0% found this document useful (0 votes)
246 views113 pages

III Year I Sem High Speed Aerodynamics

The document summarizes the course objectives, topics, and units for a High Speed Aerodynamics course offered at Malla Reddy College of Engineering and Technology. The course covers fundamental concepts of compressible flows including governing equations, normal and oblique shock waves, expansion waves, subsonic and supersonic airfoil flows, linearized supersonic flows, hypersonic flows, and flow through nozzles and ducts. The course aims to study the effects of compressibility on aerodynamic characteristics and experimental methods to analyze compressible flows.

Uploaded by

Iron Mike Vimal
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
246 views113 pages

III Year I Sem High Speed Aerodynamics

The document summarizes the course objectives, topics, and units for a High Speed Aerodynamics course offered at Malla Reddy College of Engineering and Technology. The course covers fundamental concepts of compressible flows including governing equations, normal and oblique shock waves, expansion waves, subsonic and supersonic airfoil flows, linearized supersonic flows, hypersonic flows, and flow through nozzles and ducts. The course aims to study the effects of compressibility on aerodynamic characteristics and experimental methods to analyze compressible flows.

Uploaded by

Iron Mike Vimal
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 113

AERONAUTICAL ENGINEERING – MRCET (UGC Autonomous)

HIGH SPEED AERODYNAMICS (R18A2113)


COURSE FILE
III B. Tech I Semester
(2021-2022)
Prepared By
Mr. J Sandeep, Assoc. Prof

Department of Aeronautical Engineering

MALLA REDDY COLLEGE OF ENGINEERING &


TECHNOLOGY

(Autonomous Institution – UGC, Govt. of India)


Affiliated to JNTU, Hyderabad, Approved by AICTE - Accredited by NBA & NAAC – ‘A’ Grade - ISO 9001:2015
Certified)
Maisammaguda, Dhulapally (Post Via. Kompally), Secunderabad – 500100, Telangana State, India.

1
AERONAUTICAL ENGINEERING – MRCET (UGC Autonomous)

MRCET VISION

• To become a model institution in the fields of Engineering, Technology and


Management.

• To have a perfect synchronization of the ideologies of MRCET with challenging

demands of International Pioneering Organizations.

MRCET MISSION

To establish a pedestal for the integral innovation, team spirit, originality and

competence in the students, expose them to face the global challenges and become

pioneers of Indian vision of modern society.

MRCET QUALITY POLICY.

• To pursue continual improvement of teaching learning process of Undergraduate and


Post Graduate programs in Engineering & Management vigorously.

• To provide state of art infrastructure and expertise to impart the quality education.

2
AERONAUTICAL ENGINEERING – MRCET (UGC Autonomous)

PROGRAM OUTCOMES
(PO’s)
Engineering Graduates will be able to:
1. Engineering knowledge: Apply the knowledge of mathematics, science, engineering
fundamentals, and an engineering specialization to the solution of complex engineering
problems.
2. Problem analysis: Identify, formulate, review research literature, and analyze complex
engineering problems reaching substantiated conclusions using first principles of
mathematics, natural sciences, and engineering sciences.
3. Design / development of solutions: Design solutions for complex engineering problems
and design system components or processes that meet the specified needs with
appropriate consideration for the public health and safety, and the cultural, societal,
and environmental considerations.
4. Conduct investigations of complex problems: Use research-based knowledge and
research methods including design of experiments, analysis and interpretation of data,
and synthesis of the information to provide valid conclusions.
5. Modern tool usage: Create, select, and apply appropriate techniques, resources, and
modern engineering and IT tools including prediction and modeling to complex
engineering activities with an understanding of the limitations.
6. The engineer and society: Apply reasoning informed by the contextual knowledge to
assess societal, health, safety, legal and cultural issues and the consequent
responsibilities relevant to the professional engineering practice.
7. Environment and sustainability: Understand the impact of the professional engineering
solutions in societal and environmental contexts, and demonstrate the knowledge of,
and need for sustainable development.
8. Ethics: Apply ethical principles and commit to professional ethics and responsibilities
and norms of the engineering practice.
9. Individual and team work: Function effectively as an individual, and as a member or
leader in diverse teams, and in multidisciplinary settings.
10. Communication: Communicate effectively on complex engineering activities with the
engineering community and with society at large, such as, being able to comprehend
and write effective reports and design documentation, make effective presentations,
and give and receive clear instructions.
11. Project management and finance: Demonstrate knowledge and understanding of the
engineering and management principles and apply these to one’s own work, as a
member and leader in a team, to manage projects and in multi disciplinary
environments.
12. Life- long learning: Recognize the need for, and have the preparation and ability to
engage in independent and life-long learning in the broadest context of technological
change.

3
AERONAUTICAL ENGINEERING – MRCET (UGC Autonomous)

DEPARTMENT OF AERONAUTICAL ENGINEERING


VISION

Department of Aeronautical Engineering aims to be indispensable source in Aeronautical


Engineering which has a zeal to provide the value driven platform for the students to
acquire knowledge and empower themselves to shoulder higher responsibility in building
a strong nation.

MISSION

The primary mission of the department is to promote engineering education and research.
To strive consistently to provide quality education, keeping in pace with time and
technology. Department passions to integrate the intellectual, spiritual, ethical and social
development of the students for shaping them into dynamic engineers.

QUALITY POLICY STATEMENT

Impart up-to-date knowledge to the students in Aeronautical area to make them quality
engineers. Make the students experience the applications on quality equipment and tools.
Provide systems, resources and training opportunities to achieve continuous
improvement. Maintain global standards in education, training and services.

4
AERONAUTICAL ENGINEERING – MRCET (UGC Autonomous)

PROGRAM EDUCATIONAL OBJECTIVES – Aeronautical Engineering


1. PEO1 (PROFESSIONALISM & CITIZENSHIP): To create and sustain a community of
learning in which students acquire knowledge and learn to apply it professionally with
due consideration for ethical, ecological and economic issues.
2. PEO2 (TECHNICAL ACCOMPLISHMENTS): To provide knowledge based services to satisfy
the needs of society and the industry by providing hands on experience in various
technologies in core field.
3. PEO3 (INVENTION, INNOVATION AND CREATIVITY): To make the students to design,
experiment, analyze, and interpret in the core field with the help of other multi
disciplinary concepts wherever applicable.
4. PEO4 (PROFESSIONAL DEVELOPMENT): To educate the students to disseminate
research findings with good soft skills and become a successful entrepreneur.
5. PEO5 (HUMAN RESOURCE DEVELOPMENT): To graduate the students in building
national capabilities in technology, education and research

PROGRAM SPECIFIC OUTCOMES – Aeronautical Engineering


1. To mould students to become a professional with all necessary skills, personality and
sound knowledge in basic and advance technological areas.
2. To promote understanding of concepts and develop ability in design manufacture and
maintenance of aircraft, aerospace vehicles and associated equipment and develop
application capability of the concepts sciences to engineering design and processes.
3. Understanding the current scenario in the field of aeronautics and acquire ability to
apply knowledge of engineering, science and mathematics to design and conduct
experiments in the field of Aeronautical Engineering.
4. To develop leadership skills in our students necessary to shape the social, intellectual,
business and technical worlds.

5
AERONAUTICAL ENGINEERING – MRCET (UGC Autonomous)

MALLA REDDY COLLEGE OF ENGINEERING AND TECHNOLOGY


III Year B.Tech ANE-I Sem L T/P/D C
2 1/-/- 3

(R18A2113) HIGH SPEED AERODYNAMICS

Objectives:
 Study the basic governing equations of compressible flows and its parameters.
 Study the effects of Shock and Expansion waves on aerodynamic characteristics.
 Learn about the experimental methods to study about compressible flows.

Tables: Isentropic, Normal Shock, Oblique Shock, Prandtl Meyer function.

UNIT-I ONE DIMENSIONAL COMPRESSIBLE FLOWS


Review of Thermodynamics. Definition of Compressibility, Stagnation conditions, Speed of sound, Mach number,
shock waves. One dimensional flow governing equations. Alternative forms of Energy equations, Normal shock
relations with numerical.

UNIT-II OBLIQUE SHOCK AND EXPANSION WAVES


Oblique shock waves. Supersonic flow over a wedge Θ - β - M relations strong and weak shock solutions, regular
reflection from a solid boundary. Expansion waves, Prandtl – Meyer Expansion. Shock Expansion theory.

UNIT-III
SUBSONIC COMPRESSIBLE FLOW OVER AIRFOIL
Introduction - Velocity potential equation –small perturbation equation - Prandtl-Glauert compressibility
corrections - Critical Mach number with numericals - Drag divergence Mach number - Area rule - Supercritical
airfoil.

UNIT – IV
LINEARIZED SUPERSONIC FLOWS AND HYPERSONIC FLOWS
Linearized supersonic pressure coefficient, application to airfoils, lift and drag for flat plate, comparision with shock
expansion theory.
Qualitative aspects of hypersonic flows, Newtonian theory, modified Newtonian theory, lift and drag.

UNIT- V
FLOW THROUGH NOZZLES AND VARIABLE AREA DUCTS
Quasi one dimensional flow, Area-velocity relation, Isentropic flow through Convergent – Divergent nozzles.
Choked flow conditions. Under and Over expansion conditions. Flow through diffusers – wave reflections from a
free boundary. Application to supersonic wind tunnel.

Text Books:
1. Anderson, J .D., Fundamental of Aerodynamics, Mc Graw-Hill International third edition Singapore-
2001.

Reference Books:
1. Radhakrishnan, E, E., Gas Dynamics, Prentice Hall of India, 1995.
2. Anderson, J .D., Modern Compressible Flow with Historical Perspective, Mc Graw-Hill International third edition
Singapore-2004.

Outcomes:
 Understand the compressible flow parameters shock and expansion wave effecting flow
behavior.

6
AERONAUTICAL ENGINEERING – MRCET (UGC Autonomous)

 Able to design nozzle, diffuser and variable area ducts to obtain required aerodynamic outputs.
 Able to understand hypersonic flows.

7
MALLA REDDY COLLEGE OF ENGINEERING ANDTECHNOLOGY
(UGC AUTONOMOUS)
III B.TECH I SEMESTER – AERONAUTICAL ENGINEERING
HIGH SPEED AERODYNAMICS - II (R15)
MODEL PAPER – I
MAXIMUM MARKS: 75
PART A Max Marks: 25
i. All questions in this section are compulsory
ii. Answer in TWO to FOUR sentences.

1. Define compressible and incompressible flows. (2 M)


2. Using a neat sketch show the shock pattern in supersonic flow regime and state the changes in
the flow after a shock wave. (3 M)
3. Define one – dimensional flow and quasi one – dimensional flows. Give suitable examples for
each. (3 M)
4. For a calorically perfect gas prove that the square of mach number is proportional to ratio of
kinetic and internal energy. (2 M)
5. Using necessary assumptions prove that the tangential component of flow velocity is preserved
across an oblique shock wave. (3M)
6. Using neat sketch, define Mach reflection. (2 M)
7. Define the terms choking, over expanded, under expanded nozzles. (2 M)
8. Give the relation between incompressible pressure/force coefficient and compressible
pressure/force coefficients in a linearised subsonic flow. (3 M)
9. Formulate finite difference method. (2 M)
10. Define truncation error and round – off error. (3 M)

PART B Max Marks: 50


i. Answer only one question among the two questions in choice.
ii. Each question answer (irrespective of the bits) carries 10M.

11. A pressure vessel has a volume of 10 m3 is used to store a high pressure air for operating a
supersonic wind tunnel. If the air pressure and temperature inside the vessel are 20 atm and
300 K respectively, calculate
a. Mass of the air stored inside the vessel
b. Total energy of the gas stored inside the vessel
c. If the gas in the vessel is heated, the temperature rises to 600 K calculate the change in
entropy of the air inside the vessel.
OR
12. a. State second law of thermodynamics and derive the relations for calculating the change in
entropy.
b. Derive the isentropic flow relations.

8
In either, explain the nomenclature used clearly.

13. Starting from the steady flow one dimensional energy equation derive the various alternative
forms of energy equations. Explain all the symbols used clearly.
OR
14. For the flow across a normal shock
a. Prove that a*2 = u1u2 (Prandtl’s relation)
b. The Mach number behind a normal shock is always subsonic.
c. The total temperature across a normal shock wave is constant

15. Making necessary assumptions/using required conditions derive the relation between flow
deflection angle, shock angle and upstream Mach number (θ-β-M)
OR
16. a. Derive the governing equation for Prandtl – Meyer expansion flow.
b. Consider the flow past an expansion corner of angle 30o. The upstream Mach number, pressure and
temperature are given by 2, 3 atm and 400 K respectively. Calculate the downstream Mach number,
pressure, temperature, total temperature and total pressure.

17. Consider a flat plate at with chord length c at an angle of attack α to a supersonic free stream
mach number M∞. Let L and D be lift and drag per unit span S is plan-form area of the plate
per unit span, S = c(1). Using linearised theory, derive the following expressions for lift and
drag coefficients.
4𝛼 4𝛼 2
𝐶𝐿 = ; 𝐶𝐷 =
√𝑀∞2 − 1 √𝑀∞2 − 1
OR
18. Consider a rocket engine burning Hydrogen and oxygen. The combustion chamber pressure
and temperature are 25 atm and 3571 K, respectively. The molecular weight of the chemically
reacting gas in the combustion chamber is 16. The pressure at the exit of the convergent –
divergent rocket nozzle is 1.174 x 102 atm. The throat area is 0.4 m2. Assuming a calorically
perfect gas, calculate a) the exit Mach number , b) the exit velocity , c) the mass flow through
the nozzle , and d) the area at the exit

19. a. Explain about similarities of flow to be satisfied for Model testing.


b. Illustrate the flow over a delta wing in supersonic flow.
OR
20. Write a short note on Hotwire anemometer.

9
MALLA REDDY COLLEGE OF ENGINEERING ANDTECHNOLOGY
(UGC AUTONOMOUS)
III B.TECH I SEMESTER – AERONAUTICAL ENGINEERING
HIGH SPEED AERODYNAMICS - II (R15)
MODEL PAPER – II
MAXIMUM MARKS: 75
PART A Max Marks: 25
i. All questions in this section are compulsory
ii. Answer in TWO to FOUR sentences.

1. State first and second law of Thermodynamics. Define entropy, internal energy and enthalpy.
(3M)
2. Calculate the isothermal compressibility of air at a pressure of 0.5 atm. (2 M)
3. Define characteristic speed of sound and stagnation speed of sound. (3 M)
4. Give the relations between characteristic properties and stagnation properties of a flow.(2 M)
5. Define shock strength and classify strong and weak shocks. (2 M)
6. State the advantages of graphical representation of the solution of a flow problem. (3 M)
7. Using neat schematic sketch, explain the application of nozzles. (3 M)
8. Define critical Mach number and drag – divergence Mach number. (2 M)
9. Write about advantages of delta wing. (2 M)
10. Sketch the surface stream lines on a cone at an AoA. (3 M)

PART B Max Marks: 50


i. Answer only one question among the two questions in choice.
ii. Each question answer (irrespective of the bits) carries 10M.

11. a. Define speed of sound. Derive the expressions for speed of sound I terms of pressure, density
and temperature. (5 M)
b. Define thermally perfect and calorically perfect gases. Give the equation of state for
calorically and thermally perfect gases. (5 M)
OR
12. a. Air flows through a duct. The pressure and temperature at station 1 are 0.7 atm and 300C,
respectively. At a second station, the pressure is 0.5 atm. Calculate the temperature and density
at the second station. Assume the flow to be isentropic.
b. State the limitations of air as a perfect gas.
c. Air at 300C is compressed isentropically to occupy a volume which is 1/30 of its initial
volume. Assuming air as an ideal gas, determine the final temperature.

13. Using energy equation, derive the relation between static properties and stagnation properties
of a flow making necessary assumptions.
OR

10
14. a. Derive the relation between total pressures across normal shock waves. Explain all the
symbols used clearly. (5 M)
b. A re-entry vehicle is at an altitude of 15,000 m and has a velocity of 1850 m/s. a bow shock
wave envelops the vehicle. Neglecting disassociation, determine the static and stagnation
pressure just behind the shock wave on the vehicles center line where the shock is assumed to
be normal shock. Assume that air behaves as perfect gas with γ = 1.4 and R = 287 J/kg - K. (5
M)

15. A uniform supersonic stream with M1 = 3.0, p1 = 1 atm and T1 = 288 K encounters a
compression corner which deflects the flow stream by an angle of 200C. Calculate the shock
wave angle and p2, T2,, M2, po2, T02 behind the shock wave. All the symbols used are standard.
Comment on the result if the deflection angle is increased keeping Mach number constant and
the Mach number is increased with deflection angle constant, while the remaining parameters
are the same.
OR
o
16. A flat plate is kept at 15 angle of attack to a supersonic flow at Mach number 2.4. Solve the
flow field around the plate and determine the inclination of slipstream direction using shock
expansion theory.

17. a) Define linearization. Obtain an expression for linearized pressure coefficient.


b) Obtain an expression for pressure coefficient for a linearized subsonic flow over a two
dimensional profile.( Prandtl-Glauert rule).
c) The low-speed lift coefficient for an NACA 2412 airfoil at an angle of attack of 40 is 0.65.
Using the Prandtl-Glauert rule, calculate the lift coefficient for M∞ = 0.7.
OR
18. a) What is diffuser? Sketch a nozzle with conventional supersonic diffuser
b) A supersonic wind tunnel is designed is designed to produce flow at Mach 2.4. at standard
atmospheric conditions. Calculate (i) the exit to throat area ratio of the nozzle (ii) Reservoir
pressure and temperature.

19. Describe briefly about components of wind tunnel and flow measurement devices.
OR
20. Write a short note on Laser Doppler anemometer.

11
MALLA REDDY COLLEGE OF ENGINEERING ANDTECHNOLOGY
(UGC AUTONOMOUS)
III B.TECH I SEMESTER – AERONAUTICAL ENGINEERING
HIGH SPEED AERODYNAMICS - II (R15)
MODEL PAPER – III
MAXIMUM MARKS: 75
PART A Max Marks: 25
i. All questions in this section are compulsory
ii. Answer in TWO to FOUR sentences.

1. Define isentropic flow. State the relation between flow properties in an isentropic flow. (2M)
2. At the nose of the missile in flight, the pressure and temperature are 5.6 atm and 8500C,
respectively. Calculate the density and specific volume. (3 M)
3. What are the governing equations for steady one – dimensional flow? (2 M)
4. For a flow through a variable area duct, give the relation between Area and velocity of the
flow. What are the assumptions made in deriving this equation (3 M)
5. Focus on the formation of three – dimensional shock waves. (2 M)
6. State the difference between flow over wedges and cones. (3 M)
7. Give the governing equations for quasi 1- D flow. (2 M)
8. Give the three echelons of transonic inviscid flow theory. (3 M)
9. What is kinematic similarity of flow. (2 M)
10. What types of experiments are carried out by suing wind tunnel? (3 M)

PART B Max Marks: 50


i. Answer only one question among the two questions in choice.
ii. Each question answer (irrespective of the bits) carries 10M.

11. Air flows isentropically through a nozzle. If the velocity and the temperature at the exit of the
nozzle are 390 m/s and 28oC, respectively, determine the Mach number and Stagnation
temperature at the exit. What will be the Mach number just upstream of a station where the
temperature is 92.5oC.
OR
12. Derive the normal relations for a perfect gas. Make necessary assumptions and explain the
nomenclature.
13. Consider a supersonic flow at Mach 2.8 with a static pressure and temperature of 1 atm and
5190 R, respectively. The flow passes over a compression corner with a deflection angle of
160. The oblique shock generated at the corner propagates into the flow, and is incident on a
horizontal wall. Calculate the angle Φ made by the reflected shock wave with respect to the
wall, and the Mach number, pressure and temperature behind the reflected shock. Assume that
the flow is parallel to the horizontal after moving across the reflected shock.

12
OR
14. a. Write about shock polar and pressure deflection diagrams.
b. Explain about prandtl-meyer expansion waves.

15. a) Define Area rule and its importance in designing supersonic aircraft.
b) Define critical Mach number. Obtain an expression for pressure coefficient at critical Mach
number.

OR
16. a) Derive the linearised supersonic flow governing equation.
b) At α = 00, the minimum pressure coefficient for an NACA 0009 airfoil in low-speed flow is
-0.25. Calculate the critical Mach number for this airfoil using Prandtl-Glauert rule and
Karman-Tsien rule.
17. Explain about the method of characteristics for supersonic wind tunnel design.
OR
18. Explain about Quasi one dimensional flow and the area mach relation with over and under
expanded flows.
19. Write a short note on Blow down and indraft tunnel layouts and their design features.
OR
20. Write a short note on advantages and disadvantages of wind tunnel.

13
MALLA REDDY COLLEGE OF ENGINEERING ANDTECHNOLOGY
(UGC AUTONOMOUS)
III B.TECH I SEMESTER – AERONAUTICAL ENGINEERING
HIGH SPEED AERODYNAMICS - II (R15)
MODEL PAPER – IV
MAXIMUM MARKS: 75
PART A Max Marks: 25
i. All questions in this section are compulsory
ii. Answer in TWO to FOUR sentences.

1. Define stagnation conditions and characteristic conditions. (2M)


2. Give the 3 basic governing equations of fluid flow . (3 M)
3. Give the laplace equation in terms of speed of sound. (2 M)
4. Show that the mass flow rate across a stream tube in compressible flow field is inversely
proportional to its sectional area. (3 M)
5. Consider a supersonic flow at Mach 2.8 over a compression corner with a deflection angle of
150. If the deflection angle is doubled, what is the increase in shock strength? Is it also
doubled? Comment. (2 M)
6. Give Prandtl – Meyer function and its significance. (3 M)
7. Define linearization. Give the small perturbation equation. (2 M)
8. Give the expression for Cpcr and necessary deductions. (3 M)
9. Define region of influence and domain of independence. (2 M)
10. Define dynamic similarity of flows. (3 M)

PART B Max Marks: 50


i. Answer only one question among the two questions in choice.
ii. Each question answer (irrespective of the bits) carries 10M.

11. Define Mach number and its importance. Using neat sketches, explain the flow pattern in
various flow regimes.
OR

12. a. At a given point in the high speed flow over the airplane wing, the local Mach number,
pressure and Temperature are 0.7, 0.2 atm and 250 K respectively. Calculate the values of po,
To, p*, T*, a* at this point. The symbols used are according to the standard convention. (5M)
b. Consider a normal shock wave in the flow. The upstream conditions are given by M1=3, p1
= 1 atm and ρ1 = 1.23 kg/m3. Calculate the downstream values p2, T2, M2, u2, po2, To2. The
symbols used are according to the standard convention. (5M)
13. Using neat sketches, explain the mathematical/graphical procedures for solving the flow
problem
a. When the shocks of opposite families intersect

14
b. When the shocks of same family intersect
OR
14. Consider an infinitely thin flat plate at an angle of attack of 200 in a Mach 3 free – stream.
Calculate the magnitude of flow direction angle φ downstream the trailing edge.
15. Derive the linearized pressure coefficient for supersonic flows.
OR
o
16. A flat plate is kept at 15 angle of attack to a supersonic flow at Mach number 2.4. Solve the
flow field around the plate and determine the inclination of slipstream direction using shock
expansion theory.

17. a. Derive the Area – Mach relation for the variable area ducts like a nozzle.
b. Consider the purely subsonic flow in a convergent – divergent duct. The inlet, throat and
the exit area are 1 m2, 0.7 m2 and 0.85 m2 respectively. If the inlet Mach and pressure are 0.3 and
0.8 x 105 N/m2, respectively, then calculate: M and p at the throat and exit.
OR
18. Explain about the role of leading edge extension to improve the performance of aircraft at high
angle of attack.

19. Write a short note on Non dimensional parameters and explain about its importance in wind
tunnel testing.
OR
20. Discuss briefly about schileren flow visualization technique with neat sketch.

15
MALLA REDDY COLLEGE OF ENGINEERING ANDTECHNOLOGY
(UGC AUTONOMOUS)
III B.TECH I SEMESTER – AERONAUTICAL ENGINEERING
HIGH SPEED AERODYNAMICS - II (R15)
MODEL PAPER – V
MAXIMUM MARKS: 75
PART A Max Marks: 25
i. All questions in this section are compulsory
ii. Answer in TWO to FOUR sentences.

1. Define the terms continuum flow, free – molecular flow and low density or rarefied flows.
(3 M)
2. Define the terms Universal gas constant, Gas constant and Boltzmann constant. (2 M)
3. Explain in simple steps, how supersonic stream is generated in a Convergent – divergent
nozzle. (3 M)
4. Give the relation of change in entropy of the flow across a normal shock wave. (2 M)
5. Define flow deflection angle, shock angle and mach angle. (2 M)
6. How does an expansion fan or a shock wave behave when they encounter a free boundary?
Illustrate the diamond wave pattern using neat sketch. (3 M)
7. State area – rule and define super critical airfoil. (3 M)
8. Give the expressions used for correcting Prandtl – glauret rule. (2 M)
9. Give the expression for pressure coefficient in linearised supersonic flow. (2 M)
10. Define transonic drag. (3 M)

PART B Max Marks: 50


i. Answer only one question among the two questions in choice.
ii. Each question answer (irrespective of the bits) carries 10M.
11. a. Define compressibility. (3M)
b. Explain briefly about changes in flow properties due to one dimensional flow with heat
addition and friction. (7M)
OR
12. A ramjet flies at 11 km altitude with a flight Mach number of 0.9. In the inlet diffuse, the air
is brought to the stagnation condition so that it is stationary just before the combustion
chamber. Combustion takes place at constant pressure and a temperature increase of 15000C
takes place. The combustion products are then ejected through the nozzle.
a. Calculate the stagnation pressure and temperature.
b. What will be the nozzle exit velocity? (refer RathaKrishnan, chapter 4)
13. a. Using neat sketch, explain the change of properties behind a oblique shock wave. (5 M)
b. Upstream of the oblique shock wave M1 = 3, p1 = 0.5 atm and T1 = 200 K. Calculate the
effect of wave angle on the down stream properties M2, p2, T2, u2, ρ2 for 15 and 30 degrees.
(5 M)

16
OR

14. a. Write short notes on wave reflection from free boundary.


b. Air flows at Mach 4.0 and pressure 105 N/m2 is turned abruptly by a wall into the flow with
a turning angle of 20o. If the shock is reflected by another wall determine the flow properties
M and ρ downstream of the reflected shock.
15. Derive the velocity potential equation.
OR

16. Write a short note on Critical Mach number, Drag divergence number and supercritical airfoil.
17. Derive the expression for mass flow rate of a calorically perfect gas through a choked nozzle.
𝛾+1
𝑝𝑜 𝐴∗ √𝛾 2 𝛾−1
Explain the terms used clearly.𝑚̇ = ( )
√𝑇𝑜 𝑅 𝛾+1

OR
18. a. Write a short note on vortex lift and its effect.
b. Explain briefly about flow behavior over delta wings at high angle of attack.

19. Write a short note on Shadow graph flow visualization technique with neat sketches
OR
20. Discuss briefly about the wind tunnel balances to measure the forces and moments.

17
Chapter 1
One dimensional compressible
flows

1.1 Introduction
The first and foremost point, that is related with high speed aerodynamics is
that here the speed of the fluid or air is considerably large, then, how large it is
that quantification we will be doing later on, but we can say that, the speed is
comparable to the local speed of sound. And then, in order to maintain a flow at
very high speed. Obviously, the pressure difference or the pressure changes that
will be associated are quite large. Now once, the pressure changes are large, the
gases is likely to change its density.

1.2 Compressible flows


Compressible flow is the science of fluid flow where the density change associ-
ated with pressure change is significant. Fluid mechanics is the science of fluid
flow in which the temperature changes associated with the flow are insignificant.
The simple definition of compressible flow as one in which the density is variable
requires more elaboration. Consider a small element of fluid of volume v. The
pressure exerted on the sides of the element by the neighboring fluid is p. Assume
the pressure is now increased by an infinitesimal amount dp. The volume of the
element will be correspondingly compressed by the amount dv. Since the volume
is reduced, dv is a negative quantity. The compressibility of the fluid, τ , is defined
as
1 dp
τ =− (1.1)
v dv
Fluids such as water are incompressible (i.e density does not change) under
normal conditions. But under conditions of high pressure (e.g. 1000 atm) they are
compressible. The change in volume is the characteristic feature of a compressible
medium under static conditions. Under dynamic conditions, that is when the
medium is moving, the characteristic feature for incompressible and compressible
flow situations are: the volume flow rate, Q̇ = AV = constant at any cross-
section of a streamtube for incompressible flow, and the mass flow rate, ṁ =
ρAV = constant at any cross-section of a streamtube for compressible flow. Here,
A is the cross-sectional area of the streamtube and V and ρ are the the velocity
and density of the fluid at that cross-section.
As long as a gas flows at a sufficiently low speed fromone cross-section of a pas-
sage to another the change in volume (or density) can be neglected and, therefore,
the flow can be treated as incompressible. Although the fluid is compressible, this
property may be neglected when the flow is taking place at low speeds. In other

1
1
18
Chapter 1. One dimensional compressible flows

Fig. 1.1: Streamtube

words, although there is some density change associated with every physical flow,
it is often possible (for low-speed flows) to neglect it and idealize the flow as in-
compressible. This approximation is applicable to many practical flow situations,
such as low-speed flow around an airplane and flow through a vacuum cleaner.
From the above discussion it is clear that compressibility is the phenomenon
by virtue of which the flow changes its density with changes in speed. Now, the
question is, what are the precise conditions under which density changes must be
considered?
A quantitative measure of compressibility is the volume modulus of elasticity
E, defined as

△P
E= (1.2)
△V /Vi
where △P is the change in static pressure, △V is the change in volume, and
Vi is the initial volume. For ideal gases, the equation of state is

P V = RT (1.3)
For isothermal flows, this reduces to

P V = Pi Vi = constant (1.4)
where Pi is the initial pressure.
The above equation may be written as

(Pi + △P )(Vi + △V ) = Pi Vi (1.5)


Expanding this equation, and neglecting the second-order terms, we get

△P Vi + △V Pi = 0 (1.6)

High Speed Aerodynamics (R18A2113) 2 Dr. G. Srinivasan


2
19
Compressible flows

Therefore,
 
△V
△P = −Pi (1.7)
Vi
For gases, from Eqs. 1.2 and 1.7, we get

E = Pi (1.8)
Hence, by Equ. 1.7, the compressibility may be defined as the volume modulus
of the pressure.
Further, By mass conservation, we have ṁ = ρV = constant, where ṁ is mass
flow rate per unit area, V is the flow velocity, and ρ is the corresponding density.
This can also be written as

(Vi + △V ) (ρi + △ρ) = ρi Vi (1.9)

Considering only first-order terms, this simplifies to

△ρ △V
=− (1.10)
ρi Vi
Substituting this into Equ. 1.2, we get

△ρ
△P = E (1.11)
ρi
From Equ. 1.11, it can be seen that the compressibility may also be defined
as the density modulus of the pressure.
For incompressible flows, by Bernoulli’s equation, we have

1
P + ρV 2 = constant = Pstag (1.12)
2
where the subscript “stag” refers to stagnation condition.The above equation
may also be written as

1
Pstag − P = ρV 2 (1.13)
2
that is the change of pressure from stagnation to static states is equal to 12 ρV 2 .
Using equ. 1.11, the above equation can be written as

△P △ρ ρi Vi2 qi
= = = (1.14)
E ρi 2E E
Here, qi = 21 ρi Vi2 is the dynamic pressure. Equ. 1.14 relates the density change
to the flow speed.
The compressibility effects can be neglected if the density changes are very
small, i.e. if

△ρ
≪1 (1.15)
ρi

High Speed Aerodynamics (R18A2113) 3


3
20
Chapter 1. One dimensional compressible flows

From equ. 1.14, it is seen that for neglecting compressibility


q
≪1 (1.16)
E
For gases, the speed of sound a may be expressed in terms of pressure and
density changes as
△P
a2 = (1.17)
△ρ
Using Equ. 1.11 in the above relation, we get

E
a2 = (1.18)
ρi
Using this Equ. 1.14 changes to
 2
△ρ ρi Vi2 1 V
= = (1.19)
ρi 2 E 2 2
The ratio V /a is called the Mach number M .Therefore, the condition of in-
compressibility for gases becomes

M2
≪1 (1.20)
2
Thus, the criterion determining the effect of compressibility for gases is that the
magnitude of theMach numberMshould be negligibly small. Indeed,mathematics
would stipulate this limit as M → 0. But Mach number zero corresponds to
stagnation state. Therefore, in engineering sciences flows with very small Mach
numbers are treated as incompressible. To have a quantification of this limiting
value of the Mach number to treat a flow as incompressible, a Mach number
corresponding to a 5

△ρ
≤ 0.05 or 5% (1.21)
ρ
that is when M ≤ 0.3. In other words, the flow may be treated as incompress-
ible when V ≤ 100 m/s, that is when V ≤ 360 kmph under standard sea level
conditions.

1.3 Thermodynamics concepts


The kinetic energy per unit mass, v 2 /2, of a high-speed flow is large. As the flow
moves over solid bodies or through ducts such as nozzles and diffusers, the local
velocity, hence local kinetic energy, changes. In contrast to low-speed or incom-
pressible flow, these energy changes are substantial enough to strongly interact
with other properties of the flow. Because in most cases high-speed flow and com-
pressible flow are synonymous, energy concepts play a major role in the study and
understanding of compressible flow. In turn, the science of energy (and entropy) is
thermodynamics; consequently, thermodynamics is an essential ingredient in the
study of compressible flow.

High Speed Aerodynamics (R18A2113) 4 Dr. G. Srinivasan


4
21
Thermodynamics concepts

1.3.1 Thermodynamic Systems


A thermodynamic system is a quantity of matter separated from the ”surround-
ings” or the ”environment” by an enclosure. The system is studied with the help
of measurements carried out and recorded in the surroundings. A thermometer
inserted into a system forms part of the surroundings. Work done by moving a
piston is measured by, say, the extension of a spring or the movement of a weight
in the surroundings. Heat transferred to the system is measured also by changes
in the surroundings e.g., heat may be transferred by an electrical heating coil. The
electric power is measured in the surroundngs.

Types of systems
Two types of systems can be distinguished. These are referred to, respectively, as
closed systems and open systems or control volumes. A closed system or a control
mass refers to a fixed quantity of matter, whereas a control volume is a region in
space through which mass may flow. A special type of closed system that does
not interact with its surroundings is called an Isolated system.
Two types of exchange can occur between the system and its surroundings:

❼ Energy exchange (heat or work).

❼ Exchange of matter (movement of molecules across the boundary of the


system and surroundings).

Based on the types of exchange, a system can can be called as:

❼ Isolated Systems: No exchange of matter and energy.

❼ Closed Systems: No exchange of matter but some exchange of energy.

❼ Open Systems: Exchange of both matter and energy

If the boundary does not allow heat (energy) exchange to take place it is called
adiabatic boundary.

1.3.2 Perfect gas


A perfect gas is one whose individual molecules interact only via direct collisions,
with no other intermolecular forces present. For such a perfect gas, the properties
p, ρ, and the temperature T are related by the following equation of state

p = ρRT (1.22)
where R is the specific gas constant. For air, R = 287J/Kg − K

1.3.3 Internel energy and Enthalpy


Internal energy is the sum of the kinetic and potential energies of the particles that
form the system. For an equilibrium system of a real gas where intermolecular
forces are important, and also for an equilibrium chemically reacting mixture of
perfect gases, the internal energy is a function of both temperature and volume.

High Speed Aerodynamics (R18A2113) 5


5
22
Chapter 1. One dimensional compressible flows

Let e denote the specific internal energy (internal energy per unit mass). Then,
the enthalpy, h, is defined, per unit mass, as

h = e + pv (1.23)

In many compressible flow applications, the pressures and temperatures are


moderate enough that the gas can be considered to be calorically perfect. Consis-
tent with Equ. 1.22 and the definition of enthalpy is the relation

cp − cv = R (1.24)

where the specific heats at constant pressure and constant volume are defined
as  
∂h
cp = (1.25)
∂T p
and  
∂e
cv = (1.26)
∂T v
respectively.
Equ. 1.24 can be deduced into useful forms. Dividing Equ. 1.24 by cp

cv R
1− = (1.27)
cp cp
Defining the heat capacity ratio, γ = cp /cv , Equ. 1.27 becomes

1 R
1− = (1.28)
γ cp

Solving for cp ,
γR
cp = (1.29)
γ−1
Similarly, by dividing Equ. 1.27 by cv , we find that

R
cv = (1.30)
γ−1
Equations 1.29 and 1.30 hold for a thermally or calorically perfect gas. They
will be widely used for treatment of compressible flow.

1.3.4 Laws of Thermodynamics


1.3.4.1 Zeroth law of thermodynamics

This law states that, when system A is in thermal equilibrium with system B and
system B is separately in thermal equilibrium with system C then system A and
C are also in thermal equilibrium. This law portrays temperature as a property
of the system and gives basis of temperature measurement.

High Speed Aerodynamics (R18A2113) 6 Dr. G. Srinivasan


6
23
Thermodynamics concepts

1.3.4.2 First law of thermodynamics


Consider a closed system, consisting of a certain amount of gas at rest, across
whose boundaries no transfer of mass is possible. Let ∂Q be an incremental
amount of heat added to the system across the boundary (by thermal conduction
or by direct radiation). Also, let ∂W denote the work done on the system by
the surroundings (or by the system on the surroundings). The sign convention
is positive when the work is done by the system and negative when the work is
done on the system. Owing to the molecular motion of the gas, the system has
an internal energy U. The first law of thermodynamics states that the heat added
minus work done by the system is equal to the change in the internal energy of
the system:

∂Q − ∂W = de (1.31)
This is an empirical result confirmed by laboratory experiments and practical
experience. In Equ. 1.31, the internal energy U is a state variable (thermodynamic
property). Hence, the change in internal energy de is an exact differential and its
value depends only on the initial and final states of the system. In contrast (the
non-thermodynamic properties), ∂Q and ∂W depend on the process by which the
system attained its final state from the initial state.
In general, for any given de, there are an infinite number of ways (processes)
by which heat can be added and work can be done on the system. In the present
course of study, we will mainly be concerned with the following three types of
processes only.
❼ Adiabatic process: A process in which no heat is added to or taken away
from the system.

❼ Reversible process: A process which can be reversed without leaving any


trace on the surroundings, that is both the system and the surroundings are
returned to their initial states at the end of the reverse process.

❼ Isentropic process: A process which is adiabatic and reversible.

1.3.4.3 The Second Law of Thermodynamics


Let us consider a cold body coming into contact with a hot body. From experience,
we can say that the cold body will get heated up and the hot body will cool down.
However, Equ. 1.31 does not necessarily imply that this will happen. In fact, the
first law allows the cold body to become cooler and the hot body to become hotter
as long as energy is conserved during the process. However, in practice this does
not happen; instead, the law of nature imposes another condition on the process,
a condition that stipulates the direction in which a process should take place. To
ascertain the proper direction of a process, let us define a new state variable, the
entropy, as follows.
∂qrev
ds = (1.32)
T
where s is the entropy (amount of disorder) of the system, ∂qrev is an in-
cremental amount of heat added reversibly to the system, and T is the system

High Speed Aerodynamics (R18A2113) 7


7
24
Chapter 1. One dimensional compressible flows

temperature. The above definition gives the change in entropy in terms of a re-
versible addition of heat, ∂qrev . Since entropy is a state variable, it can be used
in conjunction with any type of process, reversible or irreversible. The quantity
∂qrev is just an artifice; an effective value of ∂qrev can always be assigned to relate
the initial and final states of an irreversible process, where the actual amount of
heat added is ∂qrev . Indeed, an alternative and probably more lucid relation is

∂q
ds = + dsirrev (1.33)
T
The above equation applies to all process. It states that the change in entropy
during any process is equal to the actual heat added, ∂q, divided by the temper-
ature, ∂q/T , plus a contribution from the irreversible dissipative phenomena of
viscosity, thermal conductivity, and mass diffusion occurring within the system,
dsirrev .These dissipative phenomena always cause an increase in of entropy:

dsirrev ≥ 0 (1.34)

If ds > 0, the process is called an irreversible process, and when ds = 0, the


process is called a reversible process. A reversible and adiabatic process is called
an isentropic process. However, in a nonadiabatic process, we can extract heat
from the system and thus decrease the entropy of the system.

1.3.5 Entropy Calculation


For a reversible process, Entropy is defined as

∂q = T ds (1.35)

Using Equ. 1.31, the above equation can be written as

T ds = de + p dv (1.36)

T ds = dh − v dp (1.37)

The specific heat at constant pressure can be written as


dh
cp = (1.38)
dT

Substituting the above equation in Equ. 1.37, we get

dT v dp
ds = cp − (1.39)
T T
Substituting the perfect gas equation, pv = RT into the above equaation, we
get

dT dp
ds = cp −R (1.40)
T p

High Speed Aerodynamics (R18A2113) 8 Dr. G. Srinivasan


8
25
Thermodynamics concepts

Integrating the above equation between states 1 and 2, we get


✂ T2  
dT p2
s2 − s 1 = cp − Rln (1.41)
T1 T p1

Using de = cv dT in the above equation, the change in entropy can be expressed


as    
T2 v2
s2 − s1 = cv ln + Rln (1.42)
T1 v1

1.3.6 Isentropic relations


An adiabatic and reversible process is called an isentropic process. For an adiabatic
process, ∂q = 0, and for a reversible process, dsirrev = 0. An isentropic process
is one for which ds = 0, that is the entropy is constant. Important relations for
an isentropic process can be obtained from Equ. 1.41 and Equ. 1.42, by setting
s2 = s1 . Applying s2 = s1 in Equ. 1.41 deduces to
T2 p2
0 = cp ln − R ln
T1 p1
p2 cp T 2
ln = ln
p1 R T1
 cp /R
p2 T2
= (1.43)
p1 T1
From Equ: 1.29,
cp γ
=
R γ−1

and substituting Equ. 1.29 into Equ. 1.43, we get


 γ/(γ−1)
p2 T2
= (1.44)
p1 T1

Similarly from Equ. 1.42,

T2 v2
0 = cv ln + R ln
T1 v1

v2 cv T 2
ln = − ln
v1 R T1
 −cv /R
v2 T2
= (1.45)
v1 T1

From Equ. 1.30


cv 1
=
R γ−1

High Speed Aerodynamics (R18A2113) 9


9
26
Chapter 1. One dimensional compressible flows

Substituting the above equation in Equ. 1.45, we get


 −1/(γ−1)
v2 T2
= (1.46)
v1 T1
The above equation can also written as
 1/(γ−1)
ρ2 T2
= (1.47)
ρ1 T1
Summarizing Equ. 1.44 and Equ: 1.47,
 γ  γ/(γ−1)
p2 ρ2 T2
= = (1.48)
p1 ρ1 T1

The above Equ. 1.48 relates pressure, density and temperature for an isen-
tropic process. This relation is important and is frequently used in the analysis of
compressible flows.

1.4 One-dimensional flow governing equations


Consider the flow through ,I one-dimensional region, a replesented by the shaded
area in Fig. ??. This region may be a normal shock wave. or it may be a region
with heat addition; in either case. the flow properties change as a function of
x as the gas flows through the region. To the left of this region, the flowfield
velocity, pressure, temperature, density, and internal energy are u1 , pl , T1 , p1 , and
el respectively. To the right of this region, the properties have changed, and are
given by u2 , p2 , T2 , p2 , and e2 .

Fig. 1.2: Rectangular control volume for the one-dimensional flow

High Speed Aerodynamics (R18A2113) 10 Dr. G. Srinivasan


10
27
One-dimensional flow governing equations

1.4.0.1 1D continuity equation

✍ ✝
The continuity equation is

− ρ V.dS = ρ dV (1.49)
S ∂t V


For the steady flow, the above equation becomes

ρ V.dS = 0 (1.50)
S

Evaluating the surface integral over the left-hand side, where V and dS are
parallel but in opposite directions, we obtain −ρ1 u1 A; over the right-hand side,
where V and dS are parallel and in the same direction, we obtain −ρ2 u2 A. The
upper and lower horizontal faces of the control volume both contribute nothing to
the surface integral because V and dS are perpendicular to each other. The above
equations becomes
−ρ1 u1 A + ρ2 u2 A = 0

ρ1 u 1 = ρ 2 u 2 (1.51)
Equ. 1.51 is the continuity equation for steady one-dimensional flow.

1.4.0.2 1D momentum equation

✍ ✝ ✝ ✍
The momentum equation in integral form is

∂(ρV )
(ρV.dS)V + dV = ρf dV − p dS (1.52)
s V ∂t V S

✍ ✍
For steady flow, the above equation becomes

(ρV.dS)V = − p dS (1.53)
S S

The above equation is a vector equation. Since we are dealing in 1D flow, we

✍ ✍
will consider only the scalar x component of the equation which is

(ρV.dS)u = − (pdS)x (1.54)


s s

In the above equation, the term (pdS)x is the x component of the vector p dS.
Solving the above equation over the left and right hand sides of the dashed control
volume in Fig. ??, we get

ρ1 (−u1 A)u1 + ρ2 (u2 A)u2 = −(−p1 A + p2 A)

p1 + ρ1 u21 = p2 + ρ2 u22 (1.55)


Equ. 1.55 is the momentum equation for steady 1D flow.

High Speed Aerodynamics (R18A2113) 11


11
28
Chapter 1. One dimensional compressible flows

1.4.0.3 1D Enery Equation

The energy equation in integral form can be written as

✝ ✍ ✝ ✝ ∂
 
V2
 ✍ V2
 
q̇ρdV − pV.dS+ ρ(f.V )dV = ρ e+ dV + ρ e+ V.dS
V S V V ∂t 2 S 2
(1.56)
The first term on the left physically represents the total rate of heat added
to the gas inside the control volume. For simplicity, let us denote this volume
integral by Q̇. The third and fourth terms are zero because of zero body forces

✍ ✍
and steady flow, respectively. Hence, the above equation becomes

V2
 
Q̇ − pV.dS = ρ e+ V.dS (1.57)
S S 2

Evaluating the surface integral over the left and right hand faces of the control
volume in Fig. ??, we get

u21 u22
   
Q̇ − (−p1 u1 A + p2 u2 A) = −ρ1 e1 + u 1 A + ρ2 e 2 + u2 A
2 2

Rearranging, we get

u21 u22
   

+ p1 u1 + ρ1 e1 + u 1 = p 2 u 2 + ρ2 e 2 + u2 (1.58)
A 2 2

Dividing the above equation by Equ. 1.51, i.e. dividing the left hand side by
ρ1 u1 and the right hand side by ρ2 u2 ,

Q̇ p1 u2 p2 u2
+ + e1 + 1 = + e2 + 2 (1.59)
ρ1 u 1 A ρ 1 2 ρ2 2

Considering the 1st term in the above equation, Q̇ is the net rate of heat
(energy/s) added to the control volume, and ρ1 u1 A is the mass flow (mass/s)
through the control volume. Hence, the ratio Q̇/ρ1 u1 A is the heat added per unit
mass, q. Also, the definition of enthalpy, h = e + pv, Hence, the above equation
becomes
u2 u2
h1 + 1 + q = h2 + 2 (1.60)
2 2
Equ. 1.60 is the energy equation for steady 1D flow.

1.4.1 Speed of Sound


Soundwaves are infinitely small pressure disturbances.The speedwith which sound
propagates in a medium is called the speed of sound and is denoted by a. Consider
that the sound wave is moving with velocity a through the gas as shown in fig.
??. As the flow pass through the stationary wave front the flow ahead of it moves
toward the wave at velocity u with pressure, density, and temperature p, ρ, and T ,
respectively, and the flow behind it moves away from the wave at velocity a + da

High Speed Aerodynamics (R18A2113) 12 Dr. G. Srinivasan


12
29
One-dimensional flow governing equations

with pressure p + dp, density ρ + dρ. and temperature T + dT .

Fig. 1.3: Schematic of a sound wave

The flow through the sound wave is one-dimensional. If regions 1 and 2 are in
front of and behind the wave, respectively. Using the continuity equation we can
write,
ρa = (ρ + dρ)(a + da)

ρa = ρa + adρ + ρda + dpda (1.61)


The product of two infinitesimal quantities dp da is very small (2nd order) and
hence they can be ignored in the above equation.
da
a = −ρ (1.62)

Next the momentum equation yields

p + ρa2 = (p + dp) + (ρ + dρ)(a + da)2 (1.63)

Ignoring second order (products of differentials) terms as earlier, we get

dp = −2aρda − a2 dρ (1.64)

Solving the above equation for da gives,

dp + a2 dρ
da = (1.65)
−2aρ

Substituting Equ. 1.65 into Equ. 1.62, gives

dp/dρ + a2
 
a = −rho (1.66)
−2aρ

High Speed Aerodynamics (R18A2113) 13


13
30
Chapter 1. One dimensional compressible flows

Solving the above equation for a2 gives


dp
a2 = (1.67)

The process inside the sound wave must be isentropic. In turn, the rate of
change of pressure with respect to density, dpldp, which appears in Equ. 1.67 is
an isentropic change, and Equ. 1.67 can be written as
 
2 ∂p
a = (1.68)
∂ρ s

Equ. 1.68 is the fundamental expression for the speed of sound. It imples
that the speed of sound is a direct measure of the compressibility of a gas. Using
ρ = 1/v, dp = −dv/v 2 . Hence Equ. 1.68 can be written as
   
2 ∂p ∂p v
a = =− v2 = − (1.69)
∂ρ s ∂v s (1/v)(∂v/∂p)s

Using the definition of isentropic compressibility, τs , we find


s  r
∂p v
a= = (1.70)
∂ρ s τs

The above equation confirms the statement (τs = 0) implies an infinite speed
of sound. For very strong pressure waves, the traveling speed of a disturbance
may be greater than that of sound. The pressure can be expressed as

p = p(ρ)

For an isentropic process of a gas,

pv γ = constant

Now,  
∂p γp
= (1.71)
∂ρ s ρ

Hence Equ. 1.70 can be written as

γp
r
a= (1.72)
ρ

Using the equation of state, p/ρ = RT , the above equation can be written as
p
a= γRT (1.73)

High Speed Aerodynamics (R18A2113) 14 Dr. G. Srinivasan


14
31
One-dimensional flow governing equations

1.4.2 Alternate form of energy equation


Consider Equ. 1.60. Assuming no heat addition, this becomes

u21 u2
h1 + = h2 + 2 (1.74)
2 2
Here points 1 and 2 corresponds to regions 1 and 2 identified in Fig. ??. For
a calorically perfect gas, h = cp T , the above equations becomes

u21 u2
cp T 1 + = cp T 2 + 2 (1.75)
2 2

Using Equ. 1.29, the above equation becomes

γRT1 u21 γRT2 u22


+ = + (1.76)
γ−1 2 γ−1 2

Since a = γRT , the above equation becomes

a21 u2 a22 u2
+ 1 = + 2 (1.77)
γ−1 2 γ−1 2
p
Using a = γP/ρ, the above equation can be written as

u21 u22
   
γ p1 γ p2
+ = + (1.78)
γ−1 ρ1 2 γ−1 ρ2 2

Consider the fluid is brought to Mach 1 at point 2 then flow is sonic in region
2 and suffix 2 is replaced with prefix * in equations representing sonic conditions.
On the region 1, u1 = u and in region 2, u2 = a∗ .

a2 u2 a∗ 2 a∗ 2
+ = + (1.79)
γ−1 2 γ−1 2

a2 u2 γ + 1 ∗2
+ = a (1.80)
γ−1 2 2(γ − 1)

Now consider fluid is brought to rest isentropically i.e u2 = 0, at point 2


representing total conditions denoted by suffix o then, assuming T1 = T and
u1 = u in region 1 and u2 = 0 and T2 = To Equ. 1.75 changes to

u2
cp T + = cp T o (1.81)
2
To u2 u2 u2 γ − 1  u 2
=1+ =1+ =1+ 2 =1+
T 2cp T 2γRT /(γ − 1) 2a /(γ − 1) 2 a
Hence,
To γ−1 2
=a+ M (1.82)
T 2

High Speed Aerodynamics (R18A2113) 15


15
32
Chapter 1. One dimensional compressible flows

Equ. 1.82 gives the ratio of total to static temperature in a flow as a function
of the Mach number M at that point. Furher, for an isentropic process, Equ. 1.48
holds, such that
 γ  γ/(γ−1)
po ρo To
= = (1.83)
p ρ T
Combining Equ. 1.83 and Equ. 1.82, we get
 γ/(γ−1)
po γ−1 2
= 1+ M (1.84)
p 2

 1/(γ−1)
ρo γ−1 2
= 1+ M (1.85)
ρ 2
Equ. 1.84 and Equ. 1.85 gives the ratios of total to static pressure and density,
respectively, at a point in the flow as a function of Mach number M at that point.

1.5 Flow regimes


The Mach number (M) is defined as the ratio of the speed of an object (or of a flow)
to the speed of sound. For instance, in air at room temperature, the speed of sound
is about 340 m/s (1,100 f t/s−1 . M can range from 0 to ∞, but this broad range
falls naturally into several flow regimes. These regimes are subsonic, transonic,
supersonic, hypersonic, and hypervelocity flow. The figure below illustrates the
Mach number ”spectrum” of these flow regimes.
These flow regimes are not chosen arbitrarily, but rather arise naturally from
the strong mathematical background that underlies compressible flow (see the
cited reference textbooks). At very slow flow speeds the speed of sound is so much
faster that it is mathematically ignored, and the Mach number is irrelevant. Once
the speed of the flow approaches the speed of sound, however, the Mach num-
ber becomes all-important, and shock waves begin to appear. Thus the transonic
regime is described by a different (and much more difficult) mathematical treat-
ment. In the supersonic regime the flow is dominated by wave motion at oblique
angles similar to the Mach angle. Above about Mach 5, these wave angles grow so
small that a different mathematical approach is required, defining the Hypersonic
speed regime. Finally, at speeds comparable to that of planetary atmospheric
entry from orbit, in the range of several km/s, the speed of sound is now compar-
atively so slow that it is once again mathematically ignored in the hyper-velocity
regime.

1.6 Normal shock relations


A shock which is perpendicular to flow direction is known as normal shock. It
commonly occurs in supersonic flow changes the upstream supersonic flow to sub-
sonic.
Applying one dimensional fluid flow governing equation to flow across normal
shock by considering a control volume around it as shown in Fig. 1.4. The basic
normal shock equations are directly obtained from equations 1.51, 1.55 and 1.60
with q = 0 (shock wave is adiabatic), we have

High Speed Aerodynamics (R18A2113) 16 Dr. G. Srinivasan


16
33
Normal shock relations

Fig. 1.4: Illustration of flow conditions ahead and behind the normal Shock wave

ρ1 u 1 = ρ2 u 2
p1 + ρ1 u21
= p2 + ρ2 u22
u2 u2
h1 + 1 = h2 + 2
2 2

For a calorically perfect gas, we can immediately add the thermodynamic relations

p = ρRT (1.86)

h = cp T (1.87)
Dividing the momentum equation with continuity equation, we get
p1 p2
− = u2 − u 1 (1.88)
ρ1 u1 ρ2 u2
p
Recalling a = γp/ρ, the above equation becomes

a21 a2
− 2 = u 2 − u1 (1.89)
γu1 γu2

The alternate form of energy equation, using Equ. 1.60 and Equ. 1.80, yields
γ + 1 ∗2 γ − 1 2
a21 = a − u1 (1.90)
2 2
and
γ + 1 ∗2 γ − 1 2
a22 = a − u2 (1.91)
2 2
High Speed Aerodynamics (R18A2113) 17
17
34
Chapter 1. One dimensional compressible flows

Since the flow is adiabatic across the shock wave, a∗ in the above equations
is same constant value. Substituting both of the above alternate form energy
equation into Equ. 1.89, we get

γ + 1 a∗ 2 γ−1 γ + 1 a∗ 2 γ−1
− u1 − + u 2 = u 2 − u1
2 γu1 2γ 2 γu2 2γ
γ+1 γ−1
(u2 − u1 )a∗ 2 + (u2 − u1 ) = u2 − u1
2γu1 u2 2γ
Dividing by (u2 − u1 ),
γ + 1 ∗2 γ − 1
a + =1
2γu1 u2 2γ
Solving for a∗ , gives
a∗ 2 = u1 u2 (1.92)

Equ. 1.92 is called the Prandtl relation, which is useful for intermediate normal
shocks. Example, we can obtain
u1
1= = M1∗ M2∗
a∗

1
M2∗ = (1.93)
M1∗

Here, the flow ahead of a shock wave must be supersonic, i.e. M1 > 1. this
implies that M1∗ > 1. Thus, from Equ. 1.93, M2∗ < 1 and thus M2 < 1, Hence,
the Mach number behind the normal shock is always subsonic
Now, dividing Equ. 1.80 by u2 , we get
2
(a/u)2 1

γ+1 a∗
+ =
γ−1 2 2(γ − 1) u
2
(1/M )2

γ+1 1 1
= −
γ−1 2(γ − 1) M∗ 2

2
M2 = − (γ − 1) (1.94)
[(γ + 1)/M ∗ 2 ]

Equ. 1.94 provides the direct relation between the actual Mach number M and
the characteristc Mach number M ∗ .
Solving for M ∗ in Equ. 1.94 gives

(γ + 1)M 2
M ∗2 = (1.95)
2 + (γ − 1)M 2

Subsituting the Equ. 1.93 into the above equation yields


−1
(γ + 1)M22 (γ + 1)M12

= (1.96)
2 + (γ − 1)M22 2 + (γ − 1)M12

High Speed Aerodynamics (R18A2113) 18 Dr. G. Srinivasan


18
35
Normal shock relations

Solving the above equation for M22

1 + [(γ − 1)/2]M12
M22 = (1.97)
γM12 − (γ − 1)/2

Equ. 1.97 demonstrates that, for a calorically perfect gas with a constant value
of γ, the Mach number behind the shock is a function of only Mach number ahead
of the shock. it also shows that when M1 = 1, then M2 = 1. This is the case of
an infinitely weak normal shock, which is defined as a Mach wave. In contrast.
as M1 increases above 1, the normal shock becomes stronger and M2 becomes
progressively less than 1.
The other flow properties across a normal shock can be obtained by combining
Equ. 1.92 and Equ. 1.51 gives

ρ2 u1 u2 u2
= = 1 = ∗12 = M1∗ 2 (1.98)
ρ1 u2 u2 u1 a

Substituting Equ. 1.95 into the above equation yields

ρ2 u1 (γ + 1)M12
= = (1.99)
ρ1 u2 2 + (γ − 1)M12

To obtain the pressure ratio, the momentum Equ. 1.55 can be written as

p2 − p1 = ρ1 u21 − ρ2 u22 (1.100)

combining the above equation with 1D continuity Equ. 1.51 gives


 
2 u2
p2 − p1 = ρ1 u1 (u1 − u2 ) = ρ1 u1 1 − (1.101)
u1

Dividing the above equation by p1 , and recalling a21 = γp1 /ρ1 , we obtain
 
p2 − p1 2 u2
= γM1 1 − (1.102)
p1 u1

Substitute Equ. 1.99 for u1 /u2 into the above equation,

2 + (γ − 1)M12
 
p2 − p1 2
= γM1 1 − (1.103)
p1 (γ + 1)M12

Simplifying the above equation, we get

p2 2γ
=1+ (M 2 − 1) (1.104)
p1 γ+1 1

The temperature ratio from the equation of state p + ρRT can be written as
  
T2 p2 ρ1
= (1.105)
T1 p1 ρ2

High Speed Aerodynamics (R18A2113) 19


19
36
Chapter 1. One dimensional compressible flows

Fig. 1.5: Illustration of total conditions across a normal shock wave

Substituting Equ. 1.104 and Equ. 1.99 into the above equation, gives

2 + (γ − 1)M12
  
T2 h2 2γ 2
= = 1+ (M − 1) (1.106)
T1 h1 γ+1 1 (γ + 1)M12

Now, we will study on how the total (stagnation) conditions vary across a
normal shock wave.
Fig. 1.5 illustrates the definition of total conditions before and after the shock.
In region 1 ahead of the shock, a fluid element is moving with actual conditions of
M1 , p1 , T1 and s1 . Consider in this region the imaginary state la where the fluid
element has been brought to rest isentropically. Thus, by definition, the pressure
and temperature in state la are the total values pol ,and To1 , respectively. The
entropy at state la is still sl because the stagnating of the fluid element has been
done isentropically. In region 2 behind the shock, a fluid element is moving with
actual conditions of M2 , p2 , T2 , and s2 . Consider in this region the imaginary
state 2a where the fluid element has been brought to rest isentropically. Here,
by definition, the pressure and temperature in state 2a are the total values of
pol and To1 respectively. The entropy at state 2a is still s2 by definition. The
question is now raised how po2 and To2 , behind the shock compare with po2 and
To2 , respectively, ahead of the shock. To answer this question, consider equation

u21 u2
cp T 1 + = cp T 2 + 2
2 2

From the Equ. 1.81, the total temperature is given by

u2
cp T o = cp T +
2
High Speed Aerodynamics (R18A2113) 20 Dr. G. Srinivasan
20
37
Numerical Problems

Hence,
cp To1 = cp To2
and thus
To1 = To2 (1.107)
From Equ. 1.107 it is clear that the total temperature is constant across a
stationary normal shock wave.
Now considering Fig. 1.5 and writing Equ. 1.41 between imaginary states 1a
and 2a
T2a p2a
s2a − s1a = cp ln − Rln (1.108)
T1a p1a
However, s2a = s2 , s1a = s1 , T2a = To = T1a , p2a = po2 , and p1a = po1 . Hence
the above equation becomes
po2
s2 − s1 = −Rln (1.109)
po1
or
po2
(1.110)
po1 = e−(s2 −s1 )/R
From the above equation, we can see that the ratio of total pressure across the
normal shock depends on the M1 only. Also, because s2 > s1 , the above equations
show that po2 < po1 . The total pressure decreases across a shock wave

1.7 Numerical Problems


1. A pressure vessel that has a volume of 10m3 is used to store high-pressure air
for operating a supersonic wind tunnel. If the air pressure and temperature
inside the vessel are 20 atm and 300 K, respectively. a) what is the mass of
air stored in the vessel? b) Total energy of the gas inside the vessel. c) If
the gas in the vessel is heated, the temperature rises to 600 K calculate the
change in entropy of the air inside the vessel.
Solution: The pressure, p = 1atm = 101325 Pa
p 20 × 101325
ρ= = = 23.46 kg/m3
RT 287 × 300

The total mass stored is then

m = vρ = 10 × 23.46 = 234.6 kg

R 287
cv = = = 717.5 J/kg.K
γ−1 1.4 − 1

e = cv T = 717.5 × 300 = 2.153 × 105 J/kg

The total energy is E = me = 234.6 × 2.153 × 105 = 5.05 × 107 J

High Speed Aerodynamics (R18A2113) 21


21
38
Chapter 1. One dimensional compressible flows

From the ideal gas equation pv = RT , we can write


p2 T2 600
= = =2
p1 T1 300

The change in entropy is given by


T2 p2
s2 − s1 = cp ln − Rln
T1 p1

cp = cv + R = 717.5 + 287 = 1004.5 J/kg.K

s2 − s1 = 1004.5ln2 − 287ln2

s2 − s1 = 497.3 J/kg.K

The total change in entropy is

S2 − S1 = m (s2 − s1 ) = 234.6 × 497.3 J/kg.K

2. Calculate the isothermal compressibility for air at a pressure of 0.5 atm.


Solution: The compressibility is defined as
 
1 ∂v
τT = −
v ∂p T

RT
v=
p
Thus  
∂v RT
=−
∂p T ρ2
Hence    p   RT  1
1 ∂v
τT = − =− − 2 = (1.111)
v ∂p T RT p p
We can see that, compressibility for a perfect gas is simply the reciprocal of
the pressure:
1 1
τT = = = 2 atm−1 (1.112)
p 0.5
3. Air flows through a duct. The pressure and temperature at station 1 are 0.7
atm and 300 C, respectively. At a second station, the pressure is 0.5 atm.
Calculate the temperature and density at the second station. Assume the
flow to be isentropic.
Solution: The ideal gas equation is

pv = RT

High Speed Aerodynamics (R18A2113) 22 Dr. G. Srinivasan


22
39
Numerical Problems

p2 T2
=
p1 T1

p2 0.5
T2 = T1 = (30 + 273) = 216.43K
p1 0.7

p2 0.5 × 101325
ρ2 = = = 0.81562
RT2 287 × 216.43
4. At the nose of a missile in flight, the pressure and temperature are 5.6 atm
and 700 K, respectively. Calculate the density and specific volume.
Solution: Given,

po = 5.6 atm = 5.6 × 101325 = 567420P a

to = 700 K
po 567420
ρo = = = 2.8243 kg/m3
RTo 287 × 700
The specific volume is
1 1
v= = = 0.3541
ρo 2.8243

5. At a point in the flow over an F-15 high-performance fighter airplane, the


pressure, temperature, and Mach number are 1890 Ib/ft2, 450 R, and 1.5,
respectively. At this point, calculate To , po , T ∗, p∗, and the flow velocity.
Solution From Table A.1, for M=1.5; po /p = 3.671 and To /T = 1.45, Thus

po = 3.671 × p = 3.671 × 1890 = 6938 lb/f t2

To = 1.45 × T = 1.45 × 450 = 652.5 R


From Table A. 1, for M = 1 .O: p/p∗ = 1.893 and T /T ∗ = 1.2. Keeping
in mind that, for our imaginary process where the flow is slowed down isen-
tropically to Mach I , hence defining p*, the total pressure is constant during
this process; also, where the flow is slowed down adiabatically to Mach 1 ,
hence defining T ∗ , the total temperature is constant. Thus
p∗ po 1
p∗ = p= × 3.671 × 1890 = 3665 lb/f t2
po p 1.893

T ∗ To 1
T∗ = T = × 1.45 × 450 = 543.8 R
To T 1.2
Finally, the flow velocity is
p √
V = M a = M γRT = 1.5 × 1.4 × 1716 × 450 = 1560 f t/s

High Speed Aerodynamics (R18A2113) 23


23
40
Chapter 1. One dimensional compressible flows

6. A normal shock wave is standing in the test section of a supersonic wind


tunnel. Upstream of the wave, M1 = 3, p1 = 0.5 atm, and T1 = 200 K. Find
M2 , p2 , T2 . and u2 downstream of the wave.
Solution: From A.2, for M1 = 3: p2 /p1 = 10.33, T2 /T1 = 2.679 and
M2 = 0.4752. Hence
p2
P2 = p1 = 10.33 × 0.5 = 5.165 atm
p1

T2
T2 = T1 = 2.679 × 200 = 535.8 K
T1
p √
a2 = γRT2 = 1.4 × 287 × 535.8 = 464 m/s

u2 = M2 a2 = 0.4752 × 464 = 220 m/s

7. A blunt nosed missile is flying at Mach 2 at standard sea level. Calculate


the temperature and pressure at the nose of the missile.
Solution: The nose of the missile is a stagnation point. and the streamline
through the stagnation point has also passed through the normal portion of
the bow shock wave. Hence, the temperature and pressure at the nose are
equal to the total temperature and pressure behind a normal shock. Also,
at standard sea level, T1 = 288 K and pl = 1 atm = 101325 pa.
From Table A.1, for M1 = 2: To1 /T1 = 1.8 and po1 /p1 = 7.824. Also, for
adiabatic flow through a normal shock, To2 = To1 , Hence

To1
To2 = To1 = T1 = 1.8 × 288 = 518.4 K
T1
From Table A.2, for M1 = 2: po2 /po1 = 0.7209. Hence
p02 po1
po2 = p1 = 0.7209 × 7.824 × 101325 = 5.72 × 105 pa
po1 p1

High Speed Aerodynamics (R18A2113) 24 Dr. G. Srinivasan


24
41
Chapter 2
Oblique Shock and Expansion
Waves

2.1 Introduction
2.1.1 Waves in supersonic flow
The motion of a body in a fluid at rest creates disturbance in the fluid. The
disturbances, in general, may not be small. The disturbances in the fluid close to
the body are transmitted to other parts of the body and also to the other parts of
the fluid through propagation of the waves. The wave motion is compatible with
the motion of the body. This wave motion determines the pressures on the body
as well as the complete flow field around the body. When the flow is subsonic, it
is not essential to consider the wave motion. Particularly, if the motion is steady
it is easier to study the motion from a reference system where the body is at
rest and the fluid flows over it. However, if the relative wind is supersonic, the
waves can not propagate ahead of the immediate vicinity of the body. Thus, the
wave system travels with the body and is stationary in the reference system that
moves with the body. Limited upstream influence allows the flow to be analyzed
or constructed step by step.
Let us examine the propagation of pressure disturbances created by a moving
object, shown in Fig. 2.1.In a subsonic flow the disturbance waves reach a sta-
tionary observer before the source of disturbance could reach him, as shown in
Fig. 2.1(a) and 2.1(b). But in supersonic flows it takes a considerable amount
of time for an observer to perceive the pressure disturbance, after the source has
passed. This is one of the fundamental differences between subsonic and super-
sonic flows.Therefore, in a subsonic flow the streamlines sense the presence of any
obstacle in the flow field and adjust themselves well ahead of the obstacle and
flow around it smoothly. But in a supersonic flow, the streamlines feel the ob-
stacle only when they hit it. The obstacle acts as a source, and the streamlines
deviate at the Mach cone, as shown in Fig. 2.1(d) Thus, in a supersonic flow, the
disturbance due to an obstacle is sudden and the flow behind the obstacle has to
change abruptly.
In Fig. 2.1(d), it is shown that for supersonic motion of an object there is a
well-defined conical zone in the flow field with the object located at the nose of the
cone, and the disturbance created by the moving object is confined only to the field
included inside the cone.The flow field zone outside the cone does not even feel
the disturbance. For this reason, von-Karman termed the region inside the cone
as the zone of action, and the region outside the cone as the zone of silence.The
lines at which the pressure disturbance is concentrated and which generate the
cone are called Mach waves or Mach lines.The angle between the Mach line and

25
25
42
Chapter 2. Oblique Shock and Expansion Waves

(a)
(b)

(c)

(d)

Fig. 2.1: Propagation of disturbance waves a) V = 0, b) V = a/2, c) V = a, d )


V > a.

the direction of motion of the body is called the Mach angle µ. From Fig. 2.1(d),
we have
at a
sin µ = = (2.1)
Vt V

1
sin µ = (2.2)
M

High Speed Aerodynamics (R18A2113) 26 Dr. G. Srinivasan


26
43
Oblique shock waves

2.2 Oblique shock waves


The normal shock wave, as considered in the previous, is a special case of a more
general family of oblique waves that occur in supersonic flow. Oblique shocks
usually occur when supersonic flow is turned into itself as shown in Fig. 2.2(a)
and Fig. 2.2(b). Stationary shock waves can either be normal or oblique to the flow
direction. Necessary relations between the parameters across an oblique shock can
be obtained directly from the equations of two-dimensional motion. However, the
normal shock results can easily be transformed to obtain the appropriate relations.

(a) (b)

Fig. 2.2: Oblique shock wave produced on a) Concave corner b) Convex corner

The geometry of flow through an oblique shock is given in Fig. 2.3. The
velocity upstream of the shock is V1 , and is horizontal.The corresponding Mach
number is M1 . The oblique shock makes a wave angle β with respect to V1 . Behind
the shock, the flow is deflected toward the shock by the flow-deflection angle θ.
The velocity and Mach number behind the shock are V2 and M2 , respectively.
The components of V1 perpendicular and parallel, respectively, to the shock are
ul and w1 ; the analogous components of V2 are u2 and w2 , as shown in Fig. 2.3.
Therefore, we can consider the normal and tangential Mach numbers ahead of the
shock to be Mn1 , and Mt1 , respectively; similarly, we have Mn2 and Mt2 , behind
the shock.
Consider the control volume drawn between two streamlines through an oblique
shock, as illustrated by the dashed lines at the top of Fig. 2.3. Faces a and d are
parallel to the shock wave. Apply the integral continuity equation 1.49. The time
derivative in Equ. 1.49 is zero. The surface integral evaluated over faces a and d
of the control volume in Fig 2.3 gives

ρ1 u1 A1 = ρ2 u2 A2

Here, A1 = A2 = areas of faces a and d. The faces b, c, e, and f of the control


volume are parallel to the velocity, and hence contribute nothing to the surface
integral (i.e., V.dS = 0 for these faces). Thus, the continuity equation for an
oblique shock wave is
ρ1 u 1 = ρ2 u 2 (2.3)

High Speed Aerodynamics (R18A2113) 27


27
44
Chapter 2. Oblique Shock and Expansion Waves

Fig. 2.3: Oblique shock wave geometry

From the integral form of momentum equation Equ. 1.52, considering the
equation resolved into two components, parallel and perpendicular to the shock
wave in Fig. 2.3 Again, considering steady flow with no body forces, the tangential
component of Equ. 1.52 applied to the control surface in Fig. 2.3 yields (noting
that the tangential component of p dS is zero on faces a and d, and that the
components on b cancel those on f ; similarly with faces c and e).

(ρ1 u1 ) w1 + (ρ2 u2 ) w2 = 0 (2.4)

Deviding Equ. 2.4 by Equ. 2.4, we find that

w1 = w2 (2.5)

The above equation confirms that the tangential component of flow velocity is
preserved across an oblique shock wave
Now, applying the normal component of Equ. 1.52, we get

(−ρ1 u1 ) u1 + (ρ2 u2 ) = − (−p1 + p2 )

p1 + ρ1 u21 = p2 + ρ2 u22 (2.6)

Now considering the integral form of energy equation Equ. 1.56. Applied to
the control volume in Fig. 2.3 for a steady adiabatic flow with no body forces, it

High Speed Aerodynamics (R18A2113) 28 Dr. G. Srinivasan


28
45
Oblique shock waves

yields
V12 V22
   
− (p1 u1 + p2 u2 ) = −ρ1 e1 + u 1 + ρ2 e 2 + u2
2 2
V12 V22
   
h1 + ρ1 u 1 = h 2 + ρ2 u 2 (2.7)
2 2
Dividing the above equation by the continuity Equ. 2.3,

V12 V22
h1 + = h2 + (2.8)
2 2
However, in Fig. 2.3 we can see that V 2 = u2 + w2 and that w1 = w2 . Hence,

V12 − V22 = u21 + w12 − u22 + w22 = u21 − u22


 
(2.9)

Therefore Equ. 2.8 becomes

u21 u2
h1 + = h2 + 2 (2.10)
2 2
Observing Equ. 2.3 , Equ. 2.4 and Equ. 2.10, they are similar to the normal
shock continuity, momentum and energy equation. Therefore, the changes across
an oblique shock wave are governed by the normal component of the free-stream
velocity. Furthermore. precisely the same algebra as applied to the normal shock
equations in Sec.1.6 , when applied to Equ. 2.3 , Equ. 2.4 and Equ. 2.10. will
lead to identical expressions for changes across an oblique shock in terms of the
normal component of the upstream Mach number Mn1 . That is, for an oblique
shock wave with
Mn1 = M1 sin β (2.11)
we find the flow properties around an oblique shock wave as

ρ2 (γ + 1)Mn1 2
= (2.12)
ρ1 (γ − 1)Mn1 2 + 2

p2 2γ
Mn1 2 − 1

=1+ (2.13)
p1 γ+1

Mn1 2 + [2/(γ − 1)]


Mn2 2 = (2.14)
[2γ/(γ − 1)]Mn1 2 − 1
and
T2 p 2 ρ1
= (2.15)
T1 p 1 ρ2
The Mach number behind the oblique shock, M2 , can be found from Mn2 and
the geometry of Fig. 2.3 as
Mn2
M2 = (2.16)
sin(β − θ)

Note: In Sec.1.6, it was found that the changes across a normal shock are a
function of only one component - the upstream Mach number.

High Speed Aerodynamics (R18A2113) 29


29
46
Chapter 2. Oblique Shock and Expansion Waves

From the Equ. 2.11 through Equ. 2.15, the changes across an oblique shock
are a function of two quantitties - both M1 and β. We also see that, in reality
normal shocks are a just a special case of oblique shocks where β = π/2.
The Equ. 2.16 demonstrates that M2 cannot be found until the flow deflection
angle θ is obtained. However, θ is a unique function of M1 and β. From the
geometry in Fig. 2.3,
u1
tan β = (2.17)
w1
and
u2
tan (β − θ) = (2.18)
w2

combining Equ. 2.17 and Equ. 2.18, and noting that w1 = w2 , we get

tan (β − θ) u2
= (2.19)
tan β u1

Combining Equ. 2.19 with Equ. 2.3, Equ. 2.11 and Equ. 2.12, we get

tan (β − θ) 2 + (γ − 1)M12 sin2 β


= (2.20)
tan β (γ + 1)M12 sin2 β

Solving the above equation by conducting some trigonometric manipulation,


the above equation can be expressed as

M12 sin2 β − 1
 
tan θ = 2 cot β (2.21)
M12 (γ + cos 2β) + 2

Equ. 2.21 is called the θ − β − M relation, and specifies θ as


 a unique
 function
π −1 1
of M1 and β. For example θ = 0 at β = 2 and β = sin M1
. Within this
range θ is positive and must therefore have a maximum. For each value of M1 ,
there is a maximum value of θ. For θ < θmax , each value of θ and M corresponds
to two possible solutions, having different values of β. The larger value of β gives
a stronger shock. In the solution with strong shock, the flow becomes subsonic.
With weak shock, the flow remains supersonic except for a small range of value of
θ slightly smaller that θmax .

tan (β − θ) 2 + (γ − 1)M12 sin2 β


= (2.22)
tan β (γ + 1)M12 sin2 β

1 γ + 1 tan(β − θ) γ − 1
2 = −
M12 sin β 2 tan β 2
or
γ + 1 2 sin β sin θ
M12 sin2 β = M1
2 cos(β − θ)
 
γ+1 2
M12 sin2 β ≈ M1 tan β θ F orsmallvaluesof θ (2.23)
2

High Speed Aerodynamics (R18A2113) 30 Dr. G. Srinivasan


30
47
Oblique shock waves

Fig. 2.4: θ − β − M curves. Oblique shock properties

High Speed Aerodynamics (R18A2113) 31


31
48
Chapter 2. Oblique Shock and Expansion Waves

2.2.1 Supersonic flow over a wedge


Any streamline in inviscid flow can be replaced by a solid boundary. Thus the
oblique shock flow provides the solution to supersonic flow in a corner. For a given
values of M1 and θ, the values of β and M2 are determined. Using symmetry, the
flow over a wedge of nose angle 2θ is also obtained. The flow on each side of the
wedge is determined only by the inclination of the surface on that side. Thus, the
wedge need not be symmetric. When the shock waves are attached to the nose,
the upper and lower surfaces are independent since there is no influence on the
flow upstream of the waves.

Fig. 2.5: Oblique shock over wedge and cone

2.2.2 Mach Lines


Assuming that downstream flow remain supersonic (M2 > 1), the wave angle β
decreases with decrease in wedge angle. When θ decreases to zero, β decreases to
the limiting value µ, given by
 
2 2 −1 1
M1 sin µ − 1 = 0 (or) µ = sin (2.24)
M

The jump in the flow quantities is then zero and, hence the strength of the
wave is zero. The flow is continuous without any disturbance. There is nothing
unique about the point where this wave originates; it might be any point in the
flow. The angle µ is simply a characteristic angle associated with M1 . It is called
the ’Mach angle’. The lines of inclination µ which may be drawn at any point in
the flow-field are called ’Mach lines’ or ’Mach waves’.
In nonuniform flow µ varies with M and the Mach lines are curved. At any
point P in a 2-D flow field, there are always two lines which intersect the streamline
at the angle µ. In 3-D flow, the Mach lines or characteristics define a conical
surface with vertex at P. A 2-D supersonic flow is always associated with two
families of Mach lines called right running and left running characteristics and
are often denoted by the labels (+) and (–). Those in the (+) set run to the
right of the streamlines and those in the (–) set run to the left. They are called
‘characteristics’ from the mathematical theory of hyperbolic PDEs. These are
analogous to the two families of characteristics that trace the propagation of 1-D
waves in the x − t plane. Like the characteristics in the x-t plane, Mach lines have

High Speed Aerodynamics (R18A2113) 32 Dr. G. Srinivasan


32
49
Oblique shock waves

(a)
(b)

Fig. 2.6: Mach lines a) degeneration of Mach line as θ approaches 0 b) Left and
Right running Mach lines at an arbitary point in the flow

a distinguished direction, the direction of flow or the direction of increasing time.


This is related to the fact that there is no upstream influence in supersonic flow.

2.2.3 First-order approximation for weak oblique shocks


For small deflection angles θ, the oblique shock equations reduce to very simple
expressions. The approximate relation that can be used to derive others is
 
2 2 γ+1 2
M1 sin β(β − 1) ≈ M1 tan β θ (2.25)
2

For small θ, the value of β is close to either π2 or µ, depending on whether M2 < 1


or M2 > 1. For M2 > 1, the approximation reduces to

γ+1 M2 1
M12 sin2 β − 1 ≈ p 1 θ, as tan β ≈ tan µ = p 2 (2.26)
2 M12 − 1 M1 − 1

The pressure is then approximated to

p2 − p1 △p γM 2
= ≈ p 21 θ (2.27)
p1 p M1 − 1

The changes in other flow quantities are also proportional to the deflection
angle θ. The change of entropy is proportional to the third power of the shock
strength and hence to third power of deflection angle

△S∞θ3 (2.28)

The difference between the wave angle β and the Mach angle µ, to first order
accuracy, can be found as follows, Let β = µ + ǫ, ǫ << µ Hence,

sin β = sin(µ + ǫ) ≈ sin µ + ǫ cos µ (2.29)

High Speed Aerodynamics (R18A2113) 33


33
50
Chapter 2. Oblique Shock and Expansion Waves

By definition,
1
q
sin µ = , cot µ M12 − 1
M1
Hence, q
M1 sin β ≈ 1 + ǫ M12 − 1
q
M12 2
sin β ≈ 1 + 2ǫ M12 − 1 (2.30)

γ+1 M2
q
M12 2
sin β − 1 ≈ 2ǫ M12 − 1 ≈ p 1 θ (2.31)
2 M12 − 1
or
γ + 1 M12
ǫ= θ (2.32)
4 M12 − 1
Hence for a finite deflection angle θ, the direction of the wave differs from the
Mach direction by an amount ǫ, which is of the same order as θ.
The change in flow speed can be obtained as
u2 2

w22 u22 + v 2 v
+1 tan2 (β − θ) + 1 cos2 β
= = = = (2.33)
w12 u2 + v 2 u1 2 tan2 β + 1 cos2 (β − θ)

v
+1

Now,
cos2 β = 1 − sin2 β = (2.34)

similarly, cos2 β can be obtained by replacing ǫ with ǫ − θ. The final result


after dropping all terms of order θ2 and higher
w2 θ △w θ
≈1− p 2 or = −p 2 (2.35)
w1 M1 − 1 w1 M1 − 1

2.3 Supersonic compression by turning


A shock wave passing through a fluid increases the pressure and density of the fluid.
So, shock waves can be used to compress a flow. A simple method for compressing
a supersonic flow is to turn it through an oblique shock by deflecting the wall
through an angle θ. The turn may be subdivided into several segments which
make smaller corners of angle △θ so that compression occurs through successive
weaker oblique shocks. These shocks divide the field near the wall into segments
of uniform flow. In the near wall region each segment of the flow is independent of
the next one and may be constructed step by step proceeding downstream. This
property of limited upstream influence exists as long as the deflection does not
become so great that the flow becomes subsonic. Away from the wall the shocks
tend to intersect each other since they are convergent
For each wave in the multiple shock △p∞△θ and △s∞(△θ)3 . The overall
pressure and entropy changes are

pk − p1 ∞n△θ θ (2.36)

High Speed Aerodynamics (R18A2113) 34 Dr. G. Srinivasan


34
51
Supersonic compression by turning

Fig. 2.7: Supersonic compression by turning

sk − s1 ∞n(△θ)3 (n△θ)(△θ)2 θ(△θ)2 (2.37)


Thus, when the compression is achieved through a large number of weak shocks,
the entropy increase can be reduced significantly compared to a single shock giving
the same net deflection. It decreases as n12 . By contnuing the process of subdivi-
sion, the segments can be made vanishingly small (△θ → 0), and in the limit, the
smooth turn or isentropic compression is obtained.
When the shocks become vanishingly weak, they are almost straight Mach lines.
Each segment of uniform flow becomes vanishingly narrow and finally coincides
with a Mach line. Thus, the flow inclination and Mach number are constant
on each Mach line. Thus, in the limit of smooth flow, the flow velocities and
inclination are continuous, but their derivatives may still be discontinuous. The
approximate expression for the change of speed across a very weak shock

△w θ
= −p 2 (2.38)
w M1 − 1
becomes the differential equation
dw dθ
= −p 2 → θ = θ(M ) (2.39)
w M1 − 1

Due to the convergence of the Mach lines, the change form M1 to M2 on


the streamline b occurs in a shorter distance than on the streamline a. Hence,
the gradients of velocity and temperature on b are higher than those on a. An
intersection of Mach lines would imply an infinitely high gradient for there would
be two values of M at one point. However, this cannot occur since in the region
where Mach lines converge and the gradients become very high the conditions
are no longer isentropic. Before the Mach lines cross a shock wave is developed.
Far from the corner, there would be a simple oblique shock corresponding to M1

High Speed Aerodynamics (R18A2113) 35


35
52
Chapter 2. Oblique Shock and Expansion Waves

and θ. The convergence of Mach lines in a compression is a typical nonlinear


effect: decreasing Mach number and increasing flow inclination both tend to make
successive Mach lines steeper.
If a wall is placed along one of the streamlines, say b, where the gradients are
still small enough for the flow to be isentropic; then an isentropic compression
in a curved channel is obtained. Since this flow is isentropic, it may be reversed
without violating the second law of thermodynamics.

2.4 Supersonic Expansion by Turning


Flow round a ‘concave’ turn, that is turns in which the wall is deflected in to the
flow, undergoes compression through shock wave/Mach lines. Expansion takes
place in a flow over a convex corner. In this case a turn through a single oblique
wave is not possible.

(a)

(b)
(c)

Fig. 2.8: Supersonic expansion by turning

Since v1 = v2 , u2 must be greater than u1 → decrease in entropy. Hence


expansion shocks are not possible.
The non-linear mechanism that steepens a compression produces the opposite
effect in expansion. Instead of being convergent, the Mach lines are divergent.

High Speed Aerodynamics (R18A2113) 36 Dr. G. Srinivasan


36
53
Supersonic Expansion by Turning

Fig. 2.9: Prandtl mayer expansion geometry

Consequently, there is a tendency to decrease gradients. Thus an expansion is


isentropic throughout. The expansion at a corner occurs though a centered wave
defined by a fan of straight Mach lines. The flow up to the corner is uniform at
Mach number M1 and thus the leading Mach wave must be straight at the Mach
angle µ1 . The terminating Mach lines stands at the angle µ2 (corresponding to
M2 ) to downstream wall. This centered wave is called a Prandtl-Meyer expansion
fan.
From Fig. 2.9, for a given M1 , p1 , T1 and θ1 the M2 , p2 and T2 are needed
to be calculated or predicted. The analysis can be started by considering the
infinitesimal changes across a very weak wave produces by an infinitesimally small
flow deflection, dθ. From the law of sines,

V + dV sin(π/2 + µ)
= (2.40)
V sin(π/2 − µ − dθ

However, from trignometric identies,


π  π 
sin + µ = sin − µ = cos µ (2.41)
2 2

π 
sin − µ − dθ = cos (µ + dθ) = cos µ cos dθ − sin µ sin dθ (2.42)
2
Solving above equations two equations into 2.40, we get
dV cos µ
1+ = (2.43)
V cos µ cos dθ − sin µ sin dθ

High Speed Aerodynamics (R18A2113) 37


37
54
Chapter 2. Oblique Shock and Expansion Waves

For small dθ, we can make the small angle assumptions sin dθ ≈ dθ and cos dθ ≈ 1.
The the above equation becomes
dV cos µ 1
1+ = = (2.44)
V cos µ − dθ sin µ 1 − dθ tanµ

Recalling the series expansion (for x < 1),


1
= 1 + x + x2 + x3 + ....... (2.45)
1−x
Equ. 2.44 can be expanded as (ignoring terms of second and higher order)

dV
1+ = 1 + dθ tan µ + ...... (2.46)
V

Thus from the above equation,

dV /V
dθ = (2.47)
tan µ
However, we know that mach angle is
1
µ = sin−1 (2.48)
M
which can be written as
1
tan µ = √ (2.49)
M2 − 1
Substituting the above equation into Equ. 2.47, we get

√ dV
dθ = M2 − 1 (2.50)
V

Equ. 2.50 is the governing differential equation for Prandtl-Meyer flow.


To analyze the entire Prandtl-Meyer expansion in Fig. 2.9, Equ. 2.50 must be
integrated over the complete angle θ2 . Integrating Equ. 2.50 from regions 1 to 2,
✂ θ2 ✂ M2 √ dV
dθ = M2 − 1 (2.51)
θ1 M1 V

The integral on the right hand side can be evaluated after dV /V is obtained
in terms of M as ollows. From the definition of Mach number,

V = Ma

Hence,
ln V = ln M + ln a (2.52)
Differentiating the above equation
dV dM da
= + (2.53)
V M a
High Speed Aerodynamics (R18A2113) 38 Dr. G. Srinivasan
38
55
Simple and Non-simple regions

For a calorically perfect gas the adiabatic energy equation can be written as
 a 2 To γ−1 2
o
= =1+ M (2.54)
a T 2
Solving for a,
 −1/2
γ−1 2
ao 1+ M (2.55)
2
Differentiating the above equation,
   −1
da γ−1 γ−1 2
=− M 1+ M dM (2.56)
a 2 2

Substituting Equ. 2.56 into Equ. 2.53, we obtain


dV 1 dM
= γ−1 (2.57)
V 1+ 2 M M
2

The above equation is desired relation for dV /V in terms of M , substitute it


into Equ. 2.51, ✂ θ2 ✂ M2 √ 2
M − 1 dM
dθ = θ2 − 0 = γ−1 (2.58)
M1 1 + 2 M M
2
θ1

In the above equation, the integral


✂ √
M 2 − 1 dM
ν(M ) = (2.59)
1 + γ−1
2
M2 M

is called the Prandtl-Meyer function, and is given the symbol ν. Performing


the integration, the above equation becomes


r r
γ+1 γ+1
ν(M ) = tan−1 (M 2 − 1) − tan−1 M 2 − 1 (2.60)
γ−1 γ−1

The constant of integration that would ordinarily appear in the above equation
is not important, because it drops out when the Equ. 2.60 is substituted into
Equ. 2.58. For convenience, it is chosen as zero such that ν(M ) = 0 when M = 1.
Finally, we can now write Equ. 2.58 combined with Equ. 2.59, as

ν(M2 ) − ν(M1 ) (2.61)

From Fig. 2.9, Equ. 2.61 and Equ. 2.60 allow the calculation of a Prandtl-
Meyer expansion wave.

2.5 Simple and Non-simple regions


The isentropic compression and expansion waves are distinguished by straight
Mach lines with constant conditions on each one and by the simple relation be-
tween flow deflection and Prandtl-Meyer function. A wave belongs to one of two
families (+ or –), depending on whether the wall that produces it is to the left or

High Speed Aerodynamics (R18A2113) 39


39
56
Chapter 2. Oblique Shock and Expansion Waves

Fig. 2.10: Regions in isentropic supersonic flow

right of flow respectively. In the region where two simple waves of opposite family
interact with each other, the flow is non-simple. The relation between ν and θ
is not the simple one given by ν = ν ± θ. These regions may be treated by the
method of characteristics.

2.6 Regular reflection of oblique shocks from a solid bound-


ary
Consider an oblique shock wave incident on a solid wall, as sketched in Fig. 4.18.
Question: Does the shock wave disappear at the wall, or is it reflected downstream?
If it is reflected, at what angle and what strength? The answer lies in the physical
boundary condition at the wall, where the flow immediately adjacent to the wall
must be parallel to the wall. In Figure 2.11 the flow in region 1 with Mach
number M1 is deflected through an angle θ at point A. This creates an oblique
shock wave that impinges on the upper wall at point B. In region 2 behind this
incident shock, the streamlines are inclined at an angle θ to the upper wall. All
flow conditions in region 2 are uniquely defined by M1 and θ through the oblique
shock relations discussed earlier. At point B, in order for the flow to remain
tangent to the upperwall, the streamlines in region 2 must be deflected downward
through the angle θ. This can only be done by a second shock wave, originating at
B, with sufficient strength to turn the flow through an angle 8, with an upstream
Mach number of M2 . This second shock is called a rejected shock; its strength
is uniquely defined by M2 and θ, yielding the consequent properties in region 3.
Because M2 < M1 . the reflected shock wave is weaker than the incident shock,
and the angle Φ it makes with the upper wall is not equal to β1 (i.e., the reflected
shock wave is not specularly reflected).

2.7 Mach Reflection:


The appearance of subsonic regions in the flow complicates the problem. The
complications are also encountered in shock reflections, when they are too strong
to give the simple reflections. If M2 after the incident shock is lower than the

High Speed Aerodynamics (R18A2113) 40 Dr. G. Srinivasan


40
57
Shock-Expansion Theory

Fig. 2.11: Shock reflection

Fig. 2.12: Mach reflection

detachment Mach number for θ, then no solution with simple oblique wave is
possible. A three-shock Mach reflection appears that satisfies the downstream
conditions.
A normal, or, nearly normal, shock stem that appears near the wall forms
a triple intersection point at O with the incident and reflected shocks. Due to
the difference in entropy on streamlines above and below the triple point, the
streamline that extends downstream from the triple point is a slipstream. The
nearly normal shock is termed ‘shock stem’.
The subsonic region behind the shock stem makes a local description of the
configuration impossible. The triple point solution that occurs in a particular
problem and the location of the triple point are determined by the downstream
conditions which influence the subsonic part of the flow.

2.8 Shock-Expansion Theory


Oblique shock wave and simple isentropic wave relations can be used to analyze
many 2-D supersonic flow problems, particularly for geometries with straight seg-
ments.

High Speed Aerodynamics (R18A2113) 41


41
58
Chapter 2. Oblique Shock and Expansion Waves

Fig. 2.13: Illustration of shock expansion theory for symmetrical diamond section

2.8.1 Diamond-section airfoil:


Consider a diamond section or double-wedge section airfoil with semi-vertex angle
ǫ. Assume the semi-vertex angle to be sufficiently smaller than θmax associated
with the free stream Mach number M1 . An attached oblique shock appears at the
nose that compresses the flow to pressure p2
On the straight portion, downstream of the shock the flow remains uniform at
M2 . The centered expansion at the shoulder expands the flow to pressure p3 and
the trailing edge shock recompresses it to nearly the free stream pressure (p4 ≈ p1
). Hence, an overpressure acts on the forward face and an under-pressure acts on
the rearward face. Since the pressure on the two straight portions is unequal, a
drag force acts on the airfoil. This drag force is given by

D = (p2 − p3 ) cos ǫ.t ≈ (p2 − p3 )t perunitspan (2.62)


Here t is the section thickness at the shoulder. Pressure values p2 and p3 can
be obtained using the shock and expansion relations. This drag exists only in
supersonic flow and is called ‘supersonic wave drag’.

2.8.2 Flat plat at incidence


Consider a flat plate of chord c set at an angle of attack . Due to no upstream
influence, the streamlines ahead of the leading edge are straight and the upper
surface flow is independent of lower surface. The flow on the upper surface turns
at the nose through a centered expansion by the angle α whereas on the lower
side the flow is turned through a compression angle α by an oblique shock. The
reverse happens at the trailing edge.
From the uniform pressures on the two sides, the lift and drag forces are

L = (p3 − p2 ) c cos α (2.63)

D = (p3 − p2 ) c sin α (2.64)


The shock on the lower surface at the nose is weaker than the shock at the
trailing edge on the upper surface (shock at higher Mach number). Hence, the
increase in entropy for flow on the two sides is not same and consequently the
streamline from the trailing edge is a slipstream inclined at a small angle relative

High Speed Aerodynamics (R18A2113) 42 Dr. G. Srinivasan


42
59
Shock-Expansion Theory

Fig. 2.14: Shock expansion theory flat plate

Fig. 2.15: Illustation of shock expansion theory on curved airfoil section

to the free stream.

2.8.3 Curved airfoil section


An attached shock forms at the nose. Subsequently, continuous expansion oc-
curs along the surface. The flow leaves at the trailing edge through an oblique
shock. For the shocks to be attached, it is required that nose and tail be wedge
shaped with half angle less than θmax . Since the flow over the curved wall varies
continuously, no simple expression for lift and drag forces is obtained in this case.
If a larger portion of the flow field is considered, then the shocks and expansion
waves will interact. The expansion fans attenuate the oblique shocks, making them
weak and curved. At large distances they approach asymptotically the free-stream
Mach lines. Due to the interaction the waves will reflect. The reflected wave
system will alter the flow field. In shock-expansion theory, the reflected waves are
neglected. For a diamond airfoil and a lifting flat plate, the reflected waves do not
intercept the airfoil at all. Hence, the shock-expansion results are not affected.

High Speed Aerodynamics (R18A2113) 43


43
60
Chapter 2. Oblique Shock and Expansion Waves

Fig. 2.16: Atteneuation of wave by interaction around diamond section and flat
plate

2.9 Numerical Problems


1. A uniform supersonic stream with M1 = 3.0, p1 = 1atm, and T1 = 288 K
encounters a compression corner which deflects the stream by an angle θ =
2.0. Calculate the shock wave angle, and p2 , T2 , M2 , po2 , and To2 , behind
the shock wave.

Solution: From the θ − β − M chart, for M1 = 3 and θ = 200 :

β = 37.80

Thus
Mn1 = M1 sin β = 3 sin 37.80 = 1.839

From Table A.2, for Mn1 = 1.839:


p2 T2 po2
= 3.783, = 1.562, Mn2 = 0.6078, and = 0.7948.
p1 T1 po1

Hence,
p2
p2 = p1 = 3.783 × 1 = 3.783 atm
p1

T2
T2 = T1 = 1.562 × 288 = 449.9 K
T1

High Speed Aerodynamics (R18A2113) 44 Dr. G. Srinivasan


44
61
Numerical Problems

Mn2 0.6078
M2 = = = 1.988
sin (β − θ) sin 17.8o

From Table A.1, for M1 = 3:


po1 To1
= 36.73 and = 2.8
p1 T1

Hence

po2 po1
po2 = p1 = 0.7948 × 36.73 × 1 = 29.19 atm
po1 p1

To1
To2 = To1 = = 2.8 × 288 = 806.4 K
T1
2. Consider a horizontal supersonic flow at Mach 2.8 with a static pressure
and temperature of 1 atm and 519o R, respectively. This flow passes over a
compression corner with a defection angle of 16o . The oblique shock gener-
ated at the corner propagates into the flow, and is incident on a horizontal
wall. as shown in below figure. Calculate the angle Φ, made by the reflected
shock wave with respect to the wall, and the Mach number, pressure, and
temperature behind the reflected shock.

Solution: From the θ − β − M diagram, β1 = 35o

Mn1 = M1 sin β1 = 2.8 sin 35 = 1.606

From Table A.2, for Mn1 = 1.606:


p2 T2
= 2.82, = 1.388 and Mn2 = 0.6684
p1 T1

Hence
Mn2 0.6684
M2 = = = 2.053
sin (β1 − θ) sin (35 − 16)

From the θ − β − M chart, for M = 2.053 and θ = 16o :

β2 = 45.5o

High Speed Aerodynamics (R18A2113) 45


45
62
Chapter 2. Oblique Shock and Expansion Waves

The component of the Mach number ahead of the reflected shock normal to
the shock is Mn2 , is given by

Mn2 = M2 sin β2 = 2.053 × sin 45.5 = 1.46

From Table A.2, for Mn2 = 1.46:


p3 T3
= 2.32, = 1.294, and Mn3 = 0.715
p2 T2

Where Mn3 is the component of the Mach number behind the reflected shock
normal to the shock. The Mach number in region 3 behind the reflected
shock is given by

Mn3 0.7157
M3 = = = 1.45
sin (β2 − θ) sin (45.5 − 16)
Also,
p3 p2
p3 = p1 = 2.32 × 2.82 × 1 = 6.54 atm
p2 p1

T3 T2
T3 = T1 = 1.294 × 1.388 × 519 = 932 R
T2 T1
Φ = β2 − θ = 45.5 − 16 = 29.5

3. A uniform supersonic stream with M1 = 1.5, p1 = 17001b/f t2, and T1 = 460o


R encounters an expansion comer which deflects the stream by an angle
θ2 = 20o . Calculate M2 ,p2 , T2 ,po2 , To2 ,and the angles the forward and
rearward Mach lines make with respect to the upstream flow direction.

Solution:
From Table A.5, for M1 = 1.5:

ν1 = 11.91o and µ1 = 41.81o

So,
ν2 = ν1 + θ1 = 11.91 + 0 = 31.91o

High Speed Aerodynamics (R18A2113) 46 Dr. G. Srinivasan


46
63
Numerical Problems

From Table A.5, for ν2 = 31.91o :

M2 = 2.207 and µ2 = 26.95o

From Table A.1, for M1 = 1.5:


po1 To1
= 3.671 and = 1.45
p1 T1

From Table A.1, for M2 = 2.207:


po2 To2
= 10.81 and = 1.974
p2 T2

The flow through an expansion wave is isentropic, hence po2 = po1 and
To2 = To1 . Thus,
p2 po2 po1 1
p2 = p1 = × 1 × 3.671 × 1700 = 577.3 lb/f t2
po2 po1 p1 10.81

T2 To2 To1 1
T2 = T1 = × 1 × 1.45 × 460 = 337.9o R
To2 To1 T1 1.975

po1
po2 = po1 = p1 = 3.671 × 1700 = 6241 lb/f t2
p1

To1
To2 = To1 = = 1.45 × 460 = 667o R
T1

Finally,
Angle of f orward M ach Line = µ1 = 41.81o

Angle of rearward M ach Line = µ2 − θ2 = 26.95 − 20 = 6.95o

4. Consider an infinitely thin flat plate at a 5o angle of attack in a Mach 2.6


free stream. Calculate the lift and drag coefficients.
Solution: From Table A.5, for M1 = 2.6:

ν1 = 41.41o

From the prandtl meyer function, Equ. 2.61,

ν2 = ν1 + α = 41.41 + 5 = 46.41

From table A.5, for ν2 = 46.41o :

M2 = 2.85

High Speed Aerodynamics (R18A2113) 47


47
64
Chapter 2. Oblique Shock and Expansion Waves

From Table A.1, for M1 = 2.6:


po1
= 19.95
p1

From Table A.1, for M2 = 2.85:


po2
= 29.29
p2
Hence
p2 p2 po2 po1 1
= = × 1 × 19.95 = 0.681
p1 po2 po1 p1 29.29

From the θ − β − M chart, for M1 = 2.6 and θ = α = 5o :

β = 26.5o

Thus,
Mn1 = M1 sin β = 2.6 × sin 26.5o = 1.16

From Table A.2, for Mn1 = 1.16:


p3
= 1.403.
p1

The lift per unit span L′ is

L′ = (p3 − p2 ) c cos α

The drag per unit span D′ is

D′ = (p3 − p2 ) c sin α

Recalling, the dynamic pressure is given by q = (γ/2) p1 M12 , we have

High Speed Aerodynamics (R18A2113) 48 Dr. G. Srinivasan


48
65
Numerical Problems

 
L′ 2 p3 p2
cl = = − cos α
q1 c γM12 p1 p1

2
cl = (1.403 − 0.681) cos 5o = 0.152
1.4 × 2.62
 
D′ 2 p 3 p2
cd = = − sin α
q1 c γM12 p1 p1

2
cd = 2
(1.403 − 0.681) sin 5o = 0.0133
1.4 × 2.6

High Speed Aerodynamics (R18A2113) 49


49
66
Chapter 2. Oblique Shock and Expansion Waves

High Speed Aerodynamics (R18A2113) 50 Dr. G. Srinivasan


50
67
Chapter 3
Subsonic compressible flow over
airfoil

3.1 Irrotational flow


The vorticity is a point property of the flow, and is given by ∇ × V . Vorticity is
twice the angular velocity of a fluid element, ∇× = 2ω. A flow where ∇ × V 6= 0
throughout is called a rotaional flow. In contrast, a flow where ∇ × V = 0
everywhere is called an irrotational flow.
Irrotational flows are usally simpler to analyze than rotational flows, the ir-
rotationality condition ∇ × V = 0 adds an extra simplification to the general
equations of motion. Consider an irrotational flow in more detail. In cartesian
coordinates, the mathematical statement of irrotational flow is
 
i j k
∂ ∂ ∂ 
∇ × V =  ∂x ∂y ∂z
u v w
     
∂w ∂v ∂w ∂u ∂v ∂u
∇×V =i − −j − +k − =0
∂y ∂z ∂x ∂z ∂x ∂y
For this equality to at every point in the flow,

∂w ∂v ∂w ∂u ∂v ∂u
= = = (3.1)
∂y ∂z ∂x ∂z ∂x ∂y

Equ. 3.1 are called the irrotationality conditions. Now considering the Euler’s
equations without body forces
DV
ρ = −∇p
Dt
For steady flow, the x component of this equation is
∂u ∂u ∂u ∂p
ρu + ρv + ρw =−
∂x ∂y ∂z ∂x
or
∂p ∂u ∂u ∂u
− dx = ρu dx + ρv dx + ρw dx (3.2)
∂x ∂x ∂y ∂z
From Equ. 3.1,
∂u ∂v ∂u ∂w
= and =
∂y ∂x ∂z ∂x

51
51
68
Chapter 3. Subsonic compressible flow over airfoil

Substituting the above relations into Equ. 3.2 we get


∂p ∂u ∂v ∂w
− dx = ρu dx + ρv dx + ρw dx
∂x ∂x ∂x ∂x
or
∂p 1 ∂u2 1 ∂v 2 1 ∂w2
− dx = ρ dx + ρ dx + ρ dx (3.3)
∂x 2 ∂x 2 ∂x 2 ∂x
Similarly by considering the y and z components of Euler’s equation,

∂p 1 ∂u2 1 ∂v 2 1 ∂w2
− dy = ρ dy + ρ dy + ρ dy (3.4)
∂y 2 ∂y 2 ∂y 2 ∂y

∂p 1 ∂u2 1 ∂v 2 1 ∂w2
− dz = ρ dz + ρ dz + ρ dz (3.5)
∂z 2 ∂z 2 ∂z 2 ∂z
Adding all the above three equations, we get

1 ∂V 2 1 ∂V 2 1 ∂V 2
 
∂p ∂p
− dx + dz = ρ dx + ρ dy + ρ dz (3.6)
∂x ∂z 2 ∂x 2 ∂y 2 ∂z

Where V 2 = u2 + v 2 + w2 .
Equ. 3.6 is in the form of perfect differentials, and can be written as
1
−dp = ρd(V 2 ) (3.7)
2
or
dp = −ρV dV (3.8)

3.2 Potential flow equation


The general conservation equations derived in previous chapter are simplified for
the special case of irrotational flow. It allows the separate continuity, momen-
tum and energy equations with the requisite dependent variables ρ, p, T, V etc to
cascade into one governing equation with one dependent variable new defined as
velocity potential.
For irrorational flow, ∇ × V = 0. Hence, we can define a scalar function,
Φ = Φ(x, y, z), such that
V = ∇Φ (3.9)
where Φ is called the velocity potential. In cartesian coordinates, since

V = ui + vj + wk

and
∂Φ ∂Φ ∂Φ
∇Φ = i+ j+ k
∂x ∂y ∂z
then, by comparision,
∂Φ ∂Φ ∂Φ
u= v= w= (3.10)
∂x ∂y ∂z

High Speed Aerodynamics (R18A2113) 52 Dr. G. Srinivasan


52
69
Potential flow equation

Hence, if the velocity potential is known, the velocity can be obtained directly
from the above equations.
As derived next, the velocity potential can be obtained from a single partial
differential equation which physically describes an irrotational flow. In addition.
we will assume steady, isentropic How. For simplicity, we will adopt subscript
notation for derivatives of Φ as follows: ∂Φ/∂x = Φx , ∂Φ/∂y = Φy and ∂Φ/∂z =
Φz .
The continuity equation for steady flow is

∇.(ρV ) = 0

∂(ρu) ∂(ρv) ∂(ρw)


+ + =0
∂x ∂y ∂z
∂ ∂ ∂
ρΦx + ρΦy + ρΦz = 0
∂x ∂y ∂z
∂ρ ∂ρ ∂ρ
ρ(Φxx + Φyy + Φzz ) + Φx + Φy + Φz =0 (3.11)
∂x ∂y ∂z
Since, we are striving for an equation completely in terms of Φ, we eliminate
ρ from Equ. 3.11 by using Euler’s equation in the form of Equ. 3.8, which for an
irrotational flow applies in any direction:
ρ ρ
dp = −ρV dV = − d(V 2 ) = − d(u2 + v 2 + w2 )
2 2

Φ2x + Φ2y + Φ2z


 
dp = −ρd (3.12)
2

From the speed of sound, a2 = (∂p/∂ρ)s , Recalling the flow is isentropic, any
flow change in pressure dp in the flow is followed by a corresponding change in
density, dρ. Hence,  
dp ∂p
= = a2
dρ ∂ρ s
dp
dρ = (3.13)
a2
Combining Equ. 3.12 and Equ. 3.13,

ρ Φ2x + Φ2y + Φ2z


 
dρ = − 2 (3.14)
a 2

Considering changes in the x direction, the above equation yields,

ρ ∂ Φ2x + Φ2y + Φ2z


 
∂ρ
=− 2 (3.15)
∂x a ∂x 2
or
∂ρ ρ
= − 2 (Φx Φxx + Φy Φyx + Φz Φzx ) (3.16)
∂x a
High Speed Aerodynamics (R18A2113) 53
53
70
Chapter 3. Subsonic compressible flow over airfoil

Similarly,
∂ρ ρ
= − 2 (Φx Φxy + Φy Φyy + Φz Φzy ) (3.17)
∂y a
∂ρ ρ
= − 2 (Φx Φxz + Φy Φyz + Φz Φzz ) (3.18)
∂z a
Substituting Equ. 3.16 through Equ. 3.18 into Equ. 3.11, canceling the ρ that
appears in each term, and factoring out the second derivative of Φ, we get
!
Φ2 Φ2y Φ2
   
2Φx Φy 2Φx Φz 2Φy Φz
1− x Φxx + 1− 2 Φyy + 1 − 2z Φzz − Φxy − ΦΦxz − Φyz = 0
2 a a a2 a2 a2
(3.19)
Equ. 3.19 is called the velocity potential equation. Equ. 3.19 is not strictly in
terms of Φ only, the variable speed of a sound a still appears. We need to express
a in terms of Φ. From the energy equation,

ho = constant

For a callorically perfect gas, the energy equation can be expressed as

V2
cp T + = cp T o
2
γRT V2 γRTo
+ =
γ−1 2 γ−1
a2 V2 2
u + v 2 + w2

+
γ−1 2

γ−1 2
a2 = a2o − Φx + Φ2y + Φ2z

(3.20)
2
Since ao is a known constant of the flow, the above equation gives the speed
of sound a as a function of Φ.
Equ. 3.19 coupled with Equ. 3.20 represents a singlee equation for the un-
known variable Φ. Equ. 3.20 represents a combination of continuity, momentum
and energy equations. This leads to a general procedure for the solution of irro-
tational, isentropic flowfields:
❼ Solve for Φ from Equ. 3.19 and Equ. 3.20 for the specified boundary condi-
tions of the given problem.

❼ Calculate u, v and w from Equ. 3.10, Hence V = u2 + v 2 + w2 .

❼ Calculate a from Equ. 3.20.

❼ Calculate M = V /a.

❼ Calculate T , p and ρ from Equ. 1.82, Equ. ?? and Equ. ?? respectively.


Hence, we see that once Φ = Φ(x, y, z) is obtained, the whole flowfield is known.
This shows the importance of Φ. The Equ. 3.19 combined with Equ. 3.20 is a non-
linear partial differential equation. It applies to any irrotational, isentropic flow:

High Speed Aerodynamics (R18A2113) 54 Dr. G. Srinivasan


54
71
Linearized velocity potential

subsonic, transonic, supersonic or hypersonic. It also applies to incompressible


flow, where a → ∞, hence yielding the Laplace’s equation

Φxx + Φyy + Φzz = 0

3.3 Linearized velocity potential


Transport yourself back in time to the year 1940, and imagine that you are an aero-
dynamicist responsible for calculating the lift on the wing of a high-performance
fighter plane. You recognize that the airspeed is high enough so that the wellestab-
lished incompressible flow techniques of the day will give inaccurate results. Com-
pressibility must be taken into account. However. you also recognize that the
governing equations for compressible flow are nonlinear, and that no general so-
lution exists for these equations. Numerical solutions are out of the question
high-speed digital computers are still 15 years in the future. So, what do you do?
The only practical recourse is to seek assumptions regarding the physics of the
flow, which will allow the governing equations to become linear, but which at the
same time do not totally compromise the accuracy of the real problem. In turn.
these linear equations can be attacked by conventional mathematical techniques.
In this context, it is easy to appreciate why linear solutions to flow problems
dominated the history of aerodynamics and gasdynamics up to the middle 1950s.
In modern compressible flow, with the advent of the high-speed computer, the
importance of linearized flow has been relaxed. Linearized solutions now take their
proper role as closed-form analytic solutions useful for explicitly identifying trends
and governing parameters, for highlighting some important physical aspects of the
flow, and for providing practical formulas for the rapid estimation of aerodynamic
forces and pressure distributions. In modern practice, whenever accuracy is desired
the full nonlinear equations are solved numerically on a computer, as described in
aubsequent chapters.
Consider a slender body immersed in a uniform flow. In the uniform flow, the
velocity is V∞ and is oriented in the x direction. In the perturbed flow, the local
velocity is V , where V = Vx i + Vy j + Vz k, where Vx , Vy and Vz are now used to
denote the x, y and z components of velocity, respectively. In this chapter u′ , v ′
and w′ denote perturbations from the uniform flow, such that

V x = V ∞ + u′

Vy = v ′
Vz = w ′
Here, u′ , v ′ and w′ are the perturbation velocities in the x, y and z directions.
In the perturbed flow, the pressure, density and temperature are p, ρ and T ,
respectively. In uniform stream, Vx = V∞ , Vy = 0 and Vz = 0. Also in the unifrom
stream, the pressure, density and temperature are p∞ , ρ∞ and T∞ .
In terms of velocity potential

∇Φ = V = (V∞ + u′ ) i + v ′ j + w′ k

where Φ is now denoted as the ”total velocity potential”. The perturbation

High Speed Aerodynamics (R18A2113) 55


55
72
Chapter 3. Subsonic compressible flow over airfoil

velocity potential φ is defined as


∂φ ∂φ ∂φ
= u′ = v′ = w′
∂x ∂y ∂z
Then,
Φ(x, y, z) = V∞ x + φ(x, y, z)
where
∂Φ ∂φ
V x = V ∞ + u′ = = V∞ +
∂x ∂x
∂Φ ∂φ
Vy = v ′ = =
∂y ∂y
∂Φ ∂φ
Vz = w ′ = =
∂z ∂z
Also,
∂ 2φ
Φxx =
∂x2
∂ 2φ
Φyy = 2
∂y
∂ 2φ
Φzz =
∂z 2
Considering the velocity potential Equ. 3.19. Multiplying it by a2 and sustituting
Φ = V∞ x + φ, we get
"  2 # 2 "  2 2 # 2 "  2 2 # 2
∂φ ∂ φ ∂φ ∂ φ ∂φ ∂ φ
a2 − V∞ + 2
+ a2 − 2
+ a2 −
∂x ∂x ∂y ∂y ∂z ∂z 2
∂φ ∂φ ∂ 2 φ ∂φ ∂φ ∂ 2 φ ∂φ ∂φ ∂ 2 φ
   
−2 V∞ + − 2 V∞ + −2 =0
∂x ∂y ∂x ∂y ∂x ∂z ∂x ∂z ∂y ∂z ∂y ∂z
(3.21)

The above equation is called the perturbation-velocity potential equation. Re-


casting the above equation in terms of velocities
 ∂u′ ∂v ′ ∂w′
a2 − (V∞ + u′ )2 + [a2 − v ′2 ] + [a2 − w′2 ]

∂x ∂y ∂z
(3.22)
∂u ′
∂u ′
∂v ′
−2(V∞ + u′ )v ′ − 2(V∞ + u′ )w′ − 2v ′ w′ =0
∂y ∂z ∂z

Since the total enthalpy is constant throughout the flow,

V∞2 V2 (V∞ + u′ )2 + v ′2 + w′2


h∞ + =h+ =h+
2 2 2
or
a2∞ V2 a2 (V∞ + u′ )2 + v ′2 + w′2
+ ∞ = +
γ−1 2 γ−1 2
γ−1
a2 = a2∞ − (2u′ V∞ + u′2 + v ′2 + w′2 ) (3.23)
2
High Speed Aerodynamics (R18A2113) 56 Dr. G. Srinivasan
56
73
Linearized velocity potential

Substituting Equ. 3.23 into Equ. 3.22 and rearranging,

2 ∂u ∂v ′ ∂w′

(1 − M∞ ) + +
∂x ∂y ∂z
      
2 u ′
γ+1 u ′2
γ−1 v +w
′2 ′2
∂u′
= M∞ (γ + 1) + +
V∞ 2 V∞2 2 V∞2 ∂x
       ′
2 u ′
γ+1 v ′2
γ−1 w +u
′2 ′2
∂v
+M∞ (γ + 1) + 2
+ 2
V∞ 2 V 2 V ∂y
   ∞    ′2 ∞ ′2 
2 u ′
γ+1 w ′2
γ−1 u +v ∂w′
+M∞ (γ + 1) + +
V∞ 2 V∞2 2 V∞2 ∂z
 ′          
2 v u ′
∂u ′
∂v ′
w ′
u ′
∂u ′
∂w ′
u w ∂w
′ ′ ′
∂v ′
+M∞ 1+ + + 1+ + + 2 +
V∞ V∞ ∂y ∂x V∞ V∞ ∂z ∂x V∞ ∂y ∂z
(3.24)

Equ. 3.24 is still an exact equation for irrotational, isentropic flow. It is simply
an expanded form of the perturbation-velocity potential equation. Note that the
left-hand side of Equ. 3.24 is linear, but the right-hand side is not. Also recall
that we have not said anything about the size of the perturbation velocities u′ , v ′
and w′ . Now, assume u′ , v ′ and w′ are small compared to V∞ :
 ′  2  ′ 2  ′ 2
u′ v ′ w′ u v w
, and << 1 , and <<< 1
V∞ V∞ V∞ V∞ V∞ V∞

❼ For 0 ≤ M∞ ≤ 0.8 and M∞ ≥ 1.2, the magnitude of


  ′
2 u′ ∂u
M∞ (γ + 1) + .........
V∞ ∂x

is small in comparision to the magnitude of

2 ∂u′
(1 − M∞ )
∂x
Thus ignore the former term

❼ For M∞ ≤ 5 (approx),
  ′
2 u′ ∂v
M∞ (γ − 1) + ........
V∞ ∂y

is small in comparision to ∂v ′ /∂y,


 
2 u′ ∂w′
M∞ (γ − 1) + ............
V∞ ∂z

is small in comparision to ∂w′ /∂z, and


 ′   ′  
2 v u′ ∂u ∂v ′
M∞ 1+ + + ......... ≈ 0
V∞ V∞ ∂y ∂x

High Speed Aerodynamics (R18A2113) 57


57
74
Chapter 3. Subsonic compressible flow over airfoil

Thus, ignore these terms in comparision to the left hand side of Equ. 3.24,
with these order-of-magnitude comparisions, Equ. 3.24 reduces to

2 ∂u′ ∂v ′ ∂w′
(1 − M∞ ) + + (3.25)
∂x ∂y ∂z
or in terms of the perturbation velocity potential,

2 ∂ 2φ ∂ 2φ ∂ 2φ
(1 − M∞ ) + + 2 (3.26)
∂x2 ∂y 2 ∂z

The Equ. 3.26 is the Linearized velocity potential equation.

3.4 Linearized pressure Coefficient


The pressure co-efficient cp is defined as
p − p∞
Cp = 1
ρ V2
2 ∞ ∞

where p is the local pressure, and p∞ , ρ∞ and V∞ are the pressure, density and
velocity respectively. An alternative form of the pressure coefficient, convenient
for compressible flow, can be obtained as follows:

1 1 γp∞ γ V2 γ
ρ∞ V∞2 = ρ∞ V∞2 = p∞ 2∞ = p∞ M∞
2
(3.27)
2 2 γp∞ 2 a∞ 2

Hence Equ. 3.4 becomes

p − p∞ p∞ (p/p∞ − 1)
Cp = 2
= 2
(3.28)
(γ/2)p∞ M∞ (γ/2)p∞ M∞

Hence,
 
2 p
Cp = 2
−1 (3.29)
γM∞ p∞
We now proceed to obtain approx expression for Cp that is constant with
linearized theory. Since the total enthalpy is constant,

V2 V2
h+ = h∞ + ∞
2 2
For a calorically perfect gas, the above equation becomes

V2 V∞2
T+ = T∞ + (3.30)
2cp 2cp

V∞2 − V 2 V∞2 − V 2
T − T∞ = = (3.31)
2cp 2γR/(γ − 1)
T γ − 1 V∞2 − V 2 γ − 1 V∞2 − V 2
−1= = (3.32)
T∞ 2 γRT∞ 2 a2∞

High Speed Aerodynamics (R18A2113) 58 Dr. G. Srinivasan


58
75
Prandtl-Glauert compressibility corrections

Since
V 2 = (V∞ + u′ )2 + v ′2 + w′2
Equ. 3.32 becomes
T γ−1
=1− (2u′ V∞ + u′2 + v ′2 + w′2 ) (3.33)
T∞ 2a2∞

Since the flow is isentropic, p/p∞ = (T /T∞ )γ/(γ−1) , and the above equation
becomes
 γ/(γ−1)
 
p γ−1 ′ ′2 ′2 ′2
= 1− 2u V∞ + u + v + w (3.34)
p∞ 2a2∞
or   γ/(γ−1)
p γ − 1 2 2u′ u′2 + v ′2 + w′2
= 1− M∞ + (3.35)
p∞ 2 V∞ V∞2
Equ. 3.35 is an exact expression. However, considering small perturbations:
u /V∞ << 1 : u′2 /V∞2 , and w′2 /V∞2 <<< 1. Hence the above equation is of the

form
p
= (1 − ǫ)γ/(γ−1)
p∞
where ǫ is small. Hence from the binomial expresion, neglecting higher order
terms,
p γ
=1− ǫ + ................ (3.36)
p∞ γ−1
Thus Equ. 3.35 can be expressed in the above form of equation and neglecting
higher order terms:
 
p γ 2 2u′ u′2 + v ′2 + w′2
= 1 − M∞ + + ........... (3.37)
p∞ 2 V∞ V∞2

Substitute the aboveequation into Equ. 3.4 gives


   
2 γ 2 2u′ u′2 + v ′2 + w′2
Cp = 2
1 − M∞ + + ........... − 1 (3.38)
γM∞ 2 V∞ V∞2

2u′ u′2 + v ′2 + w′2


Cp = − − + .............. (3.39)
V∞ V∞2
Negelecting small terms
2u′
Cp = − (3.40)
V∞

3.5 Prandtl-Glauert compressibility corrections


Consider the compressible subsonic flow over a thin airfoil at small angle of attack
(hence small perturbations). The usual inviscid flow boundary condition must
hold at the surface, i.e., the flow velocity must be tangent to the surface.
df v
= = tan θ
dx V∞ + u

High Speed Aerodynamics (R18A2113) 59


59
76
Chapter 3. Subsonic compressible flow over airfoil

Fig. 3.1: Definition of critical Mach number. Point A is the location of minimum
pressure on the top surface of the airfoil.

Cp o
Cp = p (3.41)
1 − M∞2

CLo
CL = p (3.42)
2
1 − M∞

CM o
CM = p (3.43)
1 − M∞2

3.6 Critical Mach number

At high-subsonic flight speeds, the local speed of the airflow can reach the speed of
sound where the flow accelerates around the aircraft body and wings. The speed at
which this development occurs varies from aircraft to aircraft and is known as the
critical Mach number. The resulting shock waves formed at these points of sonic
flow can greatly reduce power, which is experienced by the aircraft as a sudden
and very powerful drag, called wave drag. To reduce the number and power of
these shock waves, an aerodynamic shape should change in cross sectional area as
smoothly as possible.
The critical Mach number can be evaluated as follows

!γ/(γ−1)
pA 1 + γ−1
2
2
M∞
= (3.44)
pB 1 + γ−1
2
MA2
 !γ/(γ−1) 
γ−1 2
2  1+ 2
M∞
Cp A = 2 γ−1 − 1 (3.45)
γM∞ 1+ 2
MA2

High Speed Aerodynamics (R18A2113) 60 Dr. G. Srinivasan


60
77
Drag divergence Mach number

Setting MA = 1, M∞ = Mcr , and Cp = Cp cr , we get


 !γ/(γ−1) 
2  1 + γ−1
2
Mcr
Cp cr = 2 γ−1 − 1 (3.46)
γMcr 1+ 2

3.7 Drag divergence Mach number


The drag divergence Mach number (not to be confused with critical Mach number)
is the Mach number at which the aerodynamic drag on an airfoil or airframe begins
to increase rapidly as the Mach number continues to increase. This increase can
cause the drag coefficient to rise to more than ten times its low speed value.
The value of the drag divergence Mach number is typically greater than 0.6;
therefore it is a transonic effect. The drag divergence Mach number is usually
close to, and always greater than, the critical Mach number. Generally, the drag
coefficient peaks at Mach 1.0 and begins to decrease again after the transition into
the supersonic regime above approximately Mach 1.2.
The large increase in drag is caused by the formation of a shock wave on the
upper surface of the airfoil, which can induce flow separation and adverse pressure
gradients on the aft portion of the wing. This effect requires that aircraft intended
to fly at supersonic speeds have a large amount of thrust. In early development
of transonic and supersonic aircraft, a steep dive was often used to provide extra
acceleration through the high drag region around Mach 1.0. This steep increase in
drag gave rise to the popular false notion of an unbreakable sound barrier, because
it seemed that no aircraft technology in the foreseeable future would have enough
propulsive force or control authority to overcome it. Indeed, one of the popular
analytical methods for calculating drag at high speeds, the Prandtl-Glauert rule,
predicts an infinite amount of drag at Mach 1.0. Two of the important technologi-
cal advancements that arose out of attempts to conquer the sound barrier were the
Whitcomb area rule and the supercritical airfoil. A supercritical airfoil is shaped
specifically to make the drag divergence Mach number as high as possible, allow-
ing aircraft to fly with relatively lower drag at high subsonic and low transonic
speeds.

High Speed Aerodynamics (R18A2113) 61


61
78
Chapter 3. Subsonic compressible flow over airfoil

3.8 Area rule


The Whitcomb area rule, also called the transonic area rule, is a design technique
used to reduce an aircraft’s drag at transonic and supersonic speeds, particularly
between Mach 0.75 and 1.2. The area rule says that two airplanes with the same
longitudinal cross-sectional area distribution have the same wave drag, indepen-
dent of how the area is distributed laterally (i.e. in the fuselage or in the wing).
Furthermore, to avoid the formation of strong shock waves, this total area distri-
bution must be smooth. As a result, aircraft have to be carefully arranged so that
at the location of the wing, the fuselage is narrowed or ”waisted”, so that the total
area does not change much. Similar but less pronounced fuselage waisting is used
at the location of a bubble canopy and perhaps the tail surfaces.

3.9 Supercritical airfoil


A supercritical airfoil is an airfoil designed, primarily, to delay the onset of wave
drag in the transonic speed range. Supercritical airfoils are characterized by their
flattened upper surface, highly cambered (”downward-curved”) aft section, and
larger leading edge radius compared with NACA 6-series laminar airfoil shapes.
Standard wing shapes are designed to create lower pressure over the top of the
wing. The camber of the wing determines how much the air accelerates around
the wing. As the speed of the aircraft approaches the speed of sound the air
accelerating around the wing will reach Mach 1 and shockwaves will begin to
form. The formation of these shockwaves causes wave drag. Supercritical airfoils
are designed to minimize this effect by flattening the upper surface of the wing.

3.10 Numerical Problems


1. The low-speed lift coefficient for an NACA 2412 airfoil at an angle of attack
of 40 is 0.65. Using the Prandtl-Glauert rule, calculate the lift coefficient for
M∞ = 0.7.
Solution: Given
α = 4o
M∞ = 0.65

We know,
cl,0 0.65
cl = p =√ = 1.275
1 − M∞2 1 − 0.72

High Speed Aerodynamics (R18A2113) 62 Dr. G. Srinivasan


62
79
Numerical Problems

Fig. 3.2: Conventional (1) and supercritical (2) airfoils at identical free stream
Mach number. Illustrated are: A, Supersonic flow region; B, Shock wave; C, Area
of separated flow. The supersonic flow over a supercritical airfoil terminates in a
weaker shock, thereby postponing shock-induced boundary layer separation.

High Speed Aerodynamics (R18A2113) 63


63
80
Chapter 3. Subsonic compressible flow over airfoil

Fig. 3.3: Supercritical airfoil Mach Number/pressure coefficient diagram. The


sudden increase in pressure coefficient at midchord is due to the shock. (y-
axis:Mach number (or pressure coefficient, negative up); x-axis: position along
chord, leading edge left)

High Speed Aerodynamics (R18A2113) 64 Dr. G. Srinivasan


64
81
Chapter 4
Linearized Supersonic Flows and
Hypersonic Flows

4.1 Linearized supersonic pressure co-efficient


The linearized perturbation-velocity potential equation for two dimensional flows
derived in the previous chapter is repeated here which is of the form

β 2 φxx + φyy = 0 (4.1)


p
For subsonic flow, where β = 2 , and the form of
1 − M∞

λ2 φxx − φyy = 0 (4.2)


p
for supersonic flow, where λ = M∞ 2 − 1. The difference between Eqs 4.1 and 4.2

is fundamental, for they are elliptic and hyperbolic partial differential equations,
respectively.
Consider the supersonic flow over a body or surface which introduces small
changes in the flowfield, i.e., flow over a thin airfoil, over a mildly wavy wall, or
over a small hump in a surface as sketched in Figure 4.1.

Fig. 4.1: Linearized supersonic flow over a bump

The Equ. 4.2, which governs the flow is of the form of the classical wave
equation. Its general solution is

φ = f (x − λy) + g (x + λy) (4.3)

which can be verified by direct substitution into Equ. 4.2. Examining the
particular solution where g =0, and hence φ = f (x − λy), we see that lines of

65
65
82
Chapter 4. Linearized Supersonic Flows and Hypersonic Flows

constant φ correspond to x − λy = const, or


dy 1 1
= =p (4.4)
dx λ 2 −1
M∞

Recalling that the Mach ang;e µ = arcsin (1/M∞ ), we see that lines of constant φ
are the family of left-running Mach lines as sketched in upper half of Figure 4.1.
In turn if f = 0 in Equ.4.1, then lines of constant φ are the family of right-running
Mach lines shown in lower half of Figure. 4.1.
Returning to Equ. 4.2, letting g = 0, we have

φ = f (x − λy) (4.5)

Hence,
∂φ
u′ = = f′ (4.6)
∂x
and
∂φ
v′ = = −λf ′ (4.7)
∂y

where f ′ reprents the derivative with respect to the argument, (x − λy). Comb-
ing the above two equations, we get

v′
u′ = − (4.8)
λ

The slope of the left running Mach waves can be calculated as

dy v′
tan θ = = (4.9)
dx V ∞ + u′

For small perturbations, u′ << V∞ and tan θ ≈ θ. Hence, the above equation
becomes
v ′ = V∞ θ (4.10)

Substituting Equ. 4.10 into 4.8,


V∞ θ
u′ = − (4.11)
λ

Recalling subsonic linearized pressure coefficient

2u′
Cp = − (4.12)
V∞

Therefore from Equs 4.12 and 4.11, the pressure coefficient on the surface is

2u′ 2θ
Cp = − = (4.13)
V∞ ∞

High Speed Aerodynamics (R18A2113) 66 Dr. G. Srinivasan


66
83
Hypersonic Flows

or


Cp = − p (4.14)
M∞2 −1

Equation 4.14 is an important result. It is the linearized supersonic surface


pressure coefficient, and it states that Cp is directly proportional to the local
surface inclination with respect to the free stream. It holds for any slender two-
dimensional shape. For example, consider the biconvex airfoil shown in Fig. 4.2.
At two arbitrary points A and B on the top surface,

2θA 2θB
CpA = p and CpB = − p (4.15)
M∞2 −1 M∞2 −1

respectively.

Fig. 4.2: Schematic of the lineari edp ressure coefficient over a biconvex airfoil

The contrast between subsonic and supersonic flows can be seen by comparing
Equs. 4.12 and 4.14. In subsonic flow, Equ. 4.12 shows that Cp increases when
M∞ increases. However, for supersonic flow, Equ. 4.14 shows that Cp decreases
when

4.2 Hypersonic Flows


In general words the flows with Mach number higher than 5 are categorized under
hypersonic flows. However, hypersonic flow is best defined as that regime where
certain physical flow phenornena become progressively more important as the
Mach number is increased to higher values. In some cases, one or more of these
phenomena may become important above Mach 3, whereas in other cases they
may not be compelling until Mach 7 or higher.

4.2.1 Qualitative aspects of Hypersonic flows


The qualities of hypersonic flows can be listed as below:

1. Thin Shock layers

High Speed Aerodynamics (R18A2113) 67


67
84
Chapter 4. Linearized Supersonic Flows and Hypersonic Flows

Fig. 4.3: Variation of the linearized preswre coefficient with Mach number.

2. Entropy layer

3. Viscous interaction

4. High-Temperature flows

4.2.1.1 Thin shock layer

For flow over a hypersonic body, the distance between the body and the shock
wave is very small. The flowfield between the shock wave and the body is defined
as the shock layer, and for hypersonic speeds this shock layer is usually quite
thin. For example, consider the Mach 36 flow of a calorically perfect gas with
a ratio of specific heats, γ = cp /cv = 1.4, over a wedge of 15o half-angle. From
standard oblique shock theory the shock wave angle will be only 18o as shown in
Figure. 15.3. If high-temperature, chemically reacting effects are included, the
shock wave angle will be even smaller. Clearly, this shock layer is thin. It is a
basic characteristic of hypersonic flows.

Fig. 4.4: Illustration of a thin shock layer at hypersonic Mach numbers

High Speed Aerodynamics (R18A2113) 68 Dr. G. Srinivasan


68
85
Hypersonic Flows

4.2.1.2 Entropy Layer


Consider the wedge shown in Figure. 4.4, except now with a blunt nose, as sketched
in Figure. 4.5. 15.4. At hypersonic Mach numbers, the shock layer over the
blunt nose is also very thin, with a small shock detachment distance d. In the
nose region, the shock wave is highly curved. Recall that the entropy of the
flow increases across a shock wave, and the stronger the shock, the larger the
entropy increase. A streamline passing through the strong, nearly normal portion
of the curved shock near the centerline of the flow will experience a larger entropy
increase than a neighboring streamline which passes through a weaker portion
of the shock further away from the centerline. Hence, there are strong entropy
gradients generated in the nose region; this ”entropy layer” flows downstream,
and essentially wets the body for large distances from the nose, as shown in Fig.
4.5. The boundary layer along the surface grows inside this entropy layer, and is
affected by it.

Fig. 4.5: Illustration of the entropy layer of a blunt-nosed slender body at


hypersonic speeds

4.2.1.3 Viscous Interaction


Consider a boundary layer on a flat plate in a hypersonic flow, as sketched in
Fig. 15.5. A high-velocity, hypersonic flow contains a large amount of kinetic
energy; when this flow is slowed by viscous effects within the boundary layer, the
lost kinetic energy is transformed (in part) into internal energy of the gas-this is
called viscous dissipation. In turn, the temperature increases within the boundary
layer; a typical temperature profile within the boundary layer is also sketched in
Figure. 4.6. The characteristics of hypersonic boundary layers are dominated
by such temperature increases. For example, the viscosity coefficient increases
with temperature, and this by itself will make the boundary layer thicker. In
addition, because the pressure p is constant in the normal direction through a
boundary layer, the increase in temperature T results in a decrease in density
ρ through the equation of state p = ρ/RT . In order to pass the required mass
flow through the boundary layer at reduced density, the boundary layer thickness
must be larger. Both of these phenomena combine to make hypersonic boundary
layers grow more rapidly than at slower speeds. Indeed, the flat plate compressible

High Speed Aerodynamics (R18A2113) 69


69
86
Chapter 4. Linearized Supersonic Flows and Hypersonic Flows

laminar boundary layer thickness δ grows essentially as

M2
δ∝√ ∞ (4.16)
Rex

Fig. 4.6: Schematic of a temperature profile in a hypersonic boundary layer

4.2.1.4 High-Temperature Flows

As discussed earlier, the kinetic energy of a high-speed, hypersonic flow is dissi-


pated by the influence of friction within a boundary layer. The extreme viscous
dissipation that occurs within hypersonic boundary layers can create very high
temperatures-high enough to excite vibrational energy internally within molecules,
and to cause dissociation and even ionization within the gas. If the surface of a
hypersonic vehicle is protected by an ablative heat shield, the products of ablation
are also present in the boundary layer, giving rise to complex hydrocarbon chem-
ical reactions. On both accounts, we see that the surface of a hypersonic vehicle
can be wetted by a chemically reacting boundary layer.

Fig. 4.7: Illustration of a high-temperature shock layer on a blunt body moving


at hypersonic speeds

High Speed Aerodynamics (R18A2113) 70 Dr. G. Srinivasan


70
87
Hypersonic Shock wave Relations

4.3 Hypersonic Shock wave Relations


The shock wave relations in the hypersonic flows are given in the figure. 4.8.
Although the shock wave relations for hypersonic and supersonic flows are same,
the rather different wave relations seen in figure 4.8 for hypersonic flows is basically
through the assumption that at very hypersonic sppeds, M1 → ∞ and γ → 1.

Fig. 4.8: Hypersonic Shock wave relations

For a completed detail about the derivation of the above relation refer to section
15.3 in ”Modern compressible flow book”.

4.4 Newtonian Theory


Here, we will obtain a simple expression for the pressure distribution over the
surface of a blunt body. In Propositions 34 and 35 of his Principia, Isaac Newton
considered that the force of impact between a uniform stream of particles and
a surface is obtained from the loss of momentum of the particles normal to the
surface. For example, consider a stream of particles with velocity V∞ , incident on
a flat surface inclined at the angle θ with respect to the velocity, as shown in Fig.
4.9a. Upon impact with the surface, Newton assumed that the normal momentum
of the particles is transferred to the surface, whereas the tangential momentum
is preserved. Hence, after collision with the surface, the particles move along the
surface, as sketched in Fig. 4.9a. The change in normal velocity is simply V∞ sin θ.
Now consider Fig. 4.9b. The mass flux of particles incident on a surface of area
A is ρV∞ A sin θ.
Hence, the time rate of change of momentum of this mass flux, from Newton’s
reasoning, is

M ass f lux × velocity change (4.17)


or
(ρV∞ A sin θ) (V∞ sin θ) = ρV∞2 A sin2 θ (4.18)
And in turn, from Newton’s second law, this time rate of change of momentum
is equal to the force F on the surface:

F = ρV∞2 A sin2 θ (4.19)

High Speed Aerodynamics (R18A2113) 71


71
88
Chapter 4. Linearized Supersonic Flows and Hypersonic Flows

Fig. 4.9: Schematic for Newtonian Impact theory

In turn, the pressure is force per unit area, Which from the above equation is
F
= ρV∞2 sin2 θ (4.20)
A
Newton assumed the stream of particles in Fig. 4.9b to be linear, i.e., he
assumed that the individual particles do not interact with each other, and have
no random motion. Since modern science recognizes that static pressure is due to
the random motion of the particles, and since Eq. 4.20 considers only the linear,
directed motion of the particles, the value of F/A in Eq. 4.20 must be interpreted
as the pressure difference above static pressure, namely, F/A = p − p∞ . Therefore,
from Eq. 4.20, and recalling from the definition of the pressure coefficient, Cp =
(p − p∞ ) 12 ρV∞2 , we have
p − p∞ = ρV∞2 sin2 θ (4.21)

p − p∞
1 = 2 sin2 θ (4.22)
2
ρV 2

Cp = 2 sin2 θ (4.23)
Equation 4.23 is te newtonian ”sine-squared” law for preesure distribution on
a surface inclined at an angle θ with respect to the freestream.

4.5 A Local surface Inclination Method: Modified Newto-


nian Theory
Lineari eds upersonic theory leads to a simple relation for the surface pressure
coefficient, namely Equ. 4.14, repeated here:

Cp = p (4.24)
M∞ 2

Note from Eq. 4.14 that Cp , depends only on θ, the local surface inclination
angle defined by the angle between a line tangent to the surface and the free-stream
direction. In this sense, Eq. 4.14 is an example of a ”local surface inclination
method” for linearized supersonic flow. Question: Do any local surface inclination
methods exist for hypersonic flow? The answer is yes. The oldest and most widely
used of the hypersonic local surface inclination methods is newtonian theory.

High Speed Aerodynamics (R18A2113) 72 Dr. G. Srinivasan


72
89
A Local surface Inclination Method: Modified Newtonian Theory

recall the exact oblique shock relation for Cp from Figure. 4.8,

4
Cp = sin2 β (4.25)
γ+1

Now consider the limit that as M → ∞, γ → 1, we can see the above equation
becomes
Cp → 2 sin2 β (4.26)

Let us go further. Consider the exact oblique shock relation for ρ/ρ∞ , given
by
ρ2 (γ + 1) M∞2
sin2 β
= 2 sin2 β + 2
(4.27)
ρ∞ (γ − 1) M∞

The above equation, as M∞ → ∞,


ρ2 γ+1
→ (4.28)
ρ∞ γ−1

In the additional limits as γ → 1, we find


ρ2
→∞ (4.29)
ρ∞

Here the density behind the shock is infinitely large. In turn, mass flow con-
siderations then dictate that the shock wave is coincident with the body surface.
Now considering the θ − β − M relation,

M12 sin2 β − 1
 
tan θ = 2 cot β (4.30)
M12 (γ + cos 2β) + 2

as β is small and M1 is very large, the above reduces to


β γ+1
= (4.31)
θ 2

Now as γ → 1 and M∞ → ∞,

θ=β (4.32)

i.e the shock wave lies on the body. In light of this result, the Equ. 4.26 can
be written as
Cp = 2 sin2 θ (4.33)

In the newtonian model of fluid flow, the particles in the free stream impact
only on the frontal area of the body; they cannot curl around the body and impact
on the back surface. Hence, for that portion of a body which is in the ”shadow”
of the incident flow, such as the shaded region sketched in Fig. 4.10, no impact
pressure is felt. Hence, over this shadow region it is consistent to assume that
p = p∞ and therefore Cp = 0, as indicated in Fig. 4.10.

High Speed Aerodynamics (R18A2113) 73


73
90
Chapter 4. Linearized Supersonic Flows and Hypersonic Flows

Fig. 4.10: Shadow region on the leeward side of a body, from newtonian theory.

4.6 Lift and Drag in Hypersonic flow


It is instructive to examine newtonian theory applied to a flat plate, as sketched
in Fig. 4.11. Here, a two-dimensional flat plate with chord length c is at an angle
of attack α to the free stream. Since we are not including friction, and because
surface pressure always acts normal to the surface, the resultant aerodynamic force
is perpendicular to the plate, i.e., in this case the normal force N is the resultant
aerodynamic force. (For an infinitely thin flat plate, this is a general result which
is not limited to newtonian theory, or even to hypersonic flow.) In turn, N is
resolved into lift and drag, denoted by L and D, respectively, as shown in Fig.
4.11.

Fig. 4.11: Flat plate at angle of attack. Illustration of aerodynamic forces.

According to newtonian theory, the pressure coefficient on the lower surface is

Cpl = 2 sin2 α (4.34)


and that on the upper surface, which is in the shadow region, is

Cpα = 0 (4.35)

High Speed Aerodynamics (R18A2113) 74 Dr. G. Srinivasan


74
91
Lift and Drag in Hypersonic flow

Defining the normal force coefficient as cn = N/q∞ S, where S = (c)(l), we


can readily calculate cn , by integrating the pressure coefficients over the lower and
upper surfaces ✂
1 c
cn = (Cpl − Cpα ) dx (4.36)
c 0
where x is the distance along the chord from the leading edge. Substituting Equs.
4.34 and 4.35 in the above equation, we get
1
2 sin2 α c

cn = (4.37)
c

cn = 2 sin2 α (4.38)
From the geometry of Figure. 4.11, we can see that the lift and drag coefficients,
defined as cl = L/q∞ S and cd = D/q∞ S, respectively, where S = (c)(l), are given
by
cl = cn cos α (4.39)
and
cd = cn sin α (4.40)
Substituting, Equ. 4.38 into 4.39 and 4.40, we get

cl = 2 sin2 α cos α (4.41)

and
cd = 2 sin3 α (4.42)
Finally, the lift-to-drag ratio is given by
L
= cot α (4.43)
D

High Speed Aerodynamics (R18A2113) 75


75
92
Chapter 4. Linearized Supersonic Flows and Hypersonic Flows

High Speed Aerodynamics (R18A2113) 76 Dr. G. Srinivasan


76
93
Chapter 5
Flow through nozzles and
variable area ducts

5.1 Quasi one-dimensional flow


In the early chapters the one dimensional flow was strictly treated as constant
area flow. In the present chapters, the restriction on constant area will be relaxed
by allowing the streamtube area A to vary with distance x as shown in the fig 5.1.
At the same time we will continue to assume that all flow properties are uniform
across any given cross section of the flow, and hence are functions of x only (and
time t if the flow is unsteady). Such a flow where A = A(x), p = p(x), ρ = ρ(x)
and V = u = u(x) for steady flow is defined as Quasi-one-dimensional flow.

Fig. 5.1: Quasi one-dimensional flow

5.2 Governing Equations


Algebraic equations for steady quasi-one-dimensional flow can be obtained by ap-
plying the integral form of the conservation equations to the variable-area control
volume sketched in Figure 5.2.

5.2.1 Continuity Equation

✍ ✝
The continuity equation is repeated here,

− ρV.dS = ρdV (5.1)
S ∂t v

77
77
94
Chapter 5. Flow through nozzles and variable area ducts

Fig. 5.2: Finite control volume for quasi one-dimensional flow

when integrated over the control volume in Figure 5.2, for steady flow, directly to

ρ1 u1 A1 = ρ2 u2 A2 (5.2)

5.2.2 Momentum Equation


The integral form of the momentum equation is
✍ ✝ ✝ ✍
∂ (ρV )
(ρV.dS) V + dV = ρf dV − pdS (5.3)
s ∂t V s

Applied to Figure 5.2, assuming steady flow and no body forces, it directly
becomes ✂ A2
p1 A1 + ρ1 u21 A1 + pdA = p2 A2 + ρ2 u22 A2 (5.4)
A1

This is the momentum equation for steady quasi-one-dimensional flow. Note that
it is not strictly an algebraic equation because of the integral term which represents
the pressure force on the sides of the control surface between locations 1 and 2.

5.2.3 Energy Equation


The integral form of the energy equation is
✝ ✍ ✝ ✝ ✍ 
V2 V2
   

q̇ρdV − pV.dS+ ρ (f.V ) dV = ρ e+ dV + ρ e+ V.dS
V S V V ∂t 2 S 2
(5.5)
Applied to Figure 5.2 and assuming steady adiabatic flow with no body forces,
it directly yields

u2 u2
   
− (−p1 u1 A1 + p2 u2 A2 ) = ρ1 e1 + 1 = p2 u2 A2 + ρ2 u2 A2 e2 + 2 u2 A2
2 2
(5.6)
Rearranging,

u2 u2
   
p1 u1 A1 + ρ1 u1 A1 e1 + 1 = p2 u2 A2 + ρ2 u2 A2 e2 + 2 (5.7)
2 2

High Speed Aerodynamics (R18A2113) 78 Dr. G. Srinivasan


78
95
Euler’s Equation for Quasi 1D flow

Dividing the above equation with 5.2, we get

p1 u2 p2 u2
+ e1 + 1 = + e2 + 2 (5.8)
ρ1 2 ρ2 2

Noting that h = e + p/ρ, The above equation becomes

u21 u2
h1 + = h2 + 2 (5.9)
2 2
This is the energy equation for steady adiabatic quasi-one-dimensional flow-it
states that the total enthalpy is constant along the flow.

5.3 Euler’s Equation for Quasi 1D flow


Consider an infinitesimal control volume with a variable are cross section as shown
in Figure 5.3.

Fig. 5.3: Infinitesimal control volume

The Quasi one dimensional continuity equation Equ. 5.2 can be written as

ρuA = constant (5.10)

Hence,
d (ρuA) = 0 (5.11)
To obtain a differential form of the momentum equation, apply Equ. 5.4 to the
infinitesimal control volume sketched in Figure 5.3, where the length in the x
direction is dx:

pA + ρu2 A + pdA = (p + dp) (A + dA) + (ρ + dρ) (u + du)2 (A + dA) (5.12)

Dropping all second-order terms involving products of differentials, this be-


comes
Adp + Au2 dρ + ρu2 dA + 2ρuAdu = 0 (5.13)
Expanding Equ. 5.11, and multiplying by u,

ρu2 dA + ρuAdu + Au2 dρ = 0 (5.14)

High Speed Aerodynamics (R18A2113) 79


79
96
Chapter 5. Flow through nozzles and variable area ducts

Substracting the above equation from Equ. 5.13, we get

dp = −ρudu (5.15)
Equ. 5.15 is called the Euler’s equation. Finally, a differntial form of the
energy equation is obtained from Equ. 5.9,which states that

u2
h+ = const (5.16)
2
Hence,
dh + udu = 0 (5.17)

5.4 Area Velocity Relation


From the differential form of 1D continuity equation, Equ. 5.11 can be written as
dρ du dA
+ + =0 (5.18)
ρ u A
To eliminate dρ/ρ from Equ. 5.18, consider the Euler’s equation Equ. 5.15 i.e:

dp dp dρ
= = −udu (5.19)
ρ dρ ρ
Recall that we are considering adiabatic, inviscid flow, i.e., there are no dissi-
pative mechanisms such as friction, thermal conduction, or diffusion acting on the
flow. Thus, the flow is isentropic. Hence, any change in pressure, dp, in the flow
is accompanied by a corresponding isentropic change in density, dρ. Therefore, we
can write  
dp ∂p
= = a2 (5.20)
dρ ∂ρ s
Combing Equ. 5.19 and Equ. 5.20, we get

a2 = −udu (5.21)
ρ
or
dρ udu u2 du du
= − 2 = − 2 = −M 2 (5.22)
ρ a au u
Substituting the above equation in Equ. 5.18, we get

dA  du
= M2 − 1 (5.23)
A u

Equation. 5.23 is an important relation. It is called the area-velocity rela-


tion, and it tells us this information:

1. For M → 0, which in the limit corresponds to incompressible flow, Equ.


5.23 shows that Au = const. This is the familiar continuity equation for the
incompressible flow.
2. For 0 ≤ M < 1 (subsonic flow), an increase in velocity (positive du)is

High Speed Aerodynamics (R18A2113) 80 Dr. G. Srinivasan


80
97
Area-Mach Number relation for flow inside Nozzles

Fig. 5.4: Flow in coverging and diverging ducts

associated with a decrease in area (negative dA), and vice versa. Therefore,
the familiar result from incompressible flow that the velocity increases in a
converging duct and decreases in a diverging duct still holds true for subsonic
compressible flow (see top of Figure. 5.4)

3. For M > 1 (supersonic flow), an increase in velocity is associated with an


increase in area, and vice versa. Hence, we have a striking difference in
comparison to subsonic flow. For supersonic flow, the velocity increases in
a diverging duct and decreases in a converging duct (see bottom of Figure.
5.4).

4. For M = 1 (sonic flow), Equ. 5.23 yields dA/A = 0, which mathemati-


cally corresponds to a minimum or maximum in the area distribution. The
minimum in area is the only physically realistic solution, as described next.

These results clearly show that for a gas to expand isentropically from subsonic
to supersonic speeds, it must flow through a convergent-divergent duct (or stream-
tube), as sketched at the top of Figure. 5.4. Moreover, at the minimum area that
divides the convergent and divergent sections of the duct, we know from item 4
above that the flow must be sonic. This minimum area is called a throat. Con-
versely, for a gas to compress isentropically from supersonic to subsonic speeds,
it must also flow through a convergent-divergent duct, with a throat where sonic
flow occurs, as sketched at the bottom of Figure. 5.4. From this discussion, we
recognize why rocket engines have large, bell-like nozzle shapes as sketched in
Figure. 5.6-to expand the exhaust gases to high-velocity, supersonic speeds.

5.5 Area-Mach Number relation for flow inside Nozzles


Consider the duct shown in Figure. 5.7.
At the throat, the flow is sonic. Hence, denoting conditions at sonic speed by
an asterisk, we have, at the throat, M ∗ = 1 and u∗ = a∗ . The area of the throat is
A∗ . At any other section of the duct, the local area, Mach number, and velocity
are A, M and u, respectively. Apply Equ. 5.2 between these two locations:

ρ∗ u∗ A∗ = ρuA (5.24)

High Speed Aerodynamics (R18A2113) 81


81
98
Chapter 5. Flow through nozzles and variable area ducts

Fig. 5.5: Flow in coverging and diverging ducts

Fig. 5.6: Flow in coverging and diverging ducts

Since u∗ = a∗ , The above equation becomes


A ρ∗ a∗ ρ∗ ρo a∗
= = (5.25)
A∗ ρ u ρo ρ u

where ρo is the stagnation density and is defined throughout the isentropic flow.
Recalling the isentropic relation between stagnation density and staic density at
any point in the flow,
 1/(γ−1)
ρo γ−1 2
= 1+ M (5.26)
ρ 2

High Speed Aerodynamics (R18A2113) 82 Dr. G. Srinivasan


82
99
Area-Mach Number relation for flow inside Nozzles

Fig. 5.7: Flow in coverging and diverging ducts

And apply the above equation to sonic condition where M = 1, we have


 1/(γ−1)
ρo γ+1
= (5.27)
ρ∗ 2

Also recalling the relation between the characteristic Mach number M ∗ and
the freestream Mach number M ,
γ+1
 u 2
∗2 2
M2
=M = (5.28)
a∗ 1 + γ−1
2
M2

Squaring Equ. 5.25 and substituting Equs. 5.27, 5.28 and 5.26, we get
 2  2  2  2
A ρ∗ ρo a∗
= (5.29)
A∗ ρo ρ u

2 2/(γ−1)  2/(γ−1) !
γ−1
1+ M2
 
A 2 γ−1 2 2
= 1+ M γ+1 (5.30)
A∗ γ+1 2 2
M2

 2   (γ+1)/(γ−1)
A 1 2 γ−1 2
= 2 1+ M (5.31)
A∗ M γ+1 2
Equation 5.31 is called the area-Mach number relation, and it contains a strik-
ing result. Turned inside out, Equ. 5.31 tells us that M = f (A/A∗ ), i.e., the Mach
number at any location in the duct is a function of the ratio of the local duct area
to the sonic throat area. As seen from Equ. 5.23, A must be greater than or at

High Speed Aerodynamics (R18A2113) 83


83
100
Chapter 5. Flow through nozzles and variable area ducts

least equal to A∗ ; the case where A < A∗ is physically not possible in an isentropic
flow. Also, from Equ. 5.31 there are two values of M that correspond to a given
A/A∗ > 1, a subsonic and a supersonic value. The solution of Equ. 5.31 is plotted
in Figure 5.8, which clearly delineates the subsonic and supersonic branches.

Fig. 5.8: Flow in coverging and diverging ducts

5.6 Isentropic Flow through Convergent Divergent Noz-


zles - Choked Flow, Under-expansion and Over-expansion
Consider a given convergent-divergent nozzle, as sketched in Figure. 5.9. Assume
that the area ratio ratio at the inlet Ai /A∗ is very large, Ai /A∗ → ∞, and that
the inlet is fed with gas from a large reservoir at pressure and temperature po
and To , respectively.Because of the large inlet area ratio, M ≈ 0, hence po and
To are essentially stagnation (or total values). (The Mach number cannot be
precisely zero in the reservoir or else there would be no mass flow through the
nozzle. It is a finite value, but small enough to assume that it is essentially zero.).
Furthermore, assume that the given convergent-divergent nozzle expands the flow
isentropically to supersonic speeds at the exit. For the given nozzle, there is only
one possible isentropic solution for supersonic flow, and Equ. 5.31 is the key to this
solution. In the convergent portion of the nozzle, the subsonic flow is accelerated,
with the subsonic value of M dictated by the local value of A/A∗ as given by
the lower branch of Figure. 5.8. The consequent variation of Mach number with
distance x along the nozzle is sketched in Figure. 5.9. At the throat, where the
throat area A = A∗ , M = 1. In the divergent portion of the nozzle, the flow
expands supersonically, with the supersonic value of M dictated by the local value
of A/A∗ as given by the upper branch of Figure. 5.8. This variation of M with
x in the divergent nozzle is also sketched in Figure. 5.9. Once the variation of
Mach number through the nozzle is known, the variations of static temperature,

High Speed Aerodynamics (R18A2113) 84 Dr. G. Srinivasan


84
101
The Effect of Different Pressure Ratios Across a Given Nozzle

pressure, and density follow from Isentropic relations. The resulting variations of
p and T are shown in Figure. 5.9, respectively. Note that the pressure, density,
and temperature decrease continuously throughout the nozzle. Also note that the
exit pressure, density, and temperature ratios, pe /po , ρe /ρo , and Te /To , depend
only on the exit area ratio, Ae /A∗ via Equ. 5.31.

Fig. 5.9: Flow in coverging and diverging ducts

5.7 The Effect of Different Pressure Ratios Across a Given


Nozzle
If a convergent-divergent nozzle is simply placed on a table, and nothing else is
done, obviously nothing is going to happen; the air is not going to start rushing
through the nozzle of its own accord. To accelerate a gas, a pressure difference
must be exerted, as clearly stated by Euler’s equation, Equ. 5.15. Therefore, in
order to establish a flow through any duct, the exit pressure must be lower than
the inlet pressure, i.e., pe /po < 1. Indeed, for completely shockfree isentropic
supersonic flow to exist in the nozzle of Figure. 5.9a, the exit pressure ratio must

High Speed Aerodynamics (R18A2113) 85


85
102
Chapter 5. Flow through nozzles and variable area ducts

be precisely the value of pe /po shown in Figure. 5.9c.


What happens when p,/p, is not the precise value as dictated by Fig. 5.9?
In other words, what happens when the backpressure downstream of the nozzle
exit is independently governed (say by exhausting into an infinite reservoir with
controllable pressure)? Consider a convergent-divergent nozzle as sketched in Fig.
5.10a. Assume that no flow exists in the nozzle, hence pe = po . Now assume that
pe is minutely reduced below po . This small pressure difference will cause a small
wind to blow through the duct at low subsonic speeds. The local Mach number will
increase slightly through the convergent portion of the nozzle, reaching a maximum
at the throat, as shown by curve 1 of Fig. 5.10b. This maximum will not be sonic;
indeed it will be a low subsonic value. Keep in mind that the value A∗ defined
earlier is the sonic throat area, i.e., that area where M = 1. In the case we are now
considering, where M < 1 at the minimum-area section of the duct, the real throat
area of the duct, At , is larger than A∗ , which for completely subsonic flow takes
on the character of a reference quantity different from the actual geometric throat
area. Downstream of the throat, the subsonic flow encounters a diverging duct.
and hence M decreases as shown in Fig. 5.10b. The corresponding variation of
static pressure is given by curve 1 in Fig. 5.10c. Now assume pe is further reduced.
This stronger pressure ratio between the inlet and exit will now accelerate the flow
more, and the variations of subsonic Mach number and static pressure through
the duct will be larger. as indicated by curve 2 in Figs. 5.10b and c. If pe is
further reduced, there will be some value of pe at which the flow will just barely
go sonic at the throat, as given by the curve 3 in Figs. 5.10b and c. In this case,
At = A∗ . Note that all the cases sketched in Figs 5.10b and c are subsonic flows.
Hence, for subsonic flow through the convergent-divergent nozzle shown in Fig.
5.10a, there are an infinite number of isentropic solutions, where both pe /po and
A/At are the controlling factors for the local flow properties at any given section.
This is a direct contrast with the supersonic case discussed earlier, where only
one isentropic solution exists for a given duct, and where AIA* becomes the only
controlling factor for the local flow properties (relative to reservoir properties).
For the cases shown in Figs. 5.10a, b, and c, the mass flow through the duct
increases as p, decreases. This mass flow can be calculated by evaluating Eq. (5.1)
at the throat, m = ρt At ut . When pe is reduced to pe3 , where sonic flow is attained
at the throat, then m = p∗ A∗ a∗ . If p, is now reduced further, p, < pe3 , the Mach
number at the throat cannot increase beyond M = 1; this is dictated by Equ.
5.31. Hence, the flow properties at the throat, and indeed throughout the entire
subsonic section of the duct, become ”frozen” when pe < pe3 , i.e., the subsonic
flow becomes unaffected and the mass flow remains constant for pe < pe3 . This
condition, after sonic flow is attained at the throat, is called choked flow. No
matter how low pe is made, after the flow becomes choked, the mass flow remains
constant. This phenomenon is illustrated in Fig. 5.11. Note that sonic flow at the
throat corresponds to a pressure ratio p∗ /po = 0.528 for γ = 1.4; however, because
of the divergent duct downstream of the throat, the value of pe3 /po required to
attain sonic flow at the throat is larger than 0.528, as shown in Figs. 5.10c and
5.11.
What happens in the duct when p, is reduced below pe3 ? In the convergent
portion, as we stated, nothing happens. The flow properties remain as given by
the subsonic portion of curve 3 in Fig. 5.10b and c. However, a lot happens in the

High Speed Aerodynamics (R18A2113) 86 Dr. G. Srinivasan


86
103
The Effect of Different Pressure Ratios Across a Given Nozzle

Fig. 5.10: Subsonic Flow in a coverging-diverging Nozzle

High Speed Aerodynamics (R18A2113) 87


87
104
Chapter 5. Flow through nozzles and variable area ducts

Fig. 5.11: Variation of mass flow with exit pressure; illustration of choked flow.

High Speed Aerodynamics (R18A2113) 88 Dr. G. Srinivasan


88
105
The Effect of Different Pressure Ratios Across a Given Nozzle

Fig. 5.12: Flow with a shock wave inside a convergent-divergent nozzle

divergent portion of the duct. No isentropic solution is allowed in the divergent


duct until pe is adequately reduced to the specified low value dictated by Fig. 5.9c.
or values of exit pressure above this, but below pe3 , a normal shock wave exists
inside the divergent duct. This situation is sketched in Fig. 5.12. Let the exit
pressure be given by pe4 .
There is a region of supersonic flow ahead of the shock. Behind the shock, the
flow is subsonic, hence the Mach number decreases towards the exit and the static
pressure increases to pe at the exit. The location of the normal shock wave in the
duct is determined by the requirement that the increase of static pressure across
the wave plus that in the divergent portion of the subsonic flow behind the shock
be just right to achieve pe4 at the exit. As the exit pressure is reduced further,
the normal shock wave will move downstream, closer to the nozzle exit. It will
stand precisely at the exit when pe = pe5 , where pe5 is the static pressure behind
a normal shock at the design Mach number of the nozzle. This is illustrated in
Figs. 5.13a, b, and c. In Fig. 5.13c, pe6 represents the proper isentropic value for
the design exit Mach number, which exists immediately upstream of the normal
shock wave standing at the exit. When the downstream backpressure pB is further

High Speed Aerodynamics (R18A2113) 89


89
106
Chapter 5. Flow through nozzles and variable area ducts

decreased such that pe6 < pB < pe5 , the flow inside the nozzle is fully supersonic
and isentropic, with the behavior the same as given earlier in Figs. 5.9a, b, c, and
d. The increase to the backpressure takes place across an oblique shock attached
to the nozzle exit, but outside the duct itself. This is sketched in Fig. 5.13d. If
the backpressure is further reduced below pe6 , equilibration of the flow takes place
across expansion waves outside the duct, as shown in Fig. 5.13e.
When the situation in Fig. 5.13d exists, the nozzle is said to be overex-
panded, because the pressure at the exit has expanded below the back pressure,
pe6 < pB . Conversely, when the situation in Fig. 5.13e exists, the nozzle is said to
be underexpanded, because the exit pressure is higher than the back pressure,
pe6 > pB and hence the flow is capable of additional expansion after leaving the
nozzle.

5.8 Diffusers
Difusers are the devices used to slow the flow with as samall a loss
of total pressure as possible Let us go through a small thought experiment.
Assume that we want to design a supersonic wind tunnel with a test section Mach
number of 3. Some immediate information about the nozzle is obtained from
isentropic property tables; at M = 3, Ae /A∗ = 4.23 and po /pe = 36.7. Assume
the wind tunnel exhausts to the atmosphere. What value of total pressure po
must be provided by the reservoir to drive the tunnel? There are several possible
alternatives. The first is to simply exhaust the nozzle directly to the atmosphere,
as sketched in Fig. 5.14.
In order to avoid shock or expansion waves in the test region downstream of the
exit, the exit pressure pe , must be equal to the surrounding atmospheric pressure,
i.e., pe = 1 atm. Since po /pe = 36.7, the driving reservoir pressure for this case
must be 36.7 atm. However, a second alternative is to exhaust the nozzle into a
constant-area duct which serves as the test section, and to exhaust this duct into
the atmosphere, as sketched in Fig. 5.15. In this case, because the testing area
is inside the duct, shock waves from the duct exit will not affect the test section.
Therefore, assume a normal shock stands at the duct exit. The static pressure
behind the normal shock is p2 , and because the flow is subsonic behind the shock,
p2 = p∞ = 1 atm. In this case, the reservoir pressure po is obtained from

po pe 1
po = p∞ = 367 1 = 3.55atm (5.32)
pe p2 10.33
where p2 /pe is the static pressure ratio across a normal shock at Mach 3,
obtained from Normal shock table. Note that, by the simple addition of a constant-
area duct with a normal shock at the end, the reservoir pressure required to drive
the wind tunnel has markedly dropped from 36.7 to 3.55 atm. Now, as a third
alternative, add a divergent duct behind the normal shock in Fig. 5.15 in order
to slow the already subsonic flow to a lower velocity before exhausting to the
atmosphere. This is sketched in Fig. 5.16. At the duct exit, the Mach number is
a very low subsonic value, and for all practical purposes the local total and static
pressure are the same. Moreover, assuming an isentropic flow in the divergent
duct behind the shock, the total pressure at the duct exit is equal to the total
pressure behind the normal shock. Consequently, po2 ≈ p∞ = 1atm. From the

High Speed Aerodynamics (R18A2113) 90 Dr. G. Srinivasan


90
107
Diffusers

Fig. 5.13: Flow with shock and expansion waves at the exit of a convergent-
divergent nozzle

High Speed Aerodynamics (R18A2113) 91


91
108
Chapter 5. Flow through nozzles and variable area ducts

Fig. 5.14: Nozzle exhausting directly to atmosphere

Fig. 5.15: Nozzle with a normal shock at the exit, exhausting to the atmosphere

normal shock tables, the Mach number behind the shock is M2 = 0.475, and the
ratio of total to static pressure at this Mach number (from isentropic property
tables) is po2 /p2 = 1.17, Hence
po p e p2 1 1
po = p∞ = 36.7 1 = 3.04atm (5.33)
pe p2 po2 10.33 1.17

5.9 Wave Reflection from a free boundary


The gas jet from a nozzle which exhausts into the atmosphere has a boundary
surface which interfaces with the surrounding quiescent gas. The oblique shock
waves shown in Fig. 5.13d and the expansion waves sketched in Fig. 5.13e must
reflect from the jet boundary in such a fashion as to preserve the pressure at
the boundary downstream of the nozzle exit. This jet boundary is not a solid
surface as treated earlier; rather, it is a free boundary which can change in size
and direction. Considering the overexpanded nozzle flow in Fig. 5.13d, the flow
pattern downstream of the nozzle exit will appear as sketched in Fig. 5.17. The
various reflected waves form a diamond-like pattern throughout the exhaust jet.
Such a diamond wave pattern is visible in the exhaust from the free jet.

High Speed Aerodynamics (R18A2113) 92 Dr. G. Srinivasan


92
109
Supersonic Wind Tunnel

Fig. 5.16: Nozzle with a normal-shock diffuser. The normal shock is slightly
upstream of the divergent duct.

Fig. 5.17: Schematic of the diamond wave pattern in the exhaust from a super-
sonic nozzle

5.10 Supersonic Wind Tunnel


5.10.1 Parts of supersonic wind tunnel

High Speed Aerodynamics (R18A2113) 93


93
110
Chapter 5. Flow through nozzles and variable area ducts

Fig. 5.18: Schematic of a supersonic Wind tunnel

Fig. 5.19: Norh with a conventional supersonic diffuser

5.11 Numerical Problems


1. Consider the isentropic subsonic-supersonic flow through a convergent-divergent
nozzle. The reservoir pressure and temperature are 10 atm and 300 K, re-
spectively. There are two locations in the nozzle where A/A∗ = 6: one in the
convergent section and the other in the divergent section. At each location,
calculate M. p, T, and u.
Solution: In the convergent section, the flow is subsonic. From the front of
Table A.1, for subsonic flow with A/A∗ = 6:

po To
M = 0.097 , = 1.006,
p T = 1.002
Hence,
p 1
p= po = × 10 = 9.94 atm
po 1.006

T 1
T = To = × 300 = 299.4 K
To 1.002

p √
a= γRT = 1.4 × 287 × 299.4 = 346.8 m/s

High Speed Aerodynamics (R18A2113) 94 Dr. G. Srinivasan


94
111
Numerical Problems

u = M a = 0.097 × 346.8 = 33.6 m/s

In the divergent section, the flow is supersonic. From the supersonic section
of Table A.1, for A/A∗ = 6:

po To
= 63.13, and = 3.269
p T

p 1
p= po = × 10 = 0.1584 atm
po 63.13

T 1
T = To = × 300 = 91.77 K
To 3.269
p √
a= γRT = 1.4 × 287 × 91.77 = 192.0 m/s

u = M a = 3.368 × 192.0 = 646.7 m/s

2. A supersonic wind tunnel is designed is designed to produce flow at Mach


2.4. at standard atmospheric conditions. Calculate (i) the exit to throat
area ratio of the nozzle (ii) Reservoir pressure and temperature.
Solution: From Table A.1, for Me = 2.5:
Ae po To
= 2.637′ = 17.09, = 2.25
A ∗ pe Te

At standard sea level conditions, pe = 1 atm and Te = 288 K, Hence,


po
po = pe = 17.09 × 1 = 17.09 atm
pe

To
To = Te = 2.25 × 288 = 648 K
Te

High Speed Aerodynamics (R18A2113) 95


95
112
Chapter 5. Flow through nozzles and variable area ducts

High Speed Aerodynamics (R18A2113) 96 Dr. G. Srinivasan


96
113

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy