You are on page 1of 13

8.

2 Stagnation properties 3

8.2 Stagnation properties

In this section, we will learn about the static and stagnation/total states of a fluid
in motion. The stagnation or total pressure and temperature, as we will soon see,
includes the effect of the flow velocity in addition to the thermodynamics pressure
and temperature of the gas. We will define the stagnation or total state of a gas
and show us how to interpret them physically. We will also discuss how to use the
total state of a gas to analyze different flow processes occurring in aircraft engines.
Simple examples like the flow around an airfoil are presented to connect the fluid
mechanics inside a jet engine to the thermodynamics of the gas. The objective is to
understand how different fluid dynamics processes affect the static and stagnation
properties of the working fluid. We will see an extensive use of T s diagram that
is ubiquitous in engine analysis.

In the following, we compute the stagnation state properties using the steady flow
energy equation. Just to recap, the first law of thermodynamics can be written for
a flow-through device under steady state operation as

ṁ(ht2 ht1 ) = Q̇ Ẇext (8.1)

where ht is the total enthalpy and ṁ is the mass flow rate of the working fluid. The
states 1 and 2 can be the entry and exit points of the device or they can be two
different states of the working fluid undergoing a thermodynamic process. The
equation says that the change in total enthalpy (on the left-hand side) is solely
due to the heat addition or extraction from the process and by the external work
done by/on the fluid in the device. The work can be done by/on external agencies
like rotating shaft, piston movement, etc. Flow work inherent to the flow through
device is not included on the right hand side, as it is part of the total enthalpy of
the fluid on the left-hand side.

8.2.1 Stagnation point

Let us consider the stagnation point in the flow past a body. Two examples are
shown in Fig. 8.1. The first one is the flow past a cylinder, which is the most com-
mon example in fluid mechanics. The second example is an airfoil that is more
common in aircraft applications. It also has relevance in propulsion, in the form
of the shape of compressor and turbine blades. In both examples, the undisturbed
freestream flow far ahead of the body slows down and comes to rest at the stagna-
tion point. We learn in fluid mechanics that the pressure at the stagnation point is
higher than the freestream pressure.

If we place a small pressure sensor at the stagnation point on the body, we can
measure the stagnation pressure. By comparison, if the sensor is placed, hypo-
4 8 Thermodynamics of Compressible flows

Stagnation point Stagnation


streamline stg
Free stream

Free stream
Stagnation Stagnation
streamline point
(a) (b)

Fig. 8.1: The stagnation streamline in a flow past (a) a cylinder and (b) an airfoil.
A magnified view of a pitot tube placed on the airfoil is also shown in part (b).

thetically, on a fluid element as it flows past the body, we will measure what is
called the static pressure. It is the pressure measured when we are stationary or
static with respect to the fluid, and it is the thermodynamic pressure as we know it.
The static pressure is identical to the freestream pressure when the fluid element
is far away from the body, and would change as the flow takes the fluid element
from one point to another.

For incompressible flows, Bernoulli’s equation can be used to compute the stag-
nation pressure. When applied along the stagnation streamline, connecting a point
in the freestream to the stagnation point on the body, we get
r 2 r
= constant = pstg + (02 )
A
p• + • (8.2)
2 2

The above equation neglects the effect of the gravitational potential energy. This
is in line with our assumption for propulsion applications, as described in Chap-
ter 4. Additionally, the flow is assumed to be steady (a reasonable assumption),
irrotational and inviscid. The inviscid assumption (to neglect the effect of fluid
friction) is important in the present context. The flow past an airfoil (at a small
angle of attack) can be expected to be inviscid outside of the boundary layer that
forms on the surface. The boundary layer is infinitesimally thin at the stagnation
point, and hence viscous effects can be neglected along the stagnation stream line.

Compressible flow

For compressible flows encountered in gas turbines, we cannot use Bernoulli’s


equation. We revert back to the energy conservation equation and cast it in a sim-
ilar form. The first law of thermodynamics for a flow-through device is given by
8.2 Stagnation properties 5

Eq. (8.1). If we consider a thin streamtube enclosing the stagnation stream line as
our flow-through device, fluid enters at a point in the freestream and reaches the
stagnation point. There is no flow crossing the streamtube, and therefore no work
done by external forces acting on the bounding surface of the control volume. We
have no piston or shaft work, such that Wext = 0. Additionally, there is negligible
temperature gradient (and negligible heat transfer) across the stream tube, and no
addition or extraction of heat due to combustion or cooling. The total enthalpy of
the flow is therefore constant along the streamline.
2
02
A

h• + = constant = hstg + (8.3)
2 2
where h• = c p T• is the enthalpy of the fluid at freestream conditions. Note that
the expression for the stagnation enthalpy of the fluid hstg is identical to the total
enthalpy ht defined in chapter 3.

We can now define a stagnation temperature as hstg = c p Tstg , in line with the def-
inition of enthalpy in terms of the specific heat of the gas at constant pressure
c p . The stagnation temperature will be the temperature of the gas measured at the
stagnation point, and can be computed from the above equation, if we know the
freestream quantities. Substituting
p c p = gR/(g 1) and using the expression for
the speed of sound as a• = gRT , we get

(g 1) 2
Tstg = T• 1 + M• (8.4)
2
A A
where the Mach number M = /a• and is the velocity of the fluid. To compute
the stagnation point pressure, we use the fact that the flow along the stagnation
stream line is adiabatic (Q̇ = 0) and reversible (no friction and other losses); shock
losses are not included. This allows us to use isentropic relations between the
freestream state of the gas and its properties at the stagnation point.
✓ ◆gg1
pstg Tstg
= (8.5)
p• T•
which gives us
 g
(g 1) g 1
pstg = p• 1 + M•2 (8.6)
2
We can thus compute the density, energy and other properties at the stagnation
point, by knowing the static properties of the gas and its velocity in the freestream
flow far away from the body.

Let us now plot the fluid properties corresponding to the freestream conditions on
a T s diagram; it is identified by the state • in Fig. 8.2. The freestream state has
temperature T• and lies on the constant pressure line p• . If we consider a small
fluid element that travels along the stagnation stream line, it starts at the • state far
6 8 Thermodynamics of Compressible flows

T h
pstg pstg
stg stg
Tstg hstg

p p
T h

s s
(a) (b)

Fig. 8.2: The flow process along a stagnation streamline plotted in a T s and
h s diagram to identify the freestream and stagnation states of the gas.

away from the body. It undergoes an isentropic process, as discussed above, and
reaches the stagnation state denoted by stg. The stagnation state is thus reached
via a vertical line from •, without any change in entropy. It lies on the constant
pressure line corresponding to the stagnation pressure pstg given by Eq. (8.5), and
has a temperature of Tstg as computed using Eq. (8.4).

We can easily rescale the temperature axis by multiplying with the specific heat
at constant pressure, and thus get a plot of enthalpy vs. entropy. The flow pro-
cess along the stagnation streamline looks very similar on the h s plot. The only
change is the stagnation and static enthalpies on the vertical axis. The difference
between hstg and h• is the kinetic energy of the gas in the freestream flow.
8.2 Stagnation properties 7

8.2.2 Total state

The total state of a gas in motion is analogous to the stagnation state described
above, and they are often used inter-changeably. The main difference is that a to-
tal state can be defined at any point in the flow, while the stagnation properties
occur only at the stagnation point. Let us go through the development of total
temperature and pressure.

Point 1

Freestream

Point 2

Stagnation point

Fig. 8.3: Flow streamlines around an airfoil, with the points 1 and 2 marked on
the upper and lower sides.

Total temperature

In the flow over an airfoil, we now choose a location other than the freestream
and the stagnation point. We can measure the thermodynamic pressure and tem-
perature, say, at the point 1 located in the flow above the airfoil (see Fig. 8.3).
We can also measure the flow velocity and compute the Mach number. The total
temperature at point 1 can then be defined as

(g 1) 2
Tt = T 1 + M (8.7)
2
p
Here, T is the static temperature of the fluid and M = / gRT is the local Mach
A
number at point 1. This is analogous to the relation between the stagnation and
freestream temperatures. We can interpret the total temperature as the temperature
at a hypothetical stagnation point, if the flow along the streamline passing through
point 1 is brought to rest adiabatically. This can be achieved by inserting a small
hypothetical body next to the airfoil at point 1 or by inserting a pitot tube, as
shown in Fig. 8.1(b). However, the hypothetical stagnation point, where the local
temperature is equal to the total temperature, doesn’t exist in reality. In fact, the
fluid parcel moving along this streamline flows past the body and never comes to
rest.
8 8 Thermodynamics of Compressible flows

The total state of a moving gas is more of a concept than reality. The total enthalpy
of the flow is a measure of its total energy content, as described in section 3.2.1.
It is a sum of the internal energy, kinetic and potential energies, and the pressure
potential of the fluid to do flow work. The total temperature defined above is the
corresponding temperature that characterizes the total energy content of the flow.
It is the temperature the gas will attain if all its kinetic energy and potential en-
ergy is converted into enthalpy, provided it is done adiabatically, i.e. without any
heat transfer. That is exactly what happens along a stagnation streamline. The gas
slows down to zero velocity and in the process it is compressed adiabatically and
reversibly to higher pressure and temperature.

Total pressure

We plot the thermodynamic pressure, temperature and entropy of the fluid at lo-
cation 1 on a T s diagram, and identify it as the static state 1 of the gas; see
Fig. 8.4(a). We then identify the total state using the total temperature computed
above along the vertical line originating from point 1. The total state is denoted
by t1, and it lies on an isobar with pressure value given by

 g/g 1
(g 1)
pt = p 1 + M2 (8.8)
2

This is the total pressure at point 1 computed using an isentropic relation anal-
ogous to that used for the stagnation point, where p is the corresponding static
pressure. Note that a subscript 1 is ideally required to describe the properties at
point 1. Here, it is omitted to retain the generality of the expression.

T T
pstg , pt2

pstg, pt1 stg

stg Tt2 = Tstg t2


Tt1 = Tstg p2
t1
T2
p
2 p
T T
p1
T1
1

s s
(a) (b)

Fig. 8.4: The static and total states of the fluid at points 1 and 2 in the flow past an
airfoil shown in Fig. 8.3.
8.2 Stagnation properties 9

Point 1 is located above the airfoil, and is expected to have a lower pressure than
the freestream value, i.e. p1 < p• . Another point, say 2, below the airfoil (see
Fig. 8.3) will have a higher pressure, p2 > p• > p1 . The temperature values at the
three points are also in the order T2 > T• > T1 , and the T s diagrams in Fig. 8.4
show the relative magnitude of the temperature and pressure with respect to the
freestream values. The corresponding entropy values are identical, s1 = s2 = s• ,
and this is true for points outside the boundary layer. The flow here is inviscid,
and without any significant heat transfer effect, can be treated as isentropic.

Figure 8.4 also plots the stagnation states at point 1 and 2. In the case of adiabatic
flow, the total temperature of the flow remains constant. Thus the temperature at
the total states t1 and t2 are equal, and they are identical to the stagnation point
temperature plotted in Fig. 8.2. Thus all points in the flow, except for the thin
boundary layer adjacent to the airfoil surface, are characterized by a single total
temperature and entropy. The total states t1, t2, etc. all correspond to a single
point in the T s diagram, in spite of the variation of the static pressure and
temperature from one point to the other. This can lead to a major simplification
in the thermodynamic analysis of the flow around the airfoil, and by extension,
analysis of the compressor and turbine blades, which have airfoil cross-section.

In the context of propulsion application, the flow past an airfoil can be thought
of as the flow between the compressor or turbine blades. In the absence of shock
waves and large scale flow separation, the isentropic assumption can be expected
to be approximately valid. The important thing to note here is that the rotation of
the blades about the axis adds substantially to the relative velocity of the flow. The
total state of the gas is a function of the fluid velocity and is frame dependent. For
a given static pressure and temperature of the gas, the values of total pressure and
total temperature can change, when we go from the lab coordinate frame to the
rotating frame attached to the compressor.

The majority of engine applications, except for compressor and turbine blades,
encounter flow through ducts. Examples include the air intake and nozzle, and the
combustor to some extent. Here, the flow never comes to rest, as at a stagnation
point. Nevertheless, calculating the total state of the gas, corresponding to a hypo-
thetical stagnation point is important. As we noted earlier, the total temperature is
a measure of the total energy content. It can be altered only by heat exchange and
due to work interaction with external agencies. It is thus a convenient way to keep
track of the shaft work done by or on the fluid and the heat added by combustion.
10 8 Thermodynamics of Compressible flows

T pstg

stg
Boundary
Tstg bt
layer edge at

a b p

Freestream T ,a pb
c
Tb
d b
Stagnation
point s
(a) (b)

Fig. 8.5: Thermodynamic analysis of flow past a body under adiabatic and irre-
versible conditions.

8.2.3 Flow processes in terms of total state

We will now consider a few illustrative cases from the flow around an airfoil to
analyse different flow processes in terms of the total and static states of the fluid.
The examples are representative of similar flow processes that occur in different
parts of an aircraft engine. Understanding these simple examples, and how to an-
alyze them thermodynamically, will naturally lead to the more complex engine
components subsequently.

Consider the streamline ab passing over the top of the airfoil in Fig. 8.5(a). Point a
is located in the undisturbed freestream and point b is in the inviscid flow outside
the boundary layer that forms on the top surface of the airfoil. The streamline ab
thus spans the inviscid flow region, without any wall friction and other viscous
effects. In addition, there is no heat addition or cooling in this region. A fluid
element flowing along the streamline thus undergoes an adiabatic and reversible
process, i.e. an isentropic process. The two states a and b are plotted in the T s
diagram, in terms of their static pressure and temperatures; see Fig. 8.5(b). They
are located along a vertical constant-entropy line. The static properties of the fluid
at point a are identical to the freestream conditions, denoted by •. The pressure
on the top of the airfoil is lower than the freestream value, such that it contributes
to the lift force generated by the airfoil. A lower pressure is also associated with
lower temperature and point b is located below point a on the T s diagram.

The total enthalpy of a fluid undergoing an adiabatic process is conserved in the


absence of external work. The total temperatures at points a and b, calculated us-
ing Eq. (8.7), are therefore identical and they are equal to the stagnation point
temperature. Further, both the total states at and bt are attained via hypotheti-
8.2 Stagnation properties 11

cal isentropic processes, by definition, starting from the respective static states.
They are located on the vertical line extending up from the points a and b. Match-
ing total temperature and entropy for the total states, also result in identical total
pressure (equal to pstg ) for at and bt. The difference between the static and total
temperatures is a measure of the kinetic energy of the flow. They are marked in
the figure and show that the velocity of the flow over the airfoil (point b) is, as
expected, higher than the freestream velocity at point a.

8.2.4 Effect of friction

We next consider a streamline cd that passes close to the airfoil surface and enters
the viscous boundary layer on the lower side; see Fig. 8.5(a). Studying this case
will bring out the effect of friction on total pressure and total temperature of the
gas. We will assume that the surface of the airfoil is insulated, such that there is
no significant heat transfer in the flow. The fluid element going from c to d thus
undergoes an adiabatic process between the points c and d. The process is not
reversible, owing to viscous effects, and thus it is non isentropic. Friction losses
result in an increase in entropy of the fluid element, which we identify as our
system. This means that the path cd on the T s diagram will be inclined to the
right; see Fig. 8.6(a). Also, being on the lower side of the airfoil implies that the
pressure at point d will be higher than that at c, i.e. the freestream pressure. The
isobar corresponding to pd is located above that of p• , and we place the static
state d on this line, to the right of point c.

The total state ct is identical to the stagnation state, as described above. To identify
the total state dt corresponding to the point d in the flow, we draw a vertical
isentropic line from state d. This represents the hypothetical isentropic process
between a given static state and the corresponding total state. Separately, we draw
a horizontal line (constant temperature) from the total state ct, such that the total
temperature is conserved between points c and d. The point of intersection of the
two lines is the total state dt, and it has a lower total pressure than the stagnation
point value (pdt < pstg ). This is commonly known as the total pressure loss, and
often occurs due to viscous losses. Also, it can be seen that the kinetic energy at
d is substantially lower than that in the freestream. This is due to the combined
effect of friction in the boundary layer and the slower flow below the airfoil than
above.

The flow process depicted by this test case is analogous to the flow in an air intake
or diffuser, which slows down the flow from flight velocity to compressor entry
conditions. The core flow (away from the walls) in a diffuser can be treated as
inviscid, but the boundary layers on the intake walls can lead to total pressure loss.
By comparison, the case of inviscid flow with a reduction in pressure along the
12 8 Thermodynamics of Compressible flows

T T

pstg pstg

pdt pdt
ct dt ct dt
Tstg pd Tstg pd

p p
d d

T T
c
c

s s
s s
(a) (b)

Fig. 8.6: A viscous flow process is represented in a T s diagram using (a) the
static states of the two end points, and (b) an alternate path involving the static
and stagnation states.

flow path is analogous to an ideal adiabatic nozzle, where the gas expands to reach
lower pressure and temperature, with an associated increase in its velocity. The
total pressure and total temperature remain unchanged in the absence of friction
in an ideal nozzle.

The entropy change during the thermodynamic process cd is computed in terms of


the static temperatures and static pressures at the two end points of the streamline.
Applying Gibb’s equation, we get

Td pd
D s = sd sc = c p log R log (8.9)
Tc pc

where the rise in pressure (pd > pc ) and temperature (Td > Tc ) together contribute
to the change in entropy. We can consider an alternate hypothetical process on the
T s diagram, in which the fluid element is first taken from state c to its stagnation
state ct, then from ct to the stagnation state dt, followed by a third step which takes
the system from dt to the static state d. The process is shown in Fig. 8.6(b). We
note that the first and the third legs of the process are isentropic, such that the net
entropy change during the entire process c ct dt d is given by the entropy
increase between ct and dt.
Tdt pdt
D s = sdt sct = c p log R log (8.10)
Tct pct
8.2 Stagnation properties 13

The first term on the right-hand side drops out because of the adiabatic condition
(Tdt = Tct ), and we get

pdt D s/R
= exp (8.11)
pct

Physically, the fluid element traveling along the streamline cd loses some of its
velocity and momentum due to the wall friction in the boundary layer. It thus
undergoes an irreversible process with a rise in entropy, sd > sc . This corresponds
to a decrease in total pressure of the gas, pdt < pct even though the static pressure
increases, pd > pc . In other words, the loss in momentum of the fluid appears in
the form of a loss in total pressure.

In the context of incompressible flows, the total pressure or stagnation pressure is


a sum of the static pressure and the dynamic pressure (neglecting gravity effects).
It is thus a measure of the total momentum of the fluid, including the pressure
force and the momentum of the fluid. Friction reduces the velocity and momen-
tum of the working fluid, and thus has a direct effect on the total pressure. On the
other hand, friction on the walls of a flow-through device does not alter its total
enthalpy or total temperature. Thus a loss in total pressure allows us to quantify
the effect of friction and the resulting entropy rise, in the absence of heat transfer.
More important, total pressure loss tells us about the loss in momentum encoun-
tered because of viscous effects in a device. Conserving the overall momentum of
the fluid and enhancing it in the engine are crucial for generating adequate thrust.

8.2.5 Effect of heat transfer

We finally introduce the effect of heat interaction in the form of wall cooling
on the airfoil surface. The boundary layer flow adjacent to the body would feel
the effect of the heat transfer, say along the streamline e f shown in Fig. 8.7(b).
Wall cooling is a common feature in many engine components. Examples include
combustor casing, turbine blades, nozzle walls and possibly in the later stages of
axial-flow compressors.

Figure 8.7(a) plots the T s diagram of the flow process from point e to f and it is
similar to the earlier case. The static pressure changes from the freestream value
at point e to a lower value on top of the airfoil. The static temperature follows
identical trend, and there is an increase in entropy given by,

Tf pf
sf se = c p log R log (8.12)
Te pe
14 8 Thermodynamics of Compressible flows

The decrease in pressure and temperature from e to f renders the two logarithm
terms in the above equation negative. An increase in entropy is possible only when
the pressure term is more negative than the temperature one, i.e. the pressure de-
crease supersedes the temperature drop, as explained below. This translates to the
condition ✓ ◆gg1
pe Te
>
pf Tf

Note that the pressure and temperature at point e correspond to the freestream
values, while the static pressure at point f is determined by the inviscid flow
outside the boundary layer. On the other hand, the temperature T f is influenced by
the outer inviscid flows and the boundary layer cooling. If the point f was located
in the invscid flow, the isentropic relation between pressure and temperature holds.
The inequality in the above equation will then be replaced by an equality.

The viscous effect in the boundary layer slows down the fluid and heats it up.
The reduction in the kinetic energy appears in the form an increase in its internal
energy. The effect of friction thus leads to a higher static temperature than that
for inviscid isentropic flow. On the other hand, wall cooling decreases the gas
temperature. In the present case, the combined effect of friction and cooling still
results in a net decrease in fluid temperature as it goes from point e to f .

The change in entropy between the total states f t and et can also be computed
using Gibb’s equation, akin to Eq. (8.10). The only difference is the decrease in
total temperature due to wall cooling (T f t < Tet ). There is an associated decrease
in total pressure, in addition to that due to friction effects, such that p f t < pet .
Also, the flow process from e to f is shown by a curved line to imply that the

T
pstg
et Boundary
Tet=Tstg pft
layer edge
Tft
ft f
pe
e Q
Tet =T e
pf
Tf
f
Stagnation
se sf s point
(a) (b)

Fig. 8.7: A flow process with viscous and cooling effect shown in (a) T s dia-
gram, and (b) the corresponding physical picture along the streamline e f on the
airfoil.
8.2 Stagnation properties 15

initial part of the streamline is in the inviscid region outside of the boundary layer.
The flow is isentropic up to the point where the streamline enters in to the viscous
near wall region.

You might also like