You are on page 1of 28

Thermodynamics is a branch of physics which deals with the energy and work of a system.

It
was born in the 19th century as scientists were first discovering how to build and operate steam
engines. Thermodynamics deals only with the large scale response of a system which we can
observe and measure in experiments. Small scale gas interactions are described by the kinetic
theory of gases. The methods compliment each other; some principles are more easily
understood in terms of thermodynamics and some principles are more easily explained by kinetic
theory.
There are three principal laws of thermodynamics which are described on separate slides. Each
law leads to the definition of thermodynamic properties which help us to understand and predict
the operation of a physical system. We will present some simple examples of these laws and
properties for a variety of physical systems, although we are most interested in thermodynamics
in the study of propulsion systems and high speed flows. Fortunately, many of the classical
examples of thermodynamics involve gas dynamics. Unfortunately, the numbering system for the
three laws of thermodynamics is a bit confusing. We begin with the zeroth law.
The zeroth law of thermodynamics involves some simple definitions of thermodynamic
equilibrium. Thermodynamic equilibrium leads to the large scale definition of temperature, as
opposed to the small scale definition related to the kinetic energy of the molecules. The first law
of thermodynamics relates the various forms of kinetic and potential energy in a system to the
work which a system can perform and to the transfer of heat. This law is sometimes taken as the
definition of internal energy, and introduces an additional state variable, enthalpy. The first law of
thermodynamics allows for many possible states of a system to exist. But experience indicates
that only certain states occur. This leads to the second law of thermodynamics and the definition
of another state variable called entropy. The second law stipulates that the total entropy of a
system plus its environment can not decrease; it can remain constant for a reversible process
but must always increase for an irreversible process.

Gases have various properties that we can observe with our senses, including the gas pressure
p, temperature T, mass m, and volume V that contains the gas. Careful, scientific observation
has determined that these variables are related to one another, and the values of these properties
determine the state of the gas.
If we fix any two of the properties we can determine the nature of the relationship between the
other two. You can explore the relationship between the variables at the animated gas lab. If the
pressure and temperature are held constant, the volume of the gas depends directly on the mass,
or amount of gas. This allows us to define a single additional property called the gas density r,
which is the ratio of mass to volume. If the mass and temperature are held constant, the product
of the pressure and volume are observed to be nearly constant for a real gas. The product of
pressure and volume is exactly a constant for an ideal gas. This relationship between pressure
and volume is called Boyle's Law in honor of Robert Boyle who first observed it in 1660. Finally, if
the mass and pressure are held constant, the volume is directly proportional to the temperature
for an ideal gas. This relationship is called Charles and Gay-Lussac's Law in honor of the two
French scientists who discovered the relationship.
The gas laws of Boyle and Charles and Gay-Lussac can be combined into a single equation of
state given in red at the center of the slide:
p*V/T=n*R
where * denotes multiplication and / denotes division. To account for the effects of mass, we have
defined the constant to contain two parts: a universal constant R and the mass of the gas
expressed in moles n. Performing a little algebra, we obtain the more familiar form:
p*V=n*R*T

A three dimensional graph of this equation is shown at the lower left. The intersection point of any
two lines on the graph gives a unique state for the gas.
Engineers use a slightly different form of the equation of state that is specialized for a particluar
gas. If we divide both sides of the general equation by the mass of the gas, the volume becomes
the specific volume, which is the inverse of the gas density. We also define a new gas constant R,
which is equal to the universal gas constant divided by the mass per mole of the gas. The value
of the new constant depends on the type of gas as opposed to the universal gas constant, which
is the same for all gases. The value of the equation of state for air is given on the slide as .286
kilo Joule per kilogram per degree Kelvin. The equation of state can be written in terms of the
specific volume or in terms of the air density as
p*v=R*T
p=r*R*T
Notice that the equation of state given here applies only to an ideal gas, or a real gas that
behaves like an ideal gas. There are in fact many different forms for the equation of state for
different gases. Also be aware that the temperature given in the equation of state must be an
absolute temperature that begins at absolute zero. In the metric system of units, we must specify
the temperature in degrees Kelvin (not Celsius). In the English system, absolute temperature is in
degrees Rankine (not Fahrenheit).

Thermodynamics is a branch of physics which deals with the energy and work of a system.
Thermodynamics deals only with the large scale response of a system which we can observe and
measure in experiments. In rocket science, we are most interested in thermodynamics for the role
it plays in engine design and high speed flight.
On this slide we derive some equations which relate the heat capacity of a gas to the gas
constant used in the equation of state. We are going to be using specific values of the state
variables. For a scientist, a "specific" state variable means the value of the variable divided by the
mass of the substance. This allows us to derive relations between variables without regard for the
amount of the substance that we have. We can multiply the specific variable by the quantity of the
substance at any time to determine the actual value of the flow variable. From our studies of heat
transfer, we know that the amount of heat transferred between two objects is proportional to the
temperature difference between the objects and the heat capacity of the objects. The heat
capacity is a constant that tells how much heat is added per unit temperature rise. The value of
the constant is different for different materials and depends on the process. Heat capacity is not
a state variable.
If we are dealing with a gas, it is most convenient to use forms of the thermodynamics equations
based on the enthalpy of the gas. From the definition of enthalpy:
h=e+p*v
where h in the specific enthalpy, p is the pressure, v is the specific volume, and e is the specific
internal energy. During a process, the values of these variables change. Let's denote the change
by the Greek letter delta which looks like a triangle. So "delta h" means the change of "h" from
state 1 to state 2 during a process. Then, for a constant pressure process the enthalpy
equation becomes:
delta h = delta e + p * delta v
The enthalpy, internal energy, and volume are all changed, but the pressure remains the same.
From our derivation of the enthalpy equation, the change of specific enthalpy is equal to the heat
transfer for a constant pressure process:
delta h = cp * delta T
where delta T is the change of temperature of the gas during the process,and c is the specific
heat capacity. We have added a subscript "p" to the specific heat capacity to remind us that this
value only applies to a constant pressure process.
The equation of state of a gas relates the temperature, pressure, and volume through a gas
constant R . The gas constant used by aerodynamicists is derived from the universal gas
constant, but has a unique value for every gas.
p*v=R*T
If we have a constant pressure process, then:
p * delta v = R * delta T
Now let us imagine that we have a constant volume process with our gas that produces exactly
the same temperature change as the constant pressure process that we have been discussing.
Then the first law of thermodynamics tells us:

delta e = delta q - delta w


where q is the specific heat transfer and w is the work done by the gas. For a constant volume
process, the work is equal to zero. And we can express the heat transfer as a constant times the
change in temperature. This gives:
delta e = cv * delta T
where delta T is the change of temperature of the gas during the process,and c is the specific
heat capacity. We have added a subscript "v" to the specific heat capacity to remind us that this
value only applies to a constant volume process. Even though the temperature change is the
same for this process and the constant pressure process, the value of the specific heat capacity
is different.
Because we have selected the constant volume process to give the same change in temperature
as our constant pressure process, we can substitute the expression given above for "delta e" into
the enthalpy equation. In general, you can't make this substitution because a constant pressure
process and a constant volume process produce different changes in temperature If we substitute
the expressions for "delta e", "p * delta v", and "delta h" into the enthalpy equation we obtain:
cp * delta T = cv * delta T + R * delta T
dividing by "delta T" gives the relation:
cp = cv + R
The specific heat constants for constant pressure and constant volume processes are related to
the gas constant for a given gas. This rather remarkable result has been derived from
thermodynamic relations, which are based on observations of physical systems and processes.
Using the kinetic theory of gases, this same result can be derived from considerations of the
conservation of energy at a molecular level.
We can define an additional variable called the specific heat ratio, which is given the Greek
symbol "gamma", which is equal to cp divided by cv:
gamma = cp / cv
"Gamma" is just a number whose value depends on the state of the gas. For air, gamma = 1.4 for
standard day conditions. "Gamma" appears in many fluids equations including the equation
relating pressure, temperature, and volume during a simple compression or expansion process,
the equation for the speed of sound, and all the equations for isentropic flows, and shock waves.
Because the value of "gamma" just depends on the state of the gas, there are tables of these
values for given gases. You can use the tables to solve gas dynamics problems.

Thermodynamics is a branch of physics which deals with the energy and work of a system.
Thermodynamics deals only with the large scale response of a system which we can observe and
measure in experiments. Small scale gas interactions are described by the kinetic theory of
gases. There are three principal laws of thermodynamics which are described on separate slides.
Each law leads to the definition of thermodynamic properties which help us to understand and
predict the operation of a physical system. We will present some simple examples of these laws
and properties for a variety of physical systems, although we are most interested in the
thermodynamics of propulsion systems and high speed flows. Fortunately, many of the classical
examples of thermodynamics involve gas dynamics.
In our observations of the work done on, or by a gas, we have found that the amount of work
depends not only on the initial and final states of the gas but also on the process, or path
which produces the final state. Similarly the amount of heat transferred into, or from a gas
also depends on the initial and final states and the process which produces the final state.
Many observations of real gases have shown that the difference of the heat flow into the gas and
the work done by the gas depends only on the initial and final states of the gas and does not
depend on the process or path which produces the final state. This suggests the existence of an
additional variable, called the internal energy of the gas, which depends only on the state of the
gas and not on any process. The internal energy is a state variable, just like the temperature or
the pressure. The first law of thermodynamics defines the internal energy (E) as equal to the
difference of the heat transfer (Q) into a system and the work (W) done by the system.
E2 - E1 = Q - W
We have emphasized the words "into" and "by" in the definition. Heat removed from a system
would be assigned a negative sign in the equation. Similarly work done on the system is assigned
a negative sign. The internal energy is just a form of energy like the potential energy of an object
at some height above the earth, or the kinetic energy of an object in motion. In the same way that
potential energy can be converted to kinetic energy while conserving the total energy of the
system, the internal energy of a thermodynamic system can be converted to either kinetic or
potential energy. Like potential energy, the internal energy can be stored in the system. Notice,

however, that heat and work can not be stored or conserved independently since they depend on
the process. The first law of thermodynamics allows for many possible states of a system to exist,
but only certain states are found to exist in nature. The second law of thermodynamics helps to
explain this observation. If a system is fully insulated from the outside environment, it is possible
to have a change of state in which no heat is transferred into the system. Scientists refer to a
process which does not involve heat transfer as an adiabatic process. The implementation of the
first law of thermodynamics for gases introduces another useful state variable called the enthalpy.

Thermodynamics is a branch of physics which deals with the energy and work of a system.
Thermodynamics deals only with the large scale response of a system which we can observe and
measure in experiments. In rocket science, we are most interested in thermodynamics in the
study of propulsion systems and understanding high speed flows.
The state of a gas is defined by several properties including the temperature, pressure, and
volume which the gas occupies. From a study of the first law of thermodynamics, we find that the
internal energy of a gas is also a state variable, that is, a variable which depends only on the
state of the gas and not on any process that produced that state. We are free to define additional
state variables which are combinations of existing state variables. The new variables often make
the analysis of a system much simpler. For a gas, a useful additional state variable is the
enthalpy which is defined to be the sum of the internal energy E plus the product of the pressure
p and volume V. Using the symbol H for the enthalpy:
H=E+p*V
The enthalpy can be made into an intensive, or specific , variable by dividing by the mass.
Propulsion engineers use the specific enthalpy in engine analysis more than the enthalpy itself.
How does one use this new variable called enthalpy? Let's consider the first law of
thermodynamics for a gas. For a system with heat transfer Q and work W, the change in internal
energy E from state 1 to state 2 is equal to the difference in the heat transfer into the system and
the work done by the system:

E2 - E1 = Q - W
The work and heat transfer depend on the process used to change the state. For the special case
of a constant pressure process, the work done by the gas is given as the constant pressure p
times the change in volume V:
W = p * [V2 - V1]
Substituting into the first equation, we have:
E2 - E1 = Q - p * [V2 - V1]
Let's group the conditions at state 2 and the conditions at state 1 together:
(E2 + p * V2) - (E1 + p * V1) = Q
The (E + p * V) can be replaced by the enthalpy H.
H2 - H1 = Q
From our definition of the heat transfer, we can represent Q by some heat capacity coefficient Cp
times the temperature T.
(H2 - H1) = Cp * (T2 - T1)
At the bottom of the slide, we have divided by the mass of gas to produce the specific enthalpy
equation version.
(h2 - h1) = cp * (T2 - T1)
The specific heat capacity cp is called the specific heat at constant pressure and is related to the
universal gas constant of the equation of state. This final equation is used to determine values of
specific enthalpy for a given temperature. Enthalpy is used in the energy equation for a fluid.
Across shock waves, the total enthalpy of the gas remains a constant.

Thermodynamics is a branch of physics that deals with the energy and work of a system. In
rocket science, we are most interested in thermodynamics in the study of propulsion systems and
understanding high speed flows. The first law of thermodynamics indicates that the total energy of
a system is conserved. Total energy includes the potential and kinetic energy, the work done by
the system, and the transfer of heat through the system. The second law of thermodynamics
indicates that, while many physical processes that satisfy the first law are possible, the only
processes that occur in nature are those for which the entropy of the system either remains
constant or increases.
Entropy, like temperature and pressure, can be explained on both a macro scale and a micro
scale. Since thermodynamics deals only with the macro scale, the change in entropy delta S is
defined here to be the heat transfer delta Q into the system divided by the temperature T:
delta S = delta Q / T
During a thermodynamic process, the temperature T of an object changes as heat Q is applied or
extracted. A more correct definition of the entropy S is the differential form that accounts for this
variation.
dS = dQ / T
The change in entropy is then the inverse of the temperature integrated over the change in heat
transfer. For gases, there are two possible ways to evaluate the change in entropy. We begin by
using the first law of thermodynamics:

dE = dQ - dW
where E is the internal energy and W is the work done by the system. Substituting for the
definition of work for a gas.
dQ = dE + p dV
where p is the pressure and V is the volume of the gas. If we use the definition of the enthalpy H
of a gas:
H=E+p*V
Then:
dH = dE + p dV + V dp
Substitute into the first law equation:
dQ = dH - V dp - p dV + p dV
dQ = dH - V dp
is an alternate way to present the first law of thermodynamics. For an ideal gas, the equation of
state is written:
p*V=R*T
where R is the gas constant. The heat transfer of a gas is equal to the heat capacity times the
change in temperature; in differential form:
dQ = C * dT
If we have a constant volume process, the formulation of the first law gives:
dE = dQ = C (constant volume) * dT
Similarly, for a constant pressure process, the formulation of the first law gives:
dH = dQ = C (constant pressure) * dT
If we assume that the heat capacity is constant with temperature, we can use these two
equations to define the change in enthalpy and internal energy. If we substitute the value for p
from the equation of state, and the definition of dE in the first energy equation, we obtain:
dQ = C (constant volume) * dT + R * T dV / V
Similarly substituting the value of V from the equation of state, and the definition of dH we obtain
the alternate form:
dQ = C (constant pressure) * dT - R * T dp / p

Substituting these forms for dQ into the differential form of the entropy equation gives::
dS = C (constant volume) * dT / T + R * dV / V
and
dS = C (constant pressure) * dT / T - R * dp / p
These equations can be integrated from condition "1" to condition "2" to give:
S2 - S1 = Cv * ln ( T2 / T1) + R * ln ( V2 / V1)
and
S2 - S1 = Cp * ln ( T2 / T1) - R * ln ( p2 / p1)
where Cv is the heat capacity at constant volume, Cp is the heat capacity at constant pressure,
and ln is the symbol for the logarithmic function.
If we divide both equations by the mass of gas, we can obtain intrinsic, or "specific" forms of
both equations:
s2 - s1 = cv * ln ( T2 / T1) + R * ln ( v2 / v1)
and
s2 - s1 = cp * ln ( T2 / T1) - R * ln ( p2 / p1)
where cp and cv are the specific heat capacities. Depending on the type of process we
encounter, we can now determine the change in entropy for a gas.
These equations can be a bit confusing, because we use the specific heat at constant volume
when we have a process that changes volume, and the specific heat at constant pressure when
the process changes pressure.To clarify matters, let's look at the first equation:
s2 - s1 = cv * ln ( T2 / T1) + R * ln ( v2 / v1)
If we have a constant volume process, the second term in the equation is equal to zero, since
v2/v1 = 1. We can then determine the value of the specific heat for the constant volume process.
But if we have a process that changes volume, the second term in the equation is not zero. We
can think of the first term of the equation as the contribution for a constant volume process, and
the second term as the additional change produced by the change in volume. A similar type of
argument can be made for the equation used for a change in pressure.

The Earth's atmosphere is composed of air. Air is a mixture of gases, 78% nitrogen and 21%
oxygen with traces of water vapor, carbon dioxide, argon, and various other components. We
usually model air as a uniform (no variation or fluctuation) gas with properties that are averaged
from all the individual components. Any gas has certain properties that we can detect with our
senses. The values and relations of the properties define the state of the gas.
On this slide you will find typical values of the properties of air at sea level static conditions for
a standard day. We are all aware that pressure and temperature of the air depend on your
location on the earth and the season of the year. And while it is hotter in some seasons than
others, pressure and temperature change day to day, hour to hour, sometimes even minute to
minute during severe weather. The values presented on the slide are simply average values used
by engineers to design machines. That's why they are called standard values. We also know that
all of the state-of-the-gas variables will change with altitude, which is why the typical values are
given at sea level, static conditions. Because the gravity of the Earth holds the atmosphere to the
surface, as altitude increases, air density, pressure, and temperature (for lower altitudes)
decrease. At the edge of space, the density is almost zero. The variation of the air from the
standard can be very important since it affects flow parameters like the speed of sound.
A gas is composed of a large number of molecules which are in constant, random motion. The
sum of the mass of all the molecules is equal to the mass of the gas. A gas occupies some
volume in three dimensional space. For a given pressure and temperature, the volume depends
directly on the amount of gas. Since the mass and volume are directly related, we express both
the mass and volume by a single variable. When a gas is moving, it is convenient to use the
density of a gas, which is the mass divided by the volume the gas occupies. The sea level
standard value of air density r is

r = 1.229 kilograms/cubic meters = .00237 slug/cubic feet


When working with a static or unmoving gas, it is more convenient to use specific volume, which
is the volume divided by the mass. The sea level standard value of specific volume v is
v = .814 cubic meters/kilogram = 422 cubic feet/slug
The pressure of a gas equals the perpendicular force exerted by the gas divided by the surface
area on which the force is exerted. The sea level standard value of air pressure p is
p = 101.3 kilo Newtons/square meter = 14.7 pounds/square inch
The temperature of a gas is a measure of the kinetic energy of the molecules of the gas. The sea
level standard value of air temperature T is
T = 15 degrees C = 59 degrees Fahrenheit
A gas can exert a tangential (shearing) force on a surface, which acts like friction between solid
surfaces. This "sticky" property of the gas is called the viscosity and it plays a large role in
aerodynamic drag. The sea level standard value of air viscosity mu is
mu = 1.73 x 10^-5 Newton-second/square meters = 3.62 x 10^-7 pound-second/square feet
The density (specific volume), pressure, and temperature of a gas are related to each other
through the equation of state. The state of a gas can be changed by external processes, and the
reaction of the gas can be predicted using the laws of thermodynamics. A fundamental
understanding of thermodynamics is very important in describing the operation of propulsion
systems.

The conservation of momentum is a fundamental concept of physics along with the


conservation of energy and the conservation of mass. Momentum is defined to be the mass of an
object multiplied by the velocity of the object. The conservation of momentum states that, within
some problem domain, the amount of momentum remains constant; momentum is neither
created nor destroyed, but only changed through the action of forces as described by Newton's
laws of motion. Dealing with momentum is more difficult than dealing with mass and energy
because momentum is a vector quantity having both a magnitude and a direction. Momentum is
conserved in all three physical directions at the same time. It is even more difficult when dealing
with a gas because forces in one direction can affect the momentum in another direction because
of the collisions of many molecules. On this slide, we will present a very, very simplified flow
problem where properties only change in one direction. The problem is further simplified by
considering a steady flow which does not change with time and by limiting the forces to only
those associated with the pressure.
Let us consider the flow of a gas through a domain in which flow properties only change in one
direction, which we will call "x". The gas enters the domain at station 1 with some velocity u and
some pressure p and exits at station 2 with a different value of velocity and pressure. For
simplicity, we will assume that the density r remains constant within the domain and that the area
A through which the gas flows also remains constant. The location of stations 1 and 2 are
separated by a distance called del x. (Delta is the little triangle on the slide and is the Greek letter
"d". Mathematicians often use this symbol to denote a change or variation of a quantity. The web
print font does not support the Greek letters, so we will just call it "del".) A change with distance is
referred to as a gradient to avoid confusion with a change with time which is called a rate. The
velocity gradient is indicated by del u / del x; the change in velocity per change in distance. So at
station 2, the velocity is given by the velocity at 1 plus the gradient times the distance.
u2 = u1 + (del u / del x) * del x
A similar expression gives the pressure at the exit:
p2 = p1 + (del p / del x) * del x
Newton's second law of motion states that force F is equal to the change in momentum with
respect to time. For an object with constant mass m this reduces to the mass times acceleration
a. An acceleration is a change in velocity with a change in time (del u / del t). Then:
F = m * a = m * (del u / del t)
The force in this problem comes from the pressure gradient. Since pressure is a force per unit
area, the net force on our fluid domain is the pressure times the area at the exit minus the
pressure times the area at the entrance.
F = - [(p * A)2 - (p * A)1] = m * [(u2 - u1) / del t]
The minus sign at the beginning of this expression is used because gases move from a region of
high pressure to a region of low pressure; if the pressure increases with x, the velocity will
decrease. Substituting for our expressions for velocity and pressure:
- [{(p + (del p / del x) * del x} * A) - (p * A)] = m * [(u + (del u / del x) * del x - u) / del t]
Simplify:

- (del p / del x) * del x * A = m * (del u / del x) * del x / del t


Noting that (del x / del t) is the velocity and that the mass is the density r times the volume (area
times del x):
- (del p / del x) * del x * A = r * del x * A * (del u / del x) * u
Simplify:
- (del p / del x) = r * u * (del u / del x)
The del p / del x and del u / del x represent the pressure and velocity gradients. If we shrink our
domain down to differential sizes, these gradients become differentials:
- dp/dx = r * u * du/dx
This is a one dimensional, steady form of Euler's Equation. It is interesting to note that the
pressure drop of a fluid (the term on the left) is proportional to both the value of the velocity and
the gradient of the velocity. A solution of this momentum equation gives us the form of the
dynamic pressure that appears in Bernoulli's Equation

On this slide we have two versions of the Euler Equations which describe how the velocity,
pressure and density of a moving fluid are related. The equations are named in honor of Leonard
Euler, who was a student with Daniel Bernoulli, and studied various fluid dynamics problems in
the mid-1700's. The equations are a set of coupled differential equations and they can be

solved for a given flow problem by using methods from calculus. Though the equations appear to
be very complex, they are actually simplifications of the more general Navier-Stokes equations of
fluid dynamics. The Euler equations neglect the effects of the viscosity of the fluid which are
included in the Navier-Stokes equations. A solution of the Euler equations is therefore only an
approximation to a real fluids problem. For some problems, like the lift of a thin airfoil at low angle
of attack, a solution of the Euler equations provides a good model of reality. For other problems,
like the growth of the boundary layer on a flat plate, the Euler equations do not properly model the
problem.
Our world has three spatial dimensions (up-down, left-right, fore-aft) and one time dimension. In
general, the Euler equations have a time-dependent continuity equation for conservation of mass
and three time-dependent conservation of momentum equations. At the top of the figure, we show
a simplified, two-dimensional, steady form of the Euler equations. There are two independent
variables in the problem, the x and y coordinates of some domain. There are four dependent
variables, the pressure p, density r, and two components of the velocity vector; the u component
is in the x direction, and the v component is in the y direction. All of the dependent variables are
functions of both x and y. The differential equations are therefore partial differential equations
and not ordinary differential equations that you study in a beginning calculus class. You will
notice that the differential symbol is different than the usual "d /dt" or "d /dx" that you see for
ordinary differential equations.

The symbol look like a backwards "6", and, because of font limitations, we will use the symbol "6 /
6t" on this page to indicate partial derivatives. The symbol indicates that we are to hold all of the
independent variables fixed, except the variable next to symbol, when computing a derivative.
The set of equations are:
Continuity: 6(r * u)/6x + 6(r * v)/6y = 0
X - Momentum: 6(r * u^2)/6x + 6(r * u * v)/6y = - 6(p)/6x
Y - Momentum: 6(r * u * v)/6x + 6(r * v^2)/6y = - 6(p)/6y
Although these equations appear very complex, undergraduate engineering studentsare taught
how to derive them in a process very similar to the derivation that we present on the conservation
of momentum web page. The two momentum equations are two-dimensional generalizations of
the conservation of momentum equation. The equation for the mass flow rate which was derived
on the conservation of mass web page is a one dimensional solution of the continuity equation
shown here.
Generalized solutions of these equations are difficult to obtain. Notice that all of the dependent
variables appear in each equation. To solve a flow problem, you have to solve all three equations
simultaneously; that is why we call this a coupled system of equations. There is actually
anotherequation that is required to solve this system, since we only show three equations for four
unknowns. An equation of state relates the pressure and the density of a gas. In the past,
engineers made further approximations and simplifications to the equation set until they had a
group of equations that they could solve. Recently, high speed computers have been used to
solve approximations to the equations using a variety of techniques like finite difference, finite
volume, finite element, and spectral methods. This area of study is called Computational Fluid
Dynamics or CFD.
One of the simplification methods used in the past was to assume that the gas was very low
speed and to neglect the effects of compressibility. In an incompressible flow, the density is
constant and we can remove it from the continuity equation:

Continuity: 6u/6x + 6v/6y = 0


We can then factor the momentum equations and use the continuity equation to simplify them:
X - Momentum: u * 6u/6x + v * 6u/6y = - [6(p)/6x] / r
Y - Momentum: u * 6v/6x + v * 6v/6y = - [6(p)/6y] / r

In the 1700s, Daniel Bernoulli investigated the forces present in a moving fluid. This slide shows
one of many forms of Bernoulli's equation. The equation appears in many physics, fluid
mechanics, and airplane textbooks. The equation states that the static pressure ps in the flow
plus the dynamic pressure, one half of the density r times the velocity V squared, is equal to a
constant throughout the flow. We call this constant the total pressure pt of the flow.
As discussed on the gas properties page, there are two ways to look at a fluid; from the large,
macro scale properties of the fluid that we can measure, and from the small, micro scale of the
molecular motion and interaction. On this page, we will consider Bernoulli's equation from both
standpoints.

Macro Scale Derivation


Thermodynamics is the branch of science which describes the macro scale properties of a fluid.
One of the principle results of the study of thermodynamics is the conservation of energy; within a
system, energy is neither created nor destroyed but may be converted from one form to another.
We shall derive Bernoulli's equation by starting with the conservation of energy equation. The
most general form for the conservation of energy is given on the Navier-Stokes equation page.
This formula includes the effects of unsteady flows and viscous interactions. Assuming a
steady, inviscid flow we have a simplified conservation of energy equation in terms of the
enthalpy of the fluid:
ht2 - ht1 = q - wsh
where ht is the total enthalpy of the fluid, q is the heat transfer into the fluid, and wsh is the useful
work done by the fluid.
Assuming no heat transfer into the fluid, and no work done by the fluid, we have:
ht2 = ht1
From the definition of total enthalpy:
e2 + (p * v)2 + (.5 * V^2)2 = e1 + (p * v)1 + (.5 * V^2)1
where e is the internal energy, p is the pressure, v is the specific volume, and V is the velocity of
the fluid. From the first law of thermodynamics if there is no work and no heat transfer, the
internal energy remains the same:
(p * v)2 + (.5 * V^2)2 = (p * v)1 + (.5 * V^2)1
The specific volume is the inverse of the fluid density r:
(p / r)2 + (.5 * V^2)2 = (p / r)1 + (.5 * V^2)1
Assuming that the flow is incompressible, the density is a constant. Multiplying the energy
equation by the constant density:
(ps)2 + (.5 * r * V^2)2 = (ps)1 + (.5 * r * V^2)1 = a constant = pt
This is the simplest form of Bernoulli's equation and the one most often quoted in textbooks. If we
make different assumptions in the derivation, we can derive other forms of the equation.
It is important when applying any equation that you are aware of the restrictions on its use; the
restrictions usually arise in the derivation of the equation when certain simplifying assumptions
about the nature of the problem are made. If you ignore the restrictions, you may often get an
incorrect "answer" from the equation. For instance, this form of the equation was derived while
assuming that the flow was incompressible, which means that the speed of the flow is much less
than the speed of sound. If you use this form for a supersonic flow, the answer will be wrong.

Molecular Scale Derivation


We can make another interpretation of the equation by considering the motion of the gas
molecules. The molecules within a fluid are in constant random motion and collide with each
other and with the walls of an object in the fluid. The motion of the molecules gives the molecules
a linear momentum and the fluid pressure is a measure of this momentum. If a gas is at rest, all
of the motion of the molecules is random and the pressure that we detect is the total pressure of
the gas. If the gas is set in motion or flows, some of the random components of velocity are
changed in favor of the directed motion. We call the directed motion "ordered," as opposed to the
disordered random motion.
We can associate a "pressure" with the momentum of the ordered motion of the gas. We call this
pressure the dynamic pressure. The remaining random motion of the molecules still produces a
pressure called the static pressure. At the molecular level, there is no distinction between
random and ordered motion. Each molecule has a velocity in some direction until it collides with
another molecule and the velocity is changed. But when you sum up all the velocities of all the
molecules you will detect the ordered motion. From a conservation of energy and momentum, the
static pressure plus the dynamic pressure is equal to the original total pressure in a flow
(assuming we do not add or subtract energy in the flow). The form of the dynamic pressure is the
density times the square of the velocity divided by two.
Applications of Bernoulli's Equation
The fluids problem shown on this slide is low speed flow through a tube with changing crosssectional area. For a streamline along the center of the tube, the velocity decreases from station
one to two. Bernoulli's equation describes the relation between velocity, density, and pressure for
this flow problem. Since density is a constant for a low speed problem, the equation at the bottom
of the slide relates the pressure and velocity at station two to the conditions at station one.

An important property of any gas is its pressure. Because understanding what pressure is and
how it works is so fundamental to the understanding of rocketry, we are including several slides
on pressure in the Beginner's Guide.
There are two ways to look at pressure: (1) the small scale action of individual air molecules or
(2) the large scale action of a large number of molecules. On the the small scale, from the kinetic
theory of gases, a gas is composed of a large number of molecules that are very small relative to
the distance between molecules. The molecules of a gas are in constant, random motion and
frequently collide with each other and with the walls of any container. During collisions with the
walls, there is a change in velocity and therefore a change in momentum of the molecules. The
change in momentum produces a force on the walls which is related to the gas pressure. The
pressure of a gas is a measure of the average linear momentum of the moving molecules of a
gas. On the large scale, the pressure of a gas is a state variable, like the temperature and the
density. The change in pressure during any process is governed by the laws of thermodynamics.
Although pressure itself is a scalar quantity, we can define a pressure force to be equal to the
pressure (force/area) times the surface area in a direction perpendicular to the surface. If a gas is
static and not flowing, the measured pressure is the same in all directions. But if the gas is
moving, the measured pressure depends on the direction of motion. This leads to the definition of
the dynamic pressure.
To understand dynamic pressure, we begin with a one dimensional version of the conservation of
linear momentum for a fluid.
r * u * du/dx = - dp/dx
where r is the density of the gas, p is the pressure, x is the direction of the flow, and u is the
velocity in the x direction. Performing a little algebra:
dp/dx + r * u * du/dx = 0
For a constant density (incompressible flow) we can take the "r * u" term inside the differential:
dp/dx + d(.5 * r * u^2)/dx = 0
and then gather all of the terms:
d(p + .5 * r * u^2)/dx = 0
Integrating this differential equation:
ps + .5 * r * u^2 = constant = pt
This equation looks exactly like the incompressible form of Bernoulli's equation. Each term in this
equation has the dimensions of a pressure (force/area); ps is the static pressure, the constant pt
is called the total pressure, and
.5 * r * u^2
is called the dynamic pressure because it is a pressure term associated with the velocity u of
the flow. Dynamic pressure is often assigned the letter q in aerodynamics:
q = .5 * r * u^2

The dynamic pressure is a defined property of a moving flow of gas. We have performed this
simple derivation to determine the form of the dynamic pressure, but we can use and apply the
idea of dynamic pressure in much more complex flows, like compressible flows or viscous flows.
In particular, the aerodynamic forces acting on an object as it moves through the air are directly
proportional to the dynamic pressure. The dynamic pressure is therefore used in the definition of
the lift coefficient and the drag coefficient. As we have seen, dynamic pressure appears in
Bernoulli's equation even though that relationship was originally derived using energy
conservation.
The dynamic pressure depends on both the local value of the density and the velocity of the flow,
or rocket . The density of the air decreases with altitude in a complex manner. The velocity of a
rocket during launch is constantly increasing with altitude. Therefore, the dynamic pressure on a
rocket during launch is initially zero because the velocity is zero. The dynamic pressure increases
because of the increasing velocity to some maximum value, called the maximum dynamic
pressure, or Max Q. Then the dynamic pressure decreases because of the decreasing density.
The Max Q condition is a design constraint on full scale rockets. You can investigate the variation
of dynamic pressure with altitude and velocity by using our atmosphere simulator.

The conservation of energy is a fundamental concept of physics along with the conservation of
mass and the conservation of momentum. Within some problem domain, the amount of energy
remains constant and energy is neither created nor destroyed. Energy can be converted from one
form to another (potential energy can be converted to kinetic energy) but the total energy within
the domain remains fixed.

Thermodynamics is a branch of physics which deals with the energy and work of a system. As
mentioned on the gas properties slide, thermodynamics deals only with the large scale response
of a system which we can observe and measure in experiments. In rocketry, we are most
interested in thermodynamics in the study of propulsion systems and understanding high speed
flows.
On some separate slides, we have discussed the state of a static gas, the properties which define
the state, and the first law of thermodynamics as applied to any system, in general. On this slide
we derive a useful form of the energy conservation equation for a gas beginning with the first law
of thermodynamics. If we call the internal energy of a gas E, the work done by the gas W, and the
heat transferred into the gas Q, then the first law of thermodynamics indicates that between state
"1" and state "2":
E2 - E1 = Q - W
Aerospace engineers usually simplify a thermodynamic analysis by using intensive variables;
variables that do not depend on the mass of the gas. We call these variables specific variables.
We create a "specific" variable by taking a property whose value depends on the mass of the
system and dividing it by the mass of the system. Many of the state properties listed on this slide,
such as the work and internal energy depend on the total mass of gas. We will use "specific"
versions of these variables. Engineers usually use the lower case letter for the "specific" version
of a variable. Our first law equation then becomes:
e2 - e1 = q - w
Because we are considering a moving gas, we add the specific kinetic energy term to the internal
energy on the left side. The normal kinetic energy K of a moving substance is equal to 1/2 times
the mass m times the velocity u squared:
K = (m * u^2) / 2
Then the specific kinetic energy k is given by:
k = (u^2) / 2
and the first law equation becomes:
e2 - e1 + k2 - k1 = q - w
There are two parts to the specific work for a moving gas. Some of the work, called the shaft
work (wsh) is used to move the fluid or turn a shaft, while the rest of the work goes into changing
the state of the gas. For a pressure p and specific volume v, the work is given by:
w = (p * v)2 - (p * v)1 + wsh
Substituting:
e2 - e1 + k2 - k1 = q - (p * v)2 + (p * v)1 - wsh
If we perform a little algebra on the first law of thermodynamics, we can begin to group some
terms of the equations. :

e2 + (p * v)2 - e1 - (p * v)1 + [(u^2) / 2]2 - [(u^2) / 2]1 = q - wsh


A useful additional state variable for a gas is the specific enthalpy h which is equal to:
h = e + (p * v)
Simplifying the energy equation:
h2 - h1 + [(u^2) / 2]2 - [(u^2) / 2]1 = q - wsh
or
h2 + [(u^2) / 2]2 - h1 - [(u^2) / 2]1 = q - wsh
By combining the velocity terms with the enthalpy terms to form the total specific enthalpy "ht"
we can further simplify the equation.
ht = h + u^2 / 2
The total specific enthalpy is analogous to the total pressure in Bernoulli's equation; both
expressions involve a "static" value plus one half the square of the velocity.
The final, most useful, form of the energy equation is given in the red box.
ht2 - ht1 = q - wsh

As a rocket moves through the air, the air molecules near the rocket are disturbed and move
around the rocket. If the rocket passes at a low speed, typically less than 250 mph, the density of
the air remains constant. But for higher speeds, some of the energy of the rocket goes into
compressing the air and locally changing the density of the air. This compressibility effect alters
the amount of resulting force on the rocket. The effect becomes more important as speed
increases. Near and beyond the speed of sound, about 330 m/s or 760 mph at sea level on Earth,
small disturbances in the flow are transmitted to other locations isentropically or with constant
entropy. Sharp disturbances generate shock waves that affects the drag on the rocket. Within the
nozzle of the rocket engine, gases are expelled at speeds much greater than the speed of sound
to generate thrust.
The ratio of the speed of the rocket, or the speed of the nozzle flow, to the speed of sound in the
gas determines the magnitude of many of the compressibility effects. Because of the importance
of this speed ratio, engineers give it a special name, the Mach number, in honor of Ernst Mach,
a late 19th century physicist who studied gas dynamics. The Mach number M allows us to define
flow regimes in which compressibility effects vary.
1. Subsonic conditions occur for Mach numbers less than one, M < 1 . For the lowest
subsonic conditions, compressibility can be ignored.
2. As the speed of the rocket approaches the speed of sound, the flight Mach number is
nearly equal to one, M = 1, and the flow is said to be transonic. At some places on the
rocket, the local speed exceeds the speed of sound. Compressibility effects are most
important in transonic flows and lead to the early belief in a sound barrier. Flight faster
than sound was thought to be impossible. In fact, the sound barrier was only an increase
in the drag near sonic conditions because of compressibility effects. During launch, a

rocket often encounters its highest dynamic pressure , or "Max Q", at transonic
conditions.
3. Supersonic conditions occur for Mach numbers greater than one, 1 < M < 5. The flow
through the expansion bell of the nozzle is typically in this regime. Compressibility effects
are also important for the external rocket because shock waves are generated by the
surface of the rocket. For high supersonic speeds, 3 < M < 5, aerodynamic heating also
becomes very important for rocket design.
4. For speeds greater than five times the speed of sound, M > 5, the flow is said to be
hypersonic. At these speeds, some of the energy of the rocket now goes into exciting
the chemical bonds which hold together the nitrogen and oxygen molecules of the air. At
hypersonic speeds, the chemistry of the air must be considered when determining forces
on the object. The Space Shuttle re-enters the atmosphere at high hypersonic speeds,
M ~ 25. Under these conditions, the heated air becomes an ionized plasma of gas and
the spacecraft must be insulated from the high temperatures.
For supersonic and hypersonic flows, small disturbances are transmitted downstream within a
cone. The trigonometric sine of the cone angle b is equal to the inverse of the Mach number M
and the angle is therefore called the Mach angle.
sin(b) = 1 / M
There is no upstream influence in a supersonic flow; disturbances are only transmitted
downstream within the Mach cone.
The Mach number depends on the speed of sound in the gas and the speed of sound depends
on the type of gas and the temperature of the gas. The speed of sound varies from planet to
planet. On Earth, the atmosphere is composed of mostly diatomic nitrogen and oxygen, and the
temperature depends on the altitude in a rather complex way. Scientists and engineers have
created a mathematical model of the atmosphere to help them account for the changing effects of
temperature with altitude. Mars also has an atmosphere composed of mostly carbon dioxide.

As a rocket moves through a gas, the gas molecules are deflected around the rocket. If the speed
of the rocket is much less than the speed of sound of the gas, the density of the gas remains
constant and the flow of gas can be described by conserving momentum, and energy in the flow.
As the speed of the object increases towards the speed of sound, we must consider
compressibility effects on the gas. The density of the gas varies locally as the gas is compressed
by the object. Near and beyond the speed of sound (about 330 m/s or 700 mph on Earth at sea
level), small disturbances in the flow are transmitted to other locations isentropically (with
constant entropy) as sound waves.
For supersonic and hypersonic flows, small disturbances are transmitted downstream within a
cone. The edge of the cone is depicted two-dimensionally by the blue lines on the figure at the
top of this page. The sound waves strike the edge of the cone at a right angle and the speed of
the sound wave is denoted by the letter a. The flow is moving at velocity v which is greater than
a. From trigonometry, the sine of the cone angle mu is equal to the ratio of a and v:
sin(mu) = a / v
But the ratio of v to a is the Mach number of the flow.
M=v/a
With a little algebra, we can determine that the cone angle mu is equal to the inverse sin of one
over the Mach number.
sin(mu) = 1 / M
mu = asin(1 / M)
where asin is the trigonometric inverse sine function. It is also written as shown on the slide sin^1. Mu is an angle which depends only on the Mach number and is therefore called the Mach
angle of the flow.
We are interested in determining the Mach angle because small disturbances in a supersonic
flow are confined to the cone formed by the Mach angle. There is no upstream influence in a
supersonic flow; disturbances are only transmitted downstream within the cone.

As a rocket moves through the atmosphere, the gas molecules of the atmosphere near the rocket
are disturbed and move around the rocket. Aerodynamic forces are generated between the gas
and the rocket. The magnitude of these forces depend on the shape of the rocket, the speed of
the rocket, the mass of the gas going by the rocket and on two other important properties of the
gas; the viscosity, or stickiness, of the gas and the compressibility, or springiness, of the gas.
To properly model these effects, aerodynamicists use similarity parameters, which are ratios of
these effects to other forces present in the problem. If two experiments have the same values for
the similarity parameters, then the relative importance of the forces are being correctly modeled.
Representative values for the properties of the Earth's atmosphere and the Martian atmosphere
are given on other pages, but the actual value of the parameter depends on the state of the gas
and on the altitude.
Aerodynamic forces depend in a complex way on the viscosity of the gas. As a rocket moves
through a gas, the gas molecules stick to the surface. This creates a layer of air near the surface,
called a boundary layer, which, in effect, changes the shape of the rocket. The flow of gas reacts
to the edge of the boundary layer as if it was the physical surface of the rocket. To make things
more confusing, the boundary layer may separate from the body and create an effective shape
much different from the physical shape. And to make it even more confusing, the flow conditions
in and near the boundary layer are often unsteady (changing in time). The boundary layer is very
important in determining the drag of an object. To determine and predict these conditions,
aerodynamicists rely on wind tunnel testing and very sophisticated computer analysis.
The important similarity parameter for viscosity is the Reynolds number. The Reynolds number
expresses the ratio of inertial (resistant to change or motion) forces to viscous (heavy and
gluey) forces. From a detailed analysis of the momentum conservation equation, the inertial
forces are characterized by the product of the density r times the velocity V times the gradient of
the velocity dV/dx. The viscous forces are characterized by the viscosity coefficient mu times the
second gradient of the velocity d^2V/dx^2. The Reynolds number Re then becomes:
Re = (r * V * dV/dx) / (mu * d^2V/dx^2)
Re = (r * V * L) / mu

where L is some characteristic length of the problem. If the Reynolds number of the experiment
and flight are close, then we properly model the effects of the viscous forces relative to the inertial
forces. If they are very different, we do not correctly model the physics of the real problem and
predict incorrect levels of the aerodynamic forces.
Aerodynamic forces also depend in a complex way on the compressibility of the gas. As a
rocket moves through the gas, the gas molecules move around the rocket. If the rocket passes at
a low speed (typically less than 200 mph) the density of the fluid remains constant. But for high
speeds, some of the energy of the rocket goes into compressing the fluid and changing the
density, which alters the amount of resulting force on the object. This effect becomes more
important as speed increases. Near and beyond the speed of sound (about 330 m/s or 700 mph
on earth), shock waves are produced that affect the drag of the rocket. Again, aerodynamicists
rely on wind tunnel testing and sophisticated computer analysis to predict these conditions.
The important similarity parameter for compressibility is the Mach number - M, the ratio of the
velocity of the rocket to the speed of sound a.
M=V/a
So it is completely incorrect to measure a drag coefficient at some low speed (say 200 mph) and
apply that drag coefficient at twice the speed of sound (approximately 1400 mph, Mach = 2.0).
The compressibility of the air alters the important physics between these two cases.
The effects of compressibility and viscosity on drag are contained in the drag coefficient. For
propulsion systems, compressibility affects the amount of mass that can pass through an engine
and the amount of thrust generated by a rocket nozzle.

You might also like