Sei sulla pagina 1di 35

Introduction To Aerospace

Engineering.
Part-A

Written by
Venus Kumar (B.tech Aerospace).
Supported by
Sonu Kr. Gupta (B.Tech
Aerospace).
Reference by
Google and Wikkipedia
(Sometimes).

Historical Evaluation.
History of Aviation.
The history of aviation has extended over more than two thousand years from the earliest
attempts in kites and gliders to powered heavier-than- air, supersonic and hypersonic
flight.
The first form of man-made flying objects were kites. The earliest known record of kite
flying is from around 200 BC in China, when a general flew a kite over enemy territory
to calculate the length of tunnel required to enter the region. Yuan Huangtou, a Chinese
prince, survived by tying himself to the kite.
Leonardo da Vinci's (15th c.) dream of flight found expression in several designs, but he
did not attempt to demonstrate his ideas by actually constructing them.
With the efforts to analyze the atmosphere in the 17th and 18th century, gases such as
hydrogen were discovered which in turn led to the invention of hydrogen balloons.[1]
Various theories in mechanics by physicists during the same period of time, notably fluid
dynamics and Newton's laws of motion, led to the foundation of modern aerodynamics.

Early Development of Airplanes.


Biplanes and Monoplanes.
Biplane.

A biplane is a fixed-wing aircraft with two main wings stacked one above the other.
Biplanes are distinguished from tandem wing arrangements where the wings are placed
forward and aft, instead of above and below.

Biplane

In a biplane aircraft, two wings are placed one above the other. Each provides part of the
lift, although they are not able to produce twice as much lift as a single wing of similar
size and shape because the upper and the lower are working on nearly the same portion of
the atmosphere and thus interfere with each other's behaviour. For example, in a wing of

aspect ratio 6, and a wing separation distance of one chord length, the biplane
configuration will only produce about 20 percent more lift than a single wing of the same
planform.
In the biplane configuration, the lower wing is usually attached to the fuselage, while the
upper wing is raised above the fuselage with an arrangement of cabane struts, although
other arrangements have been used. Either or both of the main wings can support
ailerons, while flaps are more usually positioned on the lower wing. Bracing is nearly
always added between the upper and lower wings, in the form of wires (tension
members) and/or slender interplane struts positioned symmetrically on either side of the
fuselage.

Advantages and disadvantages


1. A biplane uses two main wings, so it provides more lift than a monoplane.
2. On the other hand, it offers more drag as it uses two wings, as compared to a single
wing in a monoplane.
3. The interference drag between the two wings of a biplane adds up in the total drag.
4. A biplane requires less power to fly than a monoplane.

Stagger.
Biplanes were originally designed with the wings positioned directly one above the other.
Moving the upper wing forward relative to the lower one is called positive stagger or,
more often, simply stagger. It can help increase lift and reduce drag by reducing the
aerodynamic interference effects between the two wings. Example-Waco Standard Cabin
series (1930s).
It is also possible to place the lower wing's leading edge ahead of the upper wing, giving
negative stagger.Example-Sopwith Dolphin and Beechcraft Staggerwing.

Bays.
The space enclosed by a set of interplane struts is called a bay, hence a biplane or triplane
with one set of such struts connecting the wings on each side of the aircraft is a singlebay biplane. This provided sufficient strength for smaller aircraft.

Sesquiplane.
The sesquiplane is a common variation on the biplane where one wing (usually the lower)
is significantly smaller than the other either in span, chord or both. The name means
"one-and-a-half wings." The arrangement may reduce interference drag between the
wings whilst retaining the biplane's structural advantage. Example-Nieuport military
aircraft from the Nieuport 10 through to the Nieuport 27.

Sesquiplane

Monoplane.
A monoplane is a fixed-wing aircraft with one main set of wing surfaces, in contrast to a
biplane or other multiplane.

Monoplane.

Support and Weight.


The inherent efficiency of the monoplane can be realised in the unbraced cantilever wing
which carries all structural forces internally. By contrast the braced wing gains additional
drag from the exposed bracing struts and/or wires. On the other hand, the braced wing
has greater structural efficiency and can be made much lighter. This in turn means that for
a wing of a given size, bracing allows it to fly slower with a lower-powered engine, while
a heavy cantilever wing needs a more powerful engine and can fly faster.
A cantilever wing can be made lighter by making it thicker. This increases internal
storage for fuel, retractable undercarriage, armaments and in some cases even passengers
and crew.

Wings Position.
Low wing:- A low wing is one which is close to the bottom of the fuselage.
Placing the wing low down frees up the central fuselage from the wing spar carry-through
structure and also allows good visibility upwards. By reducing pendulum stability it
makes the aircraft more manoeuvrable. It also allows a lighter structure because the
fuselage sides carry no additional loads and the main undercarriage legs can be made
shorter. It is common on jet airliners.
One problem with the low wing position is that on landing the ground effect is especially
strong, giving the plane a tendency to float a long way before it can touch down.

Low wing on a Curtiss P-40

Mid Wing:- A mid wing is mounted mid-way up the fuselage. It is aerodynamically the
cleanest and most balanced, but the carry-through spar structure can reduce the useful
fuselage volume near its centre of gravity, where space is often in most demand. It is
common on high-performance types such as sailplanes.

Mid wing on a de Havilland Vampire

Shoulder Wing:-A shoulder wing is mounted between the midpoint and the top of the
fuselage. It is sometimes classed as a type of high wing.
Placing the wing above the midpoint increases the useful volume beneath and also allows
good visibility downwards. It provides pendulum stability by placing the centre of gravity
below the centre of pressure. It also gives increased ground clearance for the propellers of
multi-engined aircraft, for underwing stores and for ground handling. Many light
transport aircraft have shoulder or high wings.

Shoulder wing on an ARV Super2.

High Wing:-A high wing has its upper surface close to or above the top of the fuselage.
It has all the advantages and disadvantages of the shoulder wing, and even more so.
On light aircraft the wing is sometimes located on top of the pilot's cabin.

High wing on a de Havilland Canada Dash 8.

Parasol Wing:- A parasol wing is a high wing which forms a separate structure above
the fuselage and is not directly attached to it, with structural support being provided by
either several cabane struts or a single wider pylon. Additional bracing has to be provided
by underwing struts extending either side of the fuselage.
The supports and bracing create extra drag so the parasol wing is not often used.
Compared to a biplane it has less bracing and lower drag, and the parasol wing was
popular only during the inter-war transition years between biplanes and monoplanes.

Parasol wing on a Pietenpol Air Camper.

Monoplane v/s Biplane.


The wing of a monoplane must nominally have twice the area of the equivalent biplane wing,
because the biplane has two of them. This makes the monoplane larger and less manoeuvrable.
However in practice the biplane's wings interfere with each other and the handicap to the
monoplane is reduced.
A biplane wing is usually braced to stiffen the structure and allow it to be much lighter and to fly
slower. However even a braced monoplane will still be more efficient and human-powered

aircraft, which are among the slowest and lightest of flying machines, are monoplanes with very
large wings for their weight.

History of Spaceflight.
Development of Space Vehicle.
Classification of Duct Jet Propulsion.
Rocket Propulsion.
Rockets.
Definition. A rocket is defined as an engine or motor that develops thrust by ejecting a stream of
matter rearward, or the missile or vehicle powered by such an engine. Since the reaction
principle involved assumes a self-contained source of energy, a rocket can operate in any
medium including space outside the earths atmosphere, where there is no oxygen to support
combustion. History. The Chinese are generally given credit for inventing the rocket because
they appear to be the first to have employed black powder or solid rockets as weapons of war,
somewhere between 1150 and 1350 a.d.. The Chinese attached a small rocket to the shaft of an
arrow to extend its range (Fig. 1). The early history of rocket development is linked to their use
as weapons; references to rockets as weapons appear from the fourteenth to eighteenth centuries.
The technology spread to Europe and came to the attention of William Congreve of the Royal
Laboratory at Woolrich, England. His stick-stabilized rocket designs were adopted by the British
Navys arsenal and were employed in the attack on Fort McHenry in Baltimore, Maryland, that
gave rise to the rockets red glare phrase in the Star Spangled Banner.
Modern treatments of rockets and spaceflight focus on the contributions of four men, Konstantin
Eduardovich Tsiolkovsky (1857-1935), Dr. Robert Goddard (1882-1945), Dr. Hermann Oberth
(1894-1989) and Dr. Wernher Von Braun (1921-1977).
The application of rocket technology to concepts of spaceflight originated with Konstantin
Eduardovich Tsiolkovsky. It first appeared in 1903 in his treatise The Investigation of Outer
Space by Means of Reaction Apparatus.A theoretician, his proficiency in mathematics and
science enabled him to foresee and address such issues as escape velocities from the earths
gravitational field,gyroscopic stabilization, and the so-called rocket equation that establishes
the relationship of the velocity increment added to a vehicle in terms of the exhaust velocity of
the rocket device and the initial and final masses of the vehicle.

Figure 1. Chinese fire arrow

Dr. Robert Goddard was an American experimentalist who began his experiments in rocketry as
a doctorate student at Clark University in Worcester, Massachusetts. His work was not widely
accepted during his time. His report A Method of Reaching Extreme Altitudes, which offered
scientifically sound concepts such as travel to the Moon, was criticized by the general public and
in the press. Dr. Goddard built and successfully launched the first liquid propellant rocket on 16
March 1926. (Fig. 2). He went on to conduct additional experiments in Roswell, New Mexico,
under the sponsorship of Daniel Guggenheim. He introduced the practical application of such
concepts as gyroscopic stabilization of the rocket vehicle and movable deflector vanes in the
rocket exhaust for directional control. He held 214 patents in rocketry.
Dr. Oberth was a physicist who wrote The Rocket into Interplanetary Space in 1923 to espouse
his theories of space travel. Among his theories was the concept of staging to achieve higher
velocities. His writings inspired Wernher Von Braun who later assisted Oberth in liquid rocket
experiments. Von Braun eventually applied his knowledge to constructing liquid-fueled rocketpowered weapons during World War II. The launch of an A-4 rocket to an altitude of 50 miles
(the altitude at which space is considered to begin) on 3 October 1942 might well be considered
the beginning of the space age. Following the war, Von Braun and his Peenemunde team was
reconstituted in the United States under the U.S. Army missile program, where its focus was
again liquid-fueled rocketry. As Dr. Von Braun was developing the next generation ballistic
missile, the Redstone, the United States was investing in the development of the Vanguard space
launcher as a civilian launch vehicle. It was intended to inaugurate the U.S. exploration of space.
The government intentionally avoided the use of Von Brauns missiles as launch vehicles to
emphasize the peaceful intent of space exploration. The Vanguard experienced a spectacular
launch pad failure in December 1957 that forced the government to turn to Dr. Von Braun in the
aftermath of the successful U.S.S.R. Sputnik launch. His successful January 1958 launch of the
Explorer 1 atop a Jupiter C launch vehicle, based on missile technology, brought Dr. Von Braun
into the nations spotlight and resulted in his becoming the leading U.S. figure in space launch
technology and space exploration. The missile technology base, thus, became the foundation for
the development of space launch vehicles. As the missions became more ambitious, the need

emerged for higher energy reactants than the liquid oxygen/RP-1 (kerosene) typically used, and
the cryogenic system of liquid oxygen and liquid hydrogen became the standard for the civilian
space launch capability. The Saturn V was the first launch vehicle developed that was not based
on a vehicle or rocket engine developed for weaponry. It used liquid oxygen and liquid hydrogen
in the second and third stages. Presently, the Space Shuttle, the French Arianne V, and the
Japanese H-II vehicles all use hydrogen and oxygen for space launch applications.

Figure 2. The Goddard rocket.

Rocket-powered manned flight began in the 1930s in Germany. The Heinkel 176 was the first
aircraft solely powered by a rocket engine. It used a 1320 lbf (297 N) engine that ran on
decomposed hydrogen peroxide. In 1943, the Messerschmitt 163B rocket fighter became the first
operational rocket-powered fighter aircraft. It had dual combusters that gave a total thrust level
of 4400 lbf (989 N) and operated on hydrogen peroxide oxidizer and a fuel mixture of hydrazine
hydrate, methyl alcohol, and water. One chamber could be shut down to achieve throttling down
to 660 lbf (148 N). The U.S. developments in rocket-powered manned flight began with a rocketpowered flying wing, the Northrop MX 334 and eventually led to the Bell X-1 in which Captain
Charles Yeager broke the sound barrier in October 1947. Ever improved experimental aircraft
were flown in attempts to increase the flight speed and altitude achievable. Mach 2 was exceeded
in November 1953 in a Navy Douglas D-558-II and the record was increased to 2.5 only 22 days
later by Captain Yeager in the Bell X-1A. The North American Aviation X-15 dominated highspeed and high-altitude research in the late 1950s and 1960s. Its 60,000-lbf (13,489 N) thrust
XLR-99 engine took the X-15 to new international records for speed (Mach 6.7) and altitude
(354,200 feet). Many of the lessons learned from these experimental manned aircraft were
subsequently applied in designing, constructing and operating the Space Shuttle orbiter.
Solid propellant rocket motors evolved in a role as strap-on boosters to liquid stages for many

space launch applications. Solid propellant rocket history was closely linked to weaponry for its
development. The first large high-performance motor design was for the Polaris submarinelaunched ballistic missiles where the requirements for storability and logistics of shipboard
operations made the solid propellant rocket very attractive. These same requirements led to its
use in the Minuteman ballistic missile. These missile developments provided the base upon
which the technology for very large solid propellant rocket boosters, suitable for space launch
vehicles, was built. The industry explored yet even larger solid propellant rocket configurations
and found that their efforts were constrained by the size of the mixing facilities and the ease of
transport of the rockets. These problems were overcome by using a segmented construction
method in which the solid propellant rocket was cast in sections or segments, cured, then
assembled into the flight vehicle at the launch site. A segmented, 156-inch-diameter solid
propellant motor demonstrated the feasibility and practicality of segmented, solid propellant
rocket motors, a concept employed as a strap-on booster for the Titan III launch system. This led
the way to using large strap-on solid propellant rockets for the recoverable solid propellant
rocket motor boosters for the Space Shuttle. A parallel development of great significance was the
movable nozzle that facilitated thrust vector control for solid propellant rockets. Previously, the
thrust vector for large solid propellant motors was controlled by injecting liquids into the nozzle
sidewall, generating side forces by the shock caused by the interaction of the liquid jet with the
supersonic nozzle flow.

Governing Laws.
The operation of rocket engines and motors and the vehicles that they propel are primarily
governed by Newtons laws of motion.
Newtons first law, often called the law of inertia, states that there is no change in the motion of a
body unless a resultant force acts on it. A number of forces act on a launch vehicle throughout its
flight. The gravitational force (weight of the vehicle), lift, drag, and the thrust of the rocket
engine all act on the vehicle to cause the resultant motion. The net amount of the resultant force
and its direction determine the acceleration on the vehicle and the path of the flight trajectory, in
accordance with Newtons second law.
Newtons second law of motion states that whenever a net (unbalanced) force acts on a body, it
produces an acceleration in the direction of the force; the acceleration is directly proportional to
the force and inversely proportional to the mass of the body:

This relationship is more typically seen in the form,

As stated in the definition, a rocket develops its thrust by expelling a mass rearward. Examining

this in the context of this equation, a mass is accelerated rearward by some means that
accelerates its velocity from near zero to thousands of meters per second. The force for this
acceleration is proportional to the mass of the exhaust gases and the acceleration per Newtons
second law. The force acting on the exhaust gases is in the direction of the accelerating mass, but
produces a thrust in accordance with Newtons third law, which states that for every acting force,
there is a reacting force that is equal in magnitude but opposite in direction. Therefore, the force
of accelerating the fluid internal to the rocket has an equal but opposite external force which is
the thrust produced by the rocket.
The magnitude of this force (thrust) can be determined by examining the change of momentum
in the device and the sum of the forces that act on a closed duct or control volume, as shown in
Fig. 3. The flow internal to the rocket experiences a change of momentum that is equal to the
mass flow rate times the change in velocity of the gases:

Assuming that the inlet velocity into the device is low, the momentum at the inlet can be
considered negligible. Thus, the momentum change is

Figure 3. Pressure balance on the rocket chamber and nozzle wall.

The sum of all pressures on the surfaces perpendicular to the flow axis of the device reduces to a
resultant force due to the pressure differential between the pressure at the nozzle exit plane and
the ambient pressure that acts on the exit area of the nozzle:

The sum of the forces that act on a rocket is equal to the change of momentum in accordance
with Newtons second law. By combining Equations 4 and 5 and rearranging the terms, we
develop the following expression for the thrust of a rocket:

When pexit equals pamb, expansion is optimum and performance best. When the nozzle exit
pressure is less than ambient, the nozzle is said to be over-expanded. If exit pressure is greater
than ambient, the nozzle is said to be under-expanded. Because the rocket, generally, flies
through the atmosphere, it experiences variations in the ambient pressure, so it operates at
optimum expansion at only one altitude. The choice of the rocket exit area ratio then becomes the
result of trading off a number of design and flight considerations.
Newtons laws are applied in analyzing the acceleration of a vehicle propelled by a rocket as
well. Examining vehicle flight in a vacuum free of gravitational forces, the rocket produces an
unbalanced force and a resultant acceleration in accordance with Newtons second law. Here, it is
written in a way slightly different from that in Equation 2. The thrust is the net accelerating force
that is equal to the instantaneous mass of the vehicle, and its instantaneous acceleration is written
as the time rate of change of the velocity:

The thrust of the rocket F can also be expressed in terms of the mass flow rate m from the rocket
and its effective exhaust velocity Ve assuming that nozzle exit pressure equals the ambient from
Equation 6:

where

The negative sign indicates that the mass of the vehicle is decreasing as the propellant exits the
engine.
Substituting for thrust F from Equation 8 and mass flow rate m from Equation 9 and rearranging,

Integrating, we obtain the result that is commonly called the rocket equation that is used to
calculate the velocity increment added to a stage in terms of its effective exhaust velocity and its
initial and final masses:

This solution assumed flight in a vacuum free of any gravitational field; thus, the value
calculated is an ideal velocity increment. Introducing the effects of atmospheric drag and gravity
result in reducing the velocity increment achieved. The gravitational field has a component of
force acting along the flight path of the vehicle (g cosine 8). The net loss due to the gravity field
is computed by integrating this component along the path during the flight in the form of the
equation,

The drag component is introduced as a drag coefficient Cd that is applied to the incompressible
flow dynamic pressure 1/2PV2 and the cross-sectional area of the vehicle and evaluated along
the flight path:

The drag coefficient is a function of the vehicle flight Mach number and its angle of attack (Fig.
4).
Finally, the velocity increment added to a single stage during flight can be determined from the
equation,

Figure 4. Drag coefficient vs. Mach number as a function of angle of attack.

The initial mass of a stage includes the inert mass of the vehicle (structural dry mass plus

engines), the mass of propellants and pressurant gases, and the pay-load mass:

The final mass of the stage includes the dry weight of the vehicle, any residual (unburned)
propellant and pressurants, and the payload:

The rocket equation can be applied to multiple-stage vehicles by solving the equation for each
stage and adding the velocity increments for all stages as follows:

Care must be exercised in considering the masses of the individual stages. The payload for the
first stage of a multiple-stage vehicle is the mass of the stages above it. Only in the final stage is
the payload the mass that is placed in orbit.

Specific Impulse.
The usual measure of performance of a rocket engine is its specific impulse. This is a measure of
how much thrust (lbf or Newtons) is generated by the engine when the flow rate is 1 unit
(lbm/sec or kg/s):

The units of specific impulse are seconds in SI but lbf/lbm/sec in the U.S. Customary units
(USCS). Seconds is still used as the terminology for specific impulse in the USCS system but
is not correct technically and must have the correct units for use in solving any equations.
Specific impulse is related to the exit velocity of the nozzle Ve through the relationship,

Specific impulse can also be determined as a function of the properties of the working fluid and
the operating conditions of the rocket nozzle. In simplified terms, we see that the specific

impulse is proportional to the square root of the temperature of the working fluid divided by its
molecular weight:

The constant in this equation is a function of the thermodynamic properties of the combustion
gases and the ratio of nozzle exit pressure to combustion chamber pressure. Therefore, the
objective in any rocket engine is to achieve the highest possible temperature of the working fluid
and the lowest possible molecular weight. In chemical rockets, this is typically achieved by using
reactants that produce large quantities of hydrogen or steam at high temperature as their
products. Beamed energy and nuclear rockets choose hydrogen for the working fluid because it is
a good coolant and has the desired low molecular weight. The choice of a working fluid for
electric rockets depends, in some cases, on other properties of the fluid such as ionization
potential. Hydrogen is still of interest for an arcjet.

Vehicle Staging.
The ideal velocity increment equation can be manipulated into the following form to examine the
sensitivity of the stage performance to various aspects of rocket engine performance and stage
design:

From this equation, we can see that the higher the engine specific impulse, the more closely the
final mass approaches the initial mass of the stage (less pro-pellant is consumed).
A measure of stage design efficiency is the propellant mass fraction l defined by the equation,

The higher the value of l, the greater the structural efficiency of the stage. Now, we can
manipulate Equations 18 and 19 into a form that presents the ratio of the initial mass of the
vehicle to the payload delivered in terms of the propellant fraction, specific impulse, and ideal
velocity increment:

We can illustrate the sensitivity of stage and vehicle performance to engine specific impulse and
stage structural efficiency by conducting a parametric study of two vehicles, single-stage-to-orbit
(SSTO) and two-stage-to-orbit (TSTO). First, one must estimate the ideal velocity increment for
the mission. A low Earth orbit typically requires about 30,000 ft/sec (9144 m/s). In the SSTO
case, three specific impulses were assumed, one representative of liquid hydrogen/liquid oxygen
rocket performance (460 s), a second chosen as a conceivable increase in chemical rocket
performance (500 s) and the third representative of predicted nuclear rocket performance (850 s).
By assuming different values for the propellant fraction l, one can then solve Equation 23 and
arrive at the family of curves presented in Fig. 5. The lower the value of the lift-off mass to
payload ratio, the better the vehicle performance. The chemical rocket options approach a limit at
a value in the vicinity of 10 (10% of the vehicle liftoff mass is effective payload delivered to
orbit), whereas the nuclear option approaches a limit in the vicinity of about 3 (33% of liftoff
mass). The lower specific impulse vehicle, is seen as quite sensitive to the propellant mass
fraction. For example, at a propellant fraction of 0.89, the 40-second difference in specific
impulse between 460 s and 500 s results in a twofold difference in the initial weight to payload
ratio. So, if a vehicle design were to experience an increase in weight from its initial design to
completion of fabrication (and they most always do), the payload capacity of the vehicle will be
severely reduced and, conceivably, could become inconsequential, depending upon the
magnitude of the growth in weight. For the higher specific impulse cases, the SSTO vehicle is
seen as less sensitive to any decrease in propellant mass fraction, and the nuclear rocket is
virtually insensitive to propellant mass fraction across the range studied. This illustrates that the
higher the delivered specific impulse, the less sensitive the stage design is to its structural
efficiency.
Another way in which the initial mass/payload ratio can be made less sensitive to the propellant
mass fraction is through the concept of staging in which two (or more) rocket stages can be
coupled, either in parallel or in tandem. As one stage is jettisoned after exhausting its propellant
load, the subsequent stage ignites and continues on to complete the mission. Figure 6., presents
the results of a parametric study to illustrate the effects of specific impulse and propellant
fraction upon the lift-off mass to payload ratio for a tandem two-stage vehicle that has an ideal
velocity increment of 30,000 ft/s (9144 m/s). In this case, it was assumed that the propellant
mass fraction was the same for each stage, and the total propellant load was distributed, 80% in
the first stage and 20% in the second stage. As in the SSTO case, there is a limit of initial to
payload mass ratio of about 10. The maximum ratios calculated are 16 for a specific impulse of
460 s and 12 for a specific impulse of 500 s within the range of propellant mass fraction studied,
as contrasted with values of 74 and 25, respectively, for the SSTO. It is easily seen that multiple
stages can offer payload gains. The driver for minimizing the number of stages is cost
considerations rather than performance.

Figure 5. Payload/initial mass vs. propellant mass fraction for SSTO.

Figure 6. Payload/initial mass vs. propellant mass fraction for TSTO.

Energy and Energy Conversion.


Energy Conversion Mechanisms in a Rocket. The rocket engine is an energy conversion device
that converts potential energy to the thermal energy of a high-temperature gas and then to the
kinetic energy of a high-velocity exhaust gas. The energy sources for this conversion can be any
of several types, chemical, electrical, beamed (solar or laser), or nuclear.
Chemical Rockets. In a chemical rocket, the potential energy is in the chemical bonds of the
molecules that compose the fuel and the oxidizer. The class of reactants identified as fuels has
large concentrations of hydrogen or carbon in the molecule and light metals such as aluminum,
lithium, beryllium, magnesium, and boron. The class of reactants identified as oxidizers has large
concentrations of oxygen, fluorine, or chlorine in the molecule. When the fuel reacts with an
oxidizer in the combustion process, the resultant products are at a temperature significantly
elevated over that at which they entered. This thermal energy is extracted for propulsion (thrust)
by passing the gases through a converging-diverging (or DeLaval) nozzle that converts the
thermal energy to kinetic energy. These chemical reactants are usually in either liquid or solid

form.
A liquid-fueled rocket might be of either bipropellant or monopropellant type. In the former, fuel
and oxidizer are introduced into the combustion device separately. They are atomized, vaporized,
mixed, and combusted in the combustion chamber. The mixture ratio (proportions of oxidizer to
fuel flow rate) at which the maximum combustion temperature occurs is called the stoichiometric
ratio. This ratio does not yield the highest specific impulse (the ratio of thrust produced by the
engine to the rate at which propellant is being consumed), however; this occurs when using fuelrich mixtures. The excess fuel tends to reduce the combustion temperature somewhat, but more
importantly, the molecular weight ofthe combustion products is reduced. The net effect is to
increase the specific impulse, the measure of goodness of rocket performance. The most
common oxidizers used to date are liquid oxygen, nitric acid (HNO3), and nitrogen tetroxide
(N2O4). The most common fuels used to date have been liquid hydrogen, RP-1 (a kerosene-like
hydrocarbon), and various amine-based fuels (hydrazine, unsymmetrical dimethyl hydrazine,
monomethyl hydrazine, and mixtures of the foregoing). Liquid propellants can also be classified
as cryogens or storables. Cryogens are liquefied gases, typically oxygen and hydrogen, and are
more energetic reactants. Within the family of cryogens, there are some propellants referred to as
space storable, meaning that they are relatively mild cryogens (boiling point greater than
238F( 150 K). These propellants can be stored in space for long periods of time and have
acceptable levels of boil-off without engaging in major efforts to insulate the propellant
containers (tanks) or refrigerate the propellants. Storable propellants are those that normally are
liquids at standard temperature and pressure and are less energetic.
Monopropellants are generally introduced into a catalyst bed or pack where the propellant
molecule is broken apart, liberates energy, and consequently increases in thermal energy. These
reactions tend to be less energetic than bipropellant reactions and are frequently used in selected
applications where specific impulse is not the most important characteristic. The most commonly
used monopropellants are hydrazine (N2H4) and hydrogen peroxide (H2O2). The catalyst for
hydrazine is iridium coated on an alumina substrate. The catalyst for hydrogen peroxide is silver,
generally in the form of a screen.
The selection of a bipropellant combination depends on many factors, for example, the ignition
characteristics of the propellant combination. Hypergolic combinations (nitrogen tetroxide and
amine fuels, for example) provide a ready ignition source because the fuel and the oxidizer react
upon contact. Such combinations are amenable to applications where engine restart may be
required or for ignition at an altitude. Combinations such as hydrogen and oxygen require an
external source of energy to provide the ignition energy for the main propellant flow. Multiple
restarts are possible through appropriate design, as used on the RL-10 engine of the Centaur
upper stage vehicle.
Another consideration for propellant selection is the choice of coolant for the rocket chamber.
Specifically, the coolant should have a very high specific heat, low viscosity, and be thermally
stable. Typically, fuels are used as engine coolants. Several designs have been tested in which
oxidizers (liquid oxygen or nitrogen tetroxide) have been used as coolants.
The solid-propellant rocket motor is a device in which the fuel and oxidizer are premixed to form
a combustible mixture which, when ignited, burns until all of the propellant is consumed. It
burns at a rate that depends on the combustion pressure p and is a function of the propellant type

(burn rate exponent n) according to the following relationship:

The burning rate r generally ranges in value from 0.3-0.5 inches per second (0.762 to 1.27 cm/s),
and the burning rate exponent n ranges from 0.2-0.5 for modern propellants. For a detailed
description of solid-propellant rockets, see the article Solid Fuel Rockets by Donald Sauvigeau
on page 531 of this topic.
The hybrid is another chemical rocket embodiment. It typically consists ofa solid fuel grain and a
liquid (or gaseous) oxidizer. Its advertised advantages are that the fuel grain is less susceptible to
safety and handling problems because the grain contains no oxidizer and, thus, cannot detonate
or sustain combustion by itself. Additionally, using a liquid (or gaseous) oxidizer provides a very
simple throttling scheme through oxidizer flow regulation and also provides for thrust
termination by simply stopping the oxidizer flow. The fuel grain is typically designed to have
several flow passages or ports through it to maximize the fuel surface exposed to the oxidizer,
and, consequently, the burning surface and the attendant flow of combustion products. The
hybrid has also achieved some attention because it permits eliminating the ammonium
perchlorate oxidizer typical of most solid-fueled rockets and the attendant HCl in the exhaust
products. The HCl presents a small but persistent environmental concern. The burn rate of a
hybrid motor is similar in form to that of the solid-propellant rocket:

In this case, r is the burn rate, a is a constant that depends on the reactants, Go is the oxidizer
mass flow rate per port, and n is the burning rate exponent whose value is from 0.5 to 0.7.
Hybrid rockets have been successfully fired in motors of up to 250,000 lbf (1.11 x 106N). The
problems that exist are achieving high burning rates to minimize the size of the grain and
complete consumption of the grain. Residual propellant can amount to 5 to 30% of the initial fuel
load depending on the port configuration.
Nonchemical Rockets. In an electric rocket, the energy is collected from some power source
(nuclear or solar) and converted to electrical energy. The energy conversion for thrusting
purposes may occur by striking an arc between an anode and a cathode and passing a working
fluid through the arc to heat it (arcjet). The hot gases are then passed through a nozzle to convert
the thermal energy to kinetic energy. Another energy conversion technique might be to use the
electrical energy to ionize an easily ionized gas (xenon, for example) and accelerate the resultant
ionized particles across a potential (electrostatic thruster). Another approach may use the
interaction of the current and its induced electromagnetic field that produces Lorentz forces to
accelerate the gas (electromagnetic thruster).
Beamed energy depends on capturing a high-energy beam from outside the vehicle and
transferring its energy to a working fluid. Solar energy is one such beaming mechanism.
Concentrating lenses (or mirrors) located on the vehicle capture the solar radiation and focus it
on a blackbody absorber at the focal point. The absorbed energy is then used to heat a working
fluid that is subsequently expelled through the nozzle at a high velocity. A laser beam might

apply equally for this purpose. The working fluid also cools the absorber/thruster structure to
maintain structural integrity at elevated temperatures.
Finally, a nuclear energy source may be used in place of any of the previously mentioned energy
sources to heat the working fluid. Here also, the working fluid serves the dual purpose of cooling
the structure and generating the thrust, as it is expelled through a nozzle.

Thermodynamics of Rockets
As previously discussed, rocket exhaust velocity is related to the specific impulse through the
gravitational constant gc as follows when the nozzle exit pressure is equal to the ambient:

The exhaust velocity must be maximized to maximize the specific impulse. The first law of
thermodynamics (conservation of energy) for bulk flow requires that the total energy entering the
engine must equal the total energy leaving the engine. In equation form this is:

The average stagnation enthalpy per unit mass ho of all of the flows entering the engine equals
the kinetic energy at the nozzle exit plus the static enthalpy h of the exhaust at the exit. The static
enthalpy for a chemically reacting system is given in terms of the static temperature T, entropy s
and Gibbs free energy f by the equation,

The Gibbs free energy is a summation of the chemical binding energy. Substituting and
rearranging, the kinetic energy can be expressed as a function of the stagnation enthalpy Ts and
the Gibbs free energy:

By differentiating the kinetic energy with respect to entropy s and Gibbs free energy f, we can
find that the exhaust velocity for a given exit pressure increases as exit entropy decreases and
Gibbs free energy decreases. From the second law of thermodynamics, the minimum exit entropy
is equal to the inlet entropy for bulk flow. Therefore, the maximum exhaust velocity occurs for
an exit entropy per unit mass that is equal to the inlet entropy per unit mass. Additionally, the
minimum Gibbs free energy occurs under conditions of chemical equilibrium. Chemical
equilibrium is achieved as the chemical constituents of the exhaust gases alter their relative
proportions in response to the pressure and temperature changes that occur as the gases flow

through the nozzle. Energy is liberated to the expansion process as a result. Therefore the
maximum specific impulse is achieved for an ideal one-dimensional flow when the exhaust
products are expanded to ambient pressure in chemical equilibrium and have total enthalpy and
entropy equal to the inlet conditions.

Figure 7. Enthalpy vs. entropy for flow in a rocket engine.

The reality of rocket operation is that none of the processes occurs ideally. The combustion
process is irreversible and has entropy increases attendant to it. Viscous boundary layer effects
and nonequilibrium expansion of the combustion products also introduce irreversibilities. Figure
7 presents the enthalpy versus entropy diagram for an oxygen-hydrogen reaction in chemical
equilibrium at 3000 psia (20.7 MPa). Process 1 to 2 is the irreversible combustion of oxygenhydrogen reactants at 3000 psia (20.7 MPa). The equilibrium products are then expanded in a
constant entropy (isentropic) process to local ambient pressure (0.101 MPa) from point 2 to point
3.

Nozzle Theory
The nozzle is the mechanism for accelerating the hot gases in all of the devices described, except
those that use electrostatic or electromagnetic forces. Applying the first law of thermodynamics
across the nozzle, the change in enthalpy equals the kinetic energy of the exhaust gases at the
exit:

Applying the idealization of isentropic, one-dimensional flow, the exhaust velocity of the nozzle
is

The nozzle is typically identified by its geometric area ratio (exit area/throat area), but the
pressure ratio across the nozzle is important in determining its performance in accelerating the
gas flow, as seen by examining the equation for exit velocity. The term g is the ratio of specific
heats of the hot gases, R is the universal gas constant, and M is the molecular weight of the
gases that exit the nozzle. Optimum expansion of the gases occurs when the exit pressure of the
nozzle equals the local ambient pressure.
The typical shape of a DeLaval nozzle is shown in Fig. 8. In the subsonic region (flow velocities
less than Mach 1), the gases are accelerated by decreasing the area of the flow passage.
Continuing the decrease of the flow area increases the gas velocity until a point is reached at
which the maximum mass flow rate per unit area is achieved. At this condition, the flow is at the
speed of sound or sonic (Mach number equal to 1). This location is called the throat of the
nozzle, and the flow is referred to as choked. The ratio of chamber pressure to pressure at the
throat (critical pressure ratio) is approximately 2:1 for this condition. From that point on, the
flow passage must increase in area to permit continuing acceleration of the flow in the
supersonic regime (Mach numbers greater than 1). Once the nozzle achieves the choked
condition, the chamber pressure remains constant regardless of the back-pressure from the flight
altitude. If the exit pressure exceeds the local ambient, it is underexpanded; if it is less than the
local ambient, it is overexpanded. Selection of the nozzle area/pressure ratio is a compromise to
provide the best performance across the vehicles flight regime. One way to examine this design
choice is through the nozzle thrust coefficient Cf. The thrust coefficient is a measure of nozzle
performance and can be used to determine the thrust of a rocket engine as a function of the throat
area and the chamber pressure from the equation,

Figure 8. The DeLaval Nozzle.

For a given ratio of specific heats, the optimum geometric area ratio corresponds to a given
pressure ratio across the nozzle. For example, at a nozzle area ratio of 10, the design pressure
ratio for a ratio of specific heats of 1.3 is 100, and the corresponding thrust coefficient is 1.6. If a
rocket has a nozzle of area ratio of 10, it is operating at its optimum condition only when the
pressure ratio is 100. Because a rocket flies through the atmosphere and is subject to varying
ambient pressure, the delivered performance will be less than optimum at all other points in the
trajectory than when the pressure ratio is 100. Therefore, the choice of design area ratio (pressure
ratio) is a compromise in which the designer knowingly accepts less than optimum performance
at some portions of the trajectory.
The nozzle designer must make compromises in choosing the design pressure ratio and also in
the shape of the nozzle. However, the most common nozzle design practice uses the Rao
optimum contour nozzle also known as the bell nozzle. This design yields a shorter design than
a simple 15 half-angle conical shape and has lower divergence losses because the gases are
exiting the nozzle at a divergence angle of less than 8.
The losses in a rocket nozzle consist of the divergence loss (nonaxial velocity vector for the
exiting gases), finite-rate kinetic losses, and drag losses, in accordance with the equation,

Inserting typical values into this equation, we can find the overall efficiency of a modern rocket
nozzle design:

There are nozzle designs that are intended to compensate for altitude as the rocket flies through
the constantly varying pressure of the atmosphere. This includes mechanical means in which
nozzle skirts are moved into place at selected times in the flight profile; each increases the
area/pressure ratio to approximate optimum expansion more closely. There are also aerodynamic
designs such as the aerospike shown in Fig. 9 that have a free jet boundary that adjusts to the
local ambient pressure through a Prandtl-Meyer or corner expansion. Such an expansion process
constrains the nozzle gases to expand to the local ambient at the outside lip of the nozzle.
Consequently, the flow cannot overexpand, and the nozzle flow is always optimum until the
nozzle design pressure ratio is exceeded. In practice, these nozzle designs do not operate with
perfect altitude compensation. A procedure has been developed to determine the degree to which
the aerospike type of nozzle approaches optimum expansion. This procedure is illustrated in
Figs. 10 and 11. Figure 10 presents the thrust efficiency CT (the ratio of measured CF to ideal
CF) of both an altitude compensating and bell nozzle versus pressure ratio. Both nozzles have the
same area ratio (design pressure ratio), the same thermodynamic properties of the flowing gases,
and the same maximum thrust efficiency. At a given pressure ratio, the percentage of maximum
thrust efficiency CT achieved by the altitude compensating nozzle is compared to that of the
fully flowing bell nozzle. The results of this comparison are presented in Fig. 11. The 45 line
across the graph represents the performance of a bell nozzle in which the flowing gases do not
separate. All points above that line represent some degree of altitude compensation. The point
identified by CTADV/CTMAX = 1 and CTN-S/CTMAX = 1 represents the design point for both
nozzles. The line where CTADV/CTMAX = 1 represents perfect altitude compensation. The
crosshatched area is representative of the altitude compensation performance of the aerospike
type of nozzle. In general, the data indicate a compensation capability in excess of 50%, as
determined by the ratio of the distance from the aerospike family of curves to the line where
CTADV/ CTMAX = 1 divided by the distance from the 45 line (nonseparating bell
performance). The line identified as the 80% bell illustrates the fact that a bell nozzle can have
some separation of the flow at the low end of its operating regime, and the result is that it has
some altitude compensation. The breadth of the aerospike band results from the range of nozzle
lengths from 16% of an isentropic spike at the upper end to 0% at the lower end.

Figure 9. Aerospike nozzle and its operating modes.

Figure 10. Nozzle altitude compensation determination.

Figure 11. Nozzle altitude compensation comparison.

Additionally, the aerospike nozzle length is about 25% that of the Rao optimum nozzle length for
the same thrust capability. This allows reducing in vehicle length or including more propellant in
the vehicle.
The same Prandtl-Meyer expansion process occurs in the external flow aft of the flight vehicle
that produces a pressure at the nozzle lip (the flow controlling pressure) that is lower than the
ambient at the flight altitude. Thus, the nozzle behaves as though it were operating at an altitude
higher than actual. This is most noticeable in the Mach 1 to Mach 3 range of flight operations. In
that flight regime, the time spent in off-design conditions is considered negligible compared to
bell nozzle performance capabilities. The added performance of optimized nozzle expansion
across the full flight trajectory opens the possibility of single-stage-to orbit operations.

Rocket Engine Efficiencies.


The overall efficiency of a rocket engine is generally presented as a percentage of the theoretical
specific impulse. The calculated theoretical value is based on shifting equilibrium, a case in
which the reaction products remain in equilibrium (their proportions change) as the pressure and
temperature decrease during the expansion process in the nozzle. This efficiency is composed of
the combustion and nozzle efficiencies, generally, it is greater than 90% and can be as high as
95-96%. The nozzle efficiency depends on design, but can be expected to be close to the 98%
value previously shown. The combustion efficiency of the propellants is quite sensitive to the
propellant combination used and plays the major role in overall efficiency. The liquid oxygen/
hydrocarbon combination of the early space launch engines typically had combustion efficiencies
in the low 90s% range. The efficiencies of todays hydrogen/ oxygen engines are much higher
and approach 97%.

Advance Propulsion and Applications.

Configuration.
Anatomy of Flight Vehicles.
Components of An Airplanes and Their
Function.
Configuration of Space Vehicles.
Earths Atmosphare and gravitational
Field.
Bluff Bodies v/s Streamlined Body.
Airfoil.
An airfoil or aerofoil is the shape of a wing, blade (of a propeller, rotor, or turbine), or sail
(as seen in cross-section). Foils of similar function designed with water as the working fluid
are called hydrofoils.

Examples of airfoils in nature and within various vehicles. Though not strictly an airfoil, the dolphin flipper obeys
the same principles in a different fluid medium.

The leading edge is the point at the front of the airfoil that has maximum curvature.
The trailing edge is defined similarly as the point of maximum curvature at the rear of the
airfoil.
The chord line is a straight line connecting the leading and trailing edges of the airfoil.

The chord length , or simply chord, , is the length of the chord line and is the characteristic
dimension of the airfoil section.
Various components of the airfoil. An airfoil) is the shape of a wing or blade of a propeller, rotor
or turbine or sail as seen in cross-section. An airfoil shaped body moved through a fluid produces
a force perpendicular to the fluid called lift. Subsonic flight airfoils have a characteristic shape
with a rounded leading edge, followed by a sharp trailing edge, often with asymmetric camber.
Airfoils designed with water as the working fluid are also called hydrofoils.

Types of airfoils
1. Symmetrical airfoil :- An airfoil that has the same shape on both sides of its centerline (the
centerline is thus straight). The movement of the center of pressure is the least in this type of
airfoil. This type of airfoil is used extensively in helicopter rotors.

2. Cambered airfoil :- The camber of an aerofoil can be defined by a camber line, which is the
curve that is halfway between the upper and lower surfaces of the aerofoil. Call this function
Z(x). To fully define an aerofoil we also need a thickness function T(x), which describes the
thickness of the aerofoil at any given point. Then, the upper and lower surfaces can be defined as
follows:

Example An aerofoil with reflexed camber line.

An aerofoil with reflex camber.

An aerofoil where the camber line curves back up near the trailing edge is called a reflexed
camber aerofoil. Such an aerofoil is useful in certain situations, such as with tailless aircraft,
because the moment about the aerodynamic center of the aerofoil can be 0. A camber line for
such an aerofoil can be defined as follows (note that the lines over the variables indicates that
they have been nondimensionalized by dividing through by the chord):

An aerofoil with a reflexed camber line is shown at right. The thickness distribution for a NACA
4-series aerofoil was used, with a 12% thickness ratio. The equation for this thickness
distribution is:

Where t is the thickness ratio.

Angle of Attack

Angle of attack In this diagram is (say- Alpha), the black lines represent the flow of the wind.
The wing is shown end on. The angle is the angle of attack. Angle of attack is a term used in
aerodynamics to describe the angle between the airfoil's chord line and the direction of airflow
wind, effectively the direction in which the aircraft is currently moving. It can be described as
the angle between where the wing is pointing and where it is going

Lift Generation.

Significance of L/D Ratio.

There are four forces that act on an aircraft in flight: lift, weight, thrust, and drag. Forces
are vector quantities having both a magnitude and a direction. The motion of the aircraft
through the air depends on the relative magnitude and direction of the various forces. The
weight of an airplane is determined by the size and materials used in the airplane's
construction and on the payload and fuel that the airplane carries. The weight is always
directed towards the center of the earth. The thrust is determined by the size and type of
propulsion system used on the airplane and on the throttle setting selected by the pilot.
Thrust is normally directed forward along the center-line of the aircraft. Lift and drag are
aerodynamic forces that depend on the shape and size of the aircraft, air conditions, and
the flight velocity. Lift is directed perpendicular to the flight path and drag is directed
along the flight path.
Because lift and drag are both aerodynamic forces, the ratio of lift to drag is an indication
of the aerodynamic efficiency of the airplane. Aerodynamicists call the lift to drag ratio
the L/D ratio, pronounced "L over D ratio."
L/D Ratio :- A particularly important figure of merit in Aerodynamic is the ratio of Lift to
Drag the so called L/D Ratio.

Significance of L/D Ratio.


1. Many Aspects Of the flight performance of a vehicle are directed related to L/D
Ratio.

2. Other things beings equal, a higher L/D means better flight performance.
3. For an airfoil a configuration whose primary function is to produce lift with as little
drag as possible value of L/D are large.
4. An airplane has a high L/D ratio if it produces a large amount of lift or a small
amount of drag. Under cruise conditions lift is equal to weight.
5. A high lift aircraft can carry a large payload. Under cruise conditions thrust is equal
to drag. A low drag aircraft requires low thrust.
6. Thrust is produced by burning a fuel and a low thrust aircraft requires small amounts
of fuel be burned. Low fuel usage allows an aircraft to stay aloft for a long time, and
that means the aircraft can fly long range missions. So an aircraft with a high L/D
ratio can carry a large payload, for a long time, over a long distance.
7.

For glider aircraft with no engines, a high L/D ratio again produces a long range
aircraft by reducing the steady state glide angle at which the glider descends.

8. The L/D ratio is also equal to the ratio of the lift and drag coefficients.

The lift equation indicates that the lift L is equal to one half the air density r times the
square of the velocity V times the wing area A times the lift coefficient Cl:

L = .5 * Cl * r * V^2 * A

Similarly, the drag equation relates the aircraft drag D to a drag coefficient Cd:
D = .5 * Cd * r * V^2 * A

Dividing these two equations give:

L/D = Cl/ Cd

Lift and drag coefficients are normally determined experimentally using a wind tunnel. But
for some simple geometries, they can be determined mathematically.

Aerodynamic Forces.

Propulsion.
Classification and Essential
Features of Propulsion.
Jet Propulsion.
Jet propulsion is thrust produced by passing a jet of matter (typically air or water)
in the opposite direction to the direction of motion. By Newton's third law, the
moving body is propelled in the opposite direction to the jet. It is most commonly
used in the jet engine, but is also the favoured means of propulsion used to power
various space craft.

Physics.
Jet propulsion is most effective when the Reynolds number is high - that is, the
object being propelled is relatively large and passing through a low-viscosity
medium.
In biology, the most efficient jets are pulsed, rather than continuous: at least when
the Reynolds number is greater than 6.

Jet Engine.
A jet engine is a reaction engine that discharges a fast moving jet of fluid to
generate thrust by jet propulsion and in accordance with Newton's laws of motion.

This broad definition of jet engines includes turbojets, turbofans, rockets, ramjets,
pulse jets and pump-jets. In general, most jet engines are internal combustion
engines but non-combusting forms also exist.

General Characteristics of Rocket


Engines.
Theory of Propulsion.
Elementary Gas Dynamics.
Spacecrafts and Aircraft
Performance.

Potrebbero piacerti anche