You are on page 1of 73

NUMERICAL INVESTIGATION OF COMPRESSIBLE VORTICES USING THE

QUASI-CYLINDRICAL APPROXIMATION

DISSERTATION

Presented in Partial Fulfillment of the Requirements for


the Degree Doctor of Philosophy in the Graduate
School of the Ohio State University

By

David W. Bennett, M.S.A.E.

*****

The Ohio State University


2007

Dissertation Committee: Approved by

Professor Richard Bodonyi, Advisor

Professor Jen-Ping Chen

Professor Michael Foster

Professor Gerald Gregorek Advisor


Aeronautical and Astronautical Engineering
Graduate Program
ABSTRACT
To obtain a better understanding of compressible vortex dynamics, a numerical

investigation of laminar, compressible vortices was conducted using the quasi-cylindrical

approximation. The goal was to determine if the phenomena of vortex breakdown would

occur. Vortex breakdown is an important and unsolved problem in the field of fluid

dynamics.

The results show that two possible flow regimes can occur for a potential vortex

with a constant edge axial velocity. The first is a vortex that decays in a manner similar to

the incompressible vortices examined by Batchelor. The centerline axial velocity decay is

driven by the edge axial velocity. Far downstream, the axial velocity profile is nearly

constant with the tangential and radial velocities approaching zero.

The second flow regime occurs when the edge axial velocity is sufficiently small

and vortex breakdown occurs. For a given Mach number, the critical value of edge axial

velocity for breakdown was determined. For a set value of the edge axial velocity, the

stream wise distance to obtain vortex breakdown was a strong function of the Mach

number. As the Mach number increased, the distance to vortex breakdown was

decreased.

ii
ACKNOWLEDGMENTS
I wish to express my sincere thanks to Professor Richard Bodonyi for his consent

to be my dissertation advisor. His advice and willingness to take me on as a student have

made the completion of my doctorate possible. I will always be grateful to him for his

efforts on my behalf.

I would also like to express my thanks to Professors Michael Foster, Gerald

Gregorek, and Jen-Ping Chen for being members of my general examination and

dissertation committees. All have been very helpful in the pursuit of my doctorate, and I

will always be indebted to them.

My parents, Lawrence and Linda Bennett, have been with me since the beginning

of this process. The road was not always smooth, but they stood beside me with all the

support a son could ask. I thank them for instilling in me the desire to pursue my dreams

and supporting me through the endeavor.

My wife, Sandra Bennett, was not part of my life when my doctoral studies

began, but she has been supportive and patient with me during the final effort to complete

my studies. I was very fortunate to find Sandy, and I wish to express my love and thanks

to her.

I would also like to express my sincere thanks to Barbara Johnson of The Johns

Hopkins University Applied Physics Laboratory for her help in preparing this manuscript.

iii
VITA
October 18, 1966………………Born – Ripley, West Virginia, United States

1988……………………………B.S. Aerospace Engineering, West Virginia University

1989……………………………M.S. Aerospace Engineering, West Virginia University

FIELDS OF STUDY
Major Field: Aerospace Engineering

iv
TABLE OF CONTENTS
Abstract ............................................................................................................................... ii
Acknowledgments.............................................................................................................. iii
Vita..................................................................................................................................... iv
List of Tables ..................................................................................................................... vi
List of Figures ................................................................................................................... vii

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

1.1 Motivation...................................................................................................... 1
1.2 Objectives ...................................................................................................... 4
1.3 Previous Investigations .................................................................................. 4

2 Mathematical Formulation...................................................................................... 9

2.1 Navier-Stokes Equations................................................................................ 9


2.2 Quasi-Cylindrical Approximation ............................................................... 10
2.3 The Compressible, Quasi-Cylindrical Equations......................................... 12
2.4 Initial Profile ................................................................................................ 15

3 Solution Methodology .......................................................................................... 21

5 Results and Discussion ......................................................................................... 27

5 Conclusions........................................................................................................... 61

Bibliography ..................................................................................................................... 63

v
LIST OF TABLES
Table Page

4.1 Free Parameter Matrix for Vortex Examination ................................................... 30

vi
LIST OF FIGURES
Figure Page

1.1 Photograph of Vortex Breakdown from the HARV Program ................................ 2

2.1 Schematic of Vortex Coordinate System.............................................................. 12

4.1 Axial Marching Step Size Sensitivity Results ...................................................... 28

4.2 Radial Grid Size Sensitivity Results ..................................................................... 29

4.3 Initial Radial, Tangential, and Axial Velocity Profiles for Case A ...................... 32

4.4 Initial Density, Pressure, and Temperature profiles for Case A ........................... 33

4.5 Axial Velocity Profiles at Various Z Axial Locations for Case A ....................... 35

4.6 Tangential Velocity Profiles at Various Z Axial Locations for Case A ............... 36

4.7 Radial Velocity Profiles at Various Z Axial Locations for Case A...................... 37

4.8 Static Pressure Profiles at Various Z Axial Locations for Case A ....................... 38

4.9 Density Profiles at Various Z Axial Locations for Case A................................... 39

4.10 Temperature Profiles at Various Z Axial Locations for Case A........................... 40

4.11 Centerline Axial Velocity as a Function of Axial Location for Case A ............... 41

4.12 Centerline Axial Velocity for Large Values of Axial Location for Case 1 .......... 43

4.13 Axial and Tangential Velocity Profiles at Z=100 for Case A............................... 44

4.14 Density, Pressure, and Temperature Profiles at Z=100 for Case A ...................... 45

4.15 Centerline Axial Velocity Decay as a Function of ln(Z)/Z................................... 50

4.16 Radial, Tangential, and Axial Velocity Profiles for Case B ................................. 51

4.17 Centerline Axial Velocity as a Function of Axial Location for Case B ............... 52

vii
LIST OF FIGURES (Continued)
Figure Page

4.18 Axial Velocity Profiles Near Breakdown for Case B ........................................... 53

4.19 Tangential Velocity Profiles Near Breakdown for Case B................................... 54

4.20 Radial Velocity Profiles Near Breakdown for Case B.......................................... 55

4.21 Singularity in Centerline Axial Velocity for Case B ............................................ 56

4.22 Vortex Breakdown Location Using Richardson’s Extrapolation for Case B ....... 57

4.23 Centerline Axial Velocity for Various Mach Numbers ........................................ 58

4.24 Vortex Breakdown as a Function of Mach Number ............................................. 59

4.25 Critical Edge Axial Velocity Determination for M=0.2 ....................................... 60

viii
CHAPTER 1
INTRODUCTION
1.1 Motivation

The structure of fluid vortices is a major area of research in the field of fluid

dynamics. Vortices can form in both internal and external flow applications. Of particular

interest are vortices generated by flight vehicles where vortices are shed from wings,

control surfaces, and other protuberances. The behavior of these vortices directly affects

the performance and operation of flight vehicles.

The strength and the size of wing tip vortices of large commercial aircraft, such as

the Boeing 747, limit the minimum aircraft separation distance for airport landings and

takeoffs [1, 2]. Aircraft operation may be adversely affected if the aircraft encounters one

of these large vortices. The vortices may cause the aircraft to become uncontrollable and

possibly even crash. A better understanding of wing tip vortices could possibly reduce the

hazardous conditions and decrease the separation distance between aircraft, which could

increase aircraft throughput at airports. In particular, this would be beneficial to airports

with land restrictions where the addition of new runways is not possible.

For modern fighter aircraft, super-maneuverability is a desired performance

characteristic. The aircraft should be able to operate at high angles of attack without loss

of stability and control. As an example, the Boeing F/A-18 employs leading edge

extensions to generate large vortex structures that propagate downstream over the upper

1
surface of the wing [3]. These vortices increase the lift on the wings and allow for larger

pitch angles before the flow over the wing separates and the wing stalls. The benefit of

the leading edge vortices is reduced when the vortices dramatically change structure. This

change in structure is termed vortex breakdown or vortex bursting. Vortex breakdown

results in three phenomena that are detrimental to the performance of the aircraft. An

example of vortex breakdown is seen on Figure 1.1

Figure 1.1: Photograph of Vortex Breakdown from the HARV Program

2
NASA’s high angle of attack research vehicle (HARV) program was implemented

to study the performance characteristics of a modified F-18 fighter aircraft at large angles

of attack. As a part of the program, smoke was used to visualize the vortex emanating

from the strakes located forward of the cockpit. The location of vortex breakdown can

clearly be seen where the vortex transforms from a tight spiral structure to one that is

much larger in radius.

The location of vortex breakdown varies with flow conditions. If the breakdown

occurs over the wing, Hummel and Srinivasan [4] showed that the lift on the wing is

significantly reduced, and a nose-up pitching moment is generated. The nose-up pitching

moment is unstable in terms of static stability and may result in loss of maneuverability

and control of the vehicle.

Two leading edge vortices form on either side of the fuselage for some modern

fighter aircraft. During yawing maneuvers, vortex breakdown can occur at different

locations over the wing. Erickson [5] used water tunnel tests to study vortex breakdown.

An asymmetric vortex breakdown produces reduced lateral stability characteristics that

can result in loss of control at high angles of attack.

Modern aircraft with twin tail configurations, such as the F-14, F/A-18, F-22, and

F-35 are susceptible to fin buffeting [6, 7]. Fin buffeting is the phenomena where the

leading edge vortices breakdown forward of the fins. Thus, the fins are exposed to

separated flow and high turbulence levels resulting from the vortex breakdown. The

3
aerodynamic loading may cause the rudder to lose control effectiveness and result in a

loss of yaw control. In addition, the high-intensity turbulence may cause structural

damage due to fatigue of the rudder structure.

A better understanding of the dynamics of leading edge vortices is an active area

of investigation, and the examples cited previously show the engineering applications

where improved knowledge would prove invaluable. In particular, an understanding of

vortex breakdown and downstream propagation are critical issues.

1.2 Objectives

The focus of the current investigation is to investigate the structure of three-

dimensional vortices in a viscous, compressible fluid. Beginning with the Navier-Stokes

equations, the quasi-cylindrical approximation for slender vortices is employed to reduce

the Navier-Stokes equations to boundary layer equations, which then are solved

numerically. Key flow parameters for isolated vortices are identified, and the effect of

these parameters on the structure of the vortices will be presented.

1.3 Previous Investigations

Vortex formation and breakdown have been the focus of investigations for over

fifty years. The literature on the subject is extensive. Survey papers on vortex breakdown

[8, 9, 10, 11, 12, 13] review the multitude of investigations of the breakdown problem.

From these survey articles, three key points are made. The phenomenon of vortex

4
breakdown is still not completely understood, the majority of the investigations are

experimental, and nearly all investigations deal with incompressible flow.

Experimental studies of free vortices have dealt with the flow over delta wings in

wind and water tunnels. Confined vortices have been studied using cylindrical

enclosures. Escudier [12] summarizes the experimental investigations of confined and

free vortices.

As mentioned previously, the majority of vortex investigations have dealt with

incompressible flow, and there have been primarily four theoretical approaches to the

breakdown problem. The concept of a critical state, hydrodynamic instability theory, the

quasi-cylindrical approximation, and the solution of the axisymmetric Navier-Stokes

equations are all reviewed by Delery [13], and Escudier [12] summarizes the key

theoretical investigations.

Hall [8] proposed the quasi-cylindrical approximation. This approximation

assumes the variations of the flow components in the axial direction are small when

compared to the variations in the radial direction. This is similar to the classic boundary

layer approximation and reduces the complexity of the Navier-Stokes equations.

Beginning with the axisymmetric, compressible Navier-Stokes equations, he derived the

quasi-cylindrical equations for both incompressible and compressible flows. Hall also

presented numerical results for incompressible vortex breakdown. The governing

5
equations are parabolic and fail to converge when separation occurs. The location of the

separation defines the location for the onset of breakdown.

Burggraf and Foster [14] used the quasi-cylindrical approximation to investigate

tornado-like vortices in an incompressible medium. The authors presented numerical

results for vortex breakdown and determined a critical value of axial momentum flux for

which breakdown will occur. When breakdown does not occur, the vortex continues

indefinitely and takes the form of the Long vortex [15]. For the computations, the initial

distributions for the axial and tangential velocities were specified to be those of a

potential vortex. The radial velocity cannot be specified, and must be calculated. The

authors present an ordinary differential equation for the radial velocity that is derived

from the governing equations. This compatibility equation is critical for the success of

this method. A similar study was conducted by Crowley [16]. The main purpose of the

study was to reproduce the work of Burggraf and Foster and to investigate in more detail

the structure of the vortex breakdown. From the study, it was determined that the

singularity near breakdown is parabolic and that no flow reversal occurs.

There are few theoretical investigations of compressible vortices. Mack [17]

incorporated compressibility effects into a potential vortex. Brown [18] used the

incompressible vortex work of Hall [19] to address the effects of compressibility. These

efforts have restrictions on the Prandtl number and the ratio of specific heats. Self-similar

solutions for compressible vortices have been investigated by Orangi, Foster, and

6
Bodonyi [20] using the quasi-cylindrical approximation. In their study, the axial edge

velocity varies in powers of the axial distance.

Lui, Krause, and Menne [21] were the first to numerically solve the compressible,

quasi-cylindrical equations. The authors present compatibility conditions for vortices to

remain slender. Results are presented for compressible vortex breakdown for an inviscid,

nonconducting gas. The results show that the vortex breakdown location is delayed by

increasing the edge Mach number. Temperature effects were also examined. Heating the

core enhances breakdown while cooling the core delays breakdown. For the

computations, the axial velocity, tangential velocity, and temperature profile at the initial

station were specified as functions of radial distance. No information is given on the

determination of the initial profiles for the radial velocity or static pressure.

Kandil, Kandil, and Lui [22] presented results for free and confined compressible

vortices. Good agreement was found between solutions produced using a compressible

Navier-Stokes solver and a quasi-cylindrical compressible solver for subsonic flow

conditions. For the free vortex, no breakdown was reported, but there was breakdown of

a confined vortex in a circular cylinder. However, the breakdown occurred behind a

normal shock.

Kandil and Kandil [23] employed a quasi-cylindrical compressible solver to

calculate the effect of Mach number and the external axial pressure gradient on vortex

breakdown for a free vortex. The authors employed the compatibility conditions used by

7
Lui, Krause, and Menne [21]. The results are similar to those of Lui, Krause, and Menne,

with increasing Mach number delaying breakdown, and increasing external pressure

gradient decreasing the distance needed for breakdown. However, it is unclear how the

solutions are initialized. It appears that for the initial station, the viscosity and thermal

conductivity are zero.

The following chapters present results for the numerical investigation of isolated

compressible vortices. The Navier-Stokes equations will be presented, and the quasi-

cylindrical approximation will be discussed. Using the quasi-cylindrical approximation, a

simplified set of governing equations will be derived, and the appropriate boundary

conditions will be noted. The assumptions used for the generation of the initial dependent

variable profiles will be examined. The finite difference representation of the governing

equations will be presented, and solution methodology will be discussed and results for

representative cases will be presented. Following a discussion of the results, comments

on possible future work are made.

8
CHAPTER 2
MATHEMATICAL FORMULATION
2.1 Navier-Stokes Equations

The Navier-Stokes equations for an axisymmetric vortex in a perfect, viscous,

heat-conducting gas in cylindrical polar coordinates are

Conservation of Mass:

∂ρ u ρ u ∂ρ w
+ + =0 (2.1)
∂r r ∂z

Conservation of Axial Momentum:

∂w ∂w ∂p 1 ∂ ⎡ ⎛ ∂u ∂w ⎞⎤ 1 ∂ ⎡ ⎛ 2 ∂u 2 4 ∂w ⎞⎤
ρu + ρw =− + ⎢ µr ⎜ + ⎟⎥ + ⎢ µ ⎜ − r − u + r ⎟⎥ (2.2)
∂r ∂z ∂z r ∂r ⎣ ⎝ ∂z ∂r ⎠⎦ r ∂z ⎣ ⎝ 3 ∂r 3 3 ∂z ⎠⎦

Conservation of Radial Momentum:

∂u ∂u ρ v 2 ∂p 1 ∂ ⎡ ⎛ ∂u ∂u ⎞ ⎤
ρu + ρw − =− + µr ⎜ + ⎟
∂r ∂z r ∂r r ∂z ⎢⎣ ⎝ ∂z ∂r ⎠ ⎥⎦
(2.3)
1 ∂ ⎡ ⎛ 4 ∂u 2 2 ∂w ⎞ ⎤ ⎛ 2 ∂u 4 u 2 ∂w ⎞
+ µ ⎜ r − u − r ⎟⎥ + µ ⎜ r − 2 +

r ∂r ⎣ ⎝ 3 ∂r 3 3 ∂z ⎠ ⎦ ⎝ 3 ∂r 3 r 3r ∂z ⎠⎟

Conservation of Tangential Momentum:

∂v ∂v ρuv ∂ ⎛ v ⎞ 1 ∂ ⎡ 2 ∂ ⎛ v ⎞⎤ ∂ ⎡ ∂ ⎛ v ⎞⎤
ρu + ρw + = µ ⎜ ⎟+ µr ⎜ ⎟ + µr ⎜ ⎟ (2.4)
∂r ∂z r ∂r ⎝ r ⎠ r ∂r ⎢⎣ ∂r ⎝ r ⎠⎥⎦ ∂z ⎢⎣ ∂z ⎝ r ⎠⎥⎦

9
Conservation of Energy:

∂Θ ∂Θ ∂ ⎛ ∂Θ ⎞ 1 ∂ ⎛ ∂Θ ⎞ ∂p ∂p
ρC p w + ρC pu = ⎜k ⎟+ ⎜ kr ⎟+u + w
∂z ∂r ∂z ⎝ ∂z ⎠ r ∂r ⎝ ∂r ⎠ ∂r ∂z
⎧⎪⎛ ∂w ∂u ⎞2 4 ⎛ u ⎞2 4 ⎡ ∂u ∂w u ⎛ ∂w ∂u ⎞ ⎤
+ µ ⎨⎜ + ⎟ + ⎜ ⎟ + ⎢ − ⎜ + ⎟⎥ (2.5)
⎪⎩⎝ ∂r ∂z ⎠ 3 ⎝ r ⎠ 3 ⎣ ∂r ∂z r ⎝ ∂z ∂r ⎠ ⎦

∂ ⎛ v ⎞ ∂v ∂ ⎛v⎞ ∂ ⎛ v ⎞ ∂v ⎫
+r ⎜ ⎟ −v ⎜ ⎟+r ⎜ ⎟ ⎬
∂r ⎝ r ⎠ ∂r ∂r ⎝ r ⎠ ∂z ⎝ r ⎠ ∂z ⎭

The variables r and z represent the radial and axial coordinates. The radial,

tangential, and axial velocities are u, v, and w, respectively. The density, pressure, and

temperature are ρ, p, and Θ, respectively. The thermal conductivity and dynamic

viscosity are k and µ, respectively. The set of equations is closed by the use of the

equation of state for a perfect fluid

p = ρ RΘ (2.6)

where R is the gas constant.

2.2 Quasi-Cylindrical Approximation

In classical boundary layer theory, the assumption is made that the variations of

the flow quantities in the axial coordinate are small compared to the variations in the

normal coordinate. Similarly, the quasi-cylindrical approximation states that the varia-

tions in the axial direction are smaller than the variations in the radial direction (i.e., the

radial dimension is much smaller than the axial dimension).

10
Consider an order of magnitude analysis where

z ~ L and r ~δ

where L is a characteristic length scale in the axial direction, while δ is a characteristic

length scale in the radial direction. The following assumption is made

δ << L

The axial velocity and the density will be of the form

w ~ w* and ρ ~ ρ*

where the superscript * represents some characteristic value. An order of magnitude

analysis of the continuity equation yields the following information on the radial velocity:

⎛δ ⎞
u ~ ⎜ ⎟ w*
⎝ L⎠

Substituting these relationships into the axial momentum equation and assuming that

pressure gradient is the same order of magnitude as the inertial terms and that

µ~µ*

it can be shown that the inertial and highest order viscous terms are of the same order if

L
δ~
Re

The tangential velocity magnitude is estimated by examining the tangential momentum

equation and is found to be

v ~ w*

11
The remaining dependent variable is the static temperature, and it is assumed to have the

magnitude of a characteristic temperature. The order of magnitude analysis is applied to

the full Navier-Stokes equations, and a reduced set of equations is derived. This set of

equations is the compressible, quasi-cylindrical equations.

2.3 The Compressible, Quasi-Cylindrical Equations

A diagram of the coordinate system can be found in Figure 2.1.

R,U
Vortex Edge

W(R,Z)

Θ,V
Vortex Centerline

Z,W

Figure 2.1: Schematic of Vortex Coordinate System

12
The quasi-cylindrical governing equations are nondimensionalized using the

following variables:

r z
R= Z=
δ L

⎛L⎞ u v w
U =⎜ ⎟ * V= W=
⎝δ ⎠ w w* w*

p ρ Θ
P= ρ= T=
ρ w*2 ρ* Θ*

µ
µ=
µ*

The governing equations take the form

Conservation of Mass:

∂ρU ∂ρW
R + ρU + R =0 (2.7)
∂R ∂Z

Conservation of Axial Momentum:

∂W ∂W ∂P ∂W ∂µ ∂W ∂ 2W
Rρ U + RρW = −R +µ +R + Rµ (2.8)
∂R ∂Z ∂Z ∂R ∂R ∂R ∂R 2

Conservation of Tangential Momentum:

∂V ∂V ∂ 2V ∂V
R 2 ρU + R 2 ρW + R ρUV = R 2 µ + Rµ − µV
∂R ∂Z 2 ∂R
∂R (2.9)
∂µ ∂V ∂µ
+ R2 −R V
∂R ∂R ∂R

13
Conservation of Radial Momentum:

ρV 2 ∂P
= (2.10)
R ∂R

Conservation of Energy:

∂T ∂T ∂P
R 2 ρU + R 2 ρW = R 2 (γ − 1) M 2U
∂R ∂Z ∂R
∂P R ⎛ ∂T ∂µ ∂T ∂ 2T ⎞ (2.11)
+ R 2 (γ − 1) M 2W + ⎜µ +R + Rµ ⎟
∂Z Pr ⎜⎝ ∂R ∂R ∂R ∂R 2 ⎟⎠

⎛ ∂W ⎞
2 ⎡ 2 ⎤
2 ⎛ ∂V ⎞ ∂V
+ R 2 (γ − 1) M 2 µ ⎜ ⎟ + (γ − 1) M 2
⎢ R µ ⎜ ⎟ − 2 R µV + µV 2 ⎥
⎝ ∂R ⎠ ⎢⎣ ⎝ ∂R ⎠ ∂R ⎥⎦

Equation of State:

1
P= ρT (2.12)
γM 2

where M is the Mach number, Pr is the Prandtl number, and γ is the ratio of specific

heats. The equations are subject to the following boundary conditions. At the vortex

centerline (R=0), the following boundary conditions apply

U =0
V =0
∂W (2.13)
=0
∂R
∂T
=0
∂R

14
while the outer boundary (R→∞), the following boundary conditions are imposed

W = We ( Z )
V = Ve ( Z )
(2.14)
P = Pe ( Z )
T = Te ( Z )

2.4 Initial Profile

To begin the solution process, profiles for each dependent variable must be

specified as a function of the radial coordinate for some starting Z location. Because of

the complexity of the governing equations, a method of obtaining the initial profiles for

the compressible equations is not readily apparent. The method used in this investigation

is to use the incompressible formulation of the quasi-cylindrical problem to determine the

initial profiles as described in Burggraf and Foster [14]. These equations are given by

Conservation of Mass:

∂U ∂W
R +U + R =0 (2.15)
∂R ∂Z

Conservation of Axial Momentum:

∂W ∂W ∂P ∂W ∂ 2W
RU + RW = −R + +R (2.16)
∂R ∂Z ∂Z ∂R ∂R 2

Conservation of Tangential Momentum:

∂V ∂V ∂ 2V ∂V
R 2U + R 2W + RUV = R 2 +R −V (2.17)
∂R ∂Z ∂R 2
∂R

15
Conservation of Radial Momentum:

V 2 ∂P
= (2.18)
R ∂R

The governing equations can be manipulated to derive an ordinary differential equation

for the radial velocity, U, that is dependent on the tangential and axial velocities that are,

in turn, specified. The ordinary differential equation for the radial velocity is

Wi2 '

W
⎢ i i

W ''
+
R
1 2 1
W
2 i

R
W W
i i
' 2Vi

R 2
( RV ) '⎤
i ⎥ i

U −
R
( )
U i + RU i'' =
(2.19)
1 1 2 2 2 2
Wi'''Wi + WiWi'' − WiWi' − ViVi'' − ViVi' + Vi
2 2
R R R R R3

where ' denotes differentiation with respect to the radial dimension. The ordinary

differential equation for the radial velocity requires two boundary conditions. At the

centerline of the vortex (R=0),

U =0 (2.20)

At the outer boundary (R→∞), the boundary condition is dependent on the form of the

axial and tangential velocities for large R values. For the present investigation, the

tangential and axial velocities of a potential vortex are specified. The tangential velocity

has the form

1⎛ − R2 ⎞
Vi = ⎜ 1 − e ⎟ (2.21)
R⎝ ⎠

16
and the axial velocity has the following form

Wi =
1
2R
(
1 − e− R
2

) 2 ⎛W δ
+⎜ o −
⎝ K
1 ⎞ − R2 Weδ
⎟e + K
2⎠
(2.22)

where wc and we are the axial core velocity and axial edge velocity components,

respectively, and wo is wc minus we. K is the circulation parameter, which is equal to

Γ/2π, where Г is the circulation and δ is the vortex radius. At the outer boundary and with

we equal to a constant,

1 1
WR →∞ ≈ We + and V R →∞ ≈
2R R

The leading order term for the radial velocity boundary condition is

U R →∞ ≈ C

where C is a constant that is determined in the integration process. The initial profile is

completed by using the radial momentum equation to calculate the pressure distribution.

The density profile at the initial station is specified as a constant, and the temperature

distribution is obtained from the equation of state. Recall that the initial profiles are

generated using the analytical expressions for tangential and axial velocities of a potential

vortex (Equations 2.21 and 2.22). These equations are nondimensional, and the free

parameters are the nondimensional axial core and axial edge velocities. From the

governing equations for the compressible case (Equations 2.7 through 2.12), the free

parameters are the Mach number, Prandtl number, and ratio of specific heats. The

nondimensionalized variables used by Bruggraf and Foster [14] are different than those
17
used in the current compressible investigation. The relationships between the

incompressible and compressible velocity components are

⎛ K ⎞⎛ ν ⎞
U =⎜ ⎟ ⎜ ⎟U (2.23)
⎝ Wcδ ⎠⎝ K ⎠

⎛ K ⎞
V = ⎜⎜ ⎟⎟V (2.24)
⎝ Wcδ ⎠

⎛ K ⎞
W = ⎜⎜ ⎟⎟W (2.25)
⎝ Wcδ ⎠

where double-barred terms are the nondimensional compressible variables and single-

barred terms are the nondimensional incompressible terms. For the incompressible

formulation, an effective pressure is defined as

p
P= (2.26)
ρ

which is nondimensionalized in the incompressible formulation as follows:

2 2
⎛δ ⎞ ⎛δ ⎞ p
P=⎜ ⎟ P=⎜ ⎟
⎝K⎠ ⎝K⎠ ρ (2.27)

Rearranging

2
p ⎛K⎞
=⎜ ⎟ P (2.28)
ρ ⎝δ ⎠

18
The pressure and density are then nondimensionalized for the compressible formulation

⎛ p ⎞
ρ c wc2 ⎜⎜ ⎟
2 ⎟
ρ
⎝ c c⎠
w ⎛ p ⎞ ⎛ K ⎞2
=w ⎜ ⎟=⎜ ⎟ P
2
(2.29)
⎛ ρ ⎞
c
⎜ρ⎟ ⎝δ ⎠
ρ c ⎜⎜ ⎟⎟ ⎝ ⎠
⎝ ρc ⎠

Rearranging

2
⎛ K ⎞
p = ρ ⎜⎜ ⎟⎟ P (2.30)
⎝ wc δ ⎠

The density is assumed constant at the initial station. The density, pressure, and

temperature in the appropriate compressible variables are

ρ =1 (2.31)

2
⎛ K ⎞
p = ⎜⎜ ⎟⎟ P (2.32)
⎝ wcδ ⎠
2
⎛ K ⎞
T = γM p = γM ⎜⎜
2 2
⎟⎟ P (2.33)
⎝ wc δ ⎠

Equation 2.33 is simply the nondimensionalized equation of state.

The differential equation for the radial velocity is solved numerically using the

Thomas algorithm [24]. The equation contains the axial velocity as well as the first-,

second-, and third-order derivatives of the axial velocity. The equation also contains the

tangential velocity along with its first and second derivative. The velocity and its

19
derivatives behave smoothly sufficiently far from the centerline. However, near the

centerline, Taylor series expansions of the velocities and velocity derivatives are required

to avoid erroneous results.

20
CHAPTER 3
SOLUTION METHODOLOGY
Given the initial dependent variable profiles, the tangential and axial momentum

equations are integrated in a coupled manner using the Crank-Nicolson finite difference

method [23]. The method is implicit, second-order accurate in space, and well suited for a

parabolic marching problem. The equations of motion are marched in the axial direction

and discretized at a point (k+1/2,j) where k represents the axial (marching) coordinate

and j represents the radial direction. Derivatives are an average of the previous known

axial location and the next axial location to be calculated. For the axial derivatives, a

first-order, finite difference formula is employed. For the radial derivatives, a second-

order, central difference finite difference representation is employed. The truncation error

for the Crank-Nicolson method is O[(∆z)2,(∆r)2], where O denotes “of the order.” The

truncation error is defined as the difference between the partial differential equations to

be solved and the finite difference representation of the equations. A von Neumann

stability analysis of the Crank-Nicolson method shows that the method is unconditionally

stable. However, limitations on the marching step size do occur, and a grid sensitivity

study will determine the marching step size. Because of the nonlinearity of the governing

equations, Newtonian linearization is implemented. Newtonian linearization is a

technique where the dependent variables are replaced by an estimate of the variable at the

next axial location plus an error associated with the difference between the estimate and

21
the computed value. As an example, the δWi+1,j term represents the difference between

the predicted value of Wi +1, j and the computed value of Wi+1,j. The can be expressed as

Wi +1, j = W i +1, j + δ Wi +1, j (3.1)

where

δ Wi +1, j << 1
(3-2)

Second-order error terms are ignored. The solution process is iterated until the error terms

meet a set tolerance of 1×10-6. The coupling and linearization result in a 2×2 block

tri-diagonal system of equations for the tangential and axial velocity errors. The radial

velocity, temperature, pressure, and density are lagged one iteration. Once new estimates

for tangential velocity and axial velocity are computed, the continuity equation is

integrated using the trapezoidal rule to determine the updated radial velocity. For the first

axial marching step, the axial derivative is evaluated using a first-order finite difference

formula. After the first step, the derivative is approximated using a three-point, second-

order accurate backward differencing formula. The density is lagged in this calculation.

The pressure is calculated by integrating the radial momentum equation also using the

trapezoidal rule. The density is lagged in this calculation as well. The conservation of

energy equation is used to calculate the temperature using the updated velocities and

pressure. The temperature is found by solving a tri-diagonal system of equations using

the Thomas algorithm. The density is lagged in this calculation also. Finally, the density

22
is calculated from the equation of state with the updated variables, and the dynamic

viscosity is determined from a power law equation. The solution process is repeated until

the error for all calculated quantities meets a set tolerance of 1×10-6.

The following are the finite difference equations for the axial momentum

equation, the tangential momentum equation, and the energy equation:

Conservation of Axial Momentum

⎡ Rj 1 ⎤
µi +1, j + j 2 ( µi +1, j +1 − µi +1, j −1 ) − j 2 µi +1, j ⎥ δ Wi +1, j −1
R R
⎢− ρi +1, jU i +1, j +
⎣ 4 ∆R 4 ∆R 8∆R 2 ∆R ⎦
⎡ Rj ⎤
⎡ 2 ρi +1, jWi +1, j − ( ρi +1, j − ρi , j )Wi , j ⎤ + j 2 µi +1, j ⎥ δ Wi +1, j
R
+⎢ ⎣ ⎦
⎣ 2 ∆Z ∆R ⎦
⎡ Rj ε ⎤
µi +1, j − j 2 ( µi +1, j +1 − µi +1, j −1 ) − j 2 µi +1, j ⎥ δ Wi +1, j +1 =
R R
+⎢ ρi +1, jU i +1, j −
⎣ 4 ∆R 4 ∆R 8∆R 2 ∆R ⎦

ρi +1, jU i +1, j (Wi +1, j +1 − Wi +1, j −1 )


Rj

4 ∆R

ρi , jU i , j (Wi , j +1 − Wi , j −1 )
Rj
− (3.3)
4 ∆R
Rj
− ⎡ ρi +1, jWi +21, j − ρi , jWi ,2j − ( ρi +1, j − ρi , j )Wi , jWi +1, j ⎤⎦
2 ∆Z ⎣

( Pi+1, j − Pi, j )
Rj

∆Z
1 ⎡
+ µi +1, j (Wi +1, j +1 − Wi +1, j −1 ) + µi , j (Wi , j +1 − Wi , j ) ⎤⎦
4 ∆R ⎣
⎡( µi +1, j +1 − µi +1, j −1 )(Wi +1, j +1 − Wi +1, j −1 ) + ( µi , j +1 − µi , j −1 )(Wi , j +1 − Wi , j −1 ) ⎤
Rj
+
8 ∆R 2 ⎣ ⎦

⎡ µi +1, j (Wi +1, j +1 − 2Wi +1, j + Wi +1, j −1 ) + µi , j (Wi , j +1 − 2Wi , j + Wi , j −1 ) ⎤


Rj
+
2 ∆R 2 ⎣ ⎦

23
Conservation of Tangential Momentum

⎡ R2 R 2j R 2j ⎤
( )⎥⎥ δ Vi +1, j −1
Rj
⎢ − j ρi +1, jU i +1, j − µ + µ + µi +1, j +1 − µi +1, j −1
⎢ 4∆R 2 i +1, j 4∆R i +1, j 2
⎣ 2 ∆R 8∆ R ⎦
⎡ R2
( )
j Rj
+⎢ ρi +1, jWi +1. j + ρi, jWi, j + ρi +1, jU i +1, j
⎢ 2 ∆Z 2

⎛ R2 1 ⎞ R ⎤
+ µi +1, j ⎜
⎜ ∆R
j
2
+ ⎟+
j
2 ⎟ 4 ∆R
(
µi +1, j +1 − µi +1, j −1 )⎥⎥ δ Vi +1, j
⎝ ⎠ ⎦
⎡ R2 R 2j R 2j ⎤
( )⎥⎥ δ Vi +1, j +1
j Rj
+ ⎢ ρi +1, jU i +1, j − µi +1, j − µi +1, j − µi +1, j +1 − µi +1, j −1
⎢ 4 ∆R
⎣ 2∆R 2 4∆R 8∆R 2 ⎦
⎡ R2 ⎤
+⎢
j
⎢ 2 ∆Z
( )
ρi +1, j Vi +1, j − Vi, j ⎥ δ Wi +1, j =

⎣ ⎦
(3.4)
R 2j R 2j

4∆R
(
ρi +1, jU i +1, j Vi +1, j +1 − Vi +1, j −1 − ) 4∆R
(
ρi, jU i, j Vi, j +1 − Vi, j −1 )
R 2j

2∆Z
(Vi +1, j − Vi, j )( ρi +1, jWi +1. j + ρi, jWi, j )
Rj Rj
− ρi +1, jU i +1, jVi +1, j − ρi, jU i, jVi, j
2 2
R 2j
+ ( ) (
⎡ µi +1, j Vi +1, j +1 − 2Vi +1, j + Vi +1, j −1 + µi , j Vi, j +1 − 2Vi, j + Vi, j −1 ⎤
2∆R 2 ⎣ ⎦ )
⎡ µi +1, j Vi +1, j +1 − Vi +1, j −1 + µi, j Vi, j +1 − Vi, j −1 ⎤ − ε µi +1, jVi +1, j + µi, jVi , j
( ) ( ) ( )
Rj
+
4∆R ⎣ ⎦ 2
R 2j
+ ( )( ) (
⎡ µi +1, j +1 − µi +1, j −1 Vi +1, j +1 − Vi +1, j −1 + µi , j +1 − µi, j −1 Vi, j +1 − Vi, j −1 ⎤
8∆R 2 ⎣ ⎦ )( )
( ) ( )
Rj
− ⎡ µi +1, j +1 − µi +1, j −1 Vi +1, j + µi, j +1 − µi, j −1 Vi, j ⎤
4∆R ⎣ ⎦

24
Conservation of Energy

⎡ R 2j ρi +1, jU i +1, j R j µi +1, j R 2j R 2j µi +1, j ⎤


⎢ −
4 ∆R
+ +
4 Pr ∆R 8 Pr ∆R 2
( i +1, j +1 i+1, j −1 ) 2 Pr ∆R 2 ⎥ Ti+1, j −1
µ − µ −
⎢⎣ ⎥⎦
⎡ R 2j R 2j µi +1, j ⎤
+⎢ ( ρi+1, jWi +1, j + ρi, jWi , j ) + Pr ∆R 2 ⎥ Ti+1, j
⎣⎢ 2∆Z ⎦⎥
⎡ R 2j ρi +1, jU i +1, j R j µi +1, j R 2j R 2j µi +1, j ⎤
+⎢
4 ∆R
− −
4 Pr ∆R 8 Pr ∆R 2
( µi +1, j +1 − µi +1, j −1 ) − 2 Pr ∆R 2 ⎥ Ti+1, j +1 =
⎢⎣ ⎥⎦
R 2j ρi , jU i , j R 2j

4 ∆R
( i, j +1 i, j −1 ) 2∆Z ( ρi+1, jWi+1, j + ρi, jWi, j ) Ti, j
T − T +

R 2j (γ − 1) M 2
+ ⎡U i +1, j ( Pi +1, j +1 − Pi +1, j −1 ) + U i , j ( Pi , j +1 − Pi , j −1 ) ⎤
4 ∆R ⎣ ⎦
R 2j (γ − 1) M 2 R j µi , j (3.5)
+
2 ∆Z
(W i +1, j + Wi , j )( Pi +1, j − Pi , j ) +
4 Pr ∆R
(T i , j +1 − Ti , j −1 )

R 2j R 2j µi , j
+
8 Pr ∆R 2
( µi, j +1 − µi, j −1 )(Ti, j +1 − Ti, j −1 ) + 2 Pr ∆R 2 (Ti, j +1 − 2Ti, j + Ti, j −1 )
(γ − 1) M 2 R 2j ⎡
(Wi +1, j +1 − Wi +1, j −1 ) + µi , j (Wi , j +1 − Wi , j −1 ) ⎤
2 2
+
⎢⎣
µ +1, ⎥⎦
8 ∆R 2 i j

(γ − 1) M 2 R 2j ⎡
(Vi +1, j +1 − Vi +1, j −1 ) + µi , j (Vi , j +1 − Vi , j −1 ) ⎤
2 2
+

µ + ⎥
8 ∆R 2
⎣ i 1, j

(γ − 1) M 2 R j ⎡
⎣ µi +1, jVi +1, j (Vi +1, j +1 − Vi +1, j −1 ) + µi , jVi , j (Vi , j +1 − Vi , j −1 ) ⎦
− ⎤
2 ∆R
( γ − 1) M 2
+
2
( µi +1, jVi+21, j + µi, jVi,2j )

25
Equations 3.3, 3.4, and 3.5 are the finite difference equations for the interior

points of the computational grid. The subscript i denotes the marching location for the

finite difference terms, and the j subscript represents the radial location of the finite

difference terms. The equations are written such that the calculated step is represented by

the i+1 subscript.

26
CHAPTER 4
RESULTS AND DISCUSSION
A computer program was constructed to solve the compressible, quasi-cylindrical

equations. Before a parametric study of the breakdown problem was undertaken, a grid

sensitivity study was conducted. For the grid sensitivity studies, the Mach number,

centerline initial axial velocity, and edge axial velocity were taken to be 0.8, 1.2, and

0.045, respectively. Figure 4.1 shows the results for the sensitivity of the axial marching

step on the centerline axial velocity. Results for an axial marching step size of 0.0025 and

0.005 are presented. For both cases, the radial step size was 0.05. When the marching

step size was decreased by 50 percent, there was no change in the centerline axial

velocity. Therefore, it was determined that a marching step size of 0.005 was sufficient

for the parametric study. Similarly, the results for the radial grid sensitivity study are

found in Figure 4.2. The radial velocity profiles as a function of radial distance for one

axial location are shown for radial grid sizes of 0.025 and 0.05. From these results, a

radial grid size of 0.05 was determined to be sufficient.

27
1.4

1.2

0.8
Wc

0.6

0.4

0.2

0
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20
Z

dz=0.0025 dz=0.005

Figure 4.1: Axial Marching Step Size Sensitivity Results

28
1.20

1.00

0.80

0.60
U

0.40

0.20

0.00
0 1 2 3 4 5 6 7 8 9 10

-0.20

dr=0.05 dr=0.025

Figure 4.2: Radial Grid Size Sensitivity Results

29
A parametric study of compressible vortex dynamics has been conducted. The

free parameters chosen and range of values are found in Table 4.1.

Parameter
0.1 ≤ Wcδ/K ≤ 10.0
0.01 ≤ Weδ/K≤ 10.0
0.01 ≤ M ≤ 0.8

Table 4.1. Free Parameter Matrix for Vortex Examination

The parametric study, although not exhaustive, allowed for a determination of the

key parameters affecting the vortex dynamics. The parameters examined are the

nondimensional centerline axial velocity, nondimensional edge axial velocity, and the

Mach number. The upper bound for the Mach number was chosen as 0.8 due to the fact

that most subsonic, commercial aircraft cruise at this Mach number. Runs with a Mach

number of 0.01 are incompressible limit cases for the compressible solver and are

essentially modified forms of the Bruggraf and Foster problem [14].

Before presenting the results, a few comments about the results of the parametric

study are made. The parametric study showed that the key parameter affecting the vortex

dynamics is the edge axial velocity. For the majority of parameter combinations, no

vortex breakdown was observed. As stated earlier, vortex breakdown occurs when the

centerline axial velocity diminishes to zero. For each case, the centerline velocity never

diminished to zero, but was driven asymptotically to some nonzero value by the edge

axial velocity. The vortex propagates at nearly a constant axial velocity with a small

tangential velocity component. The vortex flow becomes essentially a plug flow with
30
rotation. For the remainder of the discussion, the flow will be called plug-like flow to

differentiate from a true plug flow which does not have rotation. The difference in

magnitude of the centerline axial velocity and the edge axial velocity simply determined

the distance needed to obtain plug-like flow. As an example, the distance to obtain plug-

like flow for centerline axial velocity of 10.0, an edge axial velocity of 1.0, and a Mach

number of 0.8 is much larger than when the centerline axial velocity of 0.5, the edge axial

velocity of 1.0, and the Mach number of 0.8. For the cases where the edge axial velocity

is larger than the centerline axial velocity, the centerline axial velocity increased and was

driven by the edge axial velocity. This case also resulted in plug-like flow, and the

solutions were marched in the axial direction to large values of axial distance to insure

the plug-like flow findings. For brevity, the results for a representative case will be

presented.

For Case A, the initial centerline and edge axial velocities were both 0.1. The

Mach number was 0.8. The initial velocity component profiles can be found on

Figure 4.3. The initial pressure, density, and temperature profiles can be found on

Figure 4.4. The maximum nondimensionalized radial distance is 20.0. The computational

grid consists of 401 grid points in the radial directions with a ∆R of 0.05. The axial

marching step size, ∆Z, was 0.005.

31
1.6

1.4

1.2

0.8

0.6

0.4

0.2

0
0 2 4 6 8 10 12 14 16 18 20
R

U V W

Figure 4.3: Initial Radial, Tangential, and Axial Velocity Profiles for Case A

32
2.5

1.5

0.5

0
0 2 4 6 8 10 12 14 16 18 20
R

Rho P T

Figure 4.4: Initial Density, Pressure, and Temperature profiles for Case A

33
Figure 4.5 contains axial velocity profiles in the radial coordinate at four axial

coordinate locations. When compared to Figure 4.3, the peak axial velocity has moved

from R=2 to R=3, and the profiles are flattening. The tangential velocity profiles are

found on Figure 4.6. Again, the peak tangential velocity is near R=3 and is decaying. The

radial velocity profiles are found in Figure 4.7. The maximum radial velocity occurs at

R=3 and the magnitude of the peak velocity is decreasing. Also, the form of the radial

velocity profile is changing. From the initial profile, the radial velocity has a positive

inflection point near R=2. From Figure 4.7, the radial velocity develops two inflection

points, but then returns to only having one positive inflection point at distances

downstream. The pressure profiles at various axial locations are found in Figure 4.8.

Variations are seen for radial location of seven and below, and these variations are small.

For radial location greater than seven, the pressure profiles are nearly identical. The

density profiles are found in Figure 4.9. The density variations are small with the

variation occurring from a radial distance between 0 and 16. The temperature profiles are

found in Figure 4.10. The temperature profiles show that the centerline temperature is

increasing as the axial location increases and is approaching the edge temperature which

is constant for the calculation.

The centerline axial velocity as a function of axial location is found in

Figure 4.11.

34
1.25

1.2

1.15
W

1.1

1.05

1
0 2 4 6 8 10 12 14 16 18 20
R

Z=0.5 Z=1.0 Z=1.5 Z=2.5

Figure 4.5: Axial Velocity Profiles at Various Z Axial Locations for Case A

35
0.4

0.35

0.3

0.25

0.2
V

0.15

0.1

0.05

0
0 2 4 6 8 10 12 14 16 18 20
R

Z=0.5 Z=1.0 Z=1.5 Z=2.5

Figure 4.6: Tangential Velocity Profiles at Various Z Axial Locations for Case A

36
0.16

0.14

0.12

0.1

0.08

0.06
U

0.04

0.02

0
0 2 4 6 8 10 12 14 16 18 20
-0.02

-0.04

-0.06
R

Z=0.5 Z=1.0 Z=1.5 Z=2.5

Figure 4.7: Radial Velocity Profiles at Various Z Axial Locations for Case A

37
2.05

1.95

1.9
P

1.85

1.8

1.75
0 2 4 6 8 10 12 14 16 18 20
R

Z=0.5 Z=1.0 Z=1.5 Z=2.5

Figure 4.8: Static Pressure Profiles at Various Z Axial Locations for Case A

38
1.012

1.01

1.008

1.006

1.004
Rho

1.002

0.998

0.996

0.994

0.992
0 2 4 6 8 10 12 14 16 18 20
R

Z=0.5 Z=1.0 Z=1.5 Z=2.5

Figure 4.9: Density Profiles at Various Z Axial Locations for Case A

39
1.82
1.8
1.78
1.76
1.74
1.72
T

1.7
1.68
1.66
1.64
1.62
1.6
0 2 4 6 8 10 12 14 16 18 20
R

Z=0.5 Z=1.0 Z=1.5 Z=2.5

Figure 4.10: Temperature Profiles at Various Z Axial Locations for Case A

40
1.16

1.14

1.12

1.1

1.08

1.06
W

1.04

1.02

0.98

0.96
0 0.5 1 1.5 2 2.5
Z

Figure 4.11: Centerline Axial Velocity as a Function of Axial Location for Case A

41
The centerline axial velocity initially decreases, and then increases to a constant

value of 1.14. From Figure 4.11, it appears that the centerline axial velocity is

asymptotically approaching a constant value. To determine the behavior of the vortex far

downstream, the solutions were marched to an axial location of 100. Figure 4.12 shows

the centerline axial velocity as a function of axial location far downstream. The centerline

axial velocity increases to a value of approximately 1.15 and then decreases for the

remainder of the calculation to a value of approximately 1.05 at Z=100. The radial

profiles for the axial and tangential velocities at Z=100 are found in Figure 4.13. It can be

seen that the axial velocity across the vortex is nearly constant. The tangential velocity

varies from zero at the centerline to a small value at the vortex edge. Figure 4.14 gives

the density, pressure, and temperature profiles at Z=100. Similar to the axial velocity, the

values are constant. From these results, the vortex flow has essentially transformed into a

plug-like flow with constant thermodynamic properties. A plug-like flow is one in which

every part of the flow propagates at the same velocity with some rotation.

42
1.16

1.14

1.12

1.1

1.08
Wc

1.06

1.04

1.02

0.98

0.96
0 10 20 30 40 50 60 70 80 90 100
Z

Figure 4.12: Centerline Axial Velocity for Large Values of Axial Location for Case 1

43
1.2

1.0

0.8

0.6

0.4

0.2

0.0
0 2 4 6 8 10 12 14 16 18 20
R

V W

Figure 4.13: Axial and Tangential Velocity Profiles at Z=100 for Case A

44
2.5

1.5

0.5

0
0 2 4 6 8 10 12 14 16 18 20
R

Rho P T

Figure 4.14: Density, Pressure, and Temperature Profiles at Z=100 for Case A

45
An estimate for the axial distance where plug-like flow has been established is

made using experimental data from Devenport, Rife, Liapis, and Follin [25]. A series of

wind tunnel tests were conducted to measure the structure of a wing-tip vortex generated

from a NACA 0012 half-wing. A four-sensor, hot-wire probe was used to measure

velocities of the vortex generated by the half-wing at various angles of attack. The wind

tunnel flow conditions were incompressible with a Reynolds number based on the half-

wing chord of 530,000, and the chord of the half-wing was 0.203 meter. For each flow

condition, the vortex core radius and core circulation Reynolds number were measured.

For an angle of attack of 5°, values for the nondimensional core radius, (r/c), and the core

circulation Reynolds number, (Γ/ν), were 0.036 and 33,900, respectively. The

incompressible formulation of the nondimensionalized axial coordinate is

z
z=
⎛ Kδ ⎞
⎜ ⎟
⎝ ν ⎠

where the circulation parameter, K, is defined as

Γ
K=

where Γ is the circulation. Rearranging

z ⎛Γ⎞
z= ⎜ ⎟δ
2π ⎝ν ⎠

Using a z of 100 and the vortex parameters measured in the wind tunnel, an axial

distance of approximately 13,000 feet (2.5 miles) is calculated. Although the number is
46
for a small-scale wing, it seems reasonable in the sense that Federal Aviation Administra-

tion rule for aircraft separation during landing is 3 miles.

Batchelor [26] determined that the asymptotic decay of the centerline axial

velocity in an incompressible vortex has the form

ln( Z )
Wc ≈ (4.1)
Z

where ln(Z) is the natural logarithmic function. On Figure 4.15, the centerline axial

velocity is plotted as a function of ln(Z)/Z. The result is a straight line. A linear curve fit

is shown along with the resulting linear equation. The results show that a compressible

vortex decays in the same manner as an incompressible vortex.

For Case B, the initial centerline and edge axial velocities were 1.2 and 0.01,

respectively, and the Mach number was 0.01. The initial velocity component profiles can

be found in Figure 4.16. The maximum nondimensionalized radial distance is 20. The

computational grid consists of 401 grid points in the radial directions with a ∆R of 0.05.

The axial marching step size, ∆Z, was 0.005. Case B is essentially the problem

investigated by Burggraf and Foster [14] solved by using a compressible formulation.

From Figure 4.17, vortex breakdown occurs at an axial location of 150. The axial

velocity profiles near vortex breakdown are found in Figure 4.18. The variations in axial

velocity are small with some differences along the centerline where the velocity is

decaying. The tangential velocity profiles are found in Figure 4.19. Similarly to the axial

velocity profiles, the variation in tangential velocity is small at the three axial locations.
47
The radial velocity profiles are found in Figure 4.20. The radial velocity drastically

increases over a small change in axial location with the maximum values near a radial

distance of 2. To examine the nature of the vortex breakdown, the axial step size was

reduced to 0.001 near the breakdown location. The centerline axial velocity for axial

locations near breakdown is found in Figure 4.21. Near breakdown, there is a rapid

decrease in the centerline axial velocity, but no flow reversal occurs. In order to

determine the axial location where the centerline axial velocity equals zero, Richardson’s

extrapolation was employed to calculate three points near breakdown using three

different axial step sizes of 0.0005, 0.001, and 0.002, the centerline velocity was plotted

as a function of Z-2, and the results are found in Figure 4.22. It is assumed that the

singularity is parabolic, and Figure 4.22 confirms this assumption. Using the linear curve

fit of the three points, it is determined that the centerline axial velocity equals zero at an

axial distance of 146.6414.

To determine the effect of the Mach number on vortex breakdown, a series of

computations with an edge axial velocity of 0.01 at various Mach numbers were

completed. The results of these computations are found in Figure 4.23. The Mach number

was varied from 0.1 to 0.8. As the Mach number increases, the axial distance where

breakdown occurs decreases. The breakdown location as a function of Mach number is

found in Figure 4.24. The figure clearly shows the Mach number dependence on the

48
breakdown location. This result is in contrast to the results presented by Lui, Krause, and

Menne [21].

For a given combination of centerline and edge axial velocities, it has been shown

that breakdown can occur. For a Mach number of 0.2, the edge axial velocity was varied

to determine if a critical edge velocity for breakdown exists. Figure 4.25 shows the

centerline axial velocity as a function of axial location for various edge axial velocities.

For edge velocities less than or equal to 0.056, breakdown occurs. For values greater than

0.056, breakdown does not occur, and the vortex behaves as in Case A.

49
1.053

1.052

1.051

1.05
Wc = 0.5264ln(z)/Z + 1.0202
1.049
R2 = 1
Wc

1.048

1.047

1.046

1.045

1.044
0.04 0.045 0.05 0.055 0.06 0.065 0.07

ln(Z)/Z
Wc Linear (Wc)

Figure 4.15: Centerline Axial Velocity Decay as a Function of ln(Z)/Z

50
2.5

2.0

1.5

1.0

0.5

0.0
0 2 4 6 8 10 12 14 16 18 20

-0.5
R

U V W

Figure 4.16: Radial, Tangential, and Axial Velocity Profiles for Case B

51
1.4

1.2

0.8
Wc

0.6

0.4

0.2

0
0 20 40 60 80 100 120 140 160
Z

Figure 4.17: Centerline Axial Velocity as a Function of Axial Location for Case B

52
0.3

0.25

0.2

0.15
W

0.1

0.05

0
0 2 4 6 8 10 12 14 16 18 20
R

Z=148 Z=148.6 Z=148.64

Figure 4.18: Axial Velocity Profiles Near Breakdown for Case B

53
0.25

0.20

0.15
V

0.10

0.05

0.00
0 2 4 6 8 10 12 14 16 18 20
R

Z=148 Z=148.6 Z=148.64

Figure 4.19: Tangential Velocity Profiles Near Breakdown for Case B

54
0.35

0.30

0.25

0.20
U

0.15

0.10

0.05

0.00
0 2 4 6 8 10 12 14 16 18 20
R

Z=148 Z=148.6 Z=148.64

Figure 4.20: Radial Velocity Profiles Near Breakdown for Case B

55
0.154

0.152

0.150

0.148
Wc

0.146

0.144

0.142

0.140

0.138
148.600 148.601 148.602 148.603 148.604 148.605 148.606 148.607 148.608 148.609 148.610
Z

Figure 4.21: Singularity in Centerline Axial Velocity for Case B

56
0.150

0.149

0.148

0.147

0.146

0.145
Wc

0.144
Wc = 7092597.1261/z**2 - 321.0152
0.143

0.142

0.141

0.140

0.139
4.52802E-05 4.52804E-05 4.52806E-05 4.52808E-05 4.52810E-05 4.52812E-05 4.52814E-05 4.52816E-05 4.52818E-05

1/z**2

Figure 4.22: Vortex Breakdown Location Using Richardson’s Extrapolation for Case B

57
1.4

1.2

0.8
Wc

0.6

0.4

0.2

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
Z

M=0.1 M=0.2 M=0.3 M=0.4 M=0.5 M=0.6 M=0.7 M=0.8

Figure 4.23: Centerline Axial Velocity for Various Mach Numbers

58
4.5

3.5

2.5
Zbreak

1.5

0.5

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Mach Number

Figure 4.24: Vortex Breakdown as a Function of Mach Number

59
0.4

0.35

0.3

0.25
Wc

0.2

0.15

0.1

0.05

0
1.0 1.5 2.0 2.5 3.0 3.5 4.0
Z

we=0.055 we=0.056 we=0.057 we=0.06

Figure 4.25: Critical Edge Axial Velocity Determination for M=0.2

60
CHAPTER 5
CONCLUSIONS
Results have been presented for a computational examination of compressible,

quasi-cylindrical vortices. The results show that the dynamics of the vortex are affected

by the imposed, edge axial velocity. The results show two possible flow regimes for a

compressible vortex. When the edge axial velocity is above some critical value, the

vortex will not break down and will continue to propagate downstream with decaying

velocity components and become plug-like flow. The results also show that the

compressible vortices behave similarly to incompressible vortices examined by Batchelor

[26]. The results show that the decay of the axial centerline velocity far downstream has

the form of ln (Z)/Z.

The results also show that for a small imposed edge axial velocity vortex

breakdown can occur. The breakdown structure is very similar to the ones seen for

incompressible flows. Also, a critical edge velocity was determined for a Mach number

of 0.2. The difference between the edge axial velocity for breakdown or for Batchelor

type flow is very small. This fact may prove important in the design of a system to

alleviate vortex breakdown for aircraft.

Although the investigation has produced significant results, there is an area that

needs future investigation. The initial dependent variable profiles were determined using

the incompressible formulation of Burggraf and Foster [14]. The incompressible

61
formulation allows for the calculation of the radial velocity and the static pressure based

on the assumed form of the tangential and axial velocity components. To determine the

values for temperature and density which are needed for the compressible solver, it was

assumed that the density was constant across the vortex. The temperature was then

calculated from the equation of state. The assumption of constant temperature could have

been made with the density calculated from the equation of state. Calculating the initial

profile in the described manner limits the ability to examine the effect of initial density

and temperature variations on the vortex dynamics. One possible method of obtaining

more realistic initial profiles would be to numerically solve the compressible Navier-

Stokes equations for initial dependent variable profiles. The solution would be obtained

for the initial axial location, and these initial profiles would then be used to initiate the

quasi-cylindrical calculations.

There are two areas of future work that may provide information on the

breakdown phenomena. The computational results provide a foundation for a

hydrodynamic stability analysis of isolated compressible vortices. Such analysis may

provide insight into the onset of vortex breakdown. Additionally, analytical work is

possible in the investigation of vortex breakdown using triple deck theory to handle the

separation singularity. The triple deck analysis may provide valuable information on the

behavior for the vortex after breakdown, separation, has occurred.

62
BIBLIOGRAPHY
1. Spalart, P. R., “Airplane Trailing Vortices,” Annual Review of Fluid Mechanics,
Vol. 30, 1998, Annual Reviews, Inc., Palo Alto, CA, pp. 107–138.

2. Thomas, G., Holzapfel, F, and Darracq, D., “Commercial Aircraft Wake Vortices,”
Annual Review of Fluid Mechanics, Vol. 38, 2002, pp. 181–208.

3. Cook, S. P., and Barlow, J, “Investigation of Effects of Leading Edge Extensions


Vents on the Lateral Characteristics of the F/A-18E/F in Power Approach
Configuration,” AIAA-2000-4510, 2000.

4. Hummel, D., and Srinivasan, P., “Vortex breakdown effects on the low speed
aerodynamic characteristics of slender delta wings in symmetrical flow,” Journal of
The Royal Aeronautical Society, Vol. 71, 1966, pp. 319–322.

5. Erickson, G., “Wind Tunnel Investigation of Vortex Flows on F/A-18 Configuration


at Subsonic Through Transonic Speeds,” NASA Technical Paper 3111, 1991.

6. Meyn, L. A., and James, K. D., “Full-Scale Wind-Tunnel Studies of F/A-18 Tail
Buffet,” AIAA Journal, Vol. 33, No. 3, May-June 1996, pp. 589-601.

7. Anderson W. D., Patel, S. R., and Black, C. L., “Low-Speed Wind Tunnel Buffet
Testing of the F-22,” Journal of Aircraft, Vol. 43, No. 4, July–August 2006, pp. 879–
885.

8. Hall, M. G., “Vortex Breakdown,” Annual Review of Fluid Mechanics, Vol. 4, 1972,
pp. 195–218.

9. Leibovich, S., “The Structure of Vortex Breakdown,” Annual Review of Fluid


Mechanics, Vol. 10, 1978, pp. 221–246.

10. Wedemeyer, E., “Vortex Breakdown,” AGARD-VKI Lecture Series 121 High Angle
of Attack Aerodynamics, 1982.

11. Leibovich S., “Vortex Stability and Breakdown: Survey and Extension,” AIAA
Journal, Vol. 22, No. 9, September 1984, pp. 1192–1205.

12. Escudier, M., “Vortex Breakdown: Observations and Explanations,” Progress in


Aerospace Sciences, Vol. 25, No. 2, 1988, pp. 180–229.

63
13. Delery, J. M., “Aspects of Vortex Breakdown,” Progress in Aerospace Sciences,
Vol. 30, No. 1, 1994, pp. 1–59.

14. Burggraf, O. R., and Foster, M. R., “Continuation or Breakdown of Tornado-Like


Vortices,” Journal of Fluid Mechanics, Vol. 80, Part 4, 1977, pp. 685–703.

15. Long, R. R., “A Vortex in an Infinite Fluid,” Journal of Fluid Mechanics, Vol. 11,
1962, pp. 611–625.

16. Crowley, S. J., “Personal communication on the topic,” 1998.

17. Mack, L. M., “The Compressible Viscous Heat Conducting Vortex,” Journal of Fluid
Mechanics, Vol. 8, 1960, p. 284.

18. Brown, S. N., “The Compressible Inviscid Leading-Edge Vortex,” Journal of Fluid
Mechanics, Vol. 22, 1965, p. 17.

19. Hall, M. G., “A Theory for the Core pf a Leading-Edge Velocity,” Journal of Fluid
Mechanics, Vol. 11, 1961, p. 209.

20. Orangi, S., Foster, M. R., and Bodonyi, R. J., “On the Structure of a Three-
Dimensional Compressible Vortex,” Computers and Fluids, Vol. 30, 2001, pp. 115–
135.

21. Liu, C. H., Krause, E., and Menne, S., “Admissible Upstream Conditions for Slender
Compressible Vortices,” AIAA-86-1093, 1986.

22. Kandil, O. A., Kandil, H. A., and Liu, C. H., “Computation of Steady and Unsteady
Compressible Quasi-Axisymmetric Vortex Flow and Breakdown,” AIAA-91-0752,
1991.

23. Kandil, O. A., and Kandil, H. A., “Computation of Compressible Quasi-


Axisymmetric Slender Vortex Flow and Breakdown,” Computer Physics
Communications, Vol. 65, 1991, pp. 164–172.

24. Tannehill, J. C., Anderson, D. A., and Pletcher, R. H., Computational Fluid
Mechanics and Heat Transfer, 2nd Edition, Taylor& Francis, Washington, DC, 1997.

64
25. Devenport, W. J., Rife, M. C., Liapis, S. I., and Follin, G. J., “The Structure and
Development of a Wing-tip Vortex,” Journal of Fluid Mechanics, Vol. 312, pp. 67–
106.

26. Batchelor, G. K., “Axial Flow in Trailing Line Vortices,” Journal of Fluid
Mechanics, Vol. 20, Part 4, 1964, pp. 645–658.

65

You might also like