For effective flow control, it is essential to thoroughly understand the vortical structures formed by the synthetic jet and boundary layer interaction (SJBLI), their effects near the surface, and their overall effectiveness in altering the flow dynamics [
106,
107]. Several experiments have demonstrated that synthetic jets effectively delay flow separation on aerodynamic bodies of various shapes [
108,
109,
110,
111,
112,
113,
114]. The complex SJBLI phenomenon is illustrated schematically in
Figure 5. An in-depth understanding of the SJBLI presents several challenges, including accurately characterizing the boundary layer, measuring the high-velocity gradients generated by the synthetic jets, and resolving small-scale rotational coherent structures [
115]. In addition to the complex flow physics, ample parameter space must be considered when designing an effective control strategy using SJAs. To establish a systematic approach for characterizing the effects of various geometrical and operational parameters, Jabbal and Zhong [
107] applied Buckingham’s
-theorem to an SJA embedded in a crossflow boundary layer. This analysis identified the following dimensionless parameters:
where
represents the Strouhal number
,
the freestream Reynolds number
,
the ratio of boundary layer thickness to orifice diameter,
the skin friction coefficient
, and
the blowing ratio
(also known as jet-to-freestream velocity ratio VR). The following inter-dependencies exist between these dimensionless groups and the key dimensionless parameters for an SJA in quiescent conditions:
It is important to note that this parametric study did not include all parameters relevant to flow control applications, essentially neglecting factors such as surface curvature, orifice shape or aspect ratio, and clustering effects [
116]. Moreover, quantifying momentum addition as a measure of the overall blowing strength was historically common, rather than using the jet-to-freestream velocity ratio [
117]. For a two-dimensional configuration, the momentum coefficient
may be defined as a ratio of the time-averaged expelled momentum by all the operating SJAs
to the momentum of the freestream:
where
is a reference area for the body under consideration and
T is the actuation period [
108,
118]. The definition of the momentum coefficient is inconsistent in the literature, with various definitions proposed by different researchers [
45,
47,
76,
119,
120]. Researchers have often assumed a uniform velocity distribution across the jet exit cross-section or ignored spatial variations by relying solely on the centerline velocity. Furthermore, these definitions often consider only the expelled momentum, integrating over the expulsion phase, which by also assuming a top-hat velocity distribution, leads to the following expression:
Also note the denominator in Equation (15b) and compare it with Equation (
16). Greenblatt and Wygnanski [
120] further simplified Equation (
16) by decomposing the jet velocity into mean and oscillatory components as
, deriving the following relation for the momentum coefficient:
The second term on the right-hand side of Equation (
17) may be neglected when it is relatively small compared to the first term. As highlighted by some researchers, the spatial velocity distribution can deviate significantly from the ideal top-hat profile under certain conditions [
58,
99,
100]. Therefore, if the top-hat profile is assumed, the uniformity of the velocity distribution should be validated.
4.1. Effects of Jet Strength
The strength of a synthetic jet relies on both
and
. As with SJAs in quiescent conditions, a threshold criterion exists for SJAs in crossflow to effectively enhance fluid mixing (see Figure 10 in Reference [
124]). Experimental evidence shows that the interaction of the vortices produced by a circular synthetic jet with a boundary layer leads to the formation of streamwise vortical structures, which can delay flow separation by drawing faster-moving fluid from the freestream into the near-wall region [
96,
132]. As these vortex rings enter the boundary layer, they experience combined effects from shear forces, boundary layer vorticity, and the Magnus force resulted from interactions with the crossflow [
112,
126,
133,
134]. These influences cause the vortex rings to tilt and deform at varying degrees, depending on their strength and residence time within the boundary layer, resulting in complex three-dimensional vortical structures.
The dye flow visualization by Zhong et al. [
126], with
ranging from 16 to 245 and
from 0.56 to 1.4, in a laminar boundary layer at blowing ratio
from 0.06 to 0.7, revealed three distinct behaviors:
At low blowing ratios and jet Reynolds numbers , the vortical structures generated by synthetic jets appeared as hairpin vortices attached to the wall.
At intermediate and values, the vortex sheet formed at the orifice rolled up into vortex rings, which experienced significant tilting and stretching as they entered the boundary layer.
At high and values, the vortex rings exhibited some tilting but little to no stretching, quickly penetrating the edge of the boundary layer.
Jabbal and Zhong [
107] observed the same three types of vortical structures using stereoscopic dye flow visualization. They also performed surface visualization of the vortex footprints by applying a temperature-sensitive liquid crystal (TLC) coating on the test surface. Interestingly, only two distinct types of thermal footprints were observed. Both hairpin vortices and stretched vortex rings produced two streamwise streaks of high heat transfer, corresponding to high-momentum fluid transfer toward the wall outboard of the streamwise counter-rotating legs based on the Reynolds analogy. In contrast, the tilted vortex rings generated only a single streamwise streak of high heat transfer, which was hypothesized to result from an induced vortex adjacent to the wall (see Figure 6 in Reference [
107] or Figure 7 in Reference [
124]). Q-criterion contours displaying the hairpin vortices, vortex rings, and near-wall vortices are presented in
Figure 6.
To understand the underlying flow mechanism, Jabbal and Zhong [
107] used two-dimensional PIV to study the flow fields of the three above-mentioned vortical structures. Hairpin vortices (
and
) and stretched vortex rings (
and
) exhibited characteristics similar to a streamwise vortex pair with a common upwash. In contrast, tilted vortex rings (
and
) featuring a common downwash induced a tertiary streamwise vortex pair in the near-wall region. Wall shear stress measurements revealed that stretched vortex rings provided the best performance in terms of higher near-wall fluid mixing, greater persistency, and reduced spatial fluctuations. The numerical work of Ho et al. [
131] investigated the effect of varying SJA jet momentum (
) in a turbulent boundary layer crossflow with three-dimensional URANS. An increase of jet momentum resulted in greater boundary layer penetration under the same actuation frequency. Note that, although in this study the increase of jet momentum resulted in a significant increase of wall shear stress along the streamwise direction, it also reduces the spanwise control authority of the SJA. Classification of structures emerging due to SJBLI is not unique and depends on the considered parameter space. For example, Zhong and Zhang [
112] described these vortical structures only based on blowing ratio
given a constant actuation frequency
f, first as hairpin-like vortices that are located close to the wall, and then, as
increases, as tilted vortex rings with a pair of trailing legs that penetrate the edge of the boundary layer shortly downstream. The numerical work of Zhou and Zhong [
123] of a rounded SJA in a laminar boundary layer provided a similar conclusion. For a constant value of orifice width and wall shear stress, parametric maps are available in the literature that classify synthetic jet or pulsed jet flow behavior based on the blowing ratio
(VR) and the jet Reynolds number
, or its alternative, the stroke ratio
[
121,
124,
135]. The generated vortex rings will undergo deformation due to the resident shear in the boundary layer for both pulsed and synthetic jets. For a synthetic jet, however, an additional suction effect is expected to be confined to the upstream branch of the vortex, leading to the formation of an asymmetric structure. According to Jabbal and Zhong [
124], a hairpin vortex forms as the upstream branch of the vortex ring produced by the synthetic jet is weakened by the suction cycle as it passes over the orifice and is then canceled out by resident vorticity of the opposite sign as it propagates downstream. Supposing the initial strength of the vortex ring at the orifice exit is relatively strong, the vortical structures may first appear as a stretched vortex ring with a weak upstream branch in the near-field before evolving into a hairpin structure further downstream. At higher blowing ratio
and stroke ratio
values, the vortex rings produced by synthetic jets can escape the influence of the suction cycle and the boundary layer’s resident shear, emerging as fully formed rings that appear tilted relative to the wall. Above a threshold stroke ratio
, the vortex rings generated by synthetic jets become increasingly incoherent over time, and secondary vortices can eventually be seen shedding from the primary vortex ring, similar to the observations by Jabbal et al. [
60] for SJAs in quiescent condition (see Figure 7 in Reference [
124]).
4.2. Effects of Orifice Shape and Orientation
Flow control with synthetic jets from non-circular orifices is of interest because of their enhanced entrainment and mixing capability compared to circular and two-dimensional synthetic jets. A schematic of an SJA with a rectangular orifice in crossflow is shown in
Figure 7. In addition to the pitch angle
, for only a non-circular orifice, a skew angle
must be defined to fully determine the orientation of the orifice with respect to the crossflow. Shuster et al. [
43] examined the orifice pitch angle for perpendicular (
) and inclined (
) circular synthetic jets subjected to a laminar boundary at two stroke ratios,
and 2. At the stroke ratio
, flow structure differences were significant between the perpendicular and inclined orifice orientations. For the perpendicular orifice, the vortex ring that formed at the orifice during the expulsion portion of the actuator cycle did not escape the near vicinity of the orifice and was subsequently ingested during the suction phase of the cycle. The incoming boundary layer diverted over and around this stationary vortex, creating a wake in the boundary layer downstream of the orifice. In contrast, for an inclined orifice at the same stroke length, a train of vortex rings originating from the orifice penetrated the crossflow (see Figures 5 and 8 in Reference [
43]). At the stroke ratio
, large discrete vortices were formed at the orifice for both orifice orientations that penetrated deep into the crossflow, well beyond the boundary layer edge. In general, the mean interaction from the perpendicular orifice was shorter in the streamwise direction but extended further into the crossflow, causing a slight deflection of the mean streamlines away from the wall. In contrast, the mean interaction from the inclined orifice extended farther downstream, though it remained confined to a region closer to the wall. Zhao et al. [
46] studied the flow over a national advisory committee for aeronautics (NACA) 0021 airfoil, instrumented with circular SJAs, at a maximum Reynolds number of
for a range of angle of attacks
. They observed that, for most configurations, when the momentum coefficient of SJAs was small, a larger pitch angle
was more effective in improving the maximum lift coefficient of the airfoil. Conversely, when the SJAs had a more significant momentum coefficient, a smaller pitch angle
was more effective (see Figure 11, Table 3, and Figure 16 in Reference [
46]).
As highlighted by Kim et al. [
137], the flow fields generated by circular orifices are significantly different from rectangular orifices, suggesting that their interactions with a boundary layer might also differ. Smith [
138] conducted a wind-tunnel experiment for an
rectangular SJA array expelled normally into a turbulent boundary layer at
, examining two orifice orientations, with the rectangular long side normal to (
) and aligned with the crossflow direction (
), respectively. The boundary layer in the former case exhibited a wake-like region due to the blockage effect, while in the latter, evidence suggested the presence of longitudinal vortices embedded in the boundary layer. Van Buren et al. [
139] used stereoscopic PIV to study
, 12, and 18 rectangular synthetic jets issued normally into a laminar boundary layer of
at
, 1.0, and 1.5. The flow field was characterized by two salient structures: a recirculation region downstream of the orifice and a steady streamwise vortex pair farther downstream, akin to the findings in Smith [
138] and Cui et al. [
140]. A conceptual model of these structures is shown in
Figure 8. These vortices interact with each other, as well as with the wall [
141], to eventually lift off of the surface. The spacing of the edgewise vortices, produced due to the finite span of the orifice, and the virtual blockage of the jet were reduced with the decreased AR.
Van Buren et al. [
136] used stereoscopic PIV to study
rectangular synthetic jets issued into a boundary layer of
at three momentum coefficients
, 0.33, and 0.75. The apparatus could accommodate different orifice pitch angles
, 45, 65, and 90 through separate inserts and could also be fixed at skew angles of
–90 every 15. It was found that orifice orientation significantly influenced both steady and unsteady flow structures. Different combinations of orifice skew and pitch angles led to the formation of various types of vortical structures downstream, including the absence of coherent vortex structures, a single strong vortex (either positive or negative), or a symmetric vortex pair. The phase-averaged Q-criterion iso-surfaces revealed that orifice pitch angles more aligned with the crossflow produced less coherent vortical structures as they contributed less to generating transverse velocities. In the time-averaged flow field, when the orifice pitch angle was less than 90, the upstream lengthwise vortex dominated the interaction, leading to a single negative vortex downstream. However, when the orifice had a wall-normal pitch, the downstream lengthwise vortex became dominant, as the upstream vortex was constrained by the jet blockage, resulting in a single positive streamwise-oriented vortex downstream (most clearly seen for
case, Figure 5 or Figure 6 in Reference [
136]). In general, the more wall-normal transverse jets had more blockage, which resulted in a large wake deficit downstream, while the lower-pitch angled jets were more aligned with the flow, generating only regions of higher velocity with no wake. Additionally, the virtual blockage of the jet decreased as the aspect ratio was reduced.
Wang and Feng [
40] employed time-resolved tomographic PIV to investigate the three-dimensional flow fields of an
rectangular synthetic jet and its interaction with a laminar boundary layer of
at a fixed blowing ratio
. Two typical orifice orientations, namely spanwise (
) and streamwise (
) configurations, were examined to analyze the evolution of the vortical structures and the relevant flow physics. The flow field was composed of three major structures: a tilted vortex ring, a secondary trailing vortex, and a tertiary near-wall vortex. This suggested different flow scenarios from those for high-AR cases in Van Buren et al. [
139] and Van Buren et al. [
136]. Compared with the tilted and distorted vortex rings for circular cases observed by Jabbal and Zhong [
107], the legs of the secondary trailing vortex were jointed in the spanwise direction at two different wall-normal heights. In contrast, the tertiary near-wall vortex was a crescent-shaped spanwise vortex instead of a streamwise vortex pair as in Jabbal and Zhong [
107] (see Figure 5 in Reference [
40]). Additionally, axis switching, a phenomenon typically observed in non-circular vortex rings as described in
Section 3.3, was also detected in low-AR rectangular synthetic jets in crossflow. The orientation of the orifice had a direct influence on the timing and location of the axis switching. As for the two different orientations, their observations suggested that the spanwise orientation is more efficient in energizing the boundary layer.
Elimelech et al. [
142] and Vasile and Amitay [
143] performed stereoscopic PIV to study the interaction of a high-AR rectangular SJA with a three-dimensional boundary layer over a finite 30 swept-back NACA 4421 airfoil at a Reynolds number of
and several angles of attack, where the SJA was operated at
and 1.2 and several frequencies. High-frequency pairs of spanwise vorticity rollers were generated and shed away from the SJA, while streamwise vortices formed at both edges of the SJA orifice, rotating in opposite directions. The streamwise vortices primarily influenced the time-averaged flow field, as the impact of the spanwise rollers was largely diminished through averaging, with the higher blowing ratio causing deeper penetration into the crossflow (see Figure 10 in Reference [
142]). The streamwise structures, along with the three-dimensional boundary layer, disrupt the coherence of the spanwise rollers, causing them to tilt and warp, eventually breaking them down into streamwise vorticity concentrations (see Figure 15 in Reference [
142]).
4.3. Effects of Forcing Frequency and Signal Waveform
The prior work indicates that active control schemes may become more effective by targeting the inherent flow instabilities [
120,
144]. For example, in the case of a flat plate, given that turbulent flow is less susceptible to separation, laminar separation can be delayed by prematurely triggering the transition to turbulence through the excitation of Tollmien-Schlichting waves [
145,
146]. As illustrated in
Figure 9, for separated flow over an airfoil, as the inherently unstable separated shear layer undergoes a laminar-to-turbulent transition, it may or may not reattach to the airfoil surface, leading to two distinct flow regimes [
147]. Hence, the post-separated flow is dominated by three instabilities: the global instability for the von Karman-like vortex shedding in the wake, the Kelvin-Helmholtz instability associated with the vortex roll-up in the separated shear layer, and the bubble flapping or shedding frequency in the event that laminar separation bubble (LSB) is present [
148,
149]. The frequencies of the shear layer and wake instabilities are coupled non-linearly and differ by an order of magnitude, which aligns with traditional scaling arguments as the characteristic length scale of the shear layer is an order of magnitude smaller than that of the wake [
150].
SJAs should ideally be operated at their optimal frequency, typically identified under quiescent conditions, to maximize jet velocity and momentum output. In many control scenarios, targeting the natural flow instabilities requires the SJA to operate at a sub-optimal excitation frequency. In such cases, a solution proposed by Amitay and Glezer [
151] is to use pulse-modulated actuation. This method allows the SJA to be driven at a carrier frequency that maximizes the jet velocity, while the modulation frequency of the source signal is used to specifically target and trigger the desired flow instabilities. It is noteworthy to mention that, although uncommon for SJAs, it is possible to modulate the carrier signal using non-square pulse envelopes or employ amplitude modulation. Examples of modulation for a sinusoidal carrier wave is shown in
Figure 10. Equations (15b) and (
16) may also be equivalently applied to modulated SJAs, though in such cases,
should be set to the active duration of the SJA. Typically,
is assigned as
, although an SJA may still generate momentum during the off-phase of the cycle, depending on its response characteristics. Overall, Equation () is more appropriate for burst modulation as Equation (
16) does not adequately account for the duty cycle. This can be easily verified by assuming a uniform velocity distribution during the active phase, which is sometimes a reasonable assumption for burst-modulated SJAs (see Figure 4 in Reference [
152] or Figure 2 in Reference [
153]). It then becomes clear that Equation (
16) remains unchanged, whereas Equation (15b) varies linearly with the duty cycle. In any case, the effect of forcing frequency is characterized by the reduced frequency
, defined below:
where
is an appropriate length scale for the flow under consideration and
f is either the excitation, or the modulation frequency if the carrier signal is modulated. As a rule of thumb, an appropriate characteristic length is typically determined by the length scale of flow instabilities [
154]. Nevertheless, it is also common to estimate the reduced frequency
using a reference geometrical length, such as the diameter of a cylinder or the chord length of an airfoil.
Goodfellow et al. [
155] considered the effects of excitation at
, corresponding to the SJA Helmholtz resonant frequency, on the separated shear layer and wake of a NACA 0025 airfoil at a chord-based Reynolds number of
and an angle of attack of
. Their results demonstrated that applying excitation above a threshold
leads to flow reattachment, which in turn significantly alters the wake topology. Although an initial increase in
above the threshold level resulted in an almost 50 reduction in drag, the positive impact on airfoil performance plateaued at higher
values, highlighting the limitations of relying solely on
as a control input. The significance of the forcing frequency and waveform in achieving optimal control becomes evident when comparing the findings of Goodfellow et al. [
155] with those of Feero et al. [
102], who studied the effects of both the momentum coefficient
and reduced frequency
of a high-AR rectangular SJA on flow separation over a NACA 0025 airfoil at
and
. Similar to Goodfellow et al. [
155], Feero et al. [
102] identified a threshold value for an effective momentum coefficient
at a reduced frequency of
, which corresponded to the SJA resonant peak. For a burst-modulated input at
and
, however, flow reattachment occurred at
, which was 63 less than the threshold
required for high-frequency harmonic excitation at
. Interestingly, as the modulation frequency was increased to
, the flow reattached at
, which was an order of magnitude smaller than what was required for high-frequency harmonic excitation. The effectiveness of low- and high-frequency forcing for an airfoil has been studied by several other researchers, in conjunction to different flow conditions, airfoil geometries, and SJA parameters [
41,
156,
157,
158,
159,
160]. Amitay and Glezer [
144] conducted wind-tunnel experiments on an unconventional airfoil model at
, studying the effects of the reduced frequency ranging from
to
on aerodynamic forces. They observed that forcing at
, corresponding to the wake shedding frequency, leads to the formation of large vortical structures that persist well beyond the trailing edge of the airfoil, resulting in unsteady reattachment and aerodynamic forces. In contrast, actuation at
, corresponding to the shear layer frequency, led to a complete flow reattachment, marked by the absence of large organized vortical structures. Similar observations were made by Amitay and Glezer [
151] and Glezer et al. [
119] who investigated the same two ranges of reduced frequency
for the same airfoil model as in Amitay and Glezer [
144]. Amitay and Glezer [
151] focused on flow transients and observed that the transients following the initiation or termination of pulse-modulated control were quite similar for cases where
and
. After the initial transition, the shedding of organized vortical structures gradually diminished for
case, whereas
case was characterized by the coherent shedding of a train of large vortices. Similar observations were reported by Amitay and Glezer [
161]. For a NACA 0025 airfoil at
, both Feero et al. [
162] and Xu et al. [
163] showed that forcing at the shear layer instability
maximizes the lift-to-drag ratio, while forcing at the wake instability
leads to maximum lift increase. Glezer et al. [
119] investigated the fundamental differences in the response of separated flow over a two-dimensional circular cylinder at a diameter-based Reynolds number of
to both low- and high-frequency actuation. By measuring the changes in circulation around the cylinder, the study confirmed the earlier findings of Amitay and Glezer [
144], Amitay and Glezer [
151]. Low-frequency actuation at
, coupled with the global shedding frequency of the wake, produced strong oscillations in circulation and, as a result, in the aerodynamic forces. In contrast, under high-frequency actuation at
, the circulation experienced a brief transient before stabilizing to a quasi-steady state, indicating that the aerodynamic forces became relatively time-invariant. These results underscored the fundamental distinction between low- and high-frequency actuation, that is the latter is decoupled from the unsteady base flow frequencies, leading to aerodynamic forces that are nearly constant and a damping of global flow oscillations.
The second part of the study by Amitay and Glezer [
151] focused specifically on the efficiency of SJAs using pulse-modulated forcing at a constant duty cycle
and across a range of reduced frequencies
, compared to time-harmonic excitation. A comparison of both methods at
, while maintaining the same
level, revealed that time-harmonic actuation did not produce the same levels of lift coefficient as pulse-modulated excitation. Remarkably, pulse modulation resulted in a 400 increase in the lift coefficient once steady state was achieved, compared to continuous high-frequency
actuation, while using only 25 of the jet momentum coefficient. Taylor and Amitay [
108] investigated the effects of pulse-modulated SJA flow at varying
and
values on the global forces and moments of a dynamically pitching finite-span national renewable energy laboratory (NREL) S809 blade at
. Remarkably, at a reduced modulation frequency of
and a duty cycle
, a 50 reduction in lift hysteresis was observed compared with the continuous sinusoidal actuation. Moreover, at a constant
, the reduction in hysteresis was very similar between
and
cases. The effectiveness of pulse-modulated forcing at a constant
and various
values was also confirmed by Rice et al. [
153], Rice et al. [
164], Rice et al. [
165] for an NREL S817 airfoil at
. Notably, these studies demonstrated the efficacy of pulse modulation in deep stall conditions, such as at
, using a duty cycle of only
, significantly reducing power consumption. All of these works highlight the potential for reducing power requirements for synthetic jets while achieving similar flow reattachment effects or hysteresis reductions through pulse modulation at lower duty cycles.
Margalit et al. [
166] conducted a series of tests on a balance-mounted 60 swept-back semi-span delta wing model with a sharp leading edge, and a
thickness-to-root-chord ratio for a range of Reynolds numbers
and post-stall angles of attack
, studying several SJA parameters, including the modulation wave form. Their analyses revealed that the most effective reduced frequencies in improving the normal force were an order of
, regardless of the modulation waveform or momentum coefficient
. Interestingly, high-frequency non-modulated signals resulted in no improvement, or even a slight degradation, of aerodynamic performance (see Figures 4 and 5 in Reference [
166]). Furthermore, the square pulse and chainsaw envelopes were somewhat superior to the triangle and sine envelopes in terms of normal force enhancement, for the range of effective frequencies considered (see Figure 6 in Reference [
166]). It was found that a square envelope with a duty cycle as low as
was more effective than an amplitude-modulated signal, despite the latter having a larger peak excitation velocity and an order of magnitude greater momentum input. The study by Margalit et al. [
166] demonstrates the role of modulation envelope in enhancing the performance of SJAs.
4.4. Effects of Actuator Location
Although not explicitly included in the dimensional analysis presented in
Section 4, the location of the SJAs, denoted by
when non-dimensionalized, implicitly influences other parameters, such as the boundary layer thickness
, wall shear stress
, and the local wall curvature. Amitay et al. [
159] investigated the effect of SJA location on the reattached flow over an unconventional airfoil across a range of angles of attack
at a Reynolds number of
. The results showed that less power was required to reattach the flow as the control was positioned closer to the natural separation point on the unforced airfoil (see Figure 7 in Reference [
159]). Surprisingly, reattachment was still achieved in some cases even when the actuators were positioned downstream of the stagnation point on the pressure side of the airfoil. Amitay et al. [
159] emphasized that, despite the relatively high levels of
necessary to affect the flow far upstream of the separation point, the interaction between the jets and the crossflow can yield a higher lift-to-drag ratio that may not be achievable when the jets are closer to the separation point. Zhao et al. [
46] considered two chordwise locations
and 40 for a NACA 0021 airfoil. They explained that the rear location was more effective than the front location in increasing the lift coefficient at pre-stall angles of attack
as little flow separation occurs near the rear location, allowing the induced jet to directly inject momentum into the separated flow. As
increased, the flow separation point moved toward the leading edge of the airfoil and, consequently, the front SJA location proved superior for controlling leading-edge flow separation, resulting in enhanced aerodynamic characteristics of the airfoil (see Figure 12 in Reference [
46]). Overall, the upstream location was reported to be more efficient in delaying the stall of airfoil. Tang et al. [
167] also studied two chordwise SJA locations
and 43 for a national aeronautics and space administration (NASA) straight-wing model at a Reynolds number of
and a range of angles of attack
, where the SJA was operated at several reduced frequencies
, reporting that the front SJA array was more effective than the rear one (See Figure 7 in Reference [
167]).
Amitay et al. [
114], Amitay et al. [
168] investigated the manipulation of global aerodynamic forces on a two-dimensional circular cylinder model having a pair of surface-mounted SJAs, for a range of Reynolds numbers up to
and SJA locations at circumferential angles
. A schematic of their experimental setup is shown in
Figure 11a. The smoke visualizations revealed that the interaction between the synthetic jet and the boundary layer led to the formation of closed recirculation regions, which could displace the local streaklines above the cylinder’s top surface, delaying flow separation (see Figures 5 and 7 in Reference [
114]). When the SJAs were positioned at
, the circumferential pressure distribution remained nearly unchanged, indicating that the momentum coefficient was too low to affect the flow significantly. As
increased, the effect of the SJAs on the pressure distribution became more pronounced, with a local minimum appearing in the pressure coefficient distribution around the SJAs location compared to the baseline flow (see Figure 2 in Reference [
168]). For
, the static pressure between the front stagnation point and the separation point continued to decrease as
increased. Of particular note were the changes in the base pressure of the cylinder between the top and bottom separation points, signifying the effect of actuation on the pressure drag (see Figure 10 in Reference [
114]). For
, downstream of the separation point of the baseline flow, the pressure distribution shifted, exhibiting a second local minimum in the static pressure upstream of the separation zone on the unforced lower half of the cylinder. When
, the two minima in the pressure distribution were nearly identical and symmetric, indicating an approximately zero lift force at this angle. For large enough values of
, the pressure distribution near the separation point on the forced side became nearly indistinguishable from that of the baseline flow. The SJAs seemed to influence only the unforced lower half of the cylinder, indicating a reversal in the direction of the lift force (see Figure 11 in Reference [
114]). Eventually, for substantially large values of
, when the SJAs were located near the wake, the pressure distributions for both the forced and unforced flows became nearly identical. A remarkable finding from the studies by Amitay et al. [
114], Amitay et al. [
168] is that the lift force can be entirely nullified or reversed simply by altering the circumferential position of the SJAs. These studies also confirm that when excitation is applied to the stable flow, the disturbances decay before becoming unstable. In contrast, applying excitation near the separation point proves to be an effective control strategy. Tensi et al. [
44] studied the flow over a two-dimensional circular cylinder at a Reynolds number of
, with two SJAs positioned at
, operating at reduced frequencies up to
, close to the natural shedding frequency. Surface oil visualizations showed that the separation line shifted downstream. When only one actuator was active, the trends in the pressure distributions were similar to those reported by Amitay et al. [
114], Amitay et al. [
168]. However, when both SJAs were activated, separation was delayed on both sides of the cylinder, resulting in two distinct local minima in the pressure distribution. These minima were nearly identical and symmetric, indicating an almost zero lift force (see Figures 7 and 11 in Reference [
44]).
The study by Feero et al. [
162] on a NACA 0025 airfoil, shown in
Figure 11b, considered four chordwise SJA locations
around the mean separation point
, specifically
,
,
, and
. The results indicated that once a certain blowing ratio was achieved, the benefits of control saturated for both drag reduction and lift increase. A monotonic decrease in the threshold blowing ratio required to fully reattach the flow, along with an increase in the lift-to-drag ratio, was observed as the slot location moved upstream, with the most upstream location proving to be the most effective, both by requiring the lowest threshold blowing ratio and producing the largest lift-to-drag ratio. Zhao et al. [
46] investigated two chordwise SJA locations,
and 40 , for a NACA 0021 airfoil, again concluding that the jet positioned closer to the leading edge of the airfoil was more effective in delaying stall. The analyses by Taylor and Amitay [
108] for the pitching NREL S809 blade, with two chordwise SJA locations at
and 20 , revealed that the forward jet location consistently performed better in reducing hysteresis at any given momentum coefficient
and reduced frequency
(see Figure 8 in Reference [
108]). Still, the rear SJA location produced a higher lift coefficient during most of the pitching cycle. Therefore, if the goal is to enhance the lift coefficient, the rear jet location was recommended (see Figure 7 in Reference [
108]).
4.5. Effects of Clustering
SJAs are typically smaller than the geometric scales of the body they are intended to control, so they are often arranged in arrays to cover longer spans for more effective flow control [
170,
171]. For an array of SJAs, the stability of flow structures is crucial for enabling effective mixing of low- and high-momentum fluid across the entire span of the controlled geometry. Feero et al. [
169] employed tuft and oil visualizations to examine the shape and spanwise extent of the reattached flow over a NACA 0025 airfoil caused by a high-AR rectangular SJA across a range of reduced excitation frequencies
and blowing ratios
at a constant
and
. The surface flow visualizations showed that the spanwise extent of the reattached flow narrows toward the trailing edge of the airfoil, a phenomenon described as flow contraction toward the airfoil centerline, which is also illustrated in
Figure 12.
In the work of Feero et al. [
169], the size and shape of the reattached region remained unchanged with varying
at
. However, at a constant
, as
increased from 1 to 2.5, the contraction of the attached flow decreased, resulting in a larger spanwise extent of the attached flow. These observations suggest that measurement techniques examining lift and drag improvements solely at the mid-span may fail to capture the full effects occurring across the entire span, which could result in inaccurate assessments of parameters. Also, note that the blowing ratio
was used by Feero et al. [
169] as a measure of jet strength instead of the momentum coefficient
since the SJA was essentially two-dimensional. Machado et al. [
49,
173] utilized smoke wire visualization to investigate the three-dimensionality of the reattached flow over the same NACA 0025 airfoil model as in Feero et al. [
169] at
and
. The airfoil was equipped with an array of circular SJAs instead, operating at a constant momentum coefficient of
and two burst-modulated reduced frequencies
and
. Although the contraction phenomenon was observed at both frequencies, in the low-frequency case, the contraction was sharper and occurred at various chordwise locations. In contrast, the high-frequency case resulted in a more gradual contraction along the span. The results indicated that the effective control length is confined to about 40 of the array width, significantly decreasing the sectional lift coefficient as the distance from the mid-span increases.
Clustering SJAs introduces additional parameters, such as phase differences and spacing between individual SJAs, further expanding the complex parameter space. The studies by Liddle et al. [
174], Liddle and Wood [
175], Wen et al. [
176], and Wen and Tang [
177] explored the impact of phase differences,
–270 in 90 increments, between in-line twin circular SJAs in laminar and turbulent boundary layers over a flat plate. These investigations employed a variety of experimental techniques, including constant temperature anemometry (CTA) [
174,
175], oil-flow visualization [
174], stereoscopic dye visualization [
176,
177], and PIV [
176,
177]. In both studies by Liddle et al. [
174] and Liddle and Wood [
175], power spectral analysis of the downstream velocity time-histories revealed a distinct difference in
case, where the prominent peak occurred at twice the actuation frequency, as opposed to the three other instances in which the prominent peak occurred at the actuation frequency. Additionally, the time-averaged streamwise velocity contours for
featured more abrupt changes, indicating greater penetration into the boundary layer (see Figure 10 in Reference [
174] or Figure 4 in Reference [
175]). These findings were confirmed later by Wen et al. [
176], Wen and Tang [
177]. Both studies identified, three types of vortex structures: one combined vortex at
, two completely separated hairpin vortices at
, and partially interacting vortex structures at
and 180 (see Figure 5 in Reference [
176] or Figure 4 in Reference [
177]). The strongest structure was the single combined vortex at
, which exhibited the greatest penetration into the boundary layer. At
, the resulting flow structures resembled a train of completely separated hairpin vortices, effectively doubling the frequency of the hairpin vortices produced by a single SJA. Overall, the results from the above studies indicate the potential for phase angle manipulation to improved flow control. It should be noted that the most effective phase difference between in-line SJAs also depends on the dimensionless jet-to-jet spacing
(usually spacing is normalized by the orifice width). For clarity, consider the study by Zhao et al. [
46], where the effects of phase delay and combinations of pitch angles for in-line dual arrays on a NACA 0021 airfoil were investigated. At the angle of attack
just after stall, the phase difference of
could significantly increase the lift coefficient and help prevent flow separation (see Figure 13 in Reference [
46]). On the other hand, the control effects due to combined pitch angles were more complex. However, overall, when the pitch angle of the upstream SJA was lower, the dual jet actuators provided better control over airfoil stall compared to a single actuator (see Figure 16 in Reference [
46] for a detailed report on combined pitch angles).
Fewer studies have explored the effects of spacing
and phase difference
in the parallel clustering of SJAs within a crossflow. For steady round jets having a parallel twin-jet configuration in a flat plate boundary layer, a detailed investigation was conducted by Zang and New [
178], examining various spacing
and velocity ratio values, using laser-induced fluorescence (LIF) and PIV. The study demonstrated that each jet in the cluster achieved greater entrainment and larger jet half-widths compared to a single jet in crossflow. Reducing
caused the twin jets to interact with each other closer to the orifice exit. Similar to the enhancement mechanism in quiescent flow [
114], the inner vortices were observed to move toward each other, neutralizing due to their oppositely signed vorticity and merging into a single counter-rotating vortex pair, which prevailed further downstream (see Figures 9 and 10 in Reference [
178]). The study by Vasile and Amitay [
143] on a 30 swept-back NACA 4421 wing equipped with three parallel rectangular SJAs with a spanwise spacing of
also demonstrated that, when the spanwise spacing between the jets is sufficiently large, there is minimal interaction between the jets. The resulting flow field, formed by the simultaneous activation of all three jets, was essentially a superposition of the flow fields produced when each jet was activated individually. More recently, Jankee and Ganapathisubramani [
179] investigated the interaction of parallel
rectangular twin jets with a turbulent boundary layer over a flat plate, for a range of spacing
and phase differences
, while keeping all other geometrical and operational parameters constant. A limit in spacing was identified, beyond which any further increase causes the twin jets to behave as two independent synthetic jets, which also aligned with the findings of Watson et al. [
180] observed under quiescent conditions. Similar to the observations of Zang and New [
178], when the spacing
was sufficiently small, the jets interacted with each other, with the inner vortex of each counter-rotating vortex pair canceling out due to their opposite vorticity signs and the remaining part of the vortices coalescing into a single vortex pair (see Figure 4 in Reference [
179]). For the smallest spacing of
with a phase difference
, the jets exhibited vectoring toward the actuator leading in phase, similar to the behavior observed in quiescent flow conditions [
114,
181,
182] (see Figure 6 in Reference [
179]). At a phase difference of
, the influence of the crossflow caused the vortical structures from the leading SJA to convect downstream before the lagging SJA reached peak blowing. As a result, the interactions were minimal, and the flow field of the twin jets in this scenario became analogous to that of a single SJA operating at twice the actuation frequency, the doubling effect. The studies mentioned above highlight the potential of clustering parameters
and
for enhanced control schemes. A notable example can be found in the work by Amitay et al. [
114] on a circular cylinder. When the SJAs were positioned on the downstream edge at
, substantial momentum was necessary to influence the wake. However, with the SJAs operated at
, the downward vectoring of the jets caused a downward shift in the entire wake and a simultaneous displacement of the front stagnation point. This led to a decrease in spacing between the streaklines above the cylinder and an increase in spacing below it, indicating a change in circulation and lift generation (see Figures 8 and 9 in Reference [
114]).