You are on page 1of 67

“STUDY OF CONVERGENT-DIVERGENT NOZZLE”

A MINOR PROJECT REPORT

Submitted by
A. RAHUL RAO 0206ME061001
ANIKET SINGH 0206ME061005
KAPIL PATEL 0206ME061019
MADAN DHAKAL 0206ME061021
MANISH R. BHATIA 0206ME073D03
NAVEEN KUMAR NAMDEO 0206ME061025

DEPARTMENT OF MECHANICAL ENGINEERING


GYAN GANGA INSTITUTE OF TECHNOLOGY & SCIENCES
JABALPUR (M.P.)

RAJIV GANDHI PRODYOGIKI VISHWAVIDYALAYA,


BHOPAL (M.P.)

DECEMBER- 2009
TABLE OF CONTENT

CHAPTER NO. TITLE PAGE NO.

CERTIFICATE I
DECLARATION II
ACKNOWLEDGEMENT III
ABSTRACT IV
LIST OF TABLES V
LIST OF FIGURES V

1. INTRODUCTION 1

2. LITERATURE REVIEW 2

2.1 FUNCTION OF PATTERNS 2


2.2 TYPES OF PATTERNS 2
2.3 PATTERN DESIGN AND
CONSIDERATIONS 8
2.4 SELECTION OF PATTERN MATERIAL 9
2.4.1 MATERIALS USED FOR PATTERNS
2.4.1.1 WOOD 9
2.4.1.2 METAL 10
2.4.1.3 POLYSTERENE 10
2.4.1.4 PLASTER 15
2.4.1.5 WAX 16
2.5 WOOD USED FOR PATTERN 16
2.6 PATTERN CONSTRUCTION AND STEPS 19
INVOLVED
2.6.1 PATTERN LAYOUT 19
2.6.2 PATTERN COLORS 19
2.7 PATTERN ALLOWANCES 20
2.8 CALCULATIONS 24
CERTIFICATE

This is to certify that the Minor Project report entitled “STUDY OF


CONVERGENT-DIVERGENT NOZZLE” is submitted by A. RAHUL
RAO, ANIKET SINGH, KAPILPATEL, MADAN DHAKAL, NAVEEN
KUMAR NAMDEO & MANISH R. BHATIA in Mechanical Engineering
from “RAJIV GANDHI PROUDYOGIKI VISHWAVIDYALAYA,
BHOPAL (M.P.)”

Prof. P. K. Sahu Prof. K.L. Kanojia


Guide HOD
Dept. of Mechanical Engg. Dept. of Mechanical Engg.
CERTIFICATE

This is to certify that the Minor Project report entitled “STUDY OF


CONVERGENT-DIVERGENT NOZZLE” is submitted by A. RAHUL
RAO,ANIKET SINGH, KAPIL PATEL, MADAN DHAKAL, NAVEEN
KUMAR NAMDEO & MANISH R. BHATIA in Mechanical Engineering
from “RAJIV GANDHI PROUDYOGIKI VISHWAVIDYALAYA,
BHOPAL (M.P.)”

Internal Examiner External Examiner

Date: Date:
DECLARATION

We hereby declare that the project entitled “STUDY OF CONVERGENT-


DIVERGENT NOZZLE” which is being submitted to “RAJIV GANDHI
PROUDYOGIKI VISHWAVIDYALAYA, BHOPAL (M.P.)” is an
authentic record of our own work done under the guidance of Prof. P. K.
SAHU, Department of Mechanical Engineering, GYAN GANGA
INSTITUTE OF TECHNOLOGY & SCIENCES, JABALPUR.
The matter reported in this Project has not been submitted earlier
for the award of any other degree.

Dated : / / A. RAHUL RAO


ANIKET SINGH
Place : Jabalpur KAPILPATEL
MADAN DHAKAL
NAVEEN KUMAR NAMDEO
MANISH R. BHATIA
ACKNOWLEDGEMENT

We sincerely express indebtedness to esteemed and revered guide “Prof. P.


K. Sahu”, Professor in Mechanical Department for his invaluable guidance,
supervision and encouragement throughout the work. Without his kind
patronage and guidance the project would not have taken shape.

We take this opportunity to express deep sense of gratitude to “Prof. K.L.


Kanojia”, Head of “Mechanical Department” for his encouragement and
kind approval. We would like to express our sincere regards to him for
advice and counseling from time to time.

We owe sincere thanks to all the lecturers in “Mechanical Department” for


their advice and counseling time to time.

Dated: / /
A. RAHUL RAO
Place: Jabalpur ANIKET SINGH
KAPILPATEL
MADAN DHAKAL
NAVEEN KUMAR NAMDEO
MANISH R. BHATIA
ABSTRACT

This project is a effort by our team towards the Study of convergent-divergent nozzle.
The purpose of this is to simulate the operation of a converging-diverging nozzle, perhaps
the most important and basic piece of engineering hardware associated with propulsion
and the high speed flow of gases. This device was invented by Carl de Laval toward the
end of the l9th century and is thus often referred to as the 'de Laval' nozzle. This applet is
intended to help students of compressible aerodynamics visualize the flow through this
type of nozzle at a range of conditions.

A de Laval nozzle (or convergent-divergent nozzle, CD nozzle or con-di nozzle) is


basically a tube that is pinched in the middle, making an hourglass-shape. It is used as a
means of accelerating the flow of a gas passing through it. It is widely used in some types
of steam turbine and is an essential part of the modern rocket engine and supersonic jet
engines. It is used in rocketry, where its use can achieve speeds of 4 km/s or more; jet
engines, and in wind tunnels where it is capable of producing smoother airflow than other
techniques.

We have also included the basics of Computational fluid Dynamics. Computational Fluid
Dynamics (CFD) is a computer-based tool for simulating the behavior of systems
involving fluid flow, heat transfer, and other related physical processes.
LIST OF FIGURES:-

• Figure1:-Diagram of Solid Pattern


• Figure2:-Diagram of Split Pattern
• Figure3:-Diagram of Match Plate Pattern
• Figure4:-Diagram of Gated Pattern
• Figure5:-Diagram of Sweep Pattern
• Figure6:-Wooden Pattern

LIST OF TABLES:

• Table1:- Strength of different woods


• Table2:- Shrinkage Allowance
Nozzle
A nozzle is a mechanical device designed to control the characteristics of a fluid flow as
it exits (or enters) an enclosed chamber or pipe via an orifice. A nozzle is often a pipe or
tube of varying cross sectional area and it can be used to direct or modify the flow of a
fluid (liquid or gas). Nozzles are frequently used to control the rate of flow, speed,
direction, mass, shape, and/or the pressure of the stream that emerges from them.

Types of nozzles
Jets

A gas jet, fluid jet, or hydro jet is a nozzle intended to eject gas or fluid in a coherent
stream into a surrounding medium. Gas jets are commonly found in gas stoves, ovens, or
barbecues. Gas jets were commonly used for light before the development of electric
light. Other types of fluid jets are found in carburetors, where smooth calibrated orifices
are used to regulate the flow of fuel into an engine, and in jacuzzis or spas. Another
specialized jet is the laminar jet. This is a water jet that contains devices to smooth out
the flow, and gives laminar flow, as its name suggests. This gives better results for
fountains. Nozzles used for feeding hot blast into a blast furnace or forge are called
tuyeres.

High velocity nozzles

Frequently the goal is to increase the kinetic energy of the flowing medium at the
expense of its pressure and internal energy.

Nozzles can be described as convergent (narrowing down from a wide diameter to a


smaller diameter in the direction of the flow) or divergent (expanding from a smaller
diameter to a larger one). A de Laval nozzle has a convergent section followed by a
divergent section and is often called a convergent-divergent nozzle ("con-di nozzle").
Convergent nozzles accelerate subsonic fluids. If the nozzle pressure ratio is high enough
the flow will reach sonic velocity at the narrowest point (i.e. the nozzle throat). In this
situation, the nozzle is said to be choked.

Increasing the nozzle pressure ratio further will not increase the throat Mach number
beyond unity. Downstream (i.e. external to the nozzle) the flow is free to expand to
supersonic velocities. Note that the Mach 1 can be a very high speed for a hot gas; since
the speed of sound varies as the square root of absolute temperature. Thus the speed
reached at a nozzle throat can be far higher than the speed of sound at sea level. This fact
is used extensively in rocketry where hypersonic flows are required, and where propellant
mixtures are deliberately chosen to further increases the sonic speed.

Divergent nozzles slow fluids, if the flow is subsonic, but accelerate sonic or supersonic
fluids. Convergent-divergent nozzles can therefore accelerate fluids that have choked in
the convergent section to supersonic speeds. This CD process is more efficient than
allowing a convergent nozzle to expand supersonically externally. The shape of the
divergent section also ensures that the direction of the escaping gases is directly
backwards, as any sideways component would not contribute to thrust.

Propelling nozzles

A jet exhaust produces a net thrust from the energy obtained from combusting fuel which
is added to the inducted air. This hot air is passed through a high speed nozzle, a
propelling nozzle which enormously increases its kinetic energy.

For a given mass flow, greater thrust is obtained with a higher exhaust velocity, but the
best energy efficiency is obtained when the exhaust speed is well matched with the
airspeed. However, no jet aircraft can sustain flight while exceeding its exhaust jet speed,
due to momentum considerations. Supersonic jet engines, like those employed in fighters
and SST aircraft (e.g. Concorde), need high exhaust speeds. Therefore supersonic aircraft
very typically use a CD nozzle despite weight and cost penalties. Subsonic jet engines
employ relatively low, subsonic, exhaust velocities. They thus employ simple convergent
nozzles. In addition, bypass nozzles are employed giving even lower speeds.
Rocket motors use convergent-divergent nozzles with very large area ratios so as to
maximize thrust and exhaust velocity and thus extremely high nozzle pressure ratios are
employed. Mass flow is at a premium since all the propulsive mass is carried with
vehicle, and very high exhaust speeds are desirable.

Magnetic nozzles

Magnetic nozzles have also been proposed for some types of propulsion, such as
VASIMR, in which the flow of plasma is directed by magnetic fields instead of walls
made of solid matter.

Spray nozzles

Many nozzles produce a very fine spray of liquids.

• Atomizer nozzles are used for spray painting, perfumes, carburettors for internal
combustion engines, spray on deodorants, antiperspirants and many other uses.
• Air-Aspirating Nozzle-uses an opening in the cone shaped nozzle to inject air
into a stream of water based foam (CAFS/AFFF/FFFP) to make the concentrate
"foam up". Most commonly found on foam extinguishers and foam handlines.
• Swirl nozzles inject the liquid in tangentially, and it spirals into the center and
then exits through the central hole. Due to the vortexing this causes the spray to
come out in a cone shape.

Vacuum nozzles

Vacuum cleaner nozzles come in several different shapes.

Shaping nozzles

Some nozzles are shaped to produce a stream that is of a particular shape. For example
Extrusion molding is a way of producing lengths of metals or plastics or other materials
with a particular cross-section. This nozzle is typically referred to as a die.
CONVERGENT-DIVERGENT NOZZELS

Laval nozzle is a Convergent-divergent nozzle in which subsonic flow prevails in the


converging section, critical od transonic conditions in the throat and supersonic flow in
the diverging section. The opening narrows then opens back up, as you can see in the
picture below. These are also known as De Leval nozzles (named after de Laval, the
Swedish scientist who invented it). This type of nozzle is an hourglass shape, which
accelerates the exhaust to very high velocities. The convergent portion of the nozzle
accelerates the exhaust. This acceleration produces much more thrust than the engine
would without a nozzle. This device was invented by Carl de Laval toward the end of the
l9th century and is thus often referred to as the 'de Laval' nozzle

It consists of three main parts which are:

a.) Short Convergent Part (Velocity increase towards the end of this part)
b.) Short throat part. (Here Maximum Mass Flow rate is achieved.)
c.) Divergent Part i.e. diffuser (It is designed in the manner that it doesn’t have effect on
the mass flow rate achieved at throat.
Mach number
Mach number (Ma or M) (is the speed of an object moving through air, or any fluid
substance, divided by the speed of sound as it is in that substance. It is commonly used to
represent an object's (such as an aircraft or missile) speed, when it is traveling at (or at
multiples of) the speed of sound.

An F/A-18 Hornet at transonic speed and displaying the Prandtl–Glauert singularity just
before reaching the speed of sound

Where,
is the Mach number
is the speed of the source (the object relative to the medium) and
is the speed of sound in the medium

The Mach number is named after Czech/Austrian physicist and philosopher Ernst Mach.
Because the Mach number is often viewed as a dimensionless quantity rather than a unit
of measure, with Mach, the number comes after the unit; the second Mach number is
"Mach 2" instead of "2 Mach" (or Machs). This is somewhat reminiscent of the early
modern ocean sounding unit "mark" (a synonym for fathom), which was also unit-first,
and may have influenced the use of the term Mach. In the decade preceding man's flying
faster than sound, aeronautical engineers referred to the speed of sound as Mach's
number, never "Mach 1".

The Mach number is commonly used both with objects traveling at high speed in a fluid,
and with high-speed fluid flows inside channels such as nozzles, diffusers or wind
tunnels. As it is defined as a ratio of two speeds, it is a dimensionless number. At a
temperature of 15 degrees Celsius, the speed of sound is 340.3 m/s (1225 km/h, or 761.2
mph, or 661.5 knots, or 1116 ft/s) in the Earth's atmosphere. The speed represented by
Mach 1 is not a constant; for example, it is dependent on temperature and atmospheric
composition. In the stratosphere it remains constant irrespective of altitude even though
the air pressure varies with altitude.

Since the speed of sound increases as the temperature increases, the actual speed of an
object traveling at Mach 1 will depend on the fluid temperature around it. Mach number
is useful because the fluid behaves in a similar way at the same Mach number. So, an
aircraft traveling at Mach 1 at sea level (340.3 m/s, 761.2 mph, 1,225 km/h) will
experience shock waves in much the same manner as when it is traveling at Mach 1 at
11,000 m (36,000 ft), even though it is traveling at 295 m/s (654.6 mph, 1,062 km/h, 86%
of its speed at sea level).

High-speed flow around objects

Flight can be roughly classified in five categories:

• Subsonic: M < 1
• Sonic: M=1
• Transonic: 0.8 < M < 1.2
• Supersonic: 1.2 < M < 5
• Hypersonic: M > 5

For comparison: the required speed for low Earth orbit is approximately 7.5 km/s = M
25.4 in air at high altitudes. The speed of light in vacuum corresponds to a Mach number
of approximately 880,000 (relative to air at sea level).

At transonic speeds, the flow field around the object includes both sub- and supersonic
parts. The transonic period begins when first zones of M>1 flow appear around the
object. In case of an airfoil (such as an aircraft's wing), this typically happens above the
wing. Supersonic flow can decelerate back to subsonic only in a normal shock; this
typically happens before the trailing edge. (Fig. a)

As the speed increases, the zone of M>1 flow increases towards both leading and trailing
edges. As M=1 is reached and passed, the normal shock reaches the trailing edge and
becomes a weak oblique shock: the flow decelerates over the shock, but remains
supersonic. A normal shock is created ahead of the object, and the only subsonic zone in
the flow field is a small area around the object's leading edge. (Fig. b)

(a) (b)

Mach number in transonic airflow around an airfoil; M<1 (a) and M>1 (b).

When an aircraft exceeds Mach 1 (i.e. the sound barrier) a large pressure difference is
created just in front of the aircraft. This abrupt pressure difference, called a shock wave,
spreads backward and outward from the aircraft in a cone shape (a so-called Mach cone).
It is this shock wave that causes the sonic boom heard as a fast moving aircraft travels
overhead. A person inside the aircraft will not hear this. The higher the speed, the more
narrow the cone; at just over M=1 it is hardly a cone at all, but closer to a slightly
concave plane.

At fully supersonic speed, the shock wave starts to take its cone shape and flow is either
completely supersonic, or (in case of a blunt object), only a very small subsonic flow area
remains between the object's nose and the shock wave it creates ahead of itself. (In the
case of a sharp object, there is no air between the nose and the shock wave: the shock
wave starts from the nose.)

As the Mach number increases, so does the strength of the shock wave and the Mach
cone become increasingly narrow? As the fluid flow crosses the shock wave, its speed is
reduced and temperature, pressure, and density increase. The stronger the shock, the
greater the changes. At high enough Mach numbers the temperature increases so much
over the shock that ionization and dissociation of gas molecules behind the shock wave
begin. Such flows are called hypersonic.
It is clear that any object traveling at hypersonic speeds will likewise be exposed to the
same extreme temperatures as the gas behind the nose shock wave, and hence choice of
heat-resistant materials becomes important.

High-speed flow in a channel


As a flow in a channel crosses M=1 becomes supersonic, one significant change takes
place. Common sense would lead one to expect that contracting the flow channel would
increase the flow speed (i.e. making the channel narrower results in faster air flow) and at
subsonic speeds this holds true. However, once the flow becomes supersonic, the
relationship of flow area and speed is reversed: expanding the channel actually increases
the speed.

The obvious result is that in order to accelerate a flow to supersonic, one needs a
convergent-divergent nozzle, where the converging section accelerates the flow to M=1,
sonic speeds, and the diverging section continues the acceleration. Such nozzles are
called de Laval nozzles and in extreme cases they are able to reach incredible, hypersonic
speeds (Mach 13 at sea level).

An aircraft Mach meter or electronic flight information system (EFIS) can display Mach
number derived from stagnation pressure (pitot tube) and static pressure.

Calculating Mach number

Assuming air to be an ideal gas, the formula to compute Mach number in a subsonic
compressible flow is derived from Bernoulli's equation for M<1:

where:

is Mach number
is impact pressure and
is static pressure
is the ratio of specific heat of a gas at a constant pressure to heat at a constant
volume (1.4 for air).

The formula to compute Mach number in a supersonic compressible flow is derived from
the Rayleigh Supersonic Pitot equation:

Where:

is now impact pressure measured behind a normal shock.

Shock wave
Whenever a supersonic flow (compressible) abruptly changes to subsonic flow, a shock
wave (analogous to hydraulic jump in an open channel) is produced, resulting in a sudden
rise in pressure, density, temperature and entropy. This occurs due to pressure
differentials and when the Mach number of the approaching flow M1>1. A shock wave is
a pressure wave of finite thickness, of the order of 10-2 to 104mm in the atmospheric
pressure. A shock wave takes place in the diverging section of a nozzle, in a diffuser,
throat of a supersonic wind tunnel, in front of sharp-nosed bodies.
Shock waves are of two types:
1. Normal shocks which are almost perpendicular to the flow.
2. Oblique shocks which are inclined to the flow direction.
Back pressure
Back pressure usually refers to the pressure exerted on a moving fluid by obstructions or
tight bends in the confinement vessel along which it is moving, such as piping or air
vents, against its direction of flow. For example, an automotive exhaust muffler with a
particularly high number of twists, bends, turns and right angles could be described as
having particularly high back pressure. Back pressure, in the exhaust sense of the term, of
a four-stroke engine is usually termed as being a "bad thing" for performance; however,
in the interest of reducing exhaust sound to levels allowable by public noise ordinances,
back pressure can be regulated using systems from simple butterfly valves to fully
computer controlled units sensing pressure in the exhaust pipe itself.

In a two-stroke engine however, a certain amount of exhaust backpressure is needed to


prevent unburned fuel/air mixture to pass right through the cylinders into the exhaust.

Fig 1.1 illustrates the effects of varying back pressure on a converging–diverging nozzle.
The series of cases labeled a through j is considered next

1. Let us first discuss the cases designated a, b, c, and d. Case a corresponds to pB = pE =


po for which there is no flow. When the back pressure is slightly less than po (case b),
there is some flow, and the flow is subsonic throughout the nozzle. In accordance with
the discussion of Fig. 1.2, the greatest velocity and lowest pressure occur at the throat,
and the diverging portion acts as a diffuser in which pressure increases and velocity
decreases in the direction of flow. If the back pressure is reduced further, corresponding
to case c, the mass flow rate and velocity at the throat are greater than before. Still, the
flow remains subsonic throughout and qua –litatively the same as case b. As the back
pressure is reduced, the Mach number at the throat increases, and eventually a Mach
number of unity is attained there (case d). As before, the greatest velocity and lowest
pressure occur at the throat, and the diverging portion remains a subsonic diffuser.
However, because the throat velocity is sonic, the nozzle is now choked: The maximum
mass flow rate has been attained for the given stagnation conditions. Further reductions in
back pressure cannot result in an increase in the mass flow rate.

2. When the back pressure is reduced below that corresponding to case d, the flow
through the converging portion and at the throat remains unchanged. Conditions within
the diverging portion can be altered, however, as illustrated by cases e, f, and g. In case e,
the fluid passing the throat continues to expand and becomes supersonic in the diverging
portion just downstream of the throat; but at a certain location an abrupt change in
properties occurs. This is called a normal shock. Across the shock, there is a rapid and
irreversible increase in pressure, accompanied by a rapid decrease from supersonic to
subsonic flow. Downstream of the shock, the diverging duct acts as a subsonic diffuser in
which the fluid continues to decelerate and the pressure increases to match the back
pressure imposed at the exit. If the back pressure is reduced further (case f), the location
of the shock moves downstream, but the flow remains qualitatively the same as in case e.
With further reductions in back pressure, the shock location moves farther downstream of
the throat until it stands at the exit (case g). In this case, the flow throughout the nozzle is
isentropic, with subsonic flow in the converging portion, M =1 at the throat, and
supersonic flow in the diverging portion. Since the fluid leaving the nozzle passes
through a shock, it is subsonic just downstream of the exit plane.

3. Finally, let us consider cases h, i, and j where the back pressure is less than that
corresponding to case g. In each of these cases, the flow through the nozzle is not
affected. The adjustment to changing back pressure occurs outside the nozzle. In case h,
the pressure decreases continuously as the fluid expands isentropically through the nozzle
and then increases to the back pressure outside the nozzle. The compression that occurs
outside the nozzle involves oblique shock waves. In case i, the fluid expands
isentropically to the back pressure and no shocks occur within or outside the nozzle. In
case j, the fluid expands isentropically through the nozzle and then expands outside the
nozzle to the back pressure through oblique expansion waves. Once M = 1 is achieved at
the throat, the mass flow rate is fixed at the maximum value for the given stagnation
conditions, so the mass flow rate is the same for back pressures corresponding to cases d
through j. The pressure variations outside the nozzle involving oblique waves cannot be
predicted using the one-dimensional flow model. In case of ideal nozzle isentropic
expansion is considered, with nozzle efficiency at unity. And flow at divergent (diffuser)
section as frictionless, but in actual it’s not feasible.

Choked flow
At the "throat", where the cross sectional area is a minimum, the gas velocity locally
becomes sonic (Mach number = 1.0), a condition called choked flow. Choked flow of a
fluid is a Fluid dynamics condition caused by the Venturi effect. When a flowing fluid at
a certain pressure and temperature flows through a restriction into a lower pressure
environment, under the conservation of mass the fluid velocity must increase for initially
subsonic upstream conditions as it flows through the smaller...
. As the nozzle cross sectional area increases the gas begins to expand and the gas flow
increases to supersonic velocities where a sound wave will not propagate backwards
through the gas as viewed in the frame of reference of the nozzle.

The Fluid Flow in Duct

During the fluid flow in a duct, the properties of fluid changes along the stream line. In
most situations this can be treated as a one-dimensional flow

*The Pressure and Velocity


From conservation of energy
dh= - cdc
du +d(pv)= - cdc
du+pdv+vdp= -cdc

From the first law of thermodynamics:


δq=du+pdv=0
Then: du+pdv+vdp= -cdc
vdp= - cdc
- vdp= cdc
To increase the velocity of fluid (dc>0), the pressure must be decreased. This kind of
duct is called Nozzle Or to decrease the velocity (dc<0) of fluid to obtain a high pressure
in a duct flow. This kind of duct is called diffuser.

Velocity and the Cross Section Area of Duct

From conservation of mass

dA dc dv
+ =
A c v
dA dv dc
= −
A v c

From the process equation


dp dv
+k =0
p v

dp dv
= −k
p v

dv dp
=−
v kp
As to a nozzle dc>0

dA dc
= ( Ma 2 − 1)
A c

(1)If the fluid velocity is subsonic, then (Ma2-1) <0


Therefore: dA<0
The nozzle’s shape should be as following:

dA dc
= ( Ma 2 −1) ( dc > 0)
A c

(2)If the fluid velocity is ultrasonic, then (Ma2-1)>0


Therefore: dA>0
The nozzle’s shape should be as following:
Ultrasonic Ultrasonic
flow flow

divergent
nozzle

(3)If the nozzle’s inlet velocity is subsonic, but outlet velocity ultrasonic, then:
dA<0 → dA=0 → dA>0
The nozzle’s shape should be as following:

Throat

Ultrasonic
Subsonic flow flow

Convergent-
divergent nozzle
dA dc
= ( Ma 2 −1) ( dc > 0)
A c

The Critical Compression Ratio


For convergent-divergent nozzle, the velocity at throat keeps as sound-velocity. This state
is called critical state. The flux of this kind of nozzle is depending on that of throat.
The Compression Ratio which is low enough to make the air flow at sound-velocity at the
exit of nozzle is called the critical compression Ratio.
It is denoted by εc
k
 2  k −1
εc =  
 k −1 

0.528 For air

εc= 0.546 For saturated steam

0.577 For superheated steam

2
k  2 k −1 p1
 = Athroat
m 2  
k +1  k +1  v1

Effect of Area Variation on Flow Properties in Isentropic Flow


In considering the effect of area variation on flow properties in isentropic flow, we shall
concern ourselves primarily with the velocity and pressure. We shall determine the effect
of change in area, A, on the velocity V, and the pressure p.

From Bernoulli's equation, we can write


Dividing by , we obtain

(19.1)

A convenient differential form of the continuity equation

Substituting from Eq. (19.1),

Invoking the relation ( ) for isentropic process in Eq. (19.2), we get

From Eq. (19.3), we see that for Ma<1 an area change causes a pressure change of the
same sign, i.e. positive dA means positive dp for Ma<1. For Ma>1, an area change causes
a pressure change of opposite sign.

Again, substituting from Eq.(19.1) into Eq. (19.3), we obtain


From Eq. (19.4), we see that Ma<1 an area change causes a velocity change of opposite
sign, i.e. positive dA means negative dV for Ma<1. For Ma>1, an area change causes a
velocity change of same sign.

These results are summarized in Fig.19.1, and the relations (19.3) and (19.4) lead to the
following important conclusions about compressible flows:

1. At subsonic speeds (Ma<1) a decrease in area increases the speed of flow. A


subsonic nozzle should have a convergent profile and a subsonic diffuser should
possess a divergent profile. The flow behavior in the regime of Ma<1 is therefore
qualitatively the same as in incompressible flows.

In supersonic flows (Ma>1), the effect of area changes are different. According to Eq.
(19.4), a supersonic nozzle must be built with an increasing area in the flow direction. A
supersonic diffuser must be a converging channel. Divergent nozzles are used to produce
supersonic flow in missiles and launch vehicles

Shapes of nozzles and diffusers in subsonic and supersonic regimes


Suppose a nozzle is used to obtain a supersonic stream staring from low speeds at the
inlet (Fig.19.2). Then the Mach number should increase from Ma=0 near the inlet to
Ma>1 at the exit. It is clear that the nozzle must converge in the subsonic portion and
diverge in the supersonic portion. Such a nozzle is called a convergent-divergent nozzle.
A convergent-divergent nozzle is also called a de Laval nozzle, after Carl G.P. de Laval
who first used such a configuration in his steam turbines in late nineteenth century (this
has already been mentioned in the introductory note). From Fig.19.2 it is clear that the
Mach number must be unity at the throat, where the area is neither increasing nor
decreasing. This is consistent with Eq. (19.4) which shows that dV can be non-zero at the
throat only if Ma=1. It also follows that the sonic velocity can be achieved only at the
throat of a nozzle or a diffuser.

A convergent-divergent nozzle

The condition, however, does not restrict that Ma must necessarily be unity at the throat,

According to Eq. (19.4), a situation is possible where at the throat if dV=0 there.
For an example, the flow in a convergent-divergent duct may be subsonic everywhere
with Ma increasing in the convergent portion and decreasing in the divergent portion with

at the throat (see Fig.19.3). The first part of the duct is acting as a nozzle, whereas
the second part is acting as a diffuser. Alternatively, we may have a convergent-divergent
duct in which the flow is supersonic everywhere with Ma decreasing in the convergent

part and increasing in the divergent part and again at the throat (see Fig. 19.4).

Convergent-divergent duct with at throat

Convergent-divergent duct with at throat

Question- At some section in the convergent-divergent nozzle in which air is flowing,


pressure, velocity, temperature and cross-sectional area are 22 kN/m2 , 170 m/s, 200º C
and 1000 mm2 respectively. If the flow conditions are iscentropic, determine:
(i) Stagnation temperature and stagnation pressure.
(ii) Sonic velocity and Mach number at this section.
(iii) Velocity, Mach number and flow area at outlet section where pressure is 110 kN/m2.
(iv) Pressure, temperature, velocity and flow at throat of the nozzle.
Take for air R=287 J/kg K, Cp=1.0 kJ/kg K and γ =1.4

Solution: Let subscripts 1, 2 and t refers to the condition at given section, outlet section
and throat section of the nozzle respectively.

Pressure in the nozzle, p1= 200 kN/m2


Velocity of air, V1=170 m/s
Temperature, T1 = 200+273=473 K
Cross sectional area, A1 = 1000 mm2 = 1000×10-6 = 0.001 m2
For air: R = 287 J/kg K, Cp=1.0 kJ/kg K , γ =1.4

(i) Stagnation temperature (Ts) and stagnation pressure (ps):

Stagnation temperature, Ts = T1+ V12


2×Cp

=473+ 1702
2×(1.0×1000)

= 487.45 K (or 214.45 ºC)

Also, PS/P1 = (TS/T1)γ/γ-1 = (487.45 / 473)1.4/1.4-1 = 1.111

Stagnation pressure, PS = 200*1.111 = 222.2 kN/m2


(ii) Sonic velocity and mach number at this section:

Sonic velocity, C1 = √γRT1 = √1.4*287*473 = 435.9 m/s


Mach number, M1 = V1/C1 = 170/435.9 = 0.39

(iii) Velocity, Mach number and flow area at outlet section where pressure is
110kN/m2
Pressure at outlet section, p2 = 110 kN/m2

PS/P2 = [1+ (1.4-1/2)M22]γ/γ-1

222.2/110 = [1+ {(1.4-1)/2} M20] 1.4/1.4-1

M2 = 1.05

(T2/TS) = (p2/pS)γ-1/γ

T2= 398.7 K

Sonic velocity at outlet section, C2 = √γRT2 = 400.25 m/s

Velocity at outlet section, V2 = M2 * C2 = 420.26 m/s

Now, mass flow at the given section = mass flow at outlet section (exit)
…..continuity equation
i.e. ρ1A1V1 = ρ2A2V2 or

p1 * A1V1 = p2 *A2V2
RT1 RT2
Flow area at the outlet section,

A2 = p1A1V1T2 = 200*0.001*170*398.7 = 6.199*10-4 m2


T1p2V2 473*110*420.26
Hence, A2= 6.199*10-4 m2 or 619.9 mm2

(iv) Pressure(pt), temperature (Tt), Velocity(Vt), and flow area (At) at throat
of the nozzle:
At throat, critical conditions prevail, i.e. the flow velocity becomes equal to the sonic
velocity and Mach number attains a unit value.

From equation, TS = [1+ (γ-1/2) M2t]


Tt

Tt = 406.2 K (or 133.2 ºC)

Also, pt = (Tt/Ts) γ/γ-1

pt = 117.32 kN / m2

Sonic velocity (corresponding to throat conditions),


Ct = √γRT1 = √1.4*287*406.2 = 404 m/s

Flow Velocity, Vt = Mt* Ct = 1*404 = 404 m/s

By continuity equation, we have : ρ1A1V1 = ρtAtVt

p1 * A1V1 = pt *AtVt
RT1 RTt

Flow area of throat, At = p1A1V1Tt = 200*0.001*170*406.2 = 6.16*10-4 m2


T1ptVt 473*117.32*404

Hence, A2= 6.16*10-4 m2 or 6.16 mm2


APPLICATION OF CONVERGENT-DIVERGENT
NOZZLES:

Rocket

Space
Shuttle

Ram-jet engine

Diffuser Nozzl
(compressor) e
Combustion
chamber
COMPUTATIONAL FLUID DYNAMICS
Computational fluid dynamics (CFD) is one of the branches of fluid mechanics that uses
numerical methods and algorithms to solve and analyze problems that involve fluid
flows. Computers are used to perform the millions of calculations required to simulate the
interaction of liquids and gases with surfaces defined by boundary conditions. Even with
high-speed supercomputers only approximate solutions can be achieved in many cases.
Ongoing research, however, may yield software that improves the accuracy and speed of
complex simulation scenarios such as transonic or turbulent flows.

Computational Fluid Dynamics (CFD) is a computer-based tool for simulating the


behavior of systems involving fluid flow, heat transfer, and other related physical
processes. It works by solving the equations of fluid flow (in a special form) over a
region of interest, with specified (known) conditions on the boundary of that region.

THE HISTORY OF COMPUTATIONAL FLUID DYNAMICS [CFD]

Computers have been used to solve fluid flow problems for many years.
Numerous programs have been written to solve either specific problems, or specific
classes of problems. From the mid-1970’s, the complex mathematics required to
generalize the algorithms began to be understood, and general purpose CFD solvers were
developed. These began to appear in the early 1980’s and required what were then very
powerful computers, as well as an in-depth knowledge of fluid dynamics, and large
amounts of time to set up simulations. Consequently, CFD was a tool used almost
exclusively in research.
Recent advances in computing power, together with powerful graphics and
interactive 3D manipulation of models have made the process of creating a CFD model
and analyzing results much less labor intensive, reducing time and, hence, cost.
Advanced solvers contain algorithms which enable robust solutions of the flow field in a
reasonable time As a result of these factors, Computational Fluid Dynamics is now an
established industrial design tool, helping to reduce design time scales and improve
processes throughout the engineering world. CFD provides a cost-effective and accurate
alternative to scale model testing, with variations on the simulation being performed
quickly, offering obvious advantages.

Uses of CFD

CFD is used by engineers and scientists in a wide range of fields. Typical applications
include:

• Process industry: Mixing vessels, chemical reactors

• Building services: Ventilation of buildings, such as atriums

• Health and safety: Investigating the effects of fire and smoke

• Motor industry: Combustion modeling, car aerodynamics

• Electronics: Heat transfer within and around circuit boards

Computational Fluid Dynamics: CFD Methodology

• Environmental: Dispersion of pollutants in air or water

• Power and energy: Optimization of combustion processes

• Medical: Blood flow through grafted blood vessels

The Mathematics of CFD

The set of equations which describe the processes of momentum, heat and mass transfer
are known as the Navier-Stokes equations. These partial differential equations were
derived in the early nineteenth century and have no known general analytical solution but
can be discretized and solved numerically .Equations describing other processes, such as
combustion, can also be solved in conjunction with the Navier-Stokes equations. Often,
an approximating model is used to derive these additional equations, turbulence models
being a particularly important example.
There are a number of different solution methods which are used in CFD codes. The most
common, and the one on which ANSYS CFX is based, is known as the finite volume
technique. In this technique, the region of interest is divided into small sub-regions,
called control volumes. The equations are discretized and solved iteratively for each
control volume. As a result, an approximation of the value of each variable at specific
points throughout the domain can be obtained. In this way, one derives a full picture of
the behavior of the flow. Additional information on the Navier-Stokes equations and
other mathematical aspects of the ANSYS CFX software suite is available

Technicalities
The most fundamental consideration in CFD is how one treats a continuous fluid in a
discretized fashion on a computer. One method is to discretize the spatial domain into
small cells to form a volume mesh or grid, and then apply a suitable algorithm to solve
the equations of motion (Euler equations for inviscid, and Navier-Stokes equations for
viscous flow). In addition, such a mesh can be either irregular (for instance consisting of
triangles in 2D, or pyramidal solids in 3D) or regular; the distinguishing characteristic of
the former is that each cell must be stored separately in memory. Where shocks or
discontinuities are present, high resolution schemes such as Total Variation Diminishing
(TVD), Flux Corrected Transport (FCT), Essentially NonOscillatory (ENO), or MUSCL
schemes are needed to avoid spurious oscillations (Gibbs phenomenon) in the solution.

If one chooses not to proceed with a mesh-based method, a number of alternatives exist,
notably :

• Smoothed particle hydrodynamics (SPH), a Lagrangian method of solving fluid


problems,
• Spectral methods, a technique where the equations are projected onto basis
functions like the spherical harmonics and Chebyshev polynomials,
• Lattice Boltzmann methods (LBM), which simulate an equivalent mesoscopic
system on a Cartesian grid, instead of solving the macroscopic system (or the real
microscopic physics).
It is possible to directly solve the Navier-Stokes equations for laminar flows and for
turbulent flows when all of the relevant length scales can be resolved by the grid (a Direct
numerical simulation). In general however, the range of length scales appropriate to the
problem is larger than even today's massively parallel computers can model. In these
cases, turbulent flow simulations require the introduction of a turbulence model. Large
eddy simulations (LES) and the Reynolds-averaged Navier-Stokes equations (RANS)
formulation, with the k-ε model or the Reynolds stress model, are two techniques for
dealing with these scales.

In many instances, other equations are solved simultaneously with the Navier-Stokes
equations. These other equations can include those describing species concentration
(mass transfer), chemical reactions, heat transfer, etc. More advanced codes allow the
simulation of more complex cases involving multi-phase flows (e.g. liquid/gas, solid/gas,
liquid/solid), non-Newtonian fluids (such as blood), or chemically reacting flows (such as
combustion).

Navier-Stokes Equations

Newton's Second Law


The integral form of the linear momentum equation was discussed in the Linear
Momentum Integral Equation section. Recall, Newton's second law on a differential fluid
element is

δF = δm a

where δF is the resultant force acting on the fluid element (mass = δm). a is the
acceleration of the fluid element, and it is given by

Expanding into its Cartesian components yields

There are two types of forces acting on the fluid element: body force (δFB) and surface
force (δFS).

δF = δFB + δFS

The only body force considered here is the weight of the fluid element. That is,

δFB = δm g = δm (gx i + gy j + gz k)

Generally, gravity only acts in one direction, but since the coordinate system is not set, all
three terms are included for the general case. Other body forces, such as those due to the
magnetic and electric fields, can be fit into this framework, but they are not covered in
this eBook.
Notations for the Stresses

Surface Forces in the x-direction

The surface forces are due to the stresses exerted on the sides of the fluid element. There
are two types of stresses applied on the surface: normal stress (σij) and shear stress (τij).
Normal stress acts perpendicular to the surface while shear stress is tangential to the
surface. The subscript i refer to the axis normal to the surface, and the subscript j
represents the direction of the stress. The notation of the stresses is illustrated further in
the animation shown on the left.
All surface forces acting in the x-direction on the fluid element are shown in the figure. A
summation of the surface forces in the x-direction yields

Note that the stresses are multiplied by the area to obtain the surface forces. Similarly, the
total surface forces in the y- and z-directions (not shown in figure) are obtained as

The resultant surface force is then given as

δFS = δFSx i + δFSy j + δFSz k

The mass of the fluid element can be expressed in terms of its volume and fluid density
(δm = ρ δxδyδz), so that the linear momentum equation in Cartesian coordinates reduces
to

For Newtonian fluids, such as water, oil and air, the shear stress field is symmetric, and it
is related to the rate of shear strain in a linear fashion. The results are (derivation details
not given here),
where μ is the viscosity of the fluid. Notice, the pressure term, p, only acts normal to the
surface for each element face. Also, the pressure is assumed to be the same on all three
faces, i.e. hydrostatic pressure.

For incompressible flow, the term is zero based on the continuity equation. The
linear momentum equations thus become
The above equations are generally referred to as the Navier-Stokes equations, and
commonly written as a single vector form,

Although the vector form looks simple, this equation is the core fluid mechanics
equations and is an unsteady, nonlinear, 2nd order, partial differential equation. It is
extremely hard to solve, and only simple 2D problems have been solved. Computational
Fluid Dynamics (CFD) is most often used to solve the Navier-Stokes equations.

Navier–Stokes equations

The Navier–Stokes equations, named after Claude-Louis Navier and George Gabriel
Stokes, describe the motion of fluid substances that is substances which can flow. These
equations arise from applying Newton's second law to fluid motion, together with the
assumption that the fluid stress is the sum of a diffusing viscous term (proportional to the
gradient of velocity), plus a pressure term.

They are exceptionally useful because they describe the physics of many things of
academic and economic interest. They may be used to model the weather, ocean currents,
water flow in a pipe, the air's flow around a wing, and motion of stars inside a galaxy.
The Navier–Stokes equations in their full and simplified forms help with the design of
aircraft and cars, the study of blood flow, the design of power stations, the analysis of
pollution, and many other things. Coupled with Maxwell's equations they can be used to
model and study magneto hydrodynamics.

The Navier–Stokes equations are also of great interest in a purely mathematical sense.
Somewhat surprisingly, given their wide range of practical uses, mathematicians have not
yet proven that in three dimensions solutions always exist (existence), or that if they do
exist they do not contain any infinities, singularities or discontinuities (smoothness).
These are called the Navier–Stokes existence and smoothness problems. The Navier–
Stokes equations dictate not position but rather velocity. A solution of the Navier–Stokes
equations is called a velocity field or flow field, which is a description of the velocity of
the fluid at a given point in space and time. Once the velocity field is solved for, other
quantities of interest (such as flow rate or drag force) may be found. This is different
from what one normally sees in classical mechanics, where solutions are typically
trajectories of position of a particle or deflection of a continuum. Studying velocity
instead of position makes more sense for a fluid; however for visualization purposes one
can compute various trajectories.

Properties

Nonlinearity

The Navier–Stokes equations are nonlinear partial differential equations in almost every
real situation. In some cases, such as one-dimensional flow and Stokes flow (or creeping
flow), the equations can be simplified to linear equations. The nonlinearity makes most
problems difficult or impossible to solve and is the main contributor to the turbulence that
the equations model.

The nonlinearity is due to convective acceleration, which is an acceleration associated


with the change in velocity over position. Hence, any convective flow, whether turbulent
or not, will involve nonlinearity, an example of convective but laminar (nonturbulent)
flow would be the passage of a viscous fluid (for example, oil) through a small
converging nozzle. Such flows, whether exactly solvable or not, can often be thoroughly
studied and understood.

Turbulence

Turbulence is the time dependent chaotic behavior seen in many fluid flows. It is
generally believed that it is due to the inertia of the fluid as a whole: the culmination of
time dependent and convective acceleration; hence flows where inertial effects are small
tend to be laminar (the Reynolds number quantifies how much the flow is affected by
inertia). It is believed, though not known with certainty, that the Navier–Stokes equations
describe turbulence properly.

The numerical solution of the Navier–Stokes equations for turbulent flow is extremely
difficult, and due to the significantly different mixing-length scales that are involved in
turbulent flow, the stable solution of this requires such a fine mesh resolution that the
computational time becomes significantly infeasible for calculation (see Direct numerical
simulation). Attempts to solve turbulent flow using a laminar solver typically result in a
time-unsteady solution, which fails to converge appropriately. To counter this, time-
averaged equations such as the Reynolds-averaged Navier-Stokes equations (RANS),
supplemented with turbulence models (such as the k-ε model), are used in practical
computational fluid dynamics (CFD) applications when modeling turbulent flows.
Another technique for solving numerically the Navier–Stokes equation is the Large-eddy
simulation (LES). This approach is computationally more expensive than the RANS
method (in time and computer memory), but produces better results since the larger
turbulent scales are explicitly resolved.

Applicability

Together with supplemental equations (for example, conservation of mass) and well
formulated boundary conditions, the Navier–Stokes equations seem to model fluid
motion accurately; even turbulent flows seem (on average) to agree with real world
observations.
The Navier–Stokes equations assume that the fluid being studied is a continuum not
moving at relativistic velocities. At very small scales or under extreme conditions, real
fluids made out of discrete molecules will produce results different from the continuous
fluids modeled by the Navier–Stokes equations. Depending on the Knudsen number of
the problem, statistical mechanics or possibly even molecular dynamics may be a more
appropriate approach.

Another limitation is very simply the complicated nature of the equations. Time tested
formulations exist for common fluid families, but the application of the Navier–Stokes
equations to less common families tends to result in very complicated formulations which
are an area of current research. For this reason, the Navier–Stokes equations are usually
written for Newtonian fluids.

Almost universally the equations are written for a simple class of fluids (which most
liquids and all known gases belong to) known as Newtonian fluids. Studying such fluids
is "simple" because the viscosity model ends up being linear; truly general models for the
flow of other kinds of fluids (such as blood) do not, as of 2009, exist.

Derivation and description

The derivation of the Navier–Stokes equations begins with an application of Newton's


second law: conservation of momentum (often alongside mass and energy conservation)
being written for an arbitrary control volume. In an inertial frame of reference, the
general form of the equations of fluid motion is:[2]

where is the flow velocity, ρ is the fluid density, p is the pressure, is the (deviatoric)
stress tensor, and represents body forces (per unit volume) acting on the fluid and is
the del operator. This is a statement of the conservation of momentum in a fluid and it is
an application of Newton's second law to a continuum; in fact this equation is applicable
to any non-relativistic continuum and is known as the Cauchy momentum equation.

This equation is often written using the substantive derivative, making it more apparent
that this is a statement of Newton's second law:

The left side of the equation describes acceleration, and may be composed of time
dependent or convective effects (also the effects of non-inertial coordinates if present).
The right side of the equation is in effect a summation of body forces (such as gravity)
and divergence of stress (pressure and stress).

Convective acceleration

An example of convection. Though the flow is steady (time independent), the fluid
decelerates as it moves down the diverging duct (when the flow is subsonic), hence there
is acceleration.

A very significant feature of the Navier–Stokes equations is the presence of convective


acceleration: the effect of time independent acceleration of a fluid with respect to space,
represented by the nonlinear quantity:
which may be interpreted either as or as with the tensor
derivative of the velocity vector Both interpretations give the same result, independent
of the coordinate system — provided is interpreted as the covariant derivative.

Interpretation as (v·∇)v

The convection term is often written as

Where the advection operator is used. Usually this representation is preferred


because it is simpler than the one in terms of the tensor derivative

Interpretation as v·(∇v)

Here is the tensor derivative of the velocity vector, equal in Cartesian coordinates to
the component by component gradient. The convection term may, by a vector calculus
identity, be expressed without a tensor derivative:[4][5]

The form has use in irrotational flow, where the curl of the velocity (called vorticity)
is equal to zero.

Regardless of what kind of fluid is being dealt with, convective acceleration is a


nonlinear effect. Convective acceleration is present in most flows (exceptions include
one-dimensional incompressible flow), but its dynamic effect is disregarded in creeping
flow (also called Stokes flow) .

Stresses

The effect of stress in the fluid is represented by the and terms, these are gradients
of surface forces, analogous to stresses in a solid. is called the pressure gradient and
arises from the isotropic part of the stress tensor. This part is given by normal stresses
that turn up in almost all situations, dynamic or not. The anisotropic part of the stress
tensor gives rise to , which conventionally describes viscous forces; for
incompressible flow, this is only a shear effect. Thus, is the deviatoric stress tensor, and
the stress tensor is equal to

where is the 3×3 identity matrix. Interestingly, only the gradient of pressure matters, not
the pressure itself. The effect of the pressure gradient is that fluid flows from high
pressure to low pressure.

The stress terms p and are yet unknown, so the general form of the equations of motion
is not usable to solve problems. Besides the equations of motion—Newton's second law
—a force model is needed relating the stresses to the fluid motion. For this reason,
assumptions on the specific behavior of a fluid are made (based on natural observations)
and applied in order to specify the stresses in terms of the other flow variables, such as
velocity and density.

The Navier–Stokes equations result from the following assumptions on the deviatoric
stress tensor :

• the deviatoric stress vanishes for a fluid at rest, and – by Galilean invariance –
also does not depend directly on the flow velocity itself, but only on spatial
derivatives of the flow velocity
• in the Navier–Stokes equations, the deviatoric stress is expressed as the product of
the tensor gradient of the flow velocity with a viscosity tensor , i.e. :

• the fluid is assumed to be isotropic, as valid for gases and simple liquids, and
consequently is an isotropic tensor; furthermore, since the deviatoric stress
tensor is symmetric, it turns out that it can be expressed in terms of two scalar

dynamic viscosities μ and μ”: where


is the rate-of-strain tensor and is the
rate of expansion of the flow
• the deviatoric stress tensor has zero trace, so for a three-dimensional flow
2μ + 3μ” = 0

As a result, in the Navier–Stokes equations the deviatoric stress tensor has the following
form:

with the quantity between brackets the non-isotropic part of the rate-of-strain tensor
The dynamic viscosity μ does not need to be constant – in general it depends on
conditions like temperature and pressure, and in turbulence modelling the concept of
eddy viscosity is used to approximate the average deviatoric stress.

The pressure p is modelled by use of an equation of state. For the special case of an
incompressible flow, the pressure constrains the flow in such a way that the volume of
fluid elements is constant: isochoric flow resulting in a solenoidal velocity field with

Incompressible flow of Newtonian fluids

A simplification of the resulting flow equations is obtained when considering an


incompressible flow of a Newtonian fluid. The assumption of incompressibility rules out
the possibility of sound or shock waves to occur; so this simplification is invalid if these
phenomena are important. The incompressible flow assumption typically holds well even
when dealing with a "compressible" fluid — such as air at room temperature — at low
Mach numbers (even when flowing up to about Mach 0.3). Taking the incompressible
flow assumption into account and assuming constant viscosity, the Navier–Stokes
equations will read, in vector form:
Here f represents "other" body forces (forces per unit volume), such as gravity or
centrifugal force. The shear stress term becomes the useful quantity when the
fluid is assumed incompressible and Newtonian, where is the dynamic viscosity.

It's well worth observing the meaning of each term (compare to the Cauchy momentum
equation):

Note that only the convective terms are nonlinear for incompressible Newtonian flow.
The convective acceleration is an acceleration caused by a (possibly steady) change in
velocity over position, for example the speeding up of fluid entering a converging nozzle.
Though individual fluid particles are being accelerated and thus are under unsteady
motion, the flow field (a velocity distribution) will not necessarily be time dependent.

Another important observation is that the viscosity is represented by the vector Laplacian
of the velocity field. This implies that Newtonian viscosity is diffusion of momentum,
this works in much the same way as the diffusion of heat seen in the heat equation (which
also involves the Laplacian).

If temperature effects are also neglected, the only "other" equation (apart from
initial/boundary conditions) needed is the mass continuity equation. Under the
incompressible assumption, density is a constant and it follows that the equation will
simplify to:

This is more specifically a statement of the conservation of volume (see divergence).


These equations are commonly used in 3 coordinates systems: Cartesian, cylindrical, and
spherical. While the Cartesian equations seem to follow directly from the vector equation
above, the vector form of the Navier–Stokes equation involves some tensor calculus
which means that writing it in other coordinate systems is not as simple as doing so for
scalar equations (such as the heat equation).

Cartesian coordinates

Writing the vector equation explicitly,

Note that gravity has been accounted for as a body force, and the values of gx,gy,gz will
depend on the orientation of gravity with respect to the chosen set of coordinates.

The continuity equation reads:

The velocity components (the dependent variables to be solved for) are typically named
u, v, w. This system of four equations comprises the most commonly used and studied
form. Though comparatively more compact than other representations, this is a nonlinear
system of partial differential equations for which solutions are difficult to obtain.
Cylindrical coordinates

A change of variables on the Cartesian equations will yield the following momentum
equations for r, φ, and z:\

The gravity components will generally not be constants, however for most applications
either the coordinates are chosen so that the gravity components are constant or else it is
assumed that gravity is counteracted by a pressure field (for example, flow in horizontal
pipe is treated normally without gravity and without a vertical pressure gradient). The
continuity equation is:

This cylindrical representation of the incompressible Navier–Stokes equations is the


second most commonly seen (the first being Cartesian above). Cylindrical coordinates are
chosen to take advantage of symmetry, so that a velocity component can disappear. A
very common case is axisymmetric flow with the assumption of no tangential velocity (uφ
= 0), and the remaining quantities are independent of φ:
Spherical coordinates

In spherical coordinates, the r, θ, and φ momentum equations are (note the convention
used: θ is colatitude):

Mass continuity will read:


These equations could be (slightly) compacted by, for example, factoring 1 / r2 from the
viscous terms. This isn't done to preserve the structure of the Laplacian and other
quantities.

Stream function formulation

Taking the curl of the Navier–Stokes equation results in the elimination of pressure. This
is especially easy to see if 2D Cartesian flow is assumed (w = 0 and no dependence of
anything on z), where the equations reduce to:

Differentiating the first with respect to y, the second with respect to x and subtracting the
resulting equations will eliminate pressure and any potential force. Defining the stream
function ψ through

results in mass continuity being unconditionally satisfied (given the stream function is
continuous), and then incompressible Newtonian 2D momentum and mass conservation
degrade into one equation:

where is the (2D) biharmonic operator and ν is the kinematic viscosity, . This
single equation together with appropriate boundary conditions describes 2D fluid flow,
taking only kinematic viscosity as a parameter. Note that the equation for creeping flow
results when the left side is assumed zero.

In axisymmetric flow another stream function formulation, called the Stokes stream
function, can be used to describe the velocity components of an incompressible flow with
one scalar function.

Compressible flow of Newtonian fluids

There are some phenomena that are closely linked with fluid compressibility. One of the
obvious examples is sound. Description of such phenomena requires more general
presentation of the Navier–Stokes equation that takes into account fluid compressibility.
If viscosity is assumed a constant, one additional term appears, as shown here:

where μv is second viscosity coefficient. It is related to volume viscosity or bulk


viscosity. This additional term disappears for incompressible fluid, when the divergence
of the flow equals zero.

Euler's Equations

For inviscid flow (μ = 0), the Navier-Stokes equations reduce to


The above equations are known as Euler's equations. Note that the equations governing
inviscid flow have been simplified tremendously compared to the Navier-Stokes
equations; however, they still cannot be solved analytically due to the complexity of the
nonlinear terms (i.e., u ∂u/∂x, v ∂u/∂y, w ∂u/∂z, etc.). Hence, in the study of fluid
mechanics, numerical methods such as the finite element and finite difference methods
(along with the use of computers) are often used to approximate the fluid flow problems.
Euler's equations can be written in vector form as

Methodology

In all of these approaches the same basic procedure is followed.

• During preprocessing
o The geometry (physical bounds) of the problem is defined.
o The volume occupied by the fluid is divided into discrete cells (the mesh).
The mesh may be uniform or non uniform.
o The physical modeling is defined - for example, the equations of motions
+ enthalpy + radiation + species conservation
o Boundary conditions are defined. This involves specifying the fluid
behaviour and properties at the boundaries of the problem. For transient
problems, the initial conditions are also defined.
• The simulation is started and the equations are solved iteratively as a steady-state
or transient.
• Finally a postprocessor is used for the analysis and visualization of the resulting
solution.

Discretization methods

The stability of the chosen discretization is generally established numerically rather than
analytically as with simple linear problems. Special care must also be taken to ensure that
the discretization handles discontinuous solutions gracefully. The Euler equations and
Navier-Stokes equations both admit shocks, and contact surfaces.

Some of the discretization methods being used are:

FINITE ELEMENT METHOD

The finite element method (FEM) (sometimes referred to as finite element analysis) is
a numerical technique for finding approximate solutions of partial differential equations
(PDE) as well as of integral equations. The solution approach is based either on
eliminating the differential equation completely (steady state problems), or rendering the
PDE into an approximating system of ordinary differential equations, which are then
numerically integrated using standard techniques such as Euler's method, Runge-Kutta,
etc.

In solving partial differential equations, the primary challenge is to create an equation


that approximates the equation to be studied, but is numerically stable, meaning that
errors in the input data and intermediate calculations do not accumulate and cause the
resulting output to be meaningless. There are many ways of doing this, all with
advantages and disadvantages. The Finite Element Method is a good choice for solving
partial differential equations over complex domains (like cars and oil pipelines), when the
domain changes (as during a solid state reaction with a moving boundary), when the
desired precision varies over the entire domain, or when the solution lacks smoothness.
For instance, in a frontal crash simulation it is possible to increase prediction accuracy in
"important" areas like the front of the car and reduce it in its rear (thus reducing cost of
the simulation); Another example would be the simulation of the weather pattern on
Earth, where it is more important to have accurate predictions over land than over the
wide-open sea.

History

The finite-element method originated from the need for solving complex elasticity and
structural analysis problems in civil and aeronautical engineering. Its development can be
traced back to the work by Alexander Hrennikoff (1941) and Richard Courant (1942).
While the approaches used by these pioneers are dramatically different, they share one
essential characteristic: mesh discretization of a continuous domain into a set of discrete
sub-domains, usually called elements.

Hrennikoff's work discretizes the domain by using a lattice analogy while Courant's
approach divides the domain into finite triangular subregions for solution of second order
elliptic partial differential equations (PDEs) that arise from the problem of torsion of a
cylinder. Courant's contribution was evolutionary, drawing on a large body of earlier
results for PDEs developed by Rayleigh, Ritz, and Galerkin.

Development of the finite element method began in earnest in the middle to late 1950s
for airframe and structural analysis and gathered momentum at the University of Stuttgart
through the work of John Argyris and at Berkeley through the work of Ray W. Clough in
the 1960s for use in civil engineering. By late 1950s, the key concepts of stiffness matrix
and element assembly existed essentially in the form used today. NASA issued request
for proposals for the development of the finite element software NASTRAN in 1965. The
method was provided with a rigorous mathematical foundation in 1973 with the
publication of Strang and Fix's An Analysis of The Finite Element Method has since been
generalized into a branch of applied mathematics for numerical modeling of physical
systems in a wide variety of engineering disciplines, e.g., electromagnetism and fluid
dynamics.

Application
A variety of specializations under the umbrella of the mechanical engineering discipline
(such as aeronautical, biomechanical, and automotive industries) commonly use
integrated FEM in design and development of their products. Several modern FEM
packages include specific components such as thermal, electromagnetic, fluid, and
structural working environments. In a structural simulation, FEM helps tremendously in
producing stiffness and strength visualizations and also in minimizing weight, materials,
and costs.

FEM allows detailed visualization of where structures bend or twist, and indicates the
distribution of stresses and displacements. FEM software provides a wide range of
simulation options for controlling the complexity of both modeling and analysis of a
system. Similarly, the desired level of accuracy required and associated computational
time requirements can be managed simultaneously to address most engineering
applications. FEM allows entire designs to be constructed, refined, and optimized before
the design is manufactured.

This powerful design tool has significantly improved both the standard of engineering
designs and the methodology of the design process in many industrial applications. The
introduction of FEM has substantially decreased the time to take products from concept
to the production line. It is primarily through improved initial prototype designs using
FEM that testing and development have been accelerated. In summary, benefits of FEM
include increased accuracy, enhanced design and better insight into critical design
parameters, virtual prototyping, fewer hardware prototypes, a faster and less expensive
design cycle, increased productivity, and increased revenue.

Finite volume method


The finite volume method is a method for representing and evaluating partial differential
equations in the form of algebraic equations. Similar to the finite difference method,
values are calculated at discrete places on a meshed geometry. "Finite volume" refers to
the small volume surrounding each node point on a mesh. In the finite volume method,
volume integrals in a partial differential equation that contain a divergence term are
converted to surface integrals, using the divergence theorem. These terms are then
evaluated as fluxes at the surfaces of each finite volume. Because the flux entering a
given volume is identical to that leaving the adjacent volume, these methods are
conservative. Another advantage of the finite volume method is that it is easily
formulated to allow for unstructured meshes. The method is used in many computational
fluid dynamics packages.

1D example
Consider a simple 1D advection problem defined by the following partial differential
equation

Here, represents the state variable and represents the


flux or flow of . Conventionally, positive represents flow to the right whilst negative
represents flow to the left. If we assume that equation (1) represents a flowing medium of
constant area, we can sub-divide the spatial domain, , into finite volumes or cells with
cell centres indexed as . For a particular cell, , we can define the volume average value

of at time and , as

and at time as,

where and represent locations of the upstream and downstream faces or edges
respectively of the cell.

Integrating equation (1) in time, we have:


where .

To obtain the volume average of at time , we integrate over the

cell volume, and divide the result by , i.e.

We assume that is well behaved and that we can reverse the order of integration. Also,
recall that flow is normal to the unit area of the cell. Now, since in one dimension

, we can apply the divergence theorem, i.e. , and


substitute for the volume integral of the divergence with the values of evaluated at
the cell surface (edges and ) of the finite volume as follows:

where .

We can therefore derive a semi-discrete numerical scheme for the above problem with

cell centres indexed as , and with cell edge fluxes indexed as , by differentiating
(6) with respect to time to obtain:

where values for the edge fluxes, , can be reconstructed by interpolation or


extrapolation of the cell averages. It should be noted that equation (7) is exact for the
volume averages; i.e., no approximations have been made during its derivation.

General hyperbolic problem


We can also consider a general hyperbolic problem, represented by the following PDE,
Here, represents a vector of states and represents the corresponding flux tensor. Again
we can sub-divide the spatial domain into finite volumes or cells. For a particular cell, ,
we take the volume integral over the total volume of the cell, , which gives,

On integrating the first term to get the volume average and applying the divergence
theorem to the second, this yields

where represents the total surface area of the cell and is a unit vector normal to the
surface and pointing outward. So, finally, we are able to present the general result
equivalent to (7), i.e.

Again, values for the edge fluxes can be reconstructed by interpolation or extrapolation of
the cell averages. The actual numerical scheme will depend upon problem geometry and
mesh construction. MUSCL reconstruction is often used in high resolution schemes
where shocks or discontinuities are present in the solution.

Finite volume schemes are conservative as cell averages change through the edge fluxes.
In other words, one cell's loss is another cell's gain!

Finite difference method.


This method has historical importance and is simple to program. It is currently only used
in few specialized codes. Modern finite difference codes make use of an embedded
boundary for handling complex geometries making these codes highly efficient and
accurate. Other ways to handle geometries are using overlapping-grids, where the
solution is interpolated across each grid.

Where Q is the vector of conserved variables, and F, G, and H are the fluxes in
the x, y, and z directions respectively.

Boundary element method. The boundary occupied by the fluid is


divided into surface mesh.
High-resolution schemes are used where shocks or discontinuities are present. To
capture sharp changes in the solution requires the use of second or higher order numerical
schemes that do not introduce spurious oscillations. This usually necessitates the
application of flux limiters to ensure that the solution is total variation diminishing.

Introduction to ANSYS CFX


ANSYS CFX is general purpose Computational Fluid Dynamics (CFD) software suite
that combines an advanced solver with powerful pre- and post-processing capabilities.

ANSYS CFX is capable of modeling:


• Steady-state and transient flows
• Laminar and turbulent flows
• Subsonic, transonic and supersonic flows
• Heat transfer and thermal radiation
• Buoyancy
• Non-Newtonian flows
• Transport of non-reacting scalar components
• Multiphase flows
• Combustion
• Flows in multiple frames of reference
• Particle tracking.

ANSYS CFX includes the following features:

• An advanced coupled solver that is both reliable and robust.


• Full integration of problem definition, analysis, and results presentation.
• An intuitive and interactive setup process, using menus and advanced graphics.

The Benefits of Fluid Simulation


from ANSYS
ANSYS fluid flow analysis technology allows for an in-depth analysis of the fluid
mechanics in many types of products and processes. Not only does it reduce the need for
expensive prototypes, it provides comprehensive data that is not easily obtainable from
experimental tests. Fluid simulation can be used to complement physical testing. Some
designers use it to analyze new systems before deciding which validation tests, and how
many, need to be performed. When troubleshooting, problems are solved faster and more
reliably because fluid dynamics analysis highlights the root cause, not just the effect.
When optimizing new equipment designs, many what-if scenarios can be analyzed in a
short time. This can result in improved performance, reliability and product consistency.
ANSYS will continue to innovate and integrate so that customers can replace more of
their traditional capital-intensive design processes with a Simulation Driven Product
Development method.

The ANSYS Advantage


ANSYS is a major product for computer based Prototyping. While material Prototyping
is as old as mankind, the computerized approach is relatively new and has on offer certain
major advantages over the classical approach. In the following some info in regard.

Before we start using or selling a actual product we want to make sure it works properly
w/o premature failures.

The earliest method to achieve a working design was (and is) 'make and brake'. If it
doesn't break it is a valid design, otherwise we need to think it over, to redesign wrt
dimensions, material, and manufacturing method - remove evident design flaws. Make
and break does have some draw backs. The make of a new prototype may be costly and
time consuming, while still not indicating with certainty were a flaw really hides..

In the last decades a new method came to maturity, which adds more flexibility, cost
efficiency and, especially, more insight into prototyping. This more recent method is
known under names like FEA (finite element analysis), simulation or virtual (digital)
prototyping, and moves the material prototype, the experimental verification, into the
computer.

Simulation is a great step ahead, allows precise prognosis and optimization of the
performance of parts and products, with all the flexibility of a computerized model -
changing dimensions, materials, loads w/o the necessity to create a new material
prototype. The range of physics problems that can be analyzed is basically unlimited - be
it mechanical or thermal loads, be it a fluid dynamic question, a acoustics setting or
electromagnetic device, simulation can handle it (keyword multiphysics).

One other great advantage, besides the flexibility of computer based prototyping is the
ability to evaluate the designs response in any location of the part. While material
prototyping needs sensors to measure quantities like strain, temperature, flow speed or
another physical quantity (if you are able to arrive with the sensor in a certain position,
and the sensor doesn't disturb the measurement by it's very presence) simulation models
provide these quantities w/o effort at every time and point on the model (section the
model at will, plot contours of quantities, graph, list, i.e. view results any way suits).

How does one build a computerized simulation model? Basically the part(s) of interest
are modeled in a 3D CAD system, usually. This CAD is then further processed in a
'preprocessor', i.e. loading and any other material data, physical data and the like are
specified, further 'solution' settings for solving the numerical model are defined. After
solving the model it gets 'post-processed' in that results are accessed for verification and
reporting.

The great advantage of computer based prototyping is evident: modifications (e.g.


updating material data, changing some dimension) and extensions (adding a new load
scenario or a result quantity to evaluate) to the actual model are possible at any stage of
simulation.

CONCLUSION
REFERENCES

1. Rajput R. K., Fluid Mechanics and Hydraulic Machines, 2006, Dhanpat Rai
Publications, Page no. 725 to 753
2. Jain R. K. , Production Technology, 2003, Khanna Publishers, Page no. 138 to 148

3. Guidelines from “Metal & Casting Industry, Madan Mehel , Jabalpur”

4. http://www.engineeringtoolbox.com/

You might also like