, such a big re-
duction takes place rarely.
the operational frequency is small (such as f=10Hz and 15 Hz) and Fig. 23 presents direct comparisons of line-averaged Nu distri-
the jet-to-target distance is large (H/d=14) at the same time, the butions on the concave target between N=6 petal-shaped orifice
spatially-averaged Nu on concave target is close to that on flat tar- and AR=2 arch-shaped orifice at two operational frequencies. Inter-
get. estingly, on the concave target, the relationship of SJ impingement

18
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

Fig. 23. Comparison of line-averaged Nu between petal-shaped orifice and arch-shaped orifice on concave target.

Fig. 24. Schematic illustration of synthetic jet impingement on concave target.

heat transfer between two types of lobe-shaped orifices is altered shaped orifice, wherein the shearing and pairing of vortices be-
with respect to that on the flat target (seen in Fig. 13). On the have most strongly. Apparently, the central part of impinging jet
flat target, the N=6 petal-shaped orifice generally produces higher issuing from the arch-shaped orifice is a little farther away from
SJ impingement heat transfer than the AR=2 arch-shaped orifice. the heated target. For this reason, two beneficial factors are con-
However, on the concave target, the contrary relationship takes on. jectured to be involved in the arch-shaped-orifice SJ impingement
A possible reason for such a change is suggested qualitatively as on a concave target. When compared to the petal-shaped orifice,
the following. As schematically illustrated in Fig. 24, the interface the formation of SJ at the central part of arch-shaped orifice is a
of the arch-shaped orifice with the working fluid at lobe valley is little more weakly affected by the warm-fluid suction. Simultane-
more tilted inwards the orifice center when compared to the petal- ously, the evolution of SJ at the central part of arch-shaped orifice

19
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

der the situations wherein the dimensionless jet-to-target


distance is departure from the ‘optimal jet-to-target dis-
tance’ in accordance with the round-orifice synthetic jet im-
pingement.
(3) When compared to the flat target, SJ impingement heat
transfer is seriously reduced on the concave target regard-
less of orifice shapes, which is conjectured to be caused by
the confinement effect of concave target to SJ. For the base-
line round orifice, at the most the spatially-averaged Nu is
reduced up to 25% on the concave target with respect to
the flat target. Only when the operational frequency is small
(such as f=10Hz and 15 Hz) and the jet-to-target distance
is large (H/d=14), the spatially-averaged Nu on the concave
target is close to that on the flat target.
(4) On the concave target, the relationship of SJ impingement
heat transfer between the N=6 petal-shaped orifice and
Fig. 25. Relative increases of spatially-averaged Nu by arch-shaped orifice with re- AR=2 arch-shaped orifice is found contrary to that on the
spect to round orifice on concave target. flat target. The N=6 petal-shaped orifice does not show its
advantages on SJ heat transfer enhancement as appeared on
the flat target. However, the AR=2 arch-shaped orifice still
is also a little more weakly affected by the circulating flow inside
shows its positive role on SJ heat transfer enhancement ob-
the concave cavity. Overall, the arch-shaped orifice is a promising
viously when compared to the baseline round orifice. The
shape that is helpful for mitigating the negative effect of concave
relative increases are mostly greater than 10%. It is suggested
confinement to the SJ impingement.
that a proper lobed-orifice shape could mitigate the confine-
In comparison with the round-orifice synthetic jet impinge-
ment effect of concave target on SJ impingement.
ment on concave target, the N=6 petal-shaped orifice generates the
nearly same spatially-averaged Nu, within a maximum difference
less than 7%. On the concave target, the AR=2 arch-shaped ori- Authors’ statement
fice still shows its positive role on enhancing SJ impingement heat
transfer enhancement obviously, as displayed in Fig. 25. In most We would like to submit the enclosed manuscript entitled “Im-
situations, the spatially-averaged Nusselt numbers are increased pingement Heat Transfer on Flat and Concave Surfaces by Piston-
up to 10% with the use of arch-shaped orifice with respect to the Driven Synthetic Jet from Planar Lobed Orifice”, which we wish to
baseline round orifice. be considered for publication in International Journal of Heat and
Mass Transfer. Publication has been approved by all authors, none
5. Conclusions of the material presented in the paper is submitted or published
elsewhere, and the paper does not contain any information with
A series of experimental tests are conducted on SJ impinge- restricted access or proprietary content.
ment, by taking two specific planar-lobed orifices (namely, petal- We confirm that there are no known conflicts of interest associ-
shaped orifice and arch-shaped orifice) and two typical targets ated with this publication and there has been no significant finan-
(flat target and semi-cylindrical concave target) into considerations. cial support for this work that could have influenced its outcome.
For the petal-shaped orifices, three configurations with different We confirm that the manuscript has been read and approved by
lobe numbers (N=4, 6 and 8 in turns) are involved. For the arch- all named authors and that there are no other persons who satis-
shaped orifices, three configurations with different lobe aspect ra- fied the criteria for authorship but are not listed. We further con-
tios (AR=1, 2 and 3 in turns) are involved. All lobed orifices have firm that the order of authors listed in the manuscript has been
the same exit area with the baseline round orifice. The experimen- approved by all of us.
tal tests are performed under a series of operational frequencies
(ranging from 10Hz to 25Hz) and dimensionless jet-to-target dis- Declaration of Competing Interest
tances (ranging from 2 to 14 for flat target, and from 6 to 14 for
concave target). In the present study, the Ta -based Nu definition is No conflict of interest is associated with this publication.
adopted. From current study, the following conclusions are brought
forward.
Acknowledgment
(1) Under the present conditions, SJ impingement heat transfer
enhancement is mostly achieved by the planar lobed ori- The authors gratefully acknowledge the financial supports for
fices when compared to the baseline round orifice, which this project from the National Natural Science Foundation of China
depends tightly on the lobe shape and target shape. Among (grant No: 51776097) and Postgraduate research and practice inno-
current petal-shaped orifices, the N=6 petal-shaped orifice is vation project of Jiangsu Province (KYCX17 0280).
demonstrated to be a most promising orifice configuration.
With regard to current arch-shaped orifices, a moderate lobe References
aspect ratio is demonstrated to be the superior.
(2) On the flat target, the N=6 petal-shaped orifice is found to [1] A. Glezer, M. Amitay, Synthetic jets, Annu. Rev. Fluid Mech. 34 (2002)
503–529.
be a little more superior to the AR=2 arch-shaped orifice in
[2] B.L. Smith, G.W. Swift, A comparison between synthetic jets and continuous
general. With respect to the baseline round orifice, a maxi- jets, Exp. Fluids 34 (2003) 467–472.
mum relative increase of 26% in the spatially-averaged Nu is [3] T.M. Crittenden, A. Glezer, A high-speed compressible synthetic jet, Phys. Flu-
achieved by the use of N=6 petal-shaped orifice, at f=15Hz ids 18 (2006) 017107.
[4] X.M. Tan, J.Z. Zhang, S. Yong, G.N. Xie, An experimental investigation on com-
and H/d=10. The active role of planar-lobed orifices on SJ parison of synthetic and continuous jets impingement heat transfer, Int. J. Heat
heat transfer enhancement behaves more pronouncedly un- Mass Transf. 90 (2015) 227–238.

20
Y.-w. Lyu, J.-z. Zhang, J.-w. Tan et al. International Journal of Heat and Mass Transfer 167 (2021) 120832

[5] G. Krishnan, K. Mohseni, An experimental study of a radial wall jet formed [34] L.D. Mangate, M.B. Chaudhari, Heat transfer and acoustic study of impinging
by the normal impingement of a round synthetic jet, Eur. J. Mech.-B/Fluids 29 synthetic jet using diamond and oval shape orifice, Int. J. Therm. Sci. 89 (2015)
(2010) 269–277. 100–109.
[6] A. McGuinn, R. Farrelly, T. Persoons, D.B. Murray, Flow regime characterisation [35] L.D. Mangate, M.B. Chaudhari, Experimental study on heat transfer character-
of an impinging axisymmetric synthetic jet, Exp. Therm. Fluid Sci. 47 (2013) istics of a heat sink with multiple-orifice synthetic jet, Int. J. Heat Mass Transf.
241–251. 103 (2016) 1181–1190.
[7] C.S. Greco, A. Ianiro, T. Astarita, G. Cardone, On the near field of single and [36] Y.H. Liu, T.H. Chang, C.C. Wang, Heat transfer enhancement of an impinging
twin circular synthetic air jets, Int. J. Heat Fluid Flow 44 (2013) 41–52. synthetic air jet using diffusion-shaped orifice, Appl. Therm. Eng. 94 (2016)
[8] C.S. Greco, G. Cardone, J. Soria, On the behaviour of impinging zero-net– 178–185.
mass-flux jets, J. Fluid Mech. 810 (2017) 25–59. [37] T. Iwana, K. Suenaga, K. Shirai, Y. Kameya, M. Motosuke, S. Honami, Heat trans-
[9] P. Valiorgue, T. Persoons, A. McGuinn, D.B. Murray, Heat transfer mechanisms fer and fluid flow characteristics of impinging jet using combined device with
in an impinging synthetic jet for a small jet-to-surface spacing, Exp. Therm. triangular tabs and synthetic jets, Exp. Therm. Fluid Sci. 68 (2015) 322–329.
Fluid Sci. 33 (2009) 597–603. [38] C.M. Crispo, C.S. Greco, F. Avallone, G. Cardone, On the flow organization of a
[10] T. Persoons, A. McGuinn, D.B. Murray, A general correlation for the stagnation chevron synthetic jet, Exp. Therm. Fluid Sci. 82 (2017) 136–146.
point Nusselt number of an axisymmetric impinging synthetic jet, Int. J. Heat [39] C.M. Crispo, C.S. Greco, G. Cardone, Convective heat transfer in circular and
Mass Transf. 54 (2011) 3900–3908. chevron impinging synthetic jets, Int. J. Heat Mass Transf. 126 (2018) 969–979.
[11] Y. Xu, J.J. Wang, Digital particle image velocimetry study on parameter influ- [40] C.M. Crispo, C.S. Greco, G. Cardone, Flow field features of chevron imping-
ence on the behavior of impinging synthetic jets, Exp. Therm. Fluid Sci. 100 ing synthetic jets at short nozzle-to-plate distance, Exp. Therm. Fluid Sci. 106
(2019) 11–32. (2019) 202–214.
[12] M.B. Chaudhari, B. Puranik, A. Agrawal, Effect of orifice shape in synthetic jet [41] Y.W. Lyu, J.Z. Zhang, X.C. Liu, X.M. Tan, Experimental investigation on convec-
based impingement cooling, Exp. Therm. Fluid Sci. 34 (2010) 246–256. tive heat transfer induced by piston-driven synthetic jet with a transmission
[13] M.B. Chaudhari, B. Puranik, A. Agrawal, Multiple orifice synthetic jet for im- pipe, Exp. Therm. Fluid Sci. 104 (2019) 26–42.
provement in impingement heat transfer, Int. J. Heat Mass Transf. 54 (2011) [42] B. Giachetti, M. Fenot, D. Couton, F. Plourde, Influence of Reynolds number
2056–2065. synthetic jet dynamic in crossflow configuration on heat transfer enhance-
[14] X.M. Tan, J.Z. Zhang, Flow and heat transfer characteristics under synthetic jets ment, Int. J. Heat Mass Transf. 118 (2018) 1–13.
impingement driven by piezoelectric actuator, Exp. Therm. Fluid Sci. 48 (2013) [43] B. Giachetti, M. Fenot, D. Couton, F. Plourde, Influence of multi-perforation syn-
134–146. thetic jet configuration on heat transfer enhancement, Int. J. Heat Mass Transf.
[15] J.Z. Zhang, S. Gao, X.M Tan, Convective heat transfer on a flat plate subjected 125 (2018) 262–273.
to normally synthetic jet and horizontal forced flow, Int. J. Heat Mass Transf. [44] M. Gillespie, W. Black, C. Rinehart, A. Glezer, Local convective heat transfer
57 (2013) 321–330. from a constant heat flux flat plate cooled by synthetic air jets, J. Heat Transf.
[16] Y.H. Liu, S.H. Tsai, C.C Wang, Effect of driven frequency on flow and heat trans- 128 (2006) 990–1000.
fer of an impinging synthetic air jet, Appl. Therm. Eng. 75 (2015) 289–297. [45] U.S. Bhapkar, A. Srivastava, A. Agrawal, Proper cavity shape can mitigate con-
[17] A. McGuinn, D.I. Rylatt, T.S. O’Donovan, Heat transfer enhancement to an ar- finement effect in synthetic jet impingement cooling, Exp. Therm. Fluid Sci. 68
ray of synthetic jets by an induced crossflow, Appl. Therm. Eng. 103 (2016) (2015) 392–401.
996–1003. [46] D.I. Rylatt, T.S. O’Donovan, Heat transfer enhancement to a confined impinging
[18] L. Silva-Llanca, A. Ortega, Vortex dynamics and mechanisms of heat trans- synthetic air jet, Appl. Therm. Eng. 51 (2013) 468–475.
fer enhancement in synthetic jet impingement, Int. J. Therm. Sci. 112 (2017) [47] D.I. Rylatt, T.S. O’Donovan, The effects of stroke length and Reynolds number
153–164. on heat transfer to a ducted confined and semi-confined synthetic air jet, J.
[19] C.S. Greco, G. Paolillo, A. Ianiro, G. Cardone, L. de Luca, Effects of the stroke Phys. Conf. Ser. 525 (2014) 012012.
length and nozzle-to-plate distance on synthetic jet impingement heat trans- [48] U.S. Bhapkar, H. Yadav, A. Agrawal, PIV study of radial wall jet formed by a nor-
fer, Int. J. Heat Mass Transf. 117 (2018) 1019–1031. mally impinging turbulent synthetic jet, J. Flow Vis. Image Process. 26 (2019)
[20] G. Krishan, K.C. Aw, R.N. Sharma, Synthetic jet impingement heat transfer en- 99–126.
hancement-a review, Appl. Therm. Eng. 149 (2019) 1305–1323. [49] X.J. Li, J.Z. Zhang, X.M. Tan, Experimental and numerical investigations on con-
[21] A. Arshad, M. Jabbal, Y.Y. Yan, Synthetic jet actuators for heat transfer enhance- vective heat transfer of dual piezoelectric fans, Sci. China Technol. Sci. 61
ment-a critical review, Int. J. Heat Mass Transf. 146 (2020) 118815. (2018) 232–241.
[22] G.M. Carlomagno, A. Ianiro, Thermo-fluid-dynamics of submerged jets imping- [50] C.S. Greco, A. Ianiro, G. Cardone, Time and phase average heat transfer in sin-
ing at short nozzle-to-plate distance: a review, Exp. Therm. Fluid Sci. 58 (2014) gle and twin circular synthetic impinging air jets, Int. J. Heat Mass Transf. 73
15–35. (2014) 776–788.
[23] D. Violato, F. Scarano, Three-dimensional evolution of flow structures in tran- [51] Y.W. Lyu, J.Z. Zhang, Y. Shan, Temperature-variation effect of piston-driven syn-
sitional circular and chevron jets, Phys. Fluids 23 (2011) 124104. thetic jet and its influence on definition of heat transfer coefficient, Int. J. Heat
[24] D. Violato, A. Ianiro, G. Cardone, F. Scarano, Three-dimensional vortex dynam- Mass Transf. 53 (2020) 1057–1069.
ics and convective heat transfer in circular and chevron impinging jets, Int. J. [52] R.J. Moffat, Describing the uncertainties in experimental results, Exp. Therm.
Heat Fluid Flow 37 (2012) 22–36. Fluid Sci. 1 (1988) 3–17.
[25] Y.Z. Yu, J.Z. Zhang, H.S. Xu, Convective heat transfer by a row of confined air [53] M.F. Reeder, M. Samimy, The evolution of a jet with vortex-generating tabs: re-
jets from round holes equipped with triangular tabs, Int. J. Heat Mass Transf. al-time visualization and quantitative measurements, J. Fluid Mech. 311 (1996)
72 (2014) 222–233. 73–118.
[26] Y.Z. Yu, J.Z. Zhang, Y. Shan, Convective heat transfer of a row of air jets im- [54] N. Gao, H. Sun, D. Ewing, Heat transfer to impinging round jets with triangular
pingement excited by triangular tabs in a confined crossflow channel, Int. J. tabs, Int. J. Heat Mass Transf. 46 (2003) 2557–2569.
Heat Mass Transf. 80 (2015) 126–138. [55] I. Nastase, A. Meslem, Vortex dynamics and mass entrainment in turbulent
[27] R. Vinze, S. Chandel, M.D. Limaye, S.V. Prabhu, Local heat transfer distribu- lobed jets with and without lobe deflection angles, Exp. Fluids 48 (2010)
tion between smooth flat surface and impinging incompressible air jet from a 693–714.
chevron nozzle, Exp. Therm. Fluid Sci. 78 (2016) 124–136. [56] Y.W. Lyu, J.Z. Zhang, B.Y. Wang, X.M. Tan, Convective heat transfer on flat and
[28] R.H. Martin, J.M. Buchlin, Jet impingement heat transfer from lobed nozzles, concave surfaces subjected to an impinging jet form lobed nozzle, Sci. China
Int. J. Therm. Sci. 50 (2011) 1199–1206. Technol. Sci. 63 (2020) 116–129.
[29] K. Sodjavi, B. Montagne, A. Meslem, P. Byrne, L. Serres, V. Sobolik, Passive con- [57] C. Cornaro, A.S. Fleischer, R.J. Goldstein, Flow visualization of a round jet im-
trol of wall shear stress and mass transfer generated by submerged lobed im- pinging on cylindrical surfaces, Exp. Therm. Fluid Sci. 20 (1999) 66–78.
pinging jet, Heat Mass Transf. 52 (2016) 925–936. [58] A. Hashiehbaf, A. Baramade, A. Agrawal, G.P. Romanoet, Experimental investi-
[30] X.T. Trinh, M. Fenot, E. Dorignac, Flow and heat transfer of hot impinging jets gation on an axisymmetric turbulent jet impinging on a concave surface, Int.
issuing from lobed nozzles, Int. J. Heat Fluid Flow 67 (2017) 185–201. J. Heat Fluid Flow 53 (2015) 167–182.
[31] C.X. He, Y.Z. Liu, Jet impingement heat transfer of a lobed nozzle: measure- [59] S.N. Yasaswy, V. Katti, S.V. Prabhu, Experimental study on local heat transfer
ments using temperature-sensitive paint and particle image velocimetry, Int. J. distribution between smooth semi-cylindrical concave surface and impinging
Heat Fluid Flow 71 (2018) 111–126. air jet from a circular straight pipe nozzle, AIAA Paper 2009-4088, 2009.
[32] C.X. He, Y.Z. Liu, Large-eddy simulation of jet impingement heat transfer using [60] M.A. Nguepnang, M. Boer, T. Kim, Stagnation heat transfer on a concave surface
a lobed nozzle, Int. J. Heat Mass Transf. 125 (2018) 828–844. cooled by unconfined slot jet, J. Thermophys. Heat Transf. 30 (2016) 558–566.
[33] U.S. Bhapkar, A. Srivastava, A. Agrawal, Acoustic and heat transfer characteris-
tics of an impinging elliptical synthetic jet generated by acoustic actuator, Int.
J. Heat Mass Transf. 79 (2014) 12–23.

21

You might also like