You are on page 1of 76

Experimental investigation of wing-tip vortex evolution

behind a half-delta wing equipped NACA0012 wing

Steve (Sunghun) Choi

M. Eng. Thesis

Department of Mechanical Engineering

McGill University

Montreal, Quebec

Thesis submitted to McGill University in partial fulfillment


of the requirements of the degree of Master of Engineering

© Steve (Sunghun) Choi 2015

i
ACKNOWLEDGEMENTS
I would like to thank my advisor, Professor Tim Lee, for his continued guidance and support
throughout my graduate research. Although he had a lot of hard time to guide and teach me, he
always pointed the clear way and encouraged me. I would also like to thank my lab mates
Lok Sun Ko, Ying Yu Su and Nirmal Raju. Lastly, I want to thank my wife for her devotion and
the support I needed to successfully finish this work effort.

ii
ABSTRACT

The modification of the tip vortex, generated by a rectangular NACA0012 wing, was attempted
using four different tip-mounted half-delta wing (HDW) models at a chord Reynolds number of
2.45 × 105. The HDW models used in this experiment were i) Ʌ=65°, cr = 1c, ii) Ʌ=50°, cr = 1c,
iii) Ʌ=50°, cr = 0.5c and iv) Ʌ=50°, cr = 0.3c where c, cr and Λ are baseline-wing chord, root
chord and slenderness, respectively. In addition to the cases listed above, the impact of the
deflection δ was also investigated. As a consequence of the vortex breakdown, the roll-up
process of the tip vortex was found to be significantly modified. Especially, HDWs with the
cr ≤ 50% exhibited a unique double-vortex pattern, one originating from the tip of the baseline
wing and the other one from the leading-edge of the HDW. HDWs with the cr ≤ 50% also caused
a rapid diffusion of vorticity and rendered a weak circulation flow-like tip vortex, suggesting an
enhanced wake-vortex decay and alleviation. The interaction and merging of double-vortex was
expedited by both increasing the angle of attack and deflecting the HDW upward. Regardless of
the vortex breakdown, the lift of the HDWs was found to increase with the increasing angle of
attack and stayed above the value of the baseline wing. The cr = 0.5c produced an improved lift-
to-drag ratio.

iii
Résumé

La modification du tourbillon marginal généré par une aile rectangulaire de profil NACA 0012 a
été essayée en utilisant quatre différentes modèles d’ailes demi-delta (ADD) à un nombre de
Reynolds de corde de 2.45 × 105. Les modèles ADD utilisés sont i) Ʌ=65°, cr = 1c, ii) Ʌ=50°,
cr = 1c, iii) Ʌ=50°, cr = 0.5c and iv) Ʌ=50°, cr = 0.3c. En conséquence de l’éclatement
tourbillonaire, le processus d’enroulement de tourbillon marginal a été modifié significativement.
Spécifiquement, les ADD avec cr≤ 50% démontrent un motif unique de double-tourbillons, un
qui origine de l’extrémité de l’aile de base et l’autre du bord d’attaque du ADD. Les ADDs avec
cr≤ 50% causent aussi une diffusion rapide de vorticité et rendent faible la circulation
d’écoulement comme le tourbillon marginal, suggérant une amélioration de sillage-tourbillion
déclin et alléviation. L’intéraction et la fusion du double-tourbillions sont accélérées par
l’augmentation d’incidence et la déviation d’ADD vers le haut. Malgré l’éclatement
tourbillionaire, la portance des ADD a augmenté avec l’augmentation de l’incidence et a resté
sous la valeur de l’aile de base. Le cas de cr = 0.5c a causéune amélioration de la finesse.

iv
Table of Contents

ACKNOWLEDGEMENTS ........................................................................................................ ii

ABSTRACT ................................................................................................................................. iii

List of symbols............................................................................................................................. vi

List of figures ............................................................................................................................. viii

1 Introduction ........................................................................................................................1

2 Literature Review ...............................................................................................................7


2.1 Static wing-tip vortex ............................................................................................7
2.2 Delta wing aerodynamics ....................................................................................13
2.3 Wing-tip vortex control techniques .....................................................................18
2.4 Objectives ............................................................................................................21

3 Experimental Procedure ...................................................................................................22


3.1 Flow facility ........................................................................................................22
3.2 Seven-hole pressure probe ..................................................................................24
3.3 Traverse mechanism ............................................................................................30
3.4 Force balance.......................................................................................................31
3.5 Lift-induced drag (Di) calculation .......................................................................32
3.6 Wing models .......................................................................................................37

4 Result and Discussion .......................................................................................................40


4.1 Flow analysis for full-chord or cr = c HDW .....................................................40
4.2 CL, CD, CL/CD and CDi analysis .........................................................................42
4.3 Flow analysis for small-chord or cr ≤ 0.5c HDW .............................................44

5 Conclusion ........................................................................................................................54

References ....................................................................................................................................55

Appendix: Normalized iso-vorticity & axial velocity with different angles of attack .............60

v
List of symbols

AR Aspect ratio, =b2/S

b semi-span

BW baseline wing

c chord of baseline wing

cr root chord of half-delta wing

CD total drag coefficient

CDi lift-induced drag coefficient

CL total lift coefficient

Cr root chord of half-delta wing

Di lift-induced drag

HDW half-delta wing

S wing surface area

s half-wing span, = b/2

SBW baseline wing area

SHDW HDW surface area

Stotal = SBW + SHDW

PSTAT static pressure

PTOT total pressure

rc core radius

ro outer radius

Re chord Reynolds number

vi
u,v,w mean axial, transverse and spanwise velocity

uc core axial velocity

u∞ freestream velocity

νθ tangential velocity

x,y,z streamwise, vertical, and spanwise direction

α angle of attack

αss static stall angle

δ deflection of half-delta wing; (+) for clockwise and


(-) for counterclockwise

ε wing apex half-angle

Ʌ sweep angle

ζ mean streamwise vorticity

Гc core circulation

Гo total circulation

ρ∞ freestream density

ψ stream function

ϕ velocity potential

ν kinematic viscosity

vii
List of figures
Figure page

1 Formation and development of a wingtip vortex (Chow et al., 1997) ............................2

2 Delta wing planform geometry (RHRC, 2000) ..............................................................3

3 Schematic drawing of the leading-edge vortex and secondary vortex formation


(McCormick, 1979) ........................................................................................................4

4 Delta wing vortex lift characteristic (RHRC, 2000) .......................................................5

5 Sketch of the leading-edge flow structure (Nelson, 2003) ...........................................13

6 Three regions within a LEV structure (Nelson, 2003) .................................................14

7 Schematic of a) bubble type and b) spiral type vortex breakdown


(Taylor et al, 2003) .......................................................................................................15

8 Visualization of a Spiral and Bubble Type Breakdown (RHRC, 2000) .......................16

9 Polhamus leading-edge suction analogy (Polhamus, 1971) .........................................18

10 The schematic diagram of J.A. Bombardier wind tunnel (Pereira, 2012) ......................22

11 J.A. Bombardier wind tunnel ........................................................................................23

12 Seven-hole pressure probe a) probe configuration b) details of the probe tip


(Pereira, 2012) ................................................................................................................24

13 Typical SHP calibration grid ........................................................................................25

14 Pressure tap numbering convention ..............................................................................27

15 Angular calibration traverse (Pereira, 2012) ................................................................29

16 Seven-hole pressure probe traverse mechanism ...........................................................29

17 Components of the force balance (Pereira, 2012) ........................................................31

viii
18 Schematics of experimental set-up ...............................................................................38

19 HDW configurations a) 50HDW, b) 65HDW, c) 0.5c 50HDW


and d) 0.3c 50HDW ......................................................................................................39

20 Wing model in the test section .....................................................................................39

21 Impact of Λ, cr and δ of the HDW on the normalized iso-vorticity contours


of the tip vortex at x/c = 2.8 for α = 6°, 8° and 10°. (a1)-(a3): BW;
(b1)-(b3): 1c65HDW; (c1)-(c3): 1c50HDW; (d1)-(d3): 0.5c65HDW;
(e1)-(e3): 0.5c50HDW; (f1)-(f3): 0.5c50HDW(+5); (g1)-(g3): 0.5c50HDW(-5);
and (h1)-(h3): 0.3c50HDW .................................................................................... 46-47

22 Growth and development of HDW vortex and BW vortex on half-delta wing


configuration and LEV on full delta wing. (a) BW at α = 10°, (b) 1c65HDW
wing at α = 10°, (c) 65°-sweep full delta wing at α = 12°, (d) 50°-sweep full
delta wing at α = 10°, (e) 0.5c50HDW at α = 10°, and (f) 0.3c50HDW at
α = 10° .................................................................................................................... 48-50

23 Variation of normalized (a) vertical position and (b) peak vorticity of LEV
with x/c on 65°-sweep full delta wing at α = 12° and 50° -sweep full delta
wing at α = 10° .............................................................................................................51

24 Effect of slenderness, root chord and deflection of HDW on aerodynamic


characteristics ...............................................................................................................52

25 Impact of Λ, cr and δ of the HDW on the trajectory and total circulation of


the tip vortex at x/c = 2.8 ..............................................................................................53

A1 Normalized iso-vorticity and axial velocity contour of BW at x/c = 2.8 .....................60

A2 Zoomed in iso-vorticity contour of BW at x/c = 2.8 ....................................................61

A3 Normalized iso-vorticity and axial velocity contour of 50HDW at x/c = 2.8 ........ 62-63

A4 Zoomed in iso-vorticity contour of 50HDW at x/c = 2.8 .............................................64

A5 Normalized iso-vorticity and axial velocity contour of 0.5c50HDW at x/c = 2.8 ... 65-66

A6 Zoomed in iso-vorticity contour of 0.5c 50HDW at x/c = 2.8 .....................................67

ix
CHAPTER 1

Introduction

On November 12th, 2001, American Airlines Flight 587, the Airbus A300-600, crashed into the
Belle Harbor near New York City, killing all 260 passengers on board and 5 people on the
ground. The National Transportation Safety Board (NTSB) proclaimed that the cause was due to
“the first officer’s overuse of rudder controls” in response to wake vortex turbulence formed by
Japan Airlines Boeing 747-400 which took off shortly before.

In the early days of aerodynamic history, tip vortices were not a big concern to the
aerodynamicists since vortices created by small aircrafts are negligible. However, with the
arrival of jumbo jets such as 747s and DC-10s, vortices produced from these large and heavy
aircrafts are so powerful that they are no longer ignorable. These vortices can pose extremely
hazardous situations to any trailing airplane directly or indirectly, causing loss of controls or
even structural failures. In addition to the wake-vortex hazard, these tip vortices also generate
lift-induced drag which limits the aerodynamic performance of the wing. Therefore, not only
because of the safety issues but also for the aerodynamic efficiencies, the suppression of the
wingtip vortex and the reduction of the lift-induced drag of aircraft wings have always been
a challenge to aerodynamicists and fluid dynamists. An excellent review on the airplane trailing
vortices can be found in the work of Spalart [1].

A vortex is generated as the flow tends to turn around the tip of a wing from lower, or
pressure, to upper, or suction, surface due to the pressure difference in between (Fig. 1).
As the flow moves further downstream, the spiral motion of the flow is fed by boundary layer
and rolls up even more. Finally, when it reaches near the trailing-edge of the wing, this tip vortex
is separated from the wing surface and becomes a free vortex. Even past the trailing-edge,

1
the tip vortex engulfs more and more of the wing wake until its circulation is nominally balanced
to that of the wing [2].

Fig. 1 Formation and development of a wingtip vortex (Chow et al., 1997)

Once the vortex is shed from the trailing-edge, it induces a downwash flow pattern behind
the wing within its span. Over time, the vortex slowly undergoes dissipation through viscous
diffusion; however, such process generally takes several tens to hundreds of chord lengths to be
completed. Due to their detrimental characteristics, numbers of studies regarding wing-tip
modifications have been continuously investigated in an attempt to control the strength and
structure, including trajectory, of the tip vortices.

The first conceptual design of a wing-tip device was adding a vertical endplate on the
wingtip to make the wing theoretically infinite (Lanchester, 1907). The basic idea was to prevent
the roll-up process by blocking the free end of the wingtip; and theoretical calculations supported
its possibility of reducing the induced drag. However, in reality, this simple endplate was not as
practical as expected. It, indeed, was able to lower the induced drag; however, a huge rise in the
viscous profile drag resulted in an increase in the total net drag. Whitcomb [3] believed that by
keeping the additional profile drag to minimum, he could still take the benefit of an endplate, the
reduced lift-induced drag. This idea was then developed into the classic winglet later on.

2
Inspired by Whitcomb’s outstanding work, various wingtip devices, or passive wing tip
vortex control techniques, including the spoiler [4], spline [5] and spiroid tip [6] had been
developed and tested by researchers elsewhere. With consistent effort to control the tip vortices
via wingtip modifications, an intriguing new idea had emerged. Staufenbiel and Vitting [7] were
the first to suggest using a fixed half-delta wing (HDW), at the tip of a rectangular wing;
however, for a better understanding of the tip-mounted half-delta wing concept, one has to be
familiar with the fundamentals of delta wing aerodynamics first.

With their capability of achieving flight at high angle of attack and delaying the onset of
compressibility drag, delta wings had been studied actively for supersonic applications in 1950s.
However, recent studies also found their possible applications at low Reynolds regime such as
MAVs and UAVs where delta wing planforms can be used to modify their aerodynamic
performance by simply changing the sweep angle. With their unique triangular shape, the flow
features of delta wings are clearly different from that of the rectangular wings. Fig. 2 illustrates
the delta wing geometry where c, b and Ʌ stands for the chord length, span and sweep angle,
respectively.

Fig. 2 Delta wing planform geometry (RHRC, 2000)

3
For the delta wings, the flow separation tends to occur at the leading-edge and rolls up to
form so-called leading-edge vortex (LEV) due to Kelvin-Helmholtz type instability.
Fig. 3 illustrates a conceptual sketch of the leading-edge vortex structure. Note that as the shear
layer rolls up to form the leading-edge vortex (LEV) or the primary vortex forms on the wing
surface which induces a spanwise flow on the upper surface. Consequently, this spanwise flow
separates from the surface and forms a smaller and weaker secondary vortex which is counter-
rotating below the primary vortex. It is also possible that there can be more than one secondary
vortices near the surface. As the flow progresses further downstream, theses vortices merge to
form a larger vortex. The rotational velocities in the core of the primary vortices can be up to
3 times the free-stream velocity. This fast rotating primary vortex core generates a lower
pressure region on the upper surface of the wing, resulting lift force known as vortex lift which
can be up to 50% of the total lift. Thus, the LEV lies just above the wing surface and increases
the wing’s lift in non-linear manner (Fig. 4).

Fig. 3 Schematic drawing of the leading-edge vortex and secondary vortex formation (McCormick, 1979)

4
Fig. 4 Delta wing vortex lift characteristic (RHRC, 2000)

The primary purpose of Staufenbiel and Vitting's work [7] was to mitigate the vortex wake
structure by enforcing the vortex breakdown via static half-delta wings which could rotate about
the main wing's mid-chord. This idea was supplemented by flow visualization techniques from
which they confirmed a premature vortex breakdown of the baseline-wing at α = 10°. They
concluded that this early vortex breakdown was caused by the additional circulation originated
from the main wing. They also observed lift increments with appropriate half-delta wing
incidence settings which was believed to be the result of lift generating capability of the half-
delta wing via leading-edge vortex (LEV).

Based on Staufenbiel and Vitting's work [7], Nikolic [8-9] suggested the use of actively
movable tip strakes, or half-delta wing attachments with adjustable deflection angle settings, on
the tip of a rectangular NACA4412 wing at Re = 2.25×105 to improve the aerodynamic
performance of the wing. They noticed the proportional relationship between the lift (and the lift-
curve slope) and the deflection of the HDW. However, the concept of HDW tip vortex control
was also accompanied by an inevitable increase in wing weight and bending moment.

5
In the current study, the passive control of the wingtip vortex had been attempted via half-
delta wings (HDWs) of different slendernesses Λ, root chords cr and deflections δ to minimize
the negative effects of the full-chord half-delta wings (cr = c). Also, growth and development of
both baseline-wing (BW) vortex and half-delta wing (HDW) vortex were investigated along
the wing models (for x/c < 1) and in the near wake region (for 1 < x/c < 4). In addition,
evolution of the leading-edge vortex developed along the slender and nonslender full delta wings
(Λ = 65° and 50°, respectively) was obtained to supplement the HDW vortex measurements.
Moreover, the force-balance data were also collected to compare the aerodynamic performance
among the HDW models tested.

6
CHAPTER 2

Literature Review

2.1 Static wing-tip vortex

Based on the characteristics of the tip vortex, the study of the tip vortex can be classified into
two categories: the near-field, which is usually defined as within seven to ten chord lengths
downstream from the leading-edge, and the far-field for beyond an order of hundreds of span.
In the near-field, there is a still ongoing roll-up process of the vortex and the entrainment of
the vorticity both from the shear layer and the wing’s wake whereas such process is already
completed and the vortex shows a quasi-steady and homogeneous behavior in the far-field region.

Towards a deeper understanding of the tip vortex, numbers of experimental and theoretical
studies have been made by researchers; however, there are comparatively few numbers of
analytical studies in the near-field due to the complex and continuous development of the vortex
structure as well as the strong dependency of the vortex on the tip condition of the wing.
Moreover, the experimental studies in the far-field are also very limited. Since the development
of the vortex can last several tens or hundreds of chord lengths, they confront the space
limitation. In addition to that, the freestream turbulence and the wall effect from the wind tunnel
can strongly affect the vortex as well.

Betz [10] was one of the pioneers who theorized the initial roll-up of the tip vortex. With
laminar and inviscid flow assumption, his theoretical work suggested that the wake should roll
up along the outer edge of the vortex via self-induction of the velocity to satisfy the Biot-Savart
law. However, as his model assumed the wing wake to be semi-infinite vortex sheet, it tended to
overestimate the rate of the roll-up process for the vortex. In the far-field, Betz model required
a condition that the moment of inertia of the vortex about its centroid to be conserved and

7
resulted in the velocity singularity at the center of the vortex. However, since this simple model
yields a reasonable accuracy, it is still applicable in some special cases.

Meanwhile, Moore and Saffman [11] developed a model for a spanwise circulation
distribution on a finite wing in the near-field under the laminar and inviscid flow assumption as:

.
𝑏 𝑛−1
Γ(y) ≡ ∮ 𝑣⃗ ∙ 𝜕𝑠⃗ = 2Γ𝑟𝑜𝑜𝑡 ( ) ; 0<𝑛<1 … (1)
𝑦
𝑠

where Г is the circulation around the closed path s, v is the tangential velocity vector, y is
the spanwise location, b is the span, and n is an arbitrary number which varies with the wing
loading condition. Moore and Saffman also expressed the spiral roll-up of the vortex sheet with
the following equation:
1
𝜉 1+𝑛 Γ𝑟𝑜𝑜𝑡 𝑏 𝑛−1
r(θ) = ( 𝑡) 𝑤ℎ𝑒𝑟𝑒 𝜉 = ( )( ) … (2)
𝜃 𝜋 𝛼𝑛

where r represents the radial distance originating from the vortex center, θ is the polar coordinate,
t is the time, αn is a constant value, and ξ is the vortex roll-up rate parameter depending on the
wing loading condition. Even though this model shows a better agreement with the experimental
data and is more widely applicable, it still employs the inviscid flow condition. In practice,
the viscosity has a strong influence on the development of the vortex characteristics.

Taking such significant role of the viscosity into an account, Batchelor [12] studied the
relationship between the axial and tangential velocity development and concluded that the
pressure gradient is the primary key in their relation. Batchelor proposed that the centrifugal
force induced by the tangential velocity is balanced due to the radial pressure gradient. Behind
the trailing-edge, the tangential velocity tends to decrease as a result of the viscosity acting on
the vortex; thus creates a positive pressure gradient and decelerates the axial flow of the vortex.
The axial velocity profile in the vicinity of the core can be then defined as:

8
1 𝜕Γ 2 𝑝 1 2
𝑢2 = 𝑢∞
2
+∫ 𝑑𝑟 − 2Δ𝐻 𝑤ℎ𝑒𝑟𝑒 𝐻 = + (𝑢 + 𝑣 2 + 𝑤 2 ) … (3)
𝑟 2 𝜕𝑟 𝜌 2

where ΔH is the total head loss and Г is the circulation around a symmetrically placed circle.
He also suggested a similarity solution of Navier-Stokes equation for the laminar flow that holds
far downstream and showed that the deficit of the axial velocity is proportional to log(x)/x.
However, this was still not enough for the practical applications since the trailing vortices tend to
have a turbulent behavior.

Hoffman and Joubert [13] considered a turbulent trailing vortex model and divided the
vortex into four regions: the viscous core, the buffer layer, the inner logarithmic region, and
the outer wake. One of their finding was that, in the inner logarithmic region, the circulation
increases logarithmically with the increasing radius and the tangential velocity reaches its peak
value. Inspired by their work, Phillips [14] studied and established a model of a turbulent trailing
vortex undergoing the roll-up process to describe the radial distribution of the circulation.
Similar to Hoffman and Joubert [13], Phillips also classified the vortex into distinct four regions :
region (I), where the viscosity is dominant and, as a result, it can be treated as a solid body
rotation; region (II), where the momentum effect takes the dominancy over from the viscous
effect and the circulation shows a logarithmical increasing trend with the radius; region (III),
where the decay of the Reynolds stresses is proportional to 1/r2; and finally the outer wake
region, where the roll-up is still incomplete according to Phillips.
By applying the boundary conditions and curve-fitting the data, Phillips proposed an empirical
third order polynomial for region (I) and (II) as a function of (r/rc)2.

Γ(𝑟) 𝑟 2 𝑟 4 𝑟 6 𝑟
= 𝐴1 ( ) + 𝐴2 ( ) + 𝐴3 ( ) 𝑓𝑜𝑟 < 1.2 … (4)
𝑟𝑐 𝑟𝑐 𝑟𝑐 𝑟𝑐 𝑟𝑐

where A1, A2, and A3 are constants. Also, he showed that the initial conditions of region (III)
only contributed to the rate of the outer wake region to attain its steady state, therefore, it was

9
predicted that every vortex would exhibit a self-similarity in the inner two regions regardless of
the initial conditions.
As opposed to the limited numbers of theoretical studies, more active experimental
investigations has been done to characterize the development of the tip vortex in terms of core
axial velocity, peak vorticity, peak tangential velocity, core radius, and core circulation.

Chigier and Corsiglia [15] studied the structure of the tip vortex in the near-field
(0.25 < x/c < 5) of a blunt-tip NACA0015 wing at Re = 9.53×105 with α = 12° with a
particular attention on the development of the trailing vortex. They observed an increase in the rc
followed by a substantial decrease in the νθ and the axial velocity were found to be jet-like
behind the trailing-edge up to 5 chord lengths downstream. Also, the maximum axial and
tangential velocities were found to be along the tip (x/c < 1). Orloff [16] investigated the effect
of the flow incidence (α = 8-12°) for a blunt-tip NACA0015 by fixing the downstream distance
and the Reynolds number to 3 and 7.5×105, respectively. With the increasing α, both νθ/u∞
(0.35-0.56) and rc/c (6.7 to 8%c) were observed to be increased. In addition to that, the data
showed the transition of the normalized axial velocity from wake-like to jet-like. This crossover
point was further examined by Brown [17].

Brown [17] demonstrated the core axial velocity as a function of the aerodynamic forces (to
be specific, the lift and the profile drag) and showed that the maximum lift-to-drag should occur
at around the crossover point where u/u∞ ≈ 1. Meanwhile, Thompson [18] considered the effect
of the tip condition by using a NACA0012 wing with square, round, and beveled tips. With
different tip conditions, the roll-up process of the vortex along the tip was modified. For the
squared and beveled tips, the flow was separated from the sharp edges on the tips and resulted in
the formation of the secondary vortices. Consequently, the generated tip vortices were more
diffused and disorganized compared to the concentrated tip vortex of the round tip.

10
Similar investigation was performed by McAlister and Takahashi [19]. They started with a
square-tipped NACA0015 wing and changed the tip condition by adding a round-lateral edge to
the wingtip. With the addition of the round end cap, they confirmed the elimination of the
secondary vortices and observed the reduction in core radius and the increase in both axial and
tangential velocities near the trailing edge. However, the modification of the tip condition
showed little impact on the trailing vortex characteristics further downstream. Moreover, they
also examined the effect of the angle of attack (4° ≤ α ≤ 12°) and the Reynolds number
(1×106 ≤ Re ≤ 3×106). The results showed that the tangential velocity was almost linearly
dependent on α whereas no significant dependence was detected with the Re. Shekkariz et al. [20]
studied the tip vortex structure of a low-aspect ratio airfoil with Re = 3.6×104, 7.2×104, and
2.2×105 at x/c = 0 – 6.7 using the particle-image velocimetry technique. Their result showed
a rapid formation of the tip vortex (within x/c less than 2) responsible for 66% of the bound
circulation originated from the wing which stayed nearly constant up to x/c = 6. They also
observed the presence of the secondary vortex structures, mostly at low Reynolds number
whereas, at high Re, they found the flow contours to be more organized and concentrated.
They believed that such difference was due to the turbulent mixing. Numerous experimental
investigations have observed the meandering behaviour of the tip vortices which often results in
the fluctuations in measured velocity fields. To address this issue, Devenport et al. [21]
investigated the trailing vortex originating from the blunt-tip NACA0012 wing in the range of
4 < x/c < 29 at Re = 5.3×105 with α = 5° and took a long-time averaged measurement of the
flow field and expressed in terms of a Gaussian distribution. However, the measured RMS values
of core radius and peak tangential velocity resulted in errors up to 64%.

To resolve the errors induced by this spatial meandering of the vortex, Devenport et al.[21]
constructed a probability density function and applied to the measured values to fit the curves.
After such correction procedures, they found that the effect of the meandering becomes small
and negligible for r/c > 0.1. Also, the normalized vortex turbulence data with the wake
measurement agreed well for all cases; thus, they concluded that it was not the vortex itself, but

11
the wake which was responsible for all the turbulence within the vortex. Chow et al. [2] studied
the near-field region of a rectangular NACA 0012 wing with a rounded tip at Re = 4.6×106 and
α = 10°. They placed a trip strip at 4%c from the leading-edge to fix the location of the
boundary layer transition. Both a seven-hole pressure probe and a triple hot-wire probe were
used to collect the flow-field data. They found that the core axial velocity to be always jet-like
throughout the entire downstream distances tested with a maximum of 1.77U∞ at x/c = 0.995
and a minimum of 1.69U∞ at x/c = 1.678. They believed that the factors attributing to such
excesses were the rounded tip condition and high Re, resulting in peak tangential velocities from
0.792 to 1.072U∞ and core radii from 1.88 to 3.06%c as well. Moreover, the peak turbulence
within the vortex showed a decreasing trend with increasing downstream distance due to the
stabilization of the solid-body rotation of the vortex core, validating the previous numerical work
of Dalces-Marianani et al. [22].

Ramaprian and Zheng [23] demonstrated the tip vortex study of a square-tipped
NACA0015 wing in the range of 0.16 < x/c < 3.33 and α = 5 and 10° at Re = 1.8×105 using
laser-doppler velocimetry technique. Interestingly, as opposed to previous studies, their axial
velocities were found to be always wake-like with 0.68U∞ near the trailing-edge and increased to
0.74U∞ within one chord length. Furthermore, the vortex became nearly axisymmetric at x/c = 1
with the maximum tangential velocity of 0.5U∞. Past x/c =1, the previously mentioned vortex
characteristics were measured to be fairly constant until x/c = 3.33. Birch et al. [24] studied the
tip vortex generated by a square-tipped NACA0015 wing at Re = 2.01×105 in the range of
0.5 < x/c < 3 and 2° < α < 18° with a particular attention on the velocity and vorticity
distributions. With the presence of the secondary vortex structures which were eventually
entrained by the main tip vortex, both magnitude and size of the main vortex grew with
increasing downstream distance and became nearly axisymmetric by x/c = 1.5. Also, they
observed the transition of the core axial velocity from wake-like to jet-like with increasing α;
however, the magnitude was found to decrease with the downstream distance until x/c = 1.5.
Moreover, the circulation distribution seemed to agree well when compared to the work of

12
Hoffman and Joubert [13] and the core vortex strength was calculated to be 74% of the total
circulation.

2.2 Delta wing aerodynamics

The original purpose of delta wing development was to minimize the negative aspects, mostly
wave drag and compressibility effects, associated with supersonic flight. Due to their high flight
performance and maneuverability over a high angle of attack regime, delta wing investigations
were mainly focused on the supersonic applications in 1950s. However, with the emerging micro
air vehicle (MAV) and unmanned air vehicle (UAV) industries, researchers found that the delta
wings, which exhibit vortex-dominated flows, can be employed by MAV and UAV applications
as well. Therefore, the flow structure of delta wings has only recently attracted more attention
[25].
With their triangular planform, the delta wings
exhibit a distinct flow feature when compared to the
rectangular wings. At moderate to high angle of
attack (α = 5-30°), the flow starts to separate along
the leading-edge. The separating shear layer results in
two vortex sheets and forms the leading-edge vortex
(LEV), or two counter-rotating vortices, the primary
and secondary vortex (Fig. 5). The fast rotating core
of the primary vortex induces a high axial velocity
region within and generates a low pressure field
above the wing surface, resulting the vortex lift.
Fig. 5 Sketch of the leading-edge flow Generally, delta wings can be classified into slender
structure (Nelson, 2003)
and non-slender delta wings based on their sweep
angle. Usually, highly swept wings are referred as slender wings whereas delta wings with the
sweep angle less than 55° are defined as non-slender wings.

13
Since both slender and nonslender delta wings share a common triangular planform shape,
they do exhibit some similar flow structures such as leading-edge vortex with primary and
secondary vortices, vortical sub-structures [26], vortex wandering [27] and vortex interactions
[28]. Researchers also have found some unique flow features of the nonslender wing different
from that of slender wing including the dual primary vortex structure at low angle of attack [29-
31] which, however, was not observed in this current investigation. Notwithstanding the
increasing interest in the nonslender delta wing aerodynamics, the published literature on low-
sweep wing is surprisingly sparse compared to that of high-sweep wing. Both Nelson and
Pelletier [32] and Gursul [25] provide excellent reviews on slender and nonslender delta wing
aerodynamics, respectively.

Earnshaw [26] was one of the first pioneers


who studied and theorized the structure of the LEV
and quantified its flow characteristics. According
to Earnshaw, the LEV structure can be divided into
three distinct substructures: the shear layer, the
rotational core, and the viscous subcore (Fig. 6).
The shear layer is the vortex sheet originating from
the wing’s leading-edge which feeds vorticity
Fig. 6 Three regions within a LEV structure
into the vortex core. The rotational core is (Nelson, 2003)
approximated to be about 30% of the local semi-span in diameter where minor perturbations
generated by the vortex sheet can be neglected. Lastly, the viscous subcore is approximately 5%
of the local semi-span in diameter and rotates as a solid body with very high core flow
parameters such as pressure and velocities. In this region, the axial velocity can go up to 2-3
times the freestream velocity.

14
Vissel and Nelson [33] also focused on the LEV structure and found the core diameter to be
5% of the local semi-span verifying Earnshaw’s model. Moreover, they observed a logarithmic
relation between the circulation and the radial distance which well- agreed with the equation that
Hoffman and Joubert [13] suggested for their turbulent vortex model.

Γ(r) 𝑟 𝑟
= 𝐵 𝑙𝑜𝑔 ( ) + 𝐶 0.5 < < 1.4 … (5)
Γ𝑐 𝑟𝑐 𝑟𝑐

1.3
𝐵≈ & 𝐶=1 … (6)
𝑠𝑖𝑛2 𝛼

Where r, rc, and Γc is the radial distance from the vortex core, the core radius, and the circulation
within the core radius respectively.

In addition to its unique vortex structure, what makes the study of the delta wing more
complicated is a phenomenon called vortex breakdown. The vortex breakdown is a sudden
diffusion of the vortex core which carries dramatic changes in the flow characteristics such as
increased vortex diameter and decreased vorticity, pressure, and axial velocity. Most of the time,
the vortex breakdown is undesirable phenomenon since it negatively affects the non-linear vortex
lift and the pitching moment of the wing; thus, limiting the delta wing’s performance in high lift
or high angle of attack conditions.

Due to extensive studies on the vortex


breakdown of the delta wings, in general, two
types of the vortex breakdown, or bursting, have
been observed; bubble and spiral type (Fig. 7 & 8).
Both types are caused by a deceleration, or a
transition of the core axial velocity from jet-like
to wake-like. These bursting types are known to Fig. 7 Schematic of a) bubble type and
be a function of Reynolds number, geometry, b) spiral type vortex breakdown
(Taylor et al., 2003)

15
and circulation; however, their nonlinearities of flow characteristics makes it hard to predict
which breakdown will occur. The bubble type breakdown is characterized by a sudden
expansion of the core size by 2-3 times and observable free-stagnation point along the vortex
centerline. On the other hand, the spiral type breakdown is characterized by a sudden
enlargement of the vortex core with a stagnation point on the vortex axis leading to an oval
shaped recirculation bubble and the vortex centerline tends to corkscrew downstream.

The spiral type breakdown is also characterized by a rapid deceleration of the core axial
flow followed by an abrupt formation of a kink from the helical vortex core. This spiral flow
remains for one or two turns and breaks into a large scale turbulence. With the increasing
incidence, the vortex breakdown point will progress upstream. As the vortex breakdown point
reaches the apex, the wing stalls. Moreover, it is possible for the vortex breakdown to transform
its type as the core rotates while proceeding downstream.

Fig. 8 Visualization of a spiral and bubble type breakdown (RHRC, 2000)

It is well known that the vortex breakdown is strongly dependent on the sweep angle of the
delta wing. Hall [34] reported that the two conditions for the onset of the vortex breakdown are
the large swirl angle and the adverse pressure gradient. In the case of the delta wing, modifying
the sweep angle has similar effects to the changing the aspect ratio of the conventional wing.
Therefore, decreasing the sweep angle induces larger swirl angle and aggravates the upstream

16
propagation of the adverse pressure gradient, which is originated from the trailing-edge. Hence,
the vortex breakdown appears earlier for the delta wing with lower sweep angle.

Furthermore, the sweep angle also affects the aerodynamic characteristics. Wentz and
Kohlman [35] focused on investigating the effect of the sweep angle on the lift by testing
numbers of delta wings with different sweep angles ranging from 55 to 85°. They concluded that
for the delta wing with sweep angle less than 70°, increasing the sweep angle resulted in
increased maximum lift coefficient and decreased lift-curve slope; also the stall was delayed.
Such conclusion is well explained in the leading-edge suction analogy by Polhamus [36].

Polhamus [36] tried to estimate the vortex-lift with a simple method. According to
Polhamus, the extra normal force generated by the vortex should be equal to the loss in the
leading-edge suction for the separated flow. In other words, the leading-edge suction for the
attached flow is transformed to the vortex-lift as the flow starts to separate and forms the
leading-edge vortex as illustrated in Fig. 9. With the assumption of no leading-edge suction or
the sharp leading-edge, Polhamus approximated the total lift as:

𝐶𝐿 = 𝐾𝑃 sin 𝛼 cos 2 𝛼 + 𝐾𝑣 sin2 𝛼 cos 𝛼 … (7)

where KV and KP, according to Traub [37], can be approximated as:

𝐾𝑃 = 4𝑡𝑎𝑛0.8 𝜀 … (8)
𝜋
𝐾𝑉 = … (9)
𝑠𝑖𝑛 Λ

17
Fig. 9 Polhamus leading-edge suction analogy (Polhamus, 1971)

However, Polhamus’s lift approximation well-agrees only until the vortex breakdown
reaches the trailing-edge due to its assumptions; and he also stated that this theoretical estimation
tends to over-predict the vortex lift for low-swept delta wing since the generated vortex cannot
provide full suction with respect to its trailing-edge.

2.3 Wing-tip vortex control techniques

In order to modify the strength, structure, and trajectory of the tip vortices, researchers have
attempted various wingtip vortex control techniques such as winglets, spoilers, stub/subwing,
and porous tips and leading-edges in the past. Such techniques can be classified into two
categories : passive and active tip vortex controls depending on whether the controller does the
work on the system or not. For the passive control methods, the modification of the vortex wake
is achieved in a steady manner without external forces. The passive wingtip vortex control
techniques are often conducted by simply changing the planform shape [38-40] or adding static
devices on the existing main wing [41-43] to modify the roll-up process of the tip vortices so that
the resulting vortex wake can be easily diffused.

18
On the other hand, the active control methods take the advantage of external forces through
the actuators to perturb the wake which will eventually promote the rapid diffusion of the
vorticity [44-46]. In addition to the above-mentioned techniques, more investigations on the
passive and active wingtip vortex control techniques including steady and oscillatory blowing
[47-48], static and rapidly actuated segmented Gurney flaps [49-50], active Gurney flaps [51],
oscillating winglet flaps [52], plasma actuators [53], and zero mass-flux fluidic perturbations [54]
had also been studied by researchers elsewhere.

Most recently, an alternative wingtip vortex control was investigated by Lee and Pereira
[55, 56] through the addition of a slender full-chord half-delta wing mounted to the tip of a static
rectangular NACA 0012 wing. Lee and Pereira [55, 56] found that the tip vortex generated
behind a rectangular NACA 0012 wing was greatly diffused and enlarged by a tip-mounted 65°-
sweep half-delta wing (HDW), as a result of the breakdown of the HDW vortex developed on the
upper surface of the half-delta wing. The HDW tip vortex control concept was inspired by the
pioneering work of [7-9] in which the HDW was attempted to improve the wing lift generating
capability. References [8-9] also reported that the lift increment and also the lift-curve slope can
be further enhanced by increasing the deflection of the HDW, relative to the main wing chord
line.
Lee and Pereira [55, 56] further noticed that, in addition to the observed changes in the size
and strength of the HDW-wing tip vortex, the lowered vorticity level also led to a much reduced
lift-induced drag coefficient CDi (=Di/½ρ∞u∞2S). The lift-induced drag Di was calculated based
on the Maskell wake integral model [57-60] (more details will be discussed in section 3.5).
The reduced CDi also translates into a smaller CD, especially at high CL range, compared to the
baseline wing at the same lift condition. Together with the HDW-induced CL increment, the zero-
deflection slender full-chord HDW was therefore capable of producing the best CL/CD
improvement, compared to the baseline wing, among all the deflections tested (for -10° ≤ δ ≤
+15°). This HDW tip vortex control concept is, however, shadowed by the undesired HDW-
induced increase in wing weight and bending moment. A 13.4% and 41.8% increase in the total

19
wing surface area and aspect ratio AR compared to the baseline wing, respectively, for the 65°-
sweep HDW wing were produced.

20
2.4 Objectives

Inspired by the work of Pereira [55, 56], this current study conducted similar experiment to
investigate the interaction between the baseline wing (BW) and the half-delta wing (HDW); and
had expanded the use of HDWs on the baseline wing with different geometric configurations.
The objective of this investigation was to examine the minimization of the above-mentioned
undesired effects incurred by the full-chord HDW (with a root chord cr equal to the baseline-
wing chord c or cr = c) through the use of small-chord HDWs with cr ≤ 0.5c at Re = 2.45×105.
Small-chord HDWs of different slenderness Λ (= 50° and 65°) and δ were also considered.
Special emphasis was placed on the impact of the small-chord HDW, of different Λ, cr, and δ, on
the near-field tip vortex at x/c = 2.8 at different angle of attack α. The growth and development
of the HDW vortex both along the tip and in the near field of the HDW wing, at α = 10°, was
also acquired to better understand the impact of small-chord HDWs on the tip vortex. The
behaviour of the leading-edge vortex developed over both the slender and nonslender full delta
wings was also characterized for a direct comparison. These tip-vortex flowfield measurements
were also supplemented by the wind-tunnel force balance measurements.

21
CHAPTER 3

Experimental Procedure

3.1 Flow facility

All experiments are conducted in Joseph Armand Bombardier open-loop wind tunnel in the
Aerodynamics Laboratory of McGill University (Fig. 10 & 11). The desired suction is achieved
by a variable speed AC motor which drives an isolated 2.5meter diameter fan consists of 16
blades. The high pitch noise is filtered using a specially designed acoustic silencer installed at the
exit of the flow. The tunnel has total length of 19m and first 3m is responsible for the contraction,
approximated to be 10:1 ratio, and conditioning of the inlet flow towards the test section. A
combination of 10mm honeycombs and 2mm screens are used to straighten the flow and the
freestream turbulence intensity measured to be within 0.05% at 35m/s. The test section has a
dimension of 0.9m×1.2m×2.7m in vertical (y), spanwise (z), and streamwise (x), respectively.

Fig. 10 The schematic diagram of J.A. Bombardier wind tunnel (Pereira, 2012)

22
Fig. 11 J.A. Bombardier wind tunnel

23
3.2 Seven-hole Pressure Probe

To investigate the vortex development, a miniature seven-hole pressure probe (SHP) is used.
The probe tip is made of brass and is machined to a cone shape with 30° angle. The outer
diameter of the probe tip is 2.7mm and has seven 0.5mm diameter pressure tap holes drilled
parallel to the probe shaft. By convention, the hole located at the center is numbered 7 and the
rest are numbered from 1 to 6 in clockwise direction and arranged in a 2.4mm diameter circle as
illustrated in Fig. 12. Each pressure tap is connected to a Honeywell DC002NDR5 differential
pressure transducer through a 1.6mm diameter Tygon tube. Due to the compressibility of the air,
the SHP is limited to time-mean measurement values. The output from the seven pressure
transducers then goes through a signal conditioner consists of a seven-channel analogue signal
differential amplifier with an external DC offset and is acquired using a 16-channel, 16-bit NI-
6259 A/D converter board. Finally, by comparing the relative pressure signals from different
pressure tabs and matching to a set of calibration data, the voltages can be converted into
velocities using a series of algorithms.

Fig. 12 Seven-hole pressure probe a) probe configuration b) details of the probe tip (Pereira, 2012)

24
The calibration of seven hole pressure probe is done on a two-degree of freedom custom
built traverse (Fig. 15) and follows the procedure of Wenger and Devenport [61]. During the
calibration procedure, the SHP is positioned over an angular range of -70° ≤ α ≤ 70° and
-70° ≤ β ≤ 70° where α is the pitch angle in x-y plane and β is the yaw angle in x-z plane.
Figure 13 is an example of a SHP calibration grid where circles refers to the location of known
parameters, such as cone angle θ, roll angle ϕ, static pressure coefficient CSTAT, and total
pressure coefficient CTOT.

2.5

1.5

0.5
C
-0.5

-1.5

-2.5
-0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

C

Fig.13 Typical SHP calibration grid

With the seven voltage outputs collected from the pressure transducers ({E}={E1, E2, ... ,
E7}), the seven hole pressure probe (SHP) is able to measures three orthogonal velocities
({v} = {u, v, w}). Since each set of {E} represents a single point in the calibration parameter
space S = {v}, there must be a unique and monotonic function f{E} so that f{E} = {v} exists.
By re-expressing the calibration parameter space S in terms of the magnitude and directions,
it can be written as S = {|v|, θ, ϕ} or S = {|v|, α, β}. Then, the magnitude |v| can be easily
eliminated from the parameter space resulting S = {θ, ϕ} or S = {α, β}, and each orthogonal
velocities can be obtained as follows :

25
𝑢 = |v| cos(β) cos(𝛼) = |v| cos(θ) … (10)

𝑣 = |v| cos(β) sin(𝛼) = |v| sin(θ) sin(𝜙) … (11)

𝑤 = |v| sin(β) = |v| sin(θ) cos(𝜙) … (12)

Now, if the flow angle respect to the pressure probe is large enough, the flow will separate
from the tip of SHP and at least one of the pressure taps will be in the wake region. However,
due to the insensitivity of the surface pressure in the wake region, any function f defined as
f{E}k = {v}k which incorporates the pressure from the tap in the wake region will not be
monotonic. Hence, the function f must be a piecewise function depending on whether the flow
separation over the probe tip exists or not, and if there is, which taps are in the separated region.
First of all, the voltage values collected from the pressure transducers have to be averaged and
converted in to pressures Pi. When the probe is aligned with (or parallel to) the flow, the flow is
assumed to be attached everywhere on the probe tip. Also, the pressure in the separated region is
known to be higher than that of the attached flow. Thus, if the reading from the center tap gives
the maximum value, the flow is assumed to be attached everywhere on the probe tip.

When the maximum pressure reading is taken from the center tap (hole 7 from Fig. 14), the
pressure difference between the top and the bottom taps (hole 1 and 4 in Fig. 14) will be
a function of pitch angle α, and the pressure differences between left and right taps will be
a function of yaw angle β. The magnitude of the pressure is a function of |v|; however, since |v|
has been eliminated from the parameter space S, the pressure is normalized by the local dynamic
pressure (= PTOT-PSTAT) where the total pressure PTOT is the direct reading from the center tap and
the static pressure PSTAT can be approximated by taking the average of the pressure readings from
peripheral taps (i.e. hole 1 to 6 from Fig. 14). Hence, the monotonic functions can be defined as
follows :

𝑃4 −𝑃1
𝑓𝛼 = ≡ 𝐶𝛼 ...(13)
𝑃7 −𝑃𝐴𝑉

26
1 1
(𝑃 −𝑃6 )− (𝑃3 −𝑃2 )
2 5 2
𝑓𝛽 = ≡ 𝐶𝛽 ...(14)
𝑃7 −𝑃𝐴𝑉

𝑃𝑖
𝑃𝐴𝑉 = ∑6𝑖=1 ≡ 𝐶𝛼 ...(15)
6

where i refers to the tap number as shown below in Fig. 14.

4
3 5
7
2 6
1

Fig.14 Pressure tap numbering convention


(Note that the size of tap holes are exaggerated for convenience)

On the other hand, if the maximum pressure reading is acquired from one of the peripheral
taps, the flow is assumed to be attached in the immediate vicinity of the stagnation point, thus the
tap with the maximum reading (tap i for convenient) and the two adjacent taps are used to define
the monotonic function. In this case, the pressure difference between tap i and the center tap 7 is
a function of the cone angle θ and the pressure differences between tap i and two other adjacent
taps are a function of the roll angle ϕ. The total pressure PTOT is again the direct reading from the
maximum pressure tap i, and the static pressure PSTAT is the average taken from the two adjacent
pressure taps PCW and PCCW. CW and CCW indicate immediate left and right holes of the
maximum pressure tap i. Thus, another case of high angle flow can be defined as follows :

𝑃𝑖 −𝑃7
𝑓𝜃𝑖 = 𝑖 = 1, 2, … , 6 ≡ 𝐶𝜃 ...(16)
𝑃𝑖 −𝑃𝐴𝑉

𝑃𝐶𝑊 −𝑃𝐶𝐶𝑊
𝑓𝛽 = ≡ 𝐶𝜙 ...(17)
𝑃𝑖 −𝑃𝐴𝑉

𝑃𝐶𝑊 −𝑃𝐶𝐶𝑊
𝑃𝐴𝑉 = ...(18)
2

27
Moreover, to reduce the error from total and static pressure approximations and gain more
accurate velocity measurements, the magnitude of the error is quantified for every point of
calibration parameter space S by introducing static and total pressure coefficients CSTAT i and
CTOT i, respectively, where i indicates the tap with maximum pressure recording.

𝑃𝑖 −𝑃𝑆𝑇𝐴𝑇
𝐶𝑆𝑇𝐴𝑇𝑖 = 𝑖 = 1, 2, … , 6 ...(19)
𝑃𝑖 −𝑃𝐴𝑉

𝑃𝑖 −𝑃𝑇𝑂𝑇
𝐶𝑇𝑂𝑇𝑖 = ...(20)
𝑃𝑖 −𝑃𝐴𝑉

where PAV is given in equation (15) or (18) depending on the case whether the maximum
pressure is recorded at the center tap or not (i.e. i = 7 or i ≠ 7).

Once, all the necessary parameters are acquired, equation (19) and (20) can be substituted
into the Bernoulli equation, or strictly speaking the steady Euler equation, to calculate the
magnitude of the velocity |v|. Considering that the vortices tend to be turbulent, the steady Euler
equation is still applicable in this study since we are taking the mean values.

2
|v| = √ (P𝑖 − PAV )(1 + CSTAT𝑖 − CTOT𝑖 ) ...(21)
ρ

where ρ is the local air density.

Finally, the velocity magnitude |v| can be substituted into equations (10), (11) and (12) to
yield each orthogonal velocities u, v and w.

28
Fig. 15 Angular calibration traverse (Pereira, 2012)

Fig. 16 Seven-hole pressure probe traverse mechanism

29
3.3 Traverse mechanism

The traversing of the SHP was achieved using a two degree of freedom custom built traverse
(Fig. 16) powered by two Sanyo-Denki model 103-718-0140 stepper motors for both vertical and
horizontal directions. Also, Labview was used for the auto-scanning in conjunction with a NI
PCI-7344 4-axis motion controller. To minimize the scanning time, an adaptive grid was
employed with the grid resolutions of Δy = Δz = 3.2mm for the inner core region,
Δy = Δz = 6.4mm for the outer core and wing’s wake region, and Δy = Δz = 12.7mm for the
freestream region. The streamwise vorticity was calculated from crossflow component and the
central difference scheme was used to evaluate the derivatives.

𝜕𝑣 𝜕𝑤 𝜕𝑣𝑗+1,𝑖 − 𝜕𝑣𝑗−1,𝑖 𝑤𝑗,𝑖+1 − 𝑤𝑗,𝑖−1


𝜁𝑖,𝑗 = ( − ) ≈ −( − ) … (22)
𝜕𝑧 𝜕𝑦 2∆𝑧 2∆𝑦

The core and outer circulations were calculated through Stokes’ theorem by integrating the
product of the vorticity and the area as follows:

Г𝑜 = ∑ ∑ 𝜁𝑖,𝑗 × ∆𝑦∆𝑧 𝑟𝑖,𝑗 < 𝑟𝑜 … (23)

Г𝑐 = ∑ ∑ 𝜁𝑖,𝑗 × ∆𝑦∆𝑧 𝑟𝑖,𝑗 < 𝑟𝑐 … (24)

where

2
𝑟𝑖,𝑗 = (𝑧𝑗 − 𝑧𝑐 ) + (𝑦𝑖 − 𝑦𝑐 )2 … (25)

𝑟𝑜 = 𝑟(𝜁 = 0.01𝜁𝑖,𝑗 𝑚𝑎𝑥 ) … (26)

30
3.4 Force Balance

A force balance is mounted on a turn-table which can be fitted into a ring installed in the test
section floor (Fig. 17). The wing models are connected to the sensor plate with a custom- built
adapting piece. The sensor plate is supported by two sets of cantilever-type spring flexures which
deflect whenever the force is applied on the sensor plate. Each individual flexure has maximum
deflection of 4mm. Each set of the flexures are aligned perpendicularly to each other, giving the
force balance two degrees of freedom, normal and tangential to the wing chord.

Fig. 17 Components of the force balance (Pereira, 2012)

The displacement of the sensor plate is then measured using two linear variable differential
transformers (LVDT), one for each direction, and converted into voltage signals. The resolution
of the LVDT was measured to be 68 and 56 Newtons per Volt in the normal and tangential
direction, respectively. Each LVDT consists of three coils and a single ferromagnetic core. The
central, or primary coil, induces a magnetic field within the core. As the core moves, it induces
voltages in the secondary coils. The voltage difference between these secondary coils is the

31
output which shows a very linear response (within 1%) in the range of calibration used. Then
this voltage difference is compared to the calibrated set of data to determine the magnitude of the
force acting on to the wing models.

3.5 Lift-induced drag (Di) calculation

The lift-induced drag Di calculation used in this experiment is based on Maskell's lift-
induced drag model [57]. In this section, a brief procedure of calculation of lift-induced drag Di
will be discussed. Further details can be reviewed in references [58-60] which extended the
Maskell wake integral model.

The derivation starts with defining the control volume V, upstream control surface S1, and
downstream control surface S2. Then, the required assumptions are as listed below :

 No mass flow ṁ across the boundaries of control volume V except S1 and S2


 Steady and incompressible flow
 Undisturbed inlet flow across control surface S1
 Steady wake in the scanning plane S2
 Adiabatic flow with in control volume V
 No external force applied to the control volume V

Given these assumptions and by applying the momentum balance on the control volume V,
the drag D can be expressed as follows :

2 )𝑑𝑦𝑑𝑧
𝐷 = ∬(𝑃∞ + 𝜌∞ 𝑈∞ − ∬(𝑃 + 𝜌𝑢2 )𝑑𝑦𝑑𝑧 … (27)
𝑆1 𝑆2

32
And from the continuity yields constant mass flow rate ṁ across control surfaces S1 and S2,

∬ 𝜌∞ 𝑈∞ 𝑑𝑦𝑑𝑧 = ∬ 𝜌𝑢 𝑑𝑦𝑑𝑧 … (28)


𝑆1 𝑆2

Hence, equation (27) can be re-arranged,

𝐷 = ∬[𝜌𝑢(𝑈∞ − 𝑢) + (𝑃∞ − 𝑃)]𝑑𝑦𝑑𝑧 … (29)


𝑆2

Kusenose [59] then re-arranged equation (29) with the definition of the speed of the sound, the
equation of state, the conservation of total enthalpy, the enthalpy change along the streamline,
and the definition of entropy, and expressed as follows :

𝐷 = 𝐷𝑃 + 𝐷𝑖 + 𝐷𝑀 + ∬ 𝑂(𝛥𝑢3 )𝑑𝑦𝑑𝑧 . . . (30)


𝑆2

where

𝑃𝑇𝑂𝑇∞
𝐷𝑃 = 𝑃∞ ∬ 𝑙𝑛 ( ) 𝑑𝑦𝑑𝑧 . . . (31)
𝑆2 𝑃𝑇𝑂𝑇

1
𝐷𝑖 = ∬ 𝜌 (𝑣 2 + 𝑤 2 )𝑑𝑦𝑑𝑧 . . . (32)
𝑆2 2 ∞

1 2
𝐷𝑀 = − ∬ 𝜌 (1 − 2𝑀∞ )𝛥𝑢2 𝑑𝑦𝑑𝑧 . . . (33)
𝑆2 2 ∞

DP is the profile drag, the pressure difference between the inlet control surface S1 and the wake
scanning plane S2. Di is the lift-induced drag associated with the cross-flow momentum. DM is the

33
Mach wave drag which considers the compressibility effect. The last is the high-order term
which can be neglected under some flow conditions. Again, the derivation of these terms,
equation (30) to (33), are discussed in great details in Kusenose [59].

The cross-flow components v and w can be decomposed into a stream function (y,z) and a
velocity potential function(y,z) to satisfy the vorticity and continuity equations identically
as follows :
𝜕𝜓 𝜕𝜙
𝑣= + . . . (34)
𝜕𝑧 𝜕𝑦

𝜕𝜓 𝜕𝜙
𝑤=− + . . . (35)
𝜕𝑦 𝜕𝑧

Then, equation (32) can be reduced to :

1 𝜕𝜙 𝜕𝜓 2 𝜕𝜙 𝜕𝜓 2
𝐷𝑖 = 𝜌 ∬ ( + ) + ( − ) 𝑑𝑦𝑑𝑧 . . . (36)
2 𝑆2 𝜕𝑦 𝜕𝑧 𝜕𝑧 𝜕𝑦

Also, by introducing a source term σ, defined as :

𝜕𝑢
σ=− . . . (37)
𝜕𝑥

which can be re-written due to the conservation of mass,

𝜕𝑣 𝜕𝑤
σ= + . . . (38)
𝜕𝑦 𝜕𝑧

Equation (32) can be further simplified to :

1
𝐷𝑖 = 𝜌 ∬ 𝜓ζ − 𝜙𝜎 𝑑𝑆 . . . (39)
2 𝑤𝑎𝑘𝑒

where ζ is the vorticity

34
Also, equation (39) requires two boundary conditions.

 Streamlined tunnel walls, (i.e. ψ(wall) = 0)


𝜕𝜙
 No flow across the walls, (i.e. 𝜕𝑛 = 0)

Hence, for the calculation of the lift-induced drag Di, one only needs to define four parameters -
vorticity ζ , source term 𝜎, stream function 𝜓, and velocity potential 𝜙.

In order to apply the Maskell's solution to calculate the lift-induced drag Di from the wake
survey measurement plane, which is a mn grid, the vorticity ζ and the source term σ are
expressed in equations (40) and (41) with the cross-flow vectors using centered finite-differences.

𝛥𝑤 𝛥𝑣 𝑤𝑖−1,𝑗 − 𝑤𝑖+1,𝑗 𝑣𝑖,𝑗+1 − 𝑣𝑖,𝑗−1


ζi,j = − =( )−( ) . . . (40)
𝛥𝑦 𝛥𝑧 2𝜂 2𝜂

𝛥𝑣 𝛥𝑤 𝑣𝑖−1,𝑗 − 𝑣𝑖+1,𝑗 𝑤𝑖,𝑗+1 − 𝑤𝑖,𝑗−1


𝜎𝑖,𝑗 = + =( )−( ) . . . (41)
𝛥𝑦 𝛥𝑧 2𝜂 2𝜂

where η = Δy = Δz, i = 2, 3, ... , n-1, j = 2, 3, ..., m-1

Similarly, the stream function 𝜓 and the velocity potential 𝜙 can be approximated as equations
(42) and (43) using the central difference method and Poisson's equation.

1
ζi,j = −𝛻 2 𝜓𝑖,𝑗 ≈ − (𝜓 + 𝜓𝑖−1,𝑗 + 4𝜓𝑖,𝑗 + 𝜓𝑖,𝑗+1 + 𝜓𝑖,𝑗−1 ) . . . (42)
𝜂2 𝑖+1,𝑗

1
𝜎𝑖,𝑗 = 𝛻 2 𝜙𝑖,𝑗 ≈ (𝜙 + 𝜙𝑖−1,𝑗 − 4𝜙𝑖,𝑗 + 𝜙𝑖,𝑗+1 + 𝜙𝑖,𝑗−1 ) . . . (43)
𝜂2 𝑖+1,𝑗

After applying the boundary conditions along the tunnel walls, these relations results in a system
of (m-2)  (n-2) equations to be solved simultaneously which also can be expressed as A𝑥⃗ = 𝑏⃗⃗

35
form. A is a sparse, square matrix with (m-2)(n-2)  (m-2)(n-2) coefficients. 𝑏⃗⃗ is the vector of the
known quantities (ζ and 𝜎) whereas 𝑥⃗ is that of unknowns (𝜓 and 𝜙).

Once, the required elements have been computed, the lift-induced drag can be approximated
based on equation (39).

m−1 n−1
1
Di ≈ ρ ∑ ∑(𝜓ji ζji − 𝜙ji 𝜎ji )η2 . . . (44)
2
j=2 i=2

where the flow is assumed to be incompressible and the source term σ is often neglected due to
its small magnitude.

36
3.6 Wing models

A rectangular, square-tipped, aluminum NACA0012 wing with chord of c = 28cm and semi-span
of s = 50.8cm with the aspect ratio AR = 3.6 was used as a baseline wing. The thickness at the
trailing-edge of this rectangular wing is 1.5mm. This wing has its pitching axis located at ¼-
chord location and is mounted vertically at the center of the test section of the wind tunnel during
the experiment. The freestream velocity was fixed at u∞ = 13.5m/s to render the chord Reynolds
number of Re = 2.45×105. The half-delta wing models with different sweeps Λ, root chords cr,
and deflections δ are attached to the square tip of the baseline wing and the parameters of the
assembled wings are provided with the Table 1. Note that for the HDW configuration, the total
wing area Stotal and total semi span s were used in the normalization of the lift and drag forces as
well as the lift-induced drag. Figure 18 illustrates the schematics of the experimental set-up. The
thickness of the HDWs are made of aluminum plate with thickness t = 3.2mm (or t/c = 1.1%).
The origin of the coordinate system is located at the leading-edge of the baseline wing. Most of
the cases, the baseline wing, equipped with HDWs, is mounted vertically (Fig. 19); however, for
specific cases, i.e., 50HDW, the wing model is mounted horizontally (Fig. 20). This is to avoid
the wall effect of the wind tunnel since when the wing model is equipped with 50HDW, there is
only a 10cm gap between the ceiling and the wing model.

Case Configuration Λ cr δ s(cm) ΔS(%) AR ΔAR(%)


A BW n/a 1c n/a 50.8 0 3.64 0
B 1c65HDW(0) 65° 1c 0° 63.9 4 5.16 41.8
C 1c50HDW(0) 50° 1c 0° 74.2 23.1 6.32 73.6
D 0.5c65HDW(0) 65° 0.5c 0° 57.4 3.2 4.48 23.3
E 0.5c50HDW(0) 50° 0.5c 0° 62.5 5.8 5.20 43.2
F 0.5c50HDW(+5) 50° 0.5c +5° 62.5 5.8 5.20 43.2
G 0.5c50HDW(-5) 50° 0.5c -5° 62.5 5.8 5.20 43.2
H 0.3c50HDW(0) 50° 0.3c 0° 57.9 2.1 4.62 26.6
(Notes: ΔS and ΔAR denote increase in percentage of S and AR, respectively.
1c65HDW (0) denotes HDW with cr = c, Λ = 65° and δ = 0°.)
Table 1 HDW wing configuration and geometric details

37
Fig. 18 Schematics of experimental set-up

38
Fig. 19 Clockwise from top left a) 50HDW, b) 65HDW, c) 0.5c 50HDW, d) 0.3c 50HDW

Fig. 20 Wing model in the test section

39
CHAPTER 4

Result and Discussion

The influence of the slenderness Ʌ, root chord cr, and deflection δ of the HDW on the tip vortex
was examined first, followed by the discussion of force-balance measurements and lift-induced
drag computation. The figures in this chapter (Fig. 21-25) can be found at the end of the chapter
(page 45-52) and the original versions of the modified figures are also provided in Appendix A
except for 1c65HDW, 0.3c50HDW and deflected HDW cases of which data were taken only at
few selective angles of attack (α = 6°, 8° and 10°) and are already included in Figures 21-25.

4.1 Flow analysis for full-chord or cr = c HDW

Figures 21b1-c3 show the impact of sweep or slenderness of the full-chord HDW on the
normalized iso-vorticity (ζc/u∞) contours of the tip vortex at x/c = 2.8 for α = 6°, 8° and 10°
(covering low-to-medium α regime). The HDW deflection was fixed at δ = 0°.
The streamwise vorticity ζ (= ∂w/∂y − ∂v/∂z) was calculated from the crossflow (vw)
measurements by using a central differencing scheme to evaluate the derivatives. The tip vortex
generated behind the baseline wing (BW) is also included in Figs. 21a1-a3 for a direct
comparison. Figures 21a1-a3 show that the BW tip vortex always remained concentrated and had
the highest normalized peak vorticity ζp (= ζpeakc/u∞; indicative of vorticity level) compared to
all the HDW wings tested in the present study. For the BW, ζp was also found to increase with α
(up to the static-stall angle αss), as a result of the interaction and merging of the secondary
vortices with the main tip vortex along the baseline-wing tip (for x/c < 1; Fig. 22a). The tip
vortex continued to progress downstream of the BW trailing edge. For x/c ≥ 2.5, the inner-flow
region of the tip vortex was found to attain nearly axisymmetric (see [60,62] for details).

40
Figures 21b1-c3 show that the addition of the full-chord HDWs, however, led to an enlarged
and diffused tip vortex with a lower ζp compared to their BW counterpart. The degree of
diffusion increased as Λ was decreased. For the zero-deflection full-chord 65°-sweep slender
HDW (designated as 1c65HDW) wing, the lower-than-BW ζp was, however, found to remain
rather insensitive to the change in α (Figs. 21b1-b3). The lowered vorticity level can be attributed
to the HDW vortex breakdown on the slender half-delta wing surface, as illustrated in Fig. 22b in
comparison to the leading-edge vortex (LEV) observed on an otherwise full delta wing of the
same sweep and at similar α (see Fig. 22c). The full delta wing and HDW wing configurations
are also illustrated in the insets at the upper left corner of Fig. 22. Figure 22b show that at
α = 10°, the HDW vortex breakdown was observed between x/c = 0.85 and 0.9 (denoted by the
large drop in the peak vorticity of the HDW vortex) on the 1c65HDW wing, resulting in the
diffused tip vortex presented in Fig. 21b3.
It should be noted that it has been shown that the presence of a probe can affect the location
of vortex breakdown, particularly when the breakdown is near the trailing edge of the wing.
However, when the probe is not in the close proximity of the breakdown location, its interference
is minimal [63]. In contrast to the HDW vortex breakdown observed at around x/c = 0.85-0.9 at
α = 10°, no LEV breakdown was observed on a slender full delta wing positioned at α = 12°
(see Fig. 22c). The observed HDW vortex breakdown, depicted in Fig. 22b, can therefore be
attributed to the separated flow originating from the junction of the baseline-wing tip and the
HDW root.

For the nonslender HDW with Λ = 50° and cr = c (designated as 1c50HDW), the tip vortex
became surprisingly weak (circulation flow-like) with a significantly enlarged size and lowered
vorticity level or ζp compared to its slender counterpart (Figs. 21c1-c3). The presence of the
circulation flow-like tip vortex can be attributed to the close proximity of the HDW vortex to the
1c50HDW surface, which in turn, led to HDW vortex breakdown at x/c ≈ 0.5 (not shown here).
Nevertheless, the closer proximity of the nonslender HDW vortex and its earlier-than-slender-
HDW vortex breakdown can be demonstrated from the LEV behavior observed on both the

41
slender and nonslender full delta wing was found to occur at around x/c ≈ 0.7 for α = 10° while
no LEV breakdown was observed over a full slender delta wing at α = 12°. The closer proximity
of the LEVs to the nonslender full delta wing surface compared to the slender delta wing is
evident in Fig. 23a. The streamwise variation of ζp of the LEVs is also summarized in Fig. 23b
for a direct comparison. Figures 23a-b clearly indicate that the LEV of the slender delta wing had
a higher vorticity level and also was located further above the wing surface compared to its
nonslender counterpart. Details of the vortex flow over slender and nonslender delta wings are
given by [25,64].
In summary, despite the observed large impact of the 1c50HDW on the tip vortex, the
addition of the 1c50HDW was, however, accompanied by a large increase in Stotal and AR
(Table 1), which can translate into an increased wing weight and bending moment. To relieve the
HDW-induced large increase Stotal and AR, HDWs of smaller chord (i.e., cr ≤ 0.5c) were
considered (see section 4.3). For now, the influence of HDW on the aerodynamic characteristics
is discussed.

4.2 CL, CD, CL/CD and CDi analysis

Figure 24a shows that the presence of full-chord HDWs, with δ = 0°, can produce an increase CL
compared to the BW. A 22.4% and 13.4% increase in CL, for example, at α = 10°
(a representative medium-α regime) was obtained for the 1c50HDW and 1c65HDW wings,
respectively, in comparison to the BW (with CL = 0.67). The observed CL increment can be
caused by the HDW-induced large increase in AR (see Table 1) and the improved wing tip
conditions. The force-balance measurements also show that full-chord HDWs also gave rise to a
reduced CD, especially for CL > 0.3 regime, regardless of Λ (see Fig. 24b), and subsequently an
improved CL/CD ratio compared to the BW at the same lift condition (see □ and ○ symbols in Fig.
24c). The 1c50HDW wing, however, produced a lower CD compared to its slender counterpart.
Also, the smaller the Λ of the full-chord HDW, the better the CL/CD ratio.

42
For the cr = 0.5c nonslender HDW (i.e., 0.5c50HDW) wing, the CL increment was
consistently lower than its full-chord counterpart while the CD reduction remained comparable.
More importantly, the 0.5c50HDW wing was found to produce the best CL/CD performance
among all HDWs tested (Fig. 24c). The cause for the observed CD reduction can be examined via
the lift-induced drag coefficient CDi computation based on the vw-velocity measurements by
using equation (39) or (44). Figures 24d shows that the CDi of the 0.5c50HDW wing always had a
lower-than-BW value, especially for the high α or lift regime. The full-chord HDWs, however,
produced a further reduction in CDi compared to the cr = 0.5c HDW wing. The 1c65HDW was
found to generate the smallest CDi or CDi/CD ratio (Fig. 24e). Figure 24e further indicates that at
higher α, the total CD is dominated by CDi.
To further explore the potential of 0.5c50HDW tip-vortex control concept, the effect of
HDW deflection was also examined. Figures 24a-d show that the δ = +5° deflection was capable
of generating a CL increment similar to the δ = 0° case, while the CD reduction was of a much
lesser extent. Meanwhile, the CDi was lower than the BW value but was higher than its δ = 0°
counterpart for CL > 0.625. The CL/CD had a near-to-BW value.

The effect of the size or root chord of the HDW on the aerodynamic characteristics was
further investigated by the use of the 0.3c50HDW wing. A 10% increase in CL and 23% decrease
in CDi at α = 10° compared to the BW were obtained. The CL/CD, however, remained somewhat
inferior to the BW value at the same lift condition. The 0.3c50HDW was also able to produce
a reduced CDi similar to that of the 0.5c50HDW wing (see Fig. 24d) and also a CDi/CD ratio
comparable to the 1c65HDW wing (see Fig. 24e).

The observed CL increment and CDi reduction achieved by the employment of the
0.3c50HDW, which rendered a 2.1% and 26.6% increase in Stotal and AR compared to the BW,
deserves some articulation. Generally, AR is illustrative of the efficiency of a wing and also the
mitigation of the 3-D effects on wing performance. This is particularly true for rectangular or

43
swept planform wings, where a large AR necessarily means a proportionally large span, thus
resulting in a minimized tip vortex effect on the total wing. This minimization comes not directly
from the attenuation of the 3-D effects but by instead reducing the proportion of impacted wing
area. When dealing with the present 0.3c50HDW wing configuration, the AR should become less
representative due to the HDW's triangular planform. The 2.1% increase in Stotal thus leads to a
26.6% increase in AR. The observed performance increases compared to the BW therefore
should not be solely attributed to the AR, since the area impacted by the 3-D effects is virtually
unchanged. Instead, the improvement can be, to a larger extent, attributed to a modification of
the tip effects. Nevertheless, to reinforce the observed changes in CL, CD, CL/CD and CDi, the tip
vortex generated by the small-chord HDW wings with cr ≤ 0.5c are discussed.

4.3 Flow analysis for small-chord or cr ≤ 0.5c HDW

Figures 21d1-e3 show that the tip vortex generated by HDW wings with cr ≤ 0.5c is
characterized by a double-vortex pattern (i.e., the coexistence of a BW vortex, originating from
the free tip portion of the rectangular NACA 0012 wing, and a HDW vortex) at low α. The BW
vortex remained concentrated and had a lower (or higher) vorticity level than the HDW vortex
for slender (or nonslender) HDWs. The BW vortex was also found to locate below the HDW
vortex. Meanwhile, the HDW vortex moved toward and engulfed the BW vortex, causing a
greatly enlarged and circulation flow-like tip vortex as α was increased.
The nonslender HDW produced a more diffused and enlarged tip vortex compared to its slender
counterpart. Typical formation and growth of the double vortex and their interaction along the
0.5c50HDW wing and its near field for 0 < x/c ≤ 2.8 are displayed in Fig. 22e.
The movement of the tip vortex trajectory as a function of α is also summarized in
Figs. 25a-b. For the HDW wing, there was always a large upward and outboard movement of the
tip vortex compared to the baseline wing. The higher the α, the more outboard vertical
displacement of the tip vortex. Meanwhile, the baseline-wing tip vortex was found to be

44
displaced slightly upward and outboard by the low-pressure viscous wake, and was also moved
downwards with increasing α.

The present measurements also show that the interaction of the double vortex was expedited
by upward HDW deflection with δ = +5°, rendering again a circulation flow-like vortex flow for
α > 8° (Figs. 21f1-f3). The double-vortex pattern, however, persisted up to x/c = 2.8 for
downward HDW deflection with δ = -5°, regardless of α (Figs. 21g1-g3). For the 0.3c50HDW
wing, the double-vortex pattern also appeared and had a lower ζp and was closer to each other
(Figs. 21h1-h3) in comparison to its cr = 0.5c counterpart for α ≤ 6°. The close proximity of the
double vortex expedited the interaction and merging for α > 6°. The merged tip vortex, however,
had a higher ζp compared to the 0.5c50HDW case. The growth, interaction and merging of the
double vortex associated with the 0.3c50HDW wing for
0 < x/c ≤ 2.8 is illustrated in Fig. 22f.

Finally, the total circulation Γo of the tip vortex was also determined and is summarized in
Fig. 25c. The Γo was calculated by summing the vorticity multiplied by the incremental area of
the measuring grid. The normalized Γo of the HDW wing-generated tip vortex was found to
increase with increasing α (for α < αss) and was always above the BW value, which is in good
agreement with the CL increment observed in Fig. 24a. The small-chord nonslender HDWs
produced a larger Γo compared to their slender counterpart.

45
46
Fig. 21 Impact of Λ, cr and δ of the HDW on the normalized iso-vorticity contours of the tip vortex at
x/c = 2.8 for α = 6°, 8° and 10°. (a1)-(a3): BW; (b1)-(b3): 1c65HDW; (c1)-(c3): 1c50HDW;
(d1)-(d3): 0.5c65HDW; (e1)-(e3): 0.5c50HDW; (f1)-(f3): 0.5c50HDW(+5);
(g1)-(g3): 0.5c50HDW(-5); and (h1)-(h3): 0.3c50HDW. ζp denotes ζpeakc/u∞.

47
48
49
Fig. 22 Growth and development of HDW vortex and BW vortex on half-delta wing configuration
and LEV on full delta wing. (a) BW at α = 10°, (b) 1c65HDW wing at α = 10°,
(c) 65°-sweep full delta wing at α = 12°, (d) 50°-sweep full delta wing at α = 10°,
(e) 0.5c50HDW at α = 10°, and (f) 0.3c50HDW at α = 10°

50
Fig. 23 Variation of normalized (a) vertical position and (b) peak vorticity of LEV with x/c on
65°-sweep full delta wing at α = 12° and 50° -sweep full delta wing at α = 10°

51
Fig. 24 Effect of slenderness, root chord and deflection of HDW on aerodynamic characteristics

52
Fig. 25 Impact of Λ, cr and δ of the HDW on the trajectory and total circulation of the tip vortex at
x/c = 2.8. Open and solid symbols denote BW vortex and HDW vortex, respectively.

53
CHAPTER 5

Conclusions

The tip vortex developed behind a rectangular NACA 0012 wing with tip-mounted HDWs, of
different Λ, cr and δ, was investigated in a subsonic wind tunnel. The results show that,
regardless of Λ, cr and δ, the addition of HDWs always led to an enlarged and diffused tip vortex
and increased CL compared to the baseline wing. Such increment in CL was mainly attributed to
the mitigated free-end effect, or improved tip conditions, via isolation of the baseline wing tip
and, to a lesser extent, the increased total surface area and AR. The near-field vortex flow behind
the small-chord HDW, however, exhibited a double-vortex pattern, which interacted and merged
with each other as it progressed downstream, forming a circulation-like flow. The mutual
interaction and merge of the double vortex was found to be expedited by upward HDW
deflection. The smaller the cr of the nonslender HDW, the lower the lift-induced drag. The zero-
deflection cr = 0.5c nonslender HDW wing produced the largest lift-to-drag ratio among all the
HDWs tested.

Towards a further understanding of the modification and decay of tip vortex using HDWs,
the investigations at higher Reynolds number conditions are required, as well as extending the
present near wake region to the intermediate and far wake regions. Current passive tip vortex
control can also be extended to active tip vortex control studies in an attempt to modify the
interaction and merge of the double vortex by putting the wing or the tip-mounted HDW alone in
motions (heaving, pitching, and etc). Therefore, the enlarged and diffused tip vortex structure
under the presence of HDWs suggests potential future researches which can employ both passive
and active HDWs for further understanding of tip vortex control.

54
References

[1] Spalart, P.R., (1998) "Airplane trailing vortices", Annual Review of Fluid Mechanics,
30, pp.107-138.

[2] Chow, J. S., Zilliac, G. G., and Bradshaw, P. (1997) "Mean and turbulence
measurements in the near field of a wingtip vortex", AIAA Journal, 35(10):1561–1567.

[3] Whitcomb, R. (1976) "A design approach and selected wind-tunnel results at high
subsonic speeds for wing-tip mounted winglets", NASA-TN-D-8260.

[4] Corsiglia, V., Jacobsen, R., and Chigier, N. (1970) "An experimental investigation of
trailing vortices behind a wing with a vortex dissipator", In Olsen, J., Goldburg, A., and
Rodgers, M., editors, Aircraft Wake Turbulence, pages 229–242. Aircraft Wake
Turbulence Symposium, Seattle Washington.

[5] Patterson, J. C. (1975) "Vortex attenuation obtained in the Langley vortex research
facility", Journal of Aircraft, 12(9):745–749.

[6] Gratzer, L. (1992)"Spiroid-tipped wing", U.S. Patent No. 5,102,068.

[7] Staufenbiel, R. and Vitting, T. (1991) "On aircraft wake properties and some methods
for stimulating decay and breakdown of tip vortices", In AGARD Conference
Proceedings 494, Vortex Flow Aerodynamics, Advisory Group for Aerospace Research
and Development, Neuilly Sur Seine, France.

[8] Nikolic, V.R . (2005) "Movable tip strakes and wing aerodynamics", Journal of Aircraft,
2(6): 1418–1426.

[9] Nikolic, V.R. (2011) "Optimal Movable Wing Tip Strake", Journal of Aircraft,
48(1): 335-341.

[10] Betz, A. (1932) "Verhalten von wirbelsystemen", Zeitschrift fuer Angewandte Mathematik
und Mechanik, 12(3):164–174. (Translation included in NACA Technical
Memorandum TM-71, 1932).

[11] Moore, D.W. and Saffman, P.G. (1973) "Axial flow in laminar trailing vortices",
Proceedings of the Royal Society. A333:491-508.

[12] Batchelor, G. (1964) "Axial flow in trailing line vortices", Journal of Fluid Mechanics,
20(4):645–658.

55
[13] Hoffman and Joubert (1963) "Turbulent line vortices", Journal of Fluid Mechanics,
16(3):395–411.

[14] Phillips, W. (1981). "The turbulent trailing vortex during roll-up", Journal of Fluid
Mechanics, 105:451–467.

[15] Chigier, N. and Corsiglia, V. (1971) "Tip vortices - velocity distributions",


NASA TM-X-62.

[16] Orloff, K. (1974) "Trailing vortex wind-tunnel diagnostics with a laser velocimeter",
Journal of Aircraft, 11(8):477–482.

[17] Brown, C. (1974) "Aerodynamics of wake vortices", AIAA Journal, 11(4):531–536.

[18] Thompson, D. (1975) "Experimental study of axial flow in wing tip vortices", Journal of
Aircraft, 12:910–911.

[19] McAlister, K. and Takahashi, R. (1991) "NACA 0015 wing pressure and trailing vortex
measurements", NASA-TP-3151.

[20] Shekkariz, A., Fu, T., and Katz, J. (1993) "Near-field behaviour of a tip vortex",
AIAA Journal, 31(2):112–118.

[21] Devenport, W., Rife, M., Liapis, S., and Follin, G. (1996) "The structure and
development of a wing-tip vortex", Journal of Fluid Mechanics, 312:67–106.

[22] Dacles-Mariani, J., Zilliac, G., Chow, J., and Bradshaw, P. (1995) "Numerical
and experimental study of a wingtip vortex in the near field", AIAA Journal,
33(9): 1561–1573.

[23] Ramaprian, B. and Zheng, Y. (1997) "Measurements in rollup region of the tip
vortex from a rectangular wing", AIAA Journal, 35(12):1837–1843.

[24] Birch, D., Lee, T., Mokhtarian, F., and Kafyeke, F. (2003) "Rollup and near-field
behaviour of a tip vortex", Journal of Aircraft, 40(6): 603–607.

[25] Gursul, I., Gordnier, R. and Visbal, M. (2005) "Unsteady aerodynamics of nonslender
delta wings", Progress in Aerospace Sciences, 41(7): 515-557.

[26] Earnshaw, P. (1961) "An experimental investigation of the structure of a leading edge
vortex", RAE TN 2740.

56
[27] Menke, M. and Gursul, I. (1997) "Unsteady nature of leading edge vortices",
Physics of Fluids, 9(10): 2960–6.

[28] Menke, M., Yang, H. and Gursul, I. (1999) "Experiments on the unsteady nature of
vortex breakdown over delta wings", Experiments in Fluids, (27): 262–72.

[29] Gordnier, R.E. and Visbal, M.R. (2003) "Higher-order compact difference scheme
applied to the simulation of a low sweep delta wing flow", AIAA 2003-0620, 41st
AIAA Aerospace Sciences Meeting and Exhibit, 6–9 January 2003, Reno, NV.

[30] Taylor, G.S., Schnorbus, T. and Gursul, I. (2003) "An investigation of vortex flows over
low sweep delta wings", AIAA-2003-4021, AIAA Fluid Dynamics Conference, 23–26
June, Orlando, FL.

[31] Yaniktepe, B. and Rockwell, D. (2004) "Flow structure on a delta wing of low sweep
angle", AIAA Journal, 42(3): 513–23.

[32] Nelson, R. and Pelletier, A. (2003) "The unsteady aerodynamics of slender wings and
aircraft undergoing large amplitude maneuvers", Progress in Aerospace Sciences,
39:185–248.

[33] Visser, K. and Nelson, R. (1993) "Measurements of circulation and vorticity in the
leading- edge vortex of a delta wing", AIAA Journal, 31(1):104–111.

[34] Hall, M. (1972) "Vortex breakdown", Annual Review of Fluid Mechanics, 4:195–217.

[35] Wentz, W. and Kohlman, D. (1971) "Vortex breakdown on slender sharp-edged wings",
Journal of Aircraft, 8(3):156–161.

[36] Polhamus, E. (1971) "Predictions of vortex-lift characteristics by a leading-edge


suction analogy", Journal of Aircraft, 8(4):193–199.

[37] Traub, L. W. (2000) "Implications of the insensitivity of vortex lift to sweep", Journal of
Aircraft, 37(3):531–533.

[38] Rossow, V. J. (1975) "Theoretical Study of Lift Generated Vortex Wakes Designed to
Avoid Rollup", AIAA Journal, 13(4): 476–484.

[39] Holbrook, G. T., Dunham, D. M. and Greene G. C. (1985) "Vortex Wake Alleviation
Studies With a Variable Twist Wing", NASA TP 2442.

[40] Graham, W. R., Park, S.-W. and Nickels, T. B. (2003) "Trailing Vortices from a Wing
with a Notched Lift Distribution", AIAA Journal, 41(9), 1835–1838.

57
[41] Croom, D. R. (1977) "Evaluation of Flight Spoilers for Vortex Alleviation", Journal of
Aircraft, 14(8), 823–825.

[42] Schell, I., Özger, E. and Jacob, D. (2000) "Influence of Different Flap Settings on the
Wake-Vortex Structure of a Rectangular Wing with Flaps and Means of Alleviation
with Wing Fins", Aerospace Science and Technology, (4): 79–90.

[43] Heyes, A. L. and Smith, D. A. R. (2005) "Modification of a Wing Tip Vortex by Vortex
Generators", Aerospace Science and Technology, (9): 469–475.

[44] Bilanin, A. J. and Widnall, S. E. (1973) "Aircraft Wake Dissipation by Sinusoidal


Instability and Vortex Breakdown", AIAA Paper 73–107.

[45] Rebours, R., Kliment, L. K. and Rokhsaz K. (2004) "Forced Response of a Vortex
Filament Pair Measured in a Water Tunnel", Journal of Aircraft, 41(5): 1163–1168.

[46] Bearman, P., Heyes, A., Lear, C. and Smith, D. (2006) "Natural and Forced Evolution
of a Counter Rotating Vortex Pair", Experiments in Fluids, 40(1): 98–105.

[47] Breitsamter, C. and Allen, A. (2009) "Transport aircraft wake influenced by oscillating
winglet flaps", Journal of Aircraft, 46(1): 175-188.

[48] Duraisamy, K. and Baeder, J.D. (2003) "Control of tip vortex structure using steady and
oscillatory blowing", AIAA Paper 2003-3407.

[49] Greenblatt, D., Vey, S., Paschereit, O.C., and Meyer, R. (2009) "Flap vortex
management using active Gurney flaps", AIAA Journal, 47(2): 2845-2856.

[50] Greenblatt, D. (2012) "Fluidic control of a wing tip vortex", AIAA Journal,
50(2): 375-386.

[51] Hasebe, H., Naka, Y. and Fukagata, K. (2011) "An attempt for suppression of wing-tip
vortex using plasma actuators", Journal of Fluid Science and Technology, 6(6): 419-437.

[52] Matalanis, C.G. and Eaton, J.K. (2007) "Wake vortex control using static segmented
Gurney flaps", AIAA Journal, 45(2): 321-328.

[53] Matalanis, C.G. and Eaton, J.K. (2007) "Wake vortex control using rapidly actuated
segmented Gurney flaps", AIAA Journal, 45(8): 1847-1884

[54] Taira, K. and Colonius, T. (2009) "Effect of tip vortices in low-Reynolds-number


poststall flow control", AIAA Journal, 47(3): 749-755.

58
[55] Pereira, J. (2012) "Experimental Investigation of tip vortex control using a half delta
shaped tip strake", McGill University, Canada.

[56] Lee, T. and Pereira, J. (2013) "Modification of static-wing tip vortex via a slender half-
delta wing", Journal of Fluids and Structures, 43:1-14.

[57] Maskell, E. (1973) "Progress towards a method for the measurement of the components
of the drag of a wing of finite span", RAE Technical Report 72232.

[58] Brune, G.W. (1994) "Quantitative low-speed wake surveys", Journal of Aircraft,
31(2): 249-255.

[59] Kusonose, K. (1998) "Drag reduction based on a wake-integral method", AIAA-98-2723.

[60] Birch, D., Lee, T., Mokhtarian, F., and Kafyeke, F. (2004) "Structure and induced drag
of a tip vortex", Journal of Aircraft, 41(5):1138–1145.

[61] Wenger, C. and Devenport, W. (1999) "Seven-hole pressure probe calibration utilizing
look-up error tables", AIAA Journal, 37:675–679.

[62] Lee, T. and Pereira, J. (2010) "On the nature of wake- and jet-like axial tip-vortex flow",
Journal of Aircraft, 47(6): 1946-1954.

[63] Payne, F.M., Ng, T.T. and Nelson, R.C. (1989) "Seven-hole probe measurement of
leading-edge vortex flows", Experiments in Fluids, 7(1): 1-8.

[64] Gursul, I. (2005) "Review of unsteady vortex flows over slender delta wings", Journal of
Aircraft, 42(2): 299–319.

[65] McCormick, B.W. (1979) "Aerodynamics, Aeronautics, and Flight Mechanics",


John Wiley and Sons, New York.

[67] RHRC. (2000) "Delta wing experiment", Rolling Hills Research Corporation, California.

[68] Taylor, G.S., Schnorbus, T., and Gursul, I. (2003) "An investigation of a vortex flow over
low sweep delta wings", 33rd AIAA Fluid Dynamics Conference and Exhibit, 2003.

59
Appendix: Normalized iso-vorticity& axial velocity
with different angles of attack

Fig. A-1 Normalized iso-vorticity and axial velocity contour of BW at x/c = 2.8

60
Fig.A-2 Zoomed in iso-vorticity contour of BW at x/c = 2.8

61
62
Fig. A-3 Normalized iso-vorticity and axial velocity contour of 50HDW at x/c = 2.8

63
Fig.A-4 Zoomed in iso-vorticity contour of 50HDW at x/c = 2.8

64
65
Fig. A-5 Normalized iso-vorticity and axial velocity contour of 0.5c 50HDW at x/c = 2.8

66
Fig.A-6 Zoomed in iso-vorticity contour of 0.5c 50HDW at x/c = 2.8

67

You might also like