Turbomachinery - Basic Theory and Applications (Mechanical - Earl Logan, JR
Turbomachinery - Basic Theory and Applications (Mechanical - Earl Logan, JR
RLLOGA N, JR
THE PENNSYLVANIA
STATE UNIVERSITY
LIBRARIES
PEROT RIMIMOTEUARA O1APE
UNIVERSITY | IBRARIFS
Digitized by the Internet Archive
in 2022 with funding from
Kahle/Austin Foundation
https://archive.org/details/turoomachineryba0O000loga
Turbomachinery
MECHANICAL ENGINEERING
A Series of Textbooks and Reference Books
EDITORS
L. L. FAULKNER S. B. MENKES
Department of Mechanical Engineering Department of Mechanical Engineering
The Ohio State University The City College of the
Columbus, Ohio City University of New York
New York, New York
Preface
Types of Turbomachines
1.1 Introduction
1.2 Geometries
1.3 Practical Uses
Basic Relations
Dimensionless Quantities
3.1 Introduction
3.2 Turbomachine Variables
3.3 Similitude
Reference
Problems
4 Centrifugal Pumps and Fans 31
4.1 Introduction ot
4.2 Impeller Flow By
4.3 Efficiency 34
4.4 Performance Characteristics 35
4.5 Design of Pumps 38
4.6 Fans 41
References 43
Problems 43
Centrifugal Compressors 45
5.1 Introduction 45
5.2 Impeller Design 47
5.3 Diffuser Design 49
5.4 Performance 50
References ae
Problems 53
6.1 Introduction 55
6.2 Stage Pressure Rise 57
6.3 Losses 60
6.4 Pump Design 61
6.5 Fan Design 65
Reference 66
Problems 66
Axial-Flow Compressors 67
7.1 Introduction 67
7.2 Basic Theory 68
7.3 Cascade Tests 69
7.4 Performance 71
References 73
Problems 74
8 Gas Turbines 75
8.1 Introduction is
8.2 Basic Theory 77
8.3 Design 81
8.4 Radial-Flow Turbine 85
References 86
Problems 86
9 Steam Turbines 87
9.1 Introduction 87
9.2 Impulse Turbines 88
9.3 Reaction Turbines 93
9.4 Design 95
References 95
Problems 95
10 Hydraulic Turbines 97
10.1 Introduction 97
10.2 Pelton Wheel 98
10.3 Francis Turbine 100
10.4 Kaplan Turbine 102
10.5 Cavitation 103
References 104
Problems 104
Index 117
Turbomachinery
l Types of Turbomachines
1.1. Introduction
1.2 Geometries
rotation
i? 2 4D
pressure drop occurs between the inlet and the outlet of the turbine; the water
exits axially and is conducted away and discharged at atmospheric pressure.
If the substance flowing through the impeller of Figure 1.1 were a gas,
then the device would be a centrifugal compressor, blower, or fan, depending
on the magnitude of the pressure rise occurring during transit from inlet to
outlet. For the reversed flow case, i.e., a radially inward flow, the machine
would be called a radial-flow gas nuine or turboexpander.
A different type of turbomachine is shown in Figure 1.2. Here the flow
. direction is generally axial, i.e., parallel to the axis of rotation. The machine
shown in this figure represents an axial-flow compressor or blower, or with a
different blade shape an axial-flow gas or steam turbine, depending on the
direction of energy flow and the kind of fluid present.
In all of the machines mentioned thus far, the working fluid undergoes
a change in pressure in flowing from inlet to outlet, or vice versa. Generally,
pressure change takes place in a diffuser _oor nozzle, and in the rotor as well.
However, there is a class of turbines in which pressure change does not occur in
the rotor. These are called impulse, or Zero-reaction, turbines, as distinquished
from the so-called reactionn turbine, which allows a pressure decrease in both
nozzle and rotor. A hydraulic turbine with zero reaction is shown in Figure
1.3, and a reaction-type hydraulic turbine appears in Figure 1.4.
Rotor vanes
Stator vanes
Water nozzle
vane :
Centrifugal machines are depicted in Figure 1.5 through 1.7, and axial-
flow turbomachines are indicated in Figures 1.8 through 1.10. A mixed-flow
pump is shown in Figure 1.11. This class of machine lies part way between
the centrifugal, or radial-flow, types and the axial-flow types.
Rotor vanes
Stationary
guide vanes
| Draft tube
‘Tail race
Casing
Guide vanes
Wi ‘Casing
Steam —>
SY
Rotor Rotor Axis
Rotor blades
Stator blades
Flow ——>
Jf Rotor axis
Impeller
Stator vanes
eae
ee Rotor axis
Sizes vary from a few inches to several feet in diameter. Fluid states vary
widely as well. Steam at near-critical conditions may enter one turbine, while
cool river water enters another. Room air may enter one compressor, while cold
refrigerant is drawn into a second. The materials encountered in the machines
are selected to suit the temperatures, pressures, and chemical natures of the
fluids handled, and manufacturing methods include welding, casting, and
machining.
Our consideration herein of the subject of turbomachines includes a
wide variety of forms and shapes, made of a variety of materials using a num-
ber of techniques. This book does not attempt to deal with all the problems
encountered by the designer or user of turbomachines,
but only with the most
general aspects of the total problem. The present treatment is concerned with
specification of principal dimensions and forms of those turbomachines en-
countered frequently in industry.
PRACTICAL USES 7
1 OD EA ee
» -
ee
a
ee
z
Help
gtae
7
a 00d gy
bg aii ce Sa he “i
: iC, aie ie yas re aeMes
j Des ae Decay) i
it ~~ ae TY ie % q~ : ' ' , SOL mi
J
se ee ee
.
7 7 : 7
% re _ ‘ ’ a bd. of
rc j ; :
A NAb Uibge Cee he |
7 i= rie ihe dilice nade U , Seth
ml r = a. © 4 7 we = eae ow
, .
'
: May i eaaes
' ;a]
: ! iy 1 Le ’ beep.
i P s
— 1
i ye ae
ty
.
oP re
, od “ 7
2
wi), | a /
re g ce " —
¥ ea
ee w ve we ‘
{pa t\ , na e a ;
- 7
ha ‘Gl be a
oa :
' a x -
Z Basic Relations
The rotor shown in cross section in Figure 2.1 will have fluid flowing in the
annulus bounded by abcda. Although fluid velocity varies radially from a tob,
it is assumed to have a single value over the entire annular section ab, namely,
the velocity V, at point 1. Similarly, at the rotor exit the velocity V, is taken
as the average of the velocity along cd. Points 1 and 2 lie on the line 1-2 which
denotes an element of a stream surface which exactly divides the flow into
two equal parts.
Figure 2.2 shows a velocity diagram at point 1. The blade, or vane, ve-
locity is calculated from
U=Nr (2.1)
for any point on the blade a radial distance r from the axis A-A of rotation.
The preilat ee vee of the rotor is denoted by N. For point 1,(2.1)
becomes ,
U,= NG (2.2)
2.2 Mass Flow Rate
The relative velocity W of the fluid, with respect to the moving vane, is added
vectorially to the blade velocity U to obtain the absolute fluid velocity V. The
relation can be expressed by
10 BASIC RELATIONS
V=Wt+U (2.3)
m=pV,A (2.5)
where A is the area normal to the flow direction. Equation (2.5) is a'statement
of conservation of mass, i.e., the mass flow rate is the same at all stations.
It is assumed that each velocity on the central streamline of the flow
passage depicted in Figure 2.1 is the average value for the entire flow area at
Ue
the position considered. The actual flow has a variable velocity across the pas-
sage. The variation can, of course, be handled mathematically through the use
of the integral form
m= f,PVm dA (2.6)
The latter form also allows for possible variations in density associated with
temperature, pressure, or concentration gradients.
Pie
Vi
age cE NG
V2
ev (2.8)
12 BASIC RELATIONS
i
h, +— =h, +—-+EV2 (2.9)
2 2D
In gas turbine or compressor applications the enthalpy and the kinetic energy
are combined to form the total enthalpy hy. Thus, (2.9) becomes
hy, = hg tE (2.10)
Compressors and pumps increase hy so that hy, >ho, , and the energy
transfer E is negative. On the other hand, turbines decrease hp, and E is posi-
tive. The work per unit mass calculated from (2.10) is also termed the head,
especially in pump or hydraulic turbine applications.
2M = Saat nS ds Ga)
icanayWS tend'y ;
where c.v. and c.s. refer to intergration over the control volume or control
surface.
Applying (2.11) to a general turbomachine, the control volume is the
volume of fluid in the casing surrounding the rotor. Forces are applied to this
fluid along the surface of the rotor, and the sum of their moments about some
point on the rotor shaftis denoted by the term on the left side of (2.11). Assum-
ing steady flow through the control volume, the first term on the right side of
(2.11) is eliminated. Noting that the quantity pV ° dS is the mass flow rate
through an elemental area dS of the control surface, and that it has a positive
sign at the outlet, a negative sign at the inlet, and is zero elsewhere, we have
2M = fa Rx V dm - fy Rx V dm (2.12)
where A, and A, refer to the flow areas at the inlet and outlet, respectively.
MOMENTUM EQUATION 13
Aligning the z axis with the rotor axis and taking the moment center at
0, as indicated in Figure 2.4, we evaluate the angular momentum term in (2.12)
by the determinant
itan) _
io]
Rx V=|r O
Vo V.
The magnitude of the momentum about the rotor axis, i.e., the z axis is the
axial component, or z component, of the moment about the origin. Thus,
Ora") et
I Do ba wesc 1) a
V; Vu Ns
Here we note that the blade speed U has been substituted for Nr. To obtain
the energy transfer E per unit mass corresponding to that in (2.7) through
_ (2.10), we simply divide (2.15) by the mass rate of flow m. Thus, the energy
transfer per unit mass from fluid to rotor, or vice versa, is given by
2.5 Applications
Ww Impulse Turbine
Flow in the impulse turbine is generally in the axial direction, and the blade
velocity is the same at the entrance and exit of the rotor. Figure 2.5 shows a
typical blade cross section and the corresponding velocity diagram. Steam or
hot gas leaves a nozzle with a velocity V, at a nozzle angle a, measured from
the tangential direction, and enters the region between the blades with rela-
tive velocity W,. Ideally, no pressure drop occurs in the blade passage, and
the relative velocity W, is equal in magnitude to W, . This is what is meant by
the term impulse turbine, also called a zero-reaction turbine. The absolute
velocity V, at the blade-passage exit is much reduced and is typically less
than half of V,. This energy, transferred from the fluid to the rotor, is found
from (2.16) by making the substitutions V,, = V, sin a,, Vy. =V, sin As ,
and UO = U>i= U; Thus,
MW,
a V :
NX U U
Figure 2.5 Velocity diagram for an impulse turbine.
APPLICATIONS 15
The law of cosines applied to the two triangles in Figure 2.5 yields two equa-
tions, which when subtracted contain the right-hand side of (2.17). Substitu-
tion into (2.17) yields
h, -h
Re ee (2.20)
he - hy
where h, is the enthalpy at the nozzle (stator) inlet, then we can write
Ve NE
WES Mesa fata tz (2.21)
V2=Nile 7B
aap 2 CP)
V2 /2-V2/2+E
Substituting (2.18) into (2.22) yields
WwW? - Ww?
Re PG)
ee ae 2
(2.23)
Vie Veo Wit
which is generally applicable to axial-flow machines. Quite commonly, in the
analysis of multistage machines it is assumed that the fluid velocity V, leav-
ing the rotor is the same as that from the stage immediately upstream, i.e.,
V, = V,. The degree of reaction would then be expressed as
16 BASIC RELATIONS
Ww? - w?
Riek Cceare sees (2.24)
v2 - v2 + w2 - w?
which we will utilize for axial-flow machines.
We have learned that the blade profile of the impulse turbine is designed
to make W, = W,. Clearly, (2.24) confirms the earlier assumption that R = 0.
Axial-Flow Compressor
An axial-flow compressor blade and velocity diagram are shown in Figure 2.6.
The fluid is deflected only slightly by the moving blade compared with the
turbine-blade deflection. Another difference is that the pressure rises in the
flow direction both in the stator and in the rotor. Pressure rise is related to
enthalpy rise, and the latter is dependent on the deflection of the fluid by the
moving blade. A relationship is obtained by eliminating E between (2.9) and
(2.18). This yields
from which the pressure ratio, and hence the pressure rise, may be determined.
It is observed that pressure rise depends on the change of relative velocity,
which is directly related to the compressor blade shape, i.e., to the angle of
deflection of the fluid.
APPLICATIONS 17
Energy transfer is also related to the deflection angle, since the applica-
tion of (2.3) gives
Centrifugal Pump
—H ake
SANG
By =
FG
The meridional component V,,, is the volume flow rate Q divided by the
flow area 271, b, , and U, = Nr,. The head is thus expressed as
Weds oe (2.32)
p
APPLICATIONS 19
ie ws
W, Vo Wo
: i
Figure 2.9 Velocity diagram for a centrifugal pump.
b= hy = Po :- Py (@333))
Hydraulic Turbine
E=gH=U,V,, (2.35)
Since V,,, = Vin, tan a, and V,,, = Q/A,, we can write the turbine head as
Further Examples
References
Allen, T., and R.L. Ditsworth. 1972. Fluid Mechanics. McGraw-Hill, New York.
Jones, J.B., and G.A. Hawkins. 1960. Engineering Thermodynamics. John
Wiley & Sons, New York.
Problems
2.1 Construct the velocity diagram for an axial-flow gas turbine having a
degree of reaction of 0.25 and minimum leaving kinetic energy «v3 /2).
2.2 Derive the relationship between torque and speed for an axial-flow im-
pulse’ turbine.
2.3 Determine energy transfer E for an axial-flow turbine in terms of blade
speed U when the degree of reaction is 0.5 and the leaving kinetic energy is
minimal.
2.4 Repeat Problem 2.3 for the impulse turbine with minimal leaving kine-
tic energy.
2.5 Sketch the head-capacity curves for centrifugal pumps having B, be-
tween 0° and 90°, equal to 0° and less than 0°.
3 Dimensionless Quantities
3.1 Introduction
=<, (3.1)
23
24 DIMENSIONLESS QUANTITIES
peFp?
a G2)
Ss
Bi tat (gH)3/4
p34
NOS (3.3)
ee Dany
st gil : Qi Co
1p
Cy ? ov im pN3D5 (3:5)
Re= = (3.6)
Another example is the inlet fluid temperature T,, or the inlet specific enthalpy
h,. Since the latter quantity contains the square of the acoustic speed in gases,
SIMILITUDE 25
M ENDS
ie (3.7)
Soul
$= (3.8)
Vip, D?
Other forms of the head coefficient are the ratio of outlet pressure to
inlet pressure p,/p,, or the ratio of outlet to inlet temperature T, /T, . Clearly,
these ratios are equivalent, since head is proportional to enthalpy difference,
which in turnis proportional to temperature difference in gas-flow machines
and pressure difference in incompressible-flow machines. In gas-flow turbo-
machines either p,/p, or T,/T, could be used, since the two are related
through isentropic or polytropic process relations.
Efficiency n has many specialized definitions, but is, in general, output
power divided by input power. It, too, can be included in the list of variables,
and since it is already dimensionless, it is also included in the list of dimen-
sionless groups.
3.3 Similitude
(0), "Sn
co CAR
-3.9
26 DIMENSIONLESS QUANTITIES
On the other hand, similar force triangles are equivalent to equal head coeffi-
cients:
(x>),
eH \ = ny a
(Ss
D2
(3.10)
Seo
Cre
NG
22
Sot
(3.11)
which is a pump law; it implies that capacity Q varies directly with speed N.
In a similar manner, we see from (3.10) that the head H or pressure rise is
governed by
H, =eHy
2) 2 (3,12)
NUNS
re
BY) Nea A scot
y
ieee
P,
Figure 3.2 Compressor map.
i.e., head varies as the square of speed. Power is the product of Q and H, and
the third pump law states that
P tore iP2
1.2 (1133)
3 3
NU
Laws for scaling up or down, i.e., varying diameter D while keeping the speed
constant, follow in a similar manner after cancelling the factors containing N.
Thus we find’
Q, Q
D>
i
D3
De
SAEY
H, 1 H 2)
ee2
pes22 (Beis)
Di Ds
Pe oP
ae (3.16)
1 Be
used to indicate the type of machine appropriate to a given service. Table 3.2
gives ranges of specific speeds corresponding to efficient operation of the turbo-
machines listed. Sizes of turbomachines required for a given service are also
determinable for correlations of groups such as those shown in Figure 3.3.
This correlation, developed by Csanady (1964), of the optimum specific
speeds of various machines as a function of specific diameter is useful in
determining an appropriate size for a given set of operating conditions. To
enter the diagram, called the Cordier diagram, a specific speed can be selected
from Table 3.2. The rotor diameter can be determined from the specific di-
ameter found from the Cordier diagram. A machine so selected or designed
would be expected to have high efficiency.
Reference
Problems
3.1 Derive expressions for specific speed, specific diameter, and power coef-
ficient by combining flow and head coefficient. Show that each is dimension-
less.
4.1 Introduction
He = WAV)
PA MAGS
i (4.1)
&
31
oo»
Suction face
Circulatory
flow
Equation (4.1) was derived earlier as (2.28). The ideal head H,; is higher than
that found in practice. Reasons for this disparity and methods for correction
will be given in subsequent sections.
Figure 4.1 shows an impeller rotating in the clockwise direction. Fluid next
to the pressure face of the vane is forced to rotate at blade speed. Motion in a
purely circular path at radius r implies a net pressure force directed radially in
inward, so that the net pressure force A dp on a differential element of cross-
sectional area A balances the centrifugal force (9 A dr)N?r; thus the radial
pressure gradient is
dp _ pU?
dr ie (4.2)
Since the fluid does not follow the impeller as in solid body rotation, but in-
stead tends to remain stationary relative to the ground, a resultant outward
flow along the vane with an accompanying adverse pressure gradient occurs.
However, the magnitude of the pressure rise across the rotor is less than that
indicated by integration of (4.2); i.e., less than
p(U3 - Uj)
Rows Pals 5) (4.3)
IMPELLER FLOW 33
p(v2 - v2)
Po — Py = PUZVy» - (4.4)
Applying the law of trigonometry to the diagrams of Figure 4.2 yields the
relations
UZ + V3 - W3
U,V = ea 300 (4.5)
and
Dy
Combining (4.4), (4.5), and (4.6) results in the pressure rise expression
_ p(UZ - UZ + Wi - W3)
PD = A Gy
Although the static pressure at the inlet and outlet of the impeller is ex-
pected to be uniform across the opening between the vanes, pressures on the
two sides of a vane are expected to be different. As the fluid moves radially
outward, its angular momentum per unit mass V,,r is clearly increased. This
means that a moment of some force has been applied to the control volume
considered. The source of such a force is obviously a pressure difference be-
tween any two points on opposite sides of the control volume at the same
radial distance from the axis of rotation. The azimuthal force resulting from
this pressure difference is the so-called Coriolis force, of magnitude |2N x WI.
This force is applied to the impeller at the pressure face and the suction face
of the vane. Equation (4.7) applied between the inlet and some intermediate
radius less than r, implies that the greater pressure rise on the pressure face is
accompanied by a lower relative velocity W on that face. Conversely, a higher
relative velocity at the suction face is indicated. Figure 4.1 shows a circulatory
flow which is radially inward on the pressure face and radially outward on the
suction face, and this is superposed on the main flow, which is radially out-
DS
Figure 4.2 Velocity diagrams at inlet and outlet.
34 CENTRIFUGAL PUMPS AND FANS
ward. The difference in pressure rise on the two sides of the passage between
vanes implies a separation, or backflow, region near the outer end of the suc-
tion face. The latter implies a flow deflection away from the suction face near
the exit of the passage. The change in V,,, associated with this flow deflec-
tion is known as slip.
The ratio of the actual V,,, to the ideal V,,, is usually known as the slip
coefficient u,. Since the slip depends on the circulation, and the circulation is
clearly dependent on the geometry of the flow passage, a theoretical relation-
ship expressing u, as a function of the number of blades ng and exit angle B, is
not surprising. Shepherd (1956) has given such a relation:
TU, sinB
Decl = MN ao coe (4.8)
Vuo np
Equation (4.8) shows that for an infinite number of vanes, i.e., perfect guidance
of the fluid, u, becomes unity. The velocity diagram of Figure 4.2 must be
modified to agree with the corrected V,,,, and the angle B, must be taken as
the fluid angle rather than the vane angle. Equation (4.1) is used to compute
the actual energy transferred to the fluid from the impeller.
4.3 Efficiency
where re is the mass rate of leakage, and m is the mass rate of flow actually
discharged from the pump.
- Because of the loss of mechanical energy by the several mechanisms
mentioned above, the head H, i.e., the net mechanical energy added to the
fluid in the pump as determined by measurement, is less than the head com-
puted from (4.1). The losses are accounted for by defining hydraulic efficiency
Nyy aS
(4.10)
es mgH
uf P (4.11)
where P is the power of the motor driving the pump as determined by dyna-
mometer test. The so-called total head H is determined from the steady-flow
energy equation after experimentally evaluating the mechanical energy terms
at the suction and discharge sides of the pump.
The mechanical efficiency n,, accounts for frictional losses occurring
between moving mechanical parts, which are typically bearings and seals, as
well as for disk friction, and is defined by
Substitution of (4.9), (4.10), and (4.11) into (4.12) yields the simple relation-
ship
= ee (4.13)
Characteristic curves for a given pump are determined by test, and they consist
primarily of a plot of head H as a function of volume rate of flow Q. A typical
characteristic curve is shown schematically in Figure 4.3. The theoretical head
from (4.1) is also shown in Figure 4.3. The actual curve is displaced downward
as a result of the losses of mechanical energy previously discussed. However,
(4.1) provides the engineer with the upper limit of performance which can be
36 CENTRIFUGAL PUMPS AND FANS
Actual
characteristics
External head
achieved, since it does not account for losses. If the speed is increased, (4.1)
indicates that the curve will shift upward, and vice versa.
Expressing (4.1) in terms of Q, we have
H; = U, : (4.14)
Dividing (4.14) by the square of twice the tip speed ND, , we obtain
H,
= Hy,
(4.16)
Ny Ny
and
Q; Q
Ni Ny (4.17)
PERFORMANCE CHARACTERISTICS 37
Equations (4.16) and (4.17) express the pump (or fan) laws. Manipulation of
these equations yields
H H
Go (4.18)
=|const. YI
/NLE
LM MALL
YiN
aa
AN
Bin
1.0 seem to have the highest maximum efficiencies (for example, see Church,
1972).
The head-capacity (H-Q) curves can be altered at the high-flow end by
the occurrence of the phenomenon of cavitation. This process consists of the
formation and collapse of vapor bubbles, which occur when the fluid pressure
falls below the vapor pressure. Outward flow in the impeller passage, which is
accompanied by pressure rise, results in a collapse of the bubble. Acceleration
of fluid surrounding the bubble, which is required to fill the void left by the va-
por, results in losses and pressure waves which cause damage to solid-boundary
materials. Since the energy transfer per unit weight is reduced by the presence
of vapor, the head-capacity curve falls off at the flow corresponding to the
beginning of cavitation.
To avoid cavitation, the net positive suction head (NPSH), defined as
the atmospheric head plus the distance of liquid level above pump centerline
minus the friction head in suction piping minus the gauge vapor pressure, is
maintained above a certain critical value. A critical specific speed S, defined as
NQ?/2
© ~ (NSPHsA (4.19)
is used to determine the lowest safe value of NSPH. For single-suction water
pumps Shepherd (1956) gives S, = 3, and for double-suction pumps he gives
S, = 4. These form useful rules of thumb for the avoidance of cavitation by
designers and users of centrifugal pumps.
| bi sora)"
Dhy =Deee
TV 9
(4.21)
The inlet velocity can be taken as 10 ft/s or less, and the leakage can be as-
sumed to be 2 or 3 percent of the capacity.
The diameter D, at the leading edge of the vane can be taken as equal
to D). Then the width b, of the leading edge can be computed using the flow
Q + Q, and a radial velocity V, slightly higher than Vg, say 1.05V9. Thus
we have
er Sy
b 4.22
: 7™D,V;, ( )
Vane angle 8, can be taken as approximately equal to the fluid angle, and
thus it is easily determined from the velocity triangle shown in Figure 4.2.
The outlet vane angle B, is usually selected as 15 to 40°, which results in
a lower energy transfer and a reduced absolute exit velocity. Csanady (1964)
has derived a formula for the determination of the optimum (minimum-loss)
ratio of whirl velocity to tip speed:
Vio (= (a a) 1/2
(4.23)
U, eae
where ¢, and §, are diffuser and impeller loss coefficients, respectively. Using
§,/§, = 1/3 and B, = 25°, we arrive at a typical relation, V,, = 0.62U, .Using
the given value of head H with (4.10) and (4.13), we can determine a value of
U, and hence D,. If the radial component V,,, of the exit velocity is too high,
it can be reduced by choosing a lower value of 8, and repeating the calcula-
tion (see Figure 4.2). Vane width b, is easily determined from
_ QQ, (4.24)
ae m™D, Vin2
The widths b, and b, can be increased slightly to account for vane thickness
in a refined calculation.
The appropriate value of B and b at stations intermediate between inlet
and exit can be worked out using velocity triangles and assuming a smooth
variation in all quantities. The optimum number of vanes is determined from
a formula of Pfleiderer (see Church, 1972), namely,
6.5(D, +D,) Me
Np = - 4.25
BD, - D,) sin [G, + 8)/21
The fluid exits the impeller with tangential and radial components of
absolute velocity and is collected and conducted to the discharge of the pump
by the volute or scroll portion of the casing (see Figure 4.6). The volute is
Vr = constant (4.26)
OQ R3
360 5
If the channel width is variable, as in a channel of circular cross section, then
the governing relation should be
OQ = R3 W
The so-called discharge nozzle, which is really a diffuser, joins the volute
to the discharge flange of the pump. For water the nozzle is typically sized to
produce a discharge velocity of 25 ft/s. A radial diffuser may be added between
the impeller and the volute for high-pressure pumps. This may take the form
of a space of constant width without vanes, or it may include vanes forming
diverging passages aligned with the absolute velocity vector.
4.6 Fans
Fans produce very small pressure heads measured in inches of water pressure
differential, and of course are employed to move air or other gases. Acom-
pressor also handles gases, but with large enough pressure rises that significant
fluid density changes occur; i.e., if density is increased by 5 percent, then the
turbomachine may be called a compressor. 3
A centrifugal fan, as compared with a pump, requires a much smaller in-
crease in impeller blade speed, i.e., a smaller radius ratio R,/R,, as may be
inferred from (4.7). It requires a volute, of course, but no diffuser is needed
to enhance pressure rise. The flow passages between impeller vanes are quite
short, as indicated in Figure 4.7.
The analysis and design of the impeller proceeds as with the centrifugal
pump. The small changes of gas density are ignored, and the incompressible
equations are applied as with pumps. Performance curves are qualitatively the
42 CENTRIFUGAL PUMPS AND FANS
Discharge
—— >
same as for pumps, except that the units of head are customarily given in
inches of water, and those of capacity are typically in cubic feet per minute.
Other differences are that both total head and static (pressure) head are
usually shown on performance curves, and a fan static efficiency, based on
(4.11), is calculated using static head (P, - P,)/pg in place of total head H.
Similarity laws for pumps are applied and are known as fan laws; these are
represented by (4.16) and (4.18).
ft
HEAD,
References
Church, A.H. 1972. Centrifugal Pumps and Blowers. Krieger, Huntington, New
York.
b& Csanady, G.T. 1964. Theory of Turbomachines. McGraw-Hill, New York.
Shepherd, D.G. 1956. Principles of Turbomachinery. MacMillan, New York.
Problems
4.1 A centrifugal water pump has the characteristic curves shown in Figure
4.8. Using the 1160-rpm curve plot the characteristics (H versus Q) for a
geometrically similar pump having twice the speed and half the diameter (of
the rotor). Show calculations that were used to obtain the coordinates plotted.
4.2 Calculate the power required to drive the original pump at 1160 rpm at
a flow rate of 3100 gal/min. Also determine the specific speed (unitless).
4.3 Fora pump impeller with a diameter D, of 1.326 ft and axial width b, =
2 in. determine the velocity diagram at the exit of the rotor for the conditions
in Problem 4.2. Vane angle B, = 25°.
4.4 Determine the principal dimensions of a centrifugal pump which can
deliver 3100 gal/min of water at a 100 ft head. The speed is 1160 rpm.
5 Centrifugal Compressors
§.1 Introduction
as well as by
Ee, (5.2)
Equations like (5.1) and (5.2) used in the same analysis require care in
handling units, since the enthalpy difference in (5.1) may carry units such as
Btu/lb, whereas (5.2) would carry units of velocity squared. Suitable conver-
45
46 CENTRIFUGAL COMPRESSORS
sion factors do not appear in the equations but must be applied in computa-
tions with them.
Since thermodynamic calculations are involved in compressor analysis
and design, the h-s diagram, such as that shown in Figure 5.1, becomes useful.
The state at the impeller inlet is indicated by point 1, and that at the impeller
outlet by point 2. The diffuser process is indicated between points 2 and 3.
The corresponding stagnation properties 01, 02, and 03 are also indicated in
Figure 5.1, since kinetic energies are usually considerable.
The expression for compressor efficiency appears to be somewhat dif-
ferent from that for pump efficiency, but, in reality, the principle of the defini-
tion is the same. Both definitions employ the ratio of the useful increase of
fluid energy divided by the actual energy input to the fluid. For the compres-
sion of a gas, the useful energy input is the work of an ideal, or isentropic,
compression to the actual final pressure P,. This is calculated from
Pee (Y-L)/Y
Ei =C,To1 722)= (5.3)
which evaluates the work of the isentropic process from state 01 to state i in
Figure 5.1. The compressor efficiency can be reduced to
Ty Tot
n —
(5.4)
i To3 = To4
Pou CyTo1
The velocity component V,,, is determined to be wU, for radial vanes (6, =
90°), as shown in Figure 5.2. Thus, the pressure ratio for a compressor stage
is expressed very simply as a function of tip speed U,, inlet temperature To, ,
slip factor u,, and compressor efficiency n,. Single stages can produce pres-
sure ratios as high as 6, which is sufficient to operate an efficient gas turbine
cycle, for example.
We
Rim of
ee impeller
Rotation
eG
Tit = Ne (5.9)
The hub radius can also be calculated from the mass flow m; thus,
ee m- \ 12 .
(5.10)
iti (3,-38.)
Referring to Figure 5.3, the fluid angle at the hub is calculated from
-1 Vy
By, = tan ue i)
The radius r, at the impeller tip can be determined from (5.5) using a
value of V,,. (u,U, for radial vanes) determined from the exit-velocity diagram.
The slip factor u, is determined as for centrifugal pumps, and the ratio of V,,5
to U, is chosen to lie in the optimal range of 0.23 to 0.35 (see Ferguson,
1963). The value of Po, is, of course, specified in advance.
The calculation of U,, and hence r,, requires the selection of a value
for compressor efficiency n,. This choice will be based on experience, but a
conservative value will probably lie in the range 70 to 80 percent (see Shepherd,
1956). Since the efficiency has been shown to vary with Reynolds number
and Mach number, both based on U,, several iterations may be required.
DIFFUSER DESIGN 49
The compressor efficiency n,, in addition to its use in (5.5), can be em-
ployed to estimate the impeller efficiency n,. The ratio x of impeller losses to
compressor losses
Mekist a= (S342)
can be estimated and lies between 0.25 and 0.50. The definition of impeller
efficiency
a i -To1
ny Cy)
To2 - tor
can be used to estimate T; (see Figure 5.1). The latter total temperature cor-
responds to the total pressure Po, , calculated from
nove (= a (S214)
Po. To,
2 _
pene ee OM! 2) Y/(Y-1) (S215)
2
The static pressure P, from (5.15) and the static temperature T, deter-
mined from
Eo) NV;
LN reece res (5.16)
C, 2C,
are used to determine density p, at the impeller exit. Finally, the axial width
b, of the impeller passage at the periphery may be found from
m rl
fie epee
UI (aig
2 2m) Vina
A vaneless diffuser, or empty space, between the leading edges of diffuser vanes
and the impeller tip allows some equalization of velocity and a reduction of
the exit Mach number. The vaneless portion, which may have a width as large
as 20 percent of the impeller diameter, also effects a rise in static pressure. As
with the pump, angular momentum rV,, is conserved, and the fluid path is ap-
proximately a logarithmic spiral. Diffuser vanes are set with the diffuser axes
50 CENTRIFUGAL COMPRESSORS
tangent to the spiral paths and with an angle of divergence between them not
exceeding 12°.
Since the addition of a vaned portion in the diffusion system results in
a small-diameter casing, vanes are preferred in instances where size limitations
are imposed. On the other hand, a completely vaneless diffuser is more efficient.
If vanes are used, then their number should generally be less than the number
of impeller vanes to ensure uniformness of flow and high diffuser efficiency
in the range of flow coefficient V,,,/U, recommended in the previous section.
Diffuser efficiency np is usually defined as (see Figure 5.1)
Np =— (5.18)
me VAS
° P3
(5.20)
T,
T 03 = T +3 V3
AS,
+ —— (Spell )
Ty NG Gy 19) 0)
Hi =(23) (5.22)
Wy)
with P93 specified, the five unknowns P,,T3, A3, V3, and T;’ are easily deter-
mined by iteration. We can omit (5.19) if the static pressure P, is specified
instead.
A volute is designed by the same methods outlined in Chapter 4. The
volute functions to collect the diffuser’s discharge around the 360° periphery
and deliver it through a single nozzle to the connecting gas-piping system or
to the inlet of the next compressor stage.
5.4 Performance
N= constant
|50
OU
1in€ —» Choke
line
the constant-speed curves at higher mass flows is due to choking in some com-
ponent of the machine. At low flows operation is limited by the phenomenon
of surge. Thus, smooth operation occurs on the compressor map at some point
between the surge line and the choke line.
The phenomenon of choking is that associated with the attainment of a
Mach number of unity. In the stationary passages of the inlet or diffuser, the
Mach number is based on the absolute velocity V. Thus for a Mach number of
unity, the absolute velocity equals the acoustic speed a, calculated from
a = (yRT)1/2 (5.23)
The temperature at this point is calculated from the total temperature Tp
using the relation
T= 1,2} (5.25)
This Mach number is found near the cross section of minimum area, or throat
(A,), so that we can estimate the choking, or maximum, mass flow rate from
1/2
th = A,P, i (5.26)
52 CENTRIFUGAL COMPRESSORS
dis TED
t in T
in
The process of estimating choked flow rate in the impeller is the same except
that relative velocity is substituted for absolute velocity. When the relative
Mach number W/a is set equal to unity in the energy equation of the rotor,
namely,
fig (5.28)
we obtain
1) 21 + U2726 519
(5:29)
To, Vier
Using the isentropic relation between pressure and temperature and substitut-
ing into the continuity relation, the mass flow rate at the throat section of
the impeller is given by
3 1/2 2 (y+1)/2(Y-1)
th= A,Poy (= : = (+ a ) farsa (5.30)
01 Fe p+o
Thus, it is clear that mass flow for choking in stationary components, given
by (5.26), is independent of impeller speed, but that mass flow for choking in
the impeller, given by (5.30), actually increases with impeller oe This is
indicated schematically in Figure 5.4.
Referring to Figure 5.4 the point A represents a point of normal opera-
tion. An increase in flow resistance in the connected external flow system
results in a decrease in V,,, at the impeller exit and a corresponding increase
in V,,,, which results in an increased head or pressure increase. However, the
surge phenomenon results when a further increase in external resistance prod-
uces a decrease in impeller flow than tends to move the point beyond
C,
where stall at some point in the impeller leads to change of direction of W,
and an accompanying decrease in the head (or pressure rise) in the impeller
A temporary flow reversal in the impeller and the ensuing buildup to the ori-
ginal flow condition is known as surging. Surging continues cyclically until
the external resistance is removed. It is an unstable and dangerous condition
and must be avoided by éareful operational planning and system design.
PROBLEMS 53
References
Ferguson,
TB. 1963. The Centrifugal Compressor Stage. Butterworths, London.
Shepherd, D.G. 1956. Principles of Turbomachinery. Macmillan, New York.
Problems
5.1 Air enters a centrifugal compressor at 1 atm, 58°F, and V, = 328 ft/s.
At the impeller exit B, = 63.4°, V,,, = 394 ft/s, and U, = 1640 ft/s. The mass
flow rate is 5.5 lb/s, the mechanical efficiency is 95 percent, and the compres-
sor efficiency is 80 percent. Determine the ratio of total pressures at outlet
and inlet and the power required to drive the machine.
5.2 Design a single-stage centrifugal compressor which will handle 2.2 lb/s of
air at a pressure ratio of 4.2:1. Use radial blades with an appropriate inducer.
Assume d,/d; = 0.361 at the inlet eye and n, = 0.78. The machine is to operate
at 60,000 rpm and to supply air to the combustion chamber of a turboject
engine. The basic design parameters required are the following:
Hub diameter
. Shroud diameter at impeller inlet
Shroud diameter at impeller exit
. Impeller inlet vane angle
Vane width at impeller exit
Velocity triangle at impeller inlet (shroud diam.)
Rwmeaoge
Velocity triangle at impeller exit
The compressor is to have no inlet guide vanes and is to draw in ambient air
at 14.7 psia and 520°R.
*
ee
—S~
< ——— ——
Tre
se
> ~
6 Axial-Flow Pumps and Fans
6.1 Introduction
Axial-flow pumps and fans move liquids and gases without significant effect
on their density. They are like propellers in that power is supplied to produce
axial motion of the fluid, but they are different in that the fluid being moved
is enclosed by a casing. Like propellers, the vanes have small curvature and
cause little deflection of the relative velocity vector of a fluid particle as it
migrates through the moving passages.
Generally, the vanes have shapes, or profiles, like that of an airfoil: they
are thin, streamlined, and cambered (see Figure 6.1). The relative velocity W,
approaches the vane at an angle a (the angle of attack) to the chord line. The
exiting fluid with relative velocity W, has been deflected slightly, and the
change of momentum results in a lift force L perpendicular to the mean direc-
tion of W, and W,,i.e., perpendicular to a mean relative velocity W,,.
The lift force L is primarily responsible for the transfer of energy, and
the drag force D, which is directed parallel to W,,, , is strongly associated with
blade losses. Lift is maximized by setting the blades at high angle of attack,
but stall occurs if the angle is too high. Such characteristics of blades are
determined in a wind tunnel using a representative set of blades arranged in
series, known as a cascade. This kind of experimentation provides informa-
tion not only about optimum incidence, but also about optimum spacing for
maximum lift and minimum drag.
55
56 AXIAL-FLOW PUMPS AND FANS
Pa L cos.
iyDsinp., “5 (6.1)
m = pV,S (6.4)
cos? B, S
Cre) scan 7 (tan B, - tan B,) (6.5)
iG
W> ey.
Axial
direction
Bo
CAS L =Hpwre
IE
(6.7)
where C is the chord, the length of a line drawn between the leading edge and
the trailing edge of the blade (see Figure 6.2). Also, in forming (6.5) we have
omitted the drag term that appears in (6.1) on the grounds that D << L. The
relationship expressed by (6.5) links the force produced by a blade with the
flow angle and the nondimensional spacing of blades, or pitch-chord ratio
S/C. For the ideal cascade, C, can be predicted from (6.5). The presence of
boundary layers and nonuniform flow deflection in actual cascades leads to
experimentally determined relationships between these variables.
A single-stage fan or pump generally consists of inlet guide vanes, rotor vanes,
and diffuser vanes. The inlet guide vanes deflect the fluid so that it enters the
moving vanes at an angle with respect to the axial direction. This is indicated
by the angle velocity V, makes with the axial direction, as shown in Figure
6.3. The kinetic energy of the fluid is increased in the rotor, so that it leaves
with velocity V, > V,. It is decelerated in the stationary blades of the dif-
fuser and deflected back to the axial direction.
Pumps and fans frequently have few blades in the rotor. The solidity,
defined as C/S, ranges from 0.4 to 0.8 or higher. For the highest specific
speeds only two blades may be used. Increasing the number of vanes increases
guidance and thereby increases head, but friction losses also are increased.
The optimum soliditiy is determined by test.
38 AXIAL-FLOW PUMPS AND FANS
ee. D
(Pht) toe aoe Bm = wee Be (6.9)
(P ~ Pi), C
= 5 (C, sin 8,, -Cp cos By) (6.10)
Yow?
a—Boundary of
/ control volume
~-@---~ Ui
ie X
Figure 6.4 Control volume for a cascade blade.
STAGE PRESSURE RISE 59
Fp,
= Fpy tan By - 9) (6.11)
um
tan B=
an B,, oe
7 ong
(6.12)
Vv a
R - ¢6
(6.13)
rN Bele eres
in which the component Fy, is replaced by using the change in tangential
momentum flow, namely,
6.3 Losses
C0102 = (6.17)
where h represents blade height and Cy is the increment to be added to the
previous drag coefficient to account for annulus losses.
The velocity variation due to boundary layers on the blades and walls of
the annulus, coupled with the curvature of the blade surfaces, results in an ad-
ditonal loss. Secondary currents are set up in a plane transverse to the flow, as
is indicated in Figure 6.5. Dissipation of the energy of these secondary currents
takes place in the blade passage and in the wakes behind the trailing edges via
vortices spawned by the interaction of neighboring secondary flow cells as
they leave the blade. Because the trailing vortices are similar to wing vortices,
it is expected that the corresponding drag is proportional to the square of the
lift coefficient. The recommended equation for drag coefficient is then
Ca 0.018C# (6.18)
The difference in pressure on the two sides of the moving blades results
in a leakage of fluid around the tip, i.e., through the narrow passage formed
between the blade tip and the casing. This loss is accounted for by the empiri-
cal formula
i
PUMP DESIGN 61
C209
Die St
acyc3/2 (6.19)
5=
Chenery eer CH
Cy (6.20)
Axial flow pumps are used for specific speeds above approximately 3, with
centrifugal pumps occupying the range below 2 and mixed-flow pumps filling
the gap between the two. They are then machines of low head, high capacity,
and a single stage. They require several well-finished rotor vanes of airfoil sec-
tion, as shown in Figure 6.1.
As a starting point in the design of an axial-flow pump we can use that
part of a Cordier diagram (Csanady) for which N, > 3. The relationship be-
tween specific speed N, and specific diameter D, is given approximately by
D, = amas (6.21)
Ss
Ce= 10 6.23
Ss (D,/D,)N3*
The solidity, calculated from (6.23), is based on a suitably chosen value of
hub-tip ratio and the required specific speed. It should lie in the range 0.4 to
1.1, and if the calculated solidity lies outside that range, a new choice of D,/D,
should be made.
The annular flow area and the required flow rate can now be used to
determine the axial velocity component V,. Thus, we have
_D.+D, re
and
ND
U= 5a ; (6.26)
where N is rotational speed in radians per second. Thus, the required fluid
angle 6, is given by
U
6, =t
=tan =1 =
(6.27)
i
U = Wu)
N ot ce (6.28)
The exit fluid angle 8, is easily found through the use of the geometric relation
U-V
6, = tan"! ve (6.29)
a
W., +W
and W,,, and W,,, are defined by Figures 6.6 and 6.7.
The hydraulic efficiency n,,, appearing in (6.28), bears the same relation-
ship to overall efficiency n as was previously indicated in connection with
centrifugal pumps, Thus, (4.9) through (4.13) all apply to the present analysis.
The overall efficiency of the single-stage pump may be estimated from (6.16),
and hydraulic efficiency 7,, will be slightly higher, as indicated by (6.13). For
example, the latter is expected to lie in the range 0.85 to 0.90.
The stagger angle i.e., the angle between the chord line of the profile
and the axial direction, is determined from the required incidence, or angle of
attack a, necessary to produce the lift coefficient C, calculated from (6.5).
Generally, we will select an airfoil section for which cascade data are available.
NACA Report 460 is an example of such a source. Figure 6.8 shows schemati-
64 AXIAL-FLOW PUMPS AND FANS
cally the sort of cascade results available in NACA reports and elsewhere. Cas-
cade results should also be checked to assure that the angle of attack chosen
does, in fact, produce the desired fluid deflection.
The blade may be twisted if the so-called free-vortex method is employed.
In this method the product of V,,, and D is kept constant, and this variation
of V,,, with radial position will result in a variation of 8, . The angle 6, varies
with U, and the blade may be twisted to provide proper guidance at the trail-
ing edge, as well as the correct incidence. Free-vortex design results in approxi-
mately uniform energy transfer at all radial positions. However, untwisted
blades may be used in the interest of economy of production.
The fluid leaving the rotor blades encounters a row of stationary vanes
(Figure 6.9). These serve to straighten the flow, i.e., to remove the whirl
component V,,,, and to increase the pressure. Referring to Figure 6.7 it is
seen that fluid enters the vanes at the angle a,, and leaves axially. The axial
component V, may be reduced by flaring the walls of the annulus by several
degrees. The angle formed by the camber line tangent at the leading edge and
the vector V, should vary from root to tip. It should be designed to provide
The design of an axial-flow fan can proceed in a manner similar to that of the
axial-flow pump. Specific speed N, can be determined from specified values
of rotational speed N in rad/s, volume flow rate Q in ft3/s, and head H in
ft? /s?.. The Cordier curve relation (6.21) may then be used to determine speci-
fic diameter D,. Finally, blade tip diameter D, is found using (6.22). The root
diameter D, is then calculated from the hub-tip ratio DID, which is chosen
to lie in the usual range of 0.25 to 0.7.
The velocity diagrams, as depicted in Figures 6.6 and 6.7, are then con-
structed as described in the previous section using (6.24) through (6.30). Fan
efficiency, estimated from (6.16), would be expected to lie in the range 0.77
to 0.86.
The fluid angles 6, and 8, required by the velocity diagrams can then
be used with cascade data such as those depicted schematically in Figure 6.10
to select a suitable solidity C/S for the mean diameter. This is expected to be
less than or equal to unity. The actual chord c would be selected to provide
an aspect ratio, i.e., blade height to chord, of 3 or more.
The angle of attack a required to produce C, calculated from (6.5) is
determined from wind tunnel cascade data for the airfoil shape and solidity
used in the design. Figure 6.8 schematically shows data of this type. The stag-
ger angle, i.e., the angle with respect to the axial direction at the chord line, is
then the difference between 8, and a.
a|”
B =Bp
Be
Figure 6.10 Cascade data.
66 AXIAL-FLOW PUMPS AND FANS
The blade is twisted to accord with the angles determined from velocity
diagrams for the tip and root diameters. Here we use the free-vortex condition,
ie., V,,.D = constant, to establish velocity triangles at the extremities of the
blade.
Reference
Stepanoff, A.J. 1957. Centrifugal and Axial Flow Pumps. Wiley, New York.
Problems
6.1 An axial-flow fan is to be designed with a tip diameter of 9.5 in. and a
hub-tip ratio of 0.5. Assuming that the fan is driven by a 1.5 hp motor and
has an overall efficiency of 80 percent, determine the flow rate and desirable
speed. The fan discharges air into the room through an exit area of 78.54 in.2.
6.2 Construct velocity diagrams for the rotor inlet and outlet at the mean
diameter for the fan considered in Problem 6.1.
, Axial-Flow Compressors
7.1 Introduction
Originally a very inefficient machine, the axial-flow compressor was not used
to compress air in the gas turbine power plant. However, the development of
the science of aerodynamics, which accompanied the development of high-
performance aircraft, made possible its present use of gas-turbine compressors.
Now a highly efficient machine, it must be studied and understood thoroughly
by engineers.
This machine resembles the axial-flow steam or gas turbine in general
appearance. Usually multistage, one observes rows of blades on a single shaft
with blade length varying monotonically as the shaft is traversed. The differ-
ence is, of course, that the blades are shorter at the outlet end of the com-
pressor, whereas the turbine receives gas or vapor on short blades and exhausts
it from long blades. A close look at the blades shows that the compressor blade
deflects the fluid through only a fraction of the angle that the turbine blade
does. This point is illustrated by Figure 7.1. Figure 7.1 also indicates that the
concave side of the blade moves ahead of the convex side; the reverse is true
of the turbine blade. Clearly, the fluid receives energy from the compressor
blade and gives up energy to the turbine blade.
Aerodynamic analysis must be carried out for compressor blades, since
flow in the boundary layer encounters an adverse pressure gradient, which
may lead to separation, stall, and the consequent surge phenomenon discussed
67
68 AXIAL-FLOW COMPRESSORS
Compressor Turbine
blade blade
The theory utilized here is similar to that presented in Chapter 6 and will not
be redeveloped. The velocity diagram in Figure 7.2 shows that the blade de-
flects the relative velocity through the angle 6, - 6. The change in tangential
component AW,,, which is equal to AV,,, is proportional to the energy transfer
E. As in (6.3), we have
E = UAV, (7.1)
From (7.2) it is clear that if the through flow velocity V, remains con-
stand, the blade speed U increases with increasing radius, and the energy transfer
per unit mass E is to remain independent of radial position, then we must
vary the fluid angles B, and 8,, and hence, we must vary the blade angles.
This is, as previously discussed, the free-vortex condition, UAV,, = constant.
As before, with the axial-flow pump, we must have twisted blades in order to
achieve this equality of energy transfer along the blade. The variation of blade
angles then implies that 6,, will vary, and hence the degree of reaction R
varies. The latter has been defined by (2.20) and can vary between zero and
unity. It is found empirically and has been shown theoretically (see Shepherd,
1956) that a value of 0.5 is a near optimum for the degree of reaction produc-
ing maximum stage efficiency. Consequently, we find this value frequently
used for a design value at the mean diameter. Another design approach is to
use a value of 0.5 for R at all radial positions. Both bases for design are used,
as well as others not discussed here.
The theory discussed in the previous section relates energy transfer to fluid
angles, blade speed, and axial velocity through the velocity diagrams drawn at
the hub, mean, and tip radii. The development of a blade design requires
the use of wind tunnel results such as those shown in Figure 7.3 (Herrig et al.,
1957). Many such results are available to the designers and they are made for
very specific blade shapes. Thus the designer will generally specify the blade
shape for which results exist, and these proportions are given in the report
of the wind tunnel results. In addition, the tests are carried out for specific
values of solidity C/S and stagger angle. For example, Figure 7.3 gives results
for the NACA 68 (18): 10 airfoil shape, a solidity 0.75, and a fixed fluid
angle 8, of 60°. The stagger angle was varied to give a range of angles of attack
from 6° to 22°. From test data such as those of Figure 7.3, the designer may
select an angle of attack. A choice of 14° in Figure 7.3 corresponds to a value
of L/D of 62 and a fluid deflection of 20°. The designer knows that this
choice of blade profile, solidity, and stagger angle will produce 6, = 60° and
B, = 40°. Use of the theory begins at this point, and the designer will thus
construct a velocity diagram at the mean diameter with these values of
70 AXIAL-FLOW COMPRESSORS
32 0.06
28 0.05
2 0.04
8, deg Cp)
20) 0.03
16 0.02
l2 OOl
8 O
Figure 7.3 Cascade results for the NACA Profile 68 (18): 10. [Source:
Herrig et al. (1957).|
B, and B,. The designer also knows that the profile drag coefficient for this
design is expected to be 0.015.
A stage design may commence after the speed N, mass noe m, and
pressure ratio P,/P, (or energy transfer E) are specified. The given variables
N, Q, and H allow the calculation of specific speed, and a Cordier diagram can
be used to find the corresponding optimum specific diameter D,; for example,
we could use (6.21). The tip diameter can then be calculated from (6.22).
A trial-and-error procedure would then ensue, in which an assumed hub-tip
ratio would lead to a velocity diagram at the mean diameter. The axial velocity
component V, from this diagram must produce the required mass flow when
substitution is made in
th = pAV, (7.3)
where A is the annular area. We may illustrate the process diagramatically,
as in Figure 7.4.
PERFORMANCE 71
Check
7.4 Performance
Ne oe ou (7.4)
03 ~ *01
This is the same definition used in (5.4) for centrifugal compressors. In
Figure 7.5, state 1 denotes conditions at the rotor inlet and state 3 those at
the stator outlet. It has been shown, however, by Cohen et al. (1974) that the
incompressible definition (6.16) predicts the stage efficiency well, because
the rise of total temperature in the stage is sufficiently small. Cascade test
results may be used to determine values of lift and drag coefficients for the
blade profile, inlet Mach number, incidence, stagger, and solidity selected for
the compressor stage design, and (6.19) and (6.20) can then be used to esti-
72 AXIAL-FLOW COMPRESSORS
mate stage efficiency. Noting that (7.4) can be rewritten in terms of the total
pressure ratio R,, we obtain
can be used for each stage with a constant value of A. The work-done factor
may be approximated from
where R, and AT in this equation denote the total pressure ratio and the
total temperature rise for the whole machine.
The compressor map for an axial-flow compressor will have the same
appearance as that shown schematically for the centrifugal compressor in
Figure 5.3. This plot of R, as a function of m with N as parameter shows
operational limits set by the phenomena of stalling at low flow rates and
choking at high flow rates. At low speeds, choking occurs in the rear stages
and stalling (due to high incidence) in the front stages, whereas the situation
is reversed at high rotor speeds. These phenomena can be predicted in advance
using indicators such as the critical Mach number M, based on inlet relative
velocity (usually M, * 0.7-0.8) to indicate the first appearance of sonic flow
in the blade passages, and the stalling incidence angle corresponding to the
maximum value of C, obtained in cascade tests.
Since temperatures increase in stages after the first, Mach numbers de-
crease. Thus the first stage will be the most likely site of shock losses. The
first stage may be designed for supersonic inlet velocities near the tips. The
leading edge of such blades will be sharp to accommodate attached oblique
shocks as discussed by Kerrebrock (1977). The blades are called transonic
in that they accommodate subsonic flow near the hub. Such a stage may be
desirable in aircraft compressors where the cross-sectional area is minimal.
References
Cohen, H., G.F.C. Rogers, and H.I.H. Saravanamuttoo. 1974. Gas Turbine
Theory. Longman, London.
Herrig, J., J.C.. Emery, and J.R. Erwin. 1957. Systematic Two-Dimensional
Cascade Tests of NACA 65-Series Compressor Blades at Low Speeds.
74 AXIAL-FLOW COMPRESSORS
NACA Tech. Note No. 3916, National Aeronautics and Space Admin-
istration, Washington, D.C.
Kerrebrock, J.L., 1977. Aircraft Engines and Gas Turbines, MIT Press, Cam-
bridge, Mass.
Shepherd, D.G. 1956. Principles of Turbomachinery, Macmillan, New York.
Problems
7.1 Use the cascade results of Figure 7.3 to determine the velocity triangles
(mean radius) and static pressure rise for a compressor stage having the fol-
lowing features: U = 1000 ft/s; ATp, = 54°F; B, = 60°; p = 0.00237 slug/
ft3.
7.2 Determine the mean radius, air angles, and blade length for the first
stage of a compressor having the following data: N = 150 rev/s; ATgg = 36°F;
m = 44 lb/s; Vi 4 Oo Nits. 590 ft/s; X = 0.96; R = 0.5 (mean radius);
Pou 1 atime, = S18cR.
; ef Gas Turbines
8.1 Introduction
The simplest gas turbine engine requires at least two major components besides
the turbine proper (see Figure 8.1). The gas (usually air) must be compressed
by a centrifugal or axial-flow compressor, and then it must be heated (usually
by burning a hydrocarbon fuel) in a combustor or heat exchanger. Gas is
delivered to the turbine inlet at an elevated pressure and temperature.
The ideal thermodynamic cycle associated with the simple gas turbine is
the Brayton cycle depicted in Figure 8.2. Process 1-2 is an isentropic compres-
sion, 2-3 an isobaric heating, and 3-4 an isentropic expansion. A more realis-
tic model of the gas processes would follow the dashed lines 1-2’ and 3+4’.
The latter processes reflect the compressor efficiency n, and turbine efficiency
n,- Besides component efficiencies n, and n,, the cycle thermal efficiency ny,
is very important. The latter efficiency is defined as
1 ae
pee We
(8.1)
where W, is turbine work per unit mass of gas, W, is compressor work, and
Q, is the heat added to the gas in process 2’-3.
Cycle thermal efficiency depends on the cycle pressure ratio P,/P, and
the cycle temperature ratio T,/T,. Usually, the design value of the peak
temperature T, is raised as high as possible, consistent with the required life
i
76 GAS TURBINES
Exhaust
Compressor
of the first-stage blades, and the cycle pressure ratio is chosen as that value
corresponding to maximum cycle thermal efficiency or maximum specific
output W, - W,. Thus determination of the optimum cycle pressure ratio is a
logical first step in the design ofa gas turbine, and it provides the turbomachine
designer with the thermodynamic state of the gas as it enters the first stage of
the turbine.
Another point that should be made is that the compressor and turbine
are interdependent. Usually mounted on the same shaft, the speed of one
must be the speed of the other. Furthermore, the pressure ratios and mass
flows are also roughly equal. The determination of a point of operation is a
process known as matching, and the two machines are matched when speeds,
mass flows, and pressure ratios are equal. Both compressor and turbine maps
are required to carry out the matching process.
In this chapter we shall consider primarily the design and performance
of axial-flow turbines, since this type is widely used in aircraft and stationary
power plants. Some space will be given to the simpler, radial-flow gas turbine
as well, for which a few applications have been found.
Val \y
Tp,01 -T 03 (8.2)
a en re
" Toi ~To3
This is simply the ratio of actual work per unit mass of gas to ideal, or isen-
tropic, work between the same total pressures.
In order to solve for stage efficiency, it is necessary to establish the
velocity diagram. This is carried out from the known upstream conditions, V, ,
P,, and T,;the design conditions, U, E, and m; and the deflection angle B,
78 GAS TURBINES
AW, =— (8.3)
Poi - Poo
Vie
S ee (8.4)
Poo P, .
for the rotor. These are useful in determining pressures P, and P,, as well as
in the calculation of stage efficiency n,.
We first modify (8.4) to read
T iy = 1) V2
oe a Z (8.8)
05 2yRT,
The stagnation temperature is constant in the nozzle, so that Tp, = Tg, , and
all inlet conditions are known at the outset. Using the experimental value of
Yg, (8.6) and (8.7) are solved for the ratio Py, /P,, and P, is found from the
known value of Pp, .The total pressure Py, is then calculated from the ratio
determined from (8.7).
The pressure P, is found in a similar manner using (8.5). Since the rotor
is moving, the relative total pressures Pp,p and Po3p are used. They are calcu-
lated from
P ~1)w2 |¥/(7-1)
ae? Gree ies (8.9)
P, 2yRT,
and
Ip = iywe |V/O-2)
JOSR 2 |efCees. (8.10)
P, 2yRT3
where T, is found from the energy relation
80 GAS TURBINES
BOE (8.11)
E
To3 = Tos Sn (8.12)
p
We must use the empirical value of Yp in (8.5) to calculate Po. p/P3. Since
Poor is known from (8.9), we can calculate P3 directly.
Finally, to obtain the stage efficiency n, using (8.2), we must calculate
a value of Tj. Referring to Figure 8.4, it is evident that P5,/P3 is equal to
Po3/P3, and thus that (V,)?/T3 is equal to Vai Therefore, we write
(y- V3
fo fo (8.13)
YR13
where
ve' & Yil /Y ican
is used to determine T,. Thus the stage efficiency is easily determined from
the cascade loss coefficients Ys and Yp. These coefficients should include
not only the profile losses obtained by direct measurement in the cascade
tunnel, but also the effects of annulus, secondary, and tip-clearance losses, as
discussed by Cohen et al. (1974). These losses are expressed in terms of Cy in
(6.17), (6.18), and (6.19). It is easily shown that they are expressible as loss
coefficients in the form
Yor +
Vee us +
Ve = (Coe
wer = , +
Ger Ce) :
um wr
aa 8 WS
Vet
vp Ye SC Cy Eee)
, wm we ’ ow wee C cos* By
(8.16)
S cos? B3 cos3 6,
for the rotor. If we denote the measured profile loss coefficients by Ymg and
Y mr for stator and rotor, respectively, then we can write
The corrected values of Yg and Yp should be used in (8.4) and (8.5) to obtain
P,,P,, and flee
8.3 Design
The stage work, or energy transfer E, is related to the turbine work through
at
E=— (8.20)
where N, is the number of stages. The mass flow rate m is related to the tur-
bine power P through
i
° P
(8.21)
We have already seen how the designer can construct a stage velocity
diagram from a knowledge of m, E, and N. We assume a specified rotational
speed N, and we have seen that E can be obtained from (8.20), providing that
a value is assumed for the number of stages N,. An approximate formula for
estimating N, initially was given by Vincent (1950) as
ee W, sin? a
3 (8.22)
@ V2 (1 - k2 sin? a)
where V,, is the assumed axial velocity, a is the nozzle angle, and ky is a blade
friction coetficient, typically having a value between 0.9 and 0.95. The stage
formula (8.22) was derived by assuming that all stage velocity diagrams are
symmetrical (see Figure 8.6) and have the same nozzle angle a and axial ve-
locity V,. The final design may utilize velocity diagrams different from this,
but this assumption is useful in the early stages of the design process.
The velocity diagram is constructed for each of the N, stages in the
manner previously discussed in connection with Figure 8.3. The pressures P,
82 GAS TURBINES
Vp = 4 (8.23)
= —
Na
(8.24)
? U
T, -T
Recess (8.25)
di eats ¢3
should have a value between 0.4 and 0.6 at the mean diameter. The Mach
numbers M, and M,, based on V, and V3, respectively, should be subsonic
and as low as possible. The nozzle exit Mach number M, , based on V, should
be subsonic or slightly supersonic, but not greater than 1.2. The correspond-
ing relative Mach number Mp,,,, based on W., should be less than the critical
Mach number, which would normally lie in the range 0.7 to 0.8. Adjustments
to the velocity diagram can be made, if necessary, to satisfy the above condi-
tions.
The free-vortex condition can be utilized to obtain velocity diagrams at
the root and tip of the blade. This is frequently done, since it yields a constant
value of energy transfer and of axial velocity, as has been pointed out pre-
viously. Twisting of blades, which is required to achieve the free-vortex con-
dition, is often avoided to reduce manufacturing expense. If a condition other
than the free-vortex condition is assumed, then the basic equations must be
utilized to determine the variation of V,, i.e., to find V, at the root and tip
of the blade. Then the velocity triangles at root, mean, and tip locations can
be constructed, and the energy transfer and mass flow rate can be determined.
The blade angles and dimensions to produce the velocity triangles must
be sought. The stator gas angles a, anda, differ from the stator blade angles
dp, and ag,, as shown in Figure 8.7. These differences in angles may be ex-
pressed as incidence i, defined by
5p = dp, - 4, (8.27)
For the moving blades we define incidence by
i= eee} (8.28)
and deviation by
55 = Bps - B3 (8.29)
ie ad eg (8,30)
84 GAS TURBINES
acs B
tana, - tana,
= 2.5 cos? a, (8.33)
ala COS Ay
c tan B, - tan B;
— = 2.5 cos? (8.34)
S Bs cos B,
for the rotor. Generally, the viaue of C/S will lie between 1.0 and 1.8.
The aerodynamic design is thus complete. However, calculation of cen-
trifugal and bending stress must also be carried out to ensure safe operation
with available materials. Stress calculations are treated by Cohen et al. (1974).
RADIAL-FLOW TURBINE 85
:
_ 2(hy= - hy)
(8.35)
2
where §, is the loss coefficient. In the rotor the loss coefficient ¢, is given by
aTea ie AW fe
The turbine efficiency is typically defined as
ciency occurs at a specific speed of 0.64, exit hub-tip ratio D,3/D,3 = 0.4,
D,3/D, = 0.7, and a, = 74°. Shepherd (1956) also presents turbine efficiency
as a function of Mach number and Reynolds number. Shepherd states that
efficiencies in the range 0.75 to 0.85 should be expected for small inward-
flow radial turbines.
Because of the uncomplicated geometry and the simple expression for
energy transfer, preliminary design calculations are extremely easy for this
machine. An estimate of turbine efficiency can be made from the literature
(e.g., see Shepherd, 1956).
References
Cohen, H., G.F.C. Rogers, and H.I.H. Saravanamuttoo. 1974. Gas Turbine
Theory. Longman, London.
Dixon, S.L. 1975. Fluid Mechanics and Thermodynamics of Turbomachinery.
Pergamon, Oxford.
Horlock, J.H. 1973. Axial Flow Turbines. Krieger, Huntington, New York.
Shepherd, D.G. 1956. Principles of Turbomachinery. Macmillan, New York.
Vincent, E.T. 1950. The Theory and Design of Gas Turbines and Jet Engines.
McGraw-Hill, New York.
Problems
8.1 Estimate the number of stages needed for a 50 percent reaction turbine
operating at standard sea level in a basic open-cycle gas-turbine plant. Turbine
inlet temperature is to be 1660°R. Assume cold-air properties throughout,
nozzle angle = 20°, axial component of velocity = 420 ft/s, minimum leaving
kinetic energy, n, = 0.90, and blade friction coefficient Kp = 0.90. Shaft speed
is 10,000 rpm and shaft power is 9000 hp. Cycle pressure ratio is 8.3.
8.2 Draw a mean velocity diagram for the turbine described in Problem 8.1.
8.3 Estimate hub and tip diameters for the last stage of the turbine of Prob-
lem 8.1.
8.4 A gas-turbine stage is designed for which To, =1100K, Po, =4 bar, V55
519 m/s, U = 340 m/s, V,. = Vaz = 272 m/s, a, = 10°, Tp, = 955 K, andB, =
20.5° (see Figure 8.6). The loss coefficients, calculated from (8.17) and (8.18),
are Y, = 0.0688 and Yp = 0.152. Determine stage efficiency.
8.5 A radial-flow gas turbine with flat radial vanes is to run at 24,200 rpm.
Gas is to enter the rotor at a radius of 6 in. and exit at a mean radius of 3 in.
Exhaust gases are to leave at 14.7 psia and 700°F and are to have a relative
Mach number of 0.75. Calculate the mass flow rate to produce 100 hp, and
estimate the blade height at the exit.
Steam Turbines
9.1 Introduction
Steam turbines, like gas turbines, are predominantly axial-flow units. They are
used extensively in power plants to drive electric generators, as are gas turbines,
but are usually much larger than gas turbines. In addition, the large units
typically use much higher pressures in the first stages and lower pressures in
their later stages than do the large gas turbines.
Steam-turbine calculations are different than gas-turbine calculations in
that tables of steam properties are substituted for simple gas relations. The
basic features of the axial-fiow turbine are the same for both steam and gas
turbines; however, the steam can contain droplets of liquid water, making the
flow a two-phase flow, and thus a further complication is added.
The simplest ideal cycle in which the steam turbine performs a function
is the Rankine cycle. This cycle is pictured on the temperature-entropy dia-
gram of Figure 9.1. The line drawn from point 1 to point 2 represents the
ideal expansion of steam in a turbine from a superheated state to a wet state.
The actual process, i.e., that with an entropy increase, is shown as a dashed
line 1-2’. The wet steam is exhausted from the turbine into a condenser where
it is condensed to a saturated liquid at point 3. A pump and boiler then act
upon the water to raise its pressure in process 3-4 and heat it in process 4-1.
The boiler, or steam generator, delivers the superheated steam to the first stage
of the turbine in state 1.
87
88 STEAM TURBINES
If the turbine stage is to be an impulse stage, the entire pressure drop must
occur in the nozzles. The purpose of the moving blade is to reduce the kinetic
energy of the steam and transfer this energy to work done on the moving
blades. The resulting energy transfer may be evaluated using (2.19):
2_y2
Ng V3
Loe aaaeT (9.1)
IMPULSE TURBINES 89
where the subscripts 2 and 3 refer to rotor inlet and exit, respectively. Typi-
cally, the velocity diagram is that shown in Figure 9.2. Note that the relative
velocity W, is equal in magnitude to W,. Thus in the ideal impulse turbine
the relative velocity changes direction from B, to B,, but the magnitude holds
constant. However, the absolute velocity leaving the rotor is much reduced
from the velocity V, exiting from the nozzles.
Since the boundary layers which form on the blade surfaces actually slow
the steam in the passage between the blades, velocity coefficients K, and Kp are
usually defined for the stator and rotor, respectively, by the following relations:
Ko = s (9 2)
: \ 2
and
Kp (9.3)
= W> ;
where the thermodynamic states 1, 2, 2’, and 3 are indicated in Figure 9.3.
These coefficients can be estimated from values given in Table 9.1. The values
in Table 9.1 were calculated from empirical data presented by Horlock (1973),
although the small effect of the Reynolds number is neglected. The quantity
AB denotes the deflection of fluid by the stator or rotor blades, and H/c,
denotes the blade aspect ratio, i.e., blade height H over axial chord c,.
The Euler turbine equation (2.16) applied to the steam turbine becomes
either
AB(deg)
20 40 60 90 100
Applying the latter form and expressing components in terms of the steam
angles B, and 6, relative to the moving blades, we have
W, sin
B, = V, cosa-U (9.8)
(9.7) becomes
Kp sin B3
E=U(V, cosa-U) (14 (9.9)
sin B,
“ BS
Baa (9.10)
V3
IMPULSE TURBINES 91
_2U U Kp sinB3
Tp
= —|\cos a= ies Cea
Vv2 Vv2 sin
B,
By differentiation with respect to U/V, we are able to find the optimum ratio
U/V.,, for maximum blade efficiency:
Ua COsia
(9712)
Ve opt 2
Since the nozzle angle a usually falls between 15° and 25°, we expect the
optimum ratio of blade speed to nozzle exit speed to lie between 0.483 and
0.453.
The actual value of np,,,, depends on the values of Kx, B,, and B;,
but it is usually quite high. It varies, however, from zero at zero blade speed
to zero at U = V,, as shown in Figure 9.4.
Torque as a function of blade speed is also shown in Figure 9.4. Torque,
denoted by T in Figure 9.4, varies in a straight line from its maximum at zero
blade speed to zero at U = V,. This easily shown by noting that torque is
power over rotational speed. Thus
ea]
T= = mE (9.13)
and
y=Nl (9.14)
U/Vp
Figure 9.4 Efficiency and torque for an impulse turbine.
92 STEAM TURBINES
fu fu
eae
Sa a
Nozzle vA Swe
mD ND Kp sinB
T=" (v,cosa- ~) (+ E (9.15)
sin B,
for the stage depicted in Figure 9.5 with Kp = Kg = 1. Thus (9.17) indicates
one-half the blade speed needed for maximum efficiency with only one row
REACTION TURBINES 93
of moving blades and thus illustrates that velocity compounding implies lower
biade speeds for the same nozzle exit velocity.
For a large steam turbine, the designer could choose to use pressure-compound-
ing, i.e., use a row of nozzles after each row of moving impulse blades. A more
effective procedure, however, is to allow the expansion to proceed in the
moving blades as well as in the nozzles. This results in a higher optimum blade
speed, a higher maximum stage efficiency than for the impulse stage, and a
broader range of blade speeds corresponding to high efficiency.
Referring to the velocity diagram in Figure 9.2 again, we see that a
reaction turbine diagram would be similar, but would have W, considerably
longer than W.. This follows from the steady-flow energy equation (2.9)
and the Euler turbine equation (2.18). Eliminating the energy transfer results
in the form
Nimes:
W> W3 (9.18)
v3 (V5?
ho, = h, ey + (9.19)
2
where the states 2 and 2’ are those depicted in Figure 9.7. Using (9.2) to sub-
stitute for V, , we have
SCRE AOD:
hy - hy = 5 (9.20)
Thus, knowing pressures P, , P,, and P3, we can easily determine V, and W;,
and we can then use (9.20) and (9.21) to estimate the actual enthalpies h,
and h,. The velocities V, and W, are also easily obtained from the velocity
coefficient.
With the velocities V, and W, known, a velocity diagram, such as that
shown in Figure 9.6, is easily constructed for a given nozzle angle a and blade
speed U. The axial and tangential components of V, are easily calcualted and
used to find W,. Assuming that a prior knowledge of A is required to select
K,, we thus know 83, and we can determine W,,, from W3, and finally V,,,
and V,. Thus all the velocities and angles may be determined using the velocity
coefficient K, to determine the enthalpy rise due to friction.
Energy transfer can be found with the help of Figure 9.6. The Euler
turbine equation becomes
=cosa (9.25)
V2 opt
which is twice the optimum blade speed found for the impulse turbine. As
mentioned previously the resulting stage efficiency is higher and the curve
flatter than for the impulse turbine, implying better off-design performance
as well. The range of speed ratios for high efficiency is roughly 0.7 < U/V, <
ey
9.4 Design
The design of steam turbines will proceed along the same lines previously
outlined in Chapter 8 for gas turbines. Of course, steam tables or Mollier
charts must be applied in lieu of perfect-gas relations for the determination of
properties, but otherwise the same methods apply. For a detailed account of
the details of steam-turbine construction and design, the reader is referred to
a book by Church (1950).
References
Problems
9.2 Determine the power produced by the turbine in Problem 9.1, assuming
full admission through all nozzles in the annulus.
10.1 Introduction
The oldest form of power-producing turbine is that utilizing the motive power
of water. Flowing rivers, streams, and waterfalls have had their energy extracted
by vanes or buckets fixed to the circumference of rotating wheels. Although
these waterwheels were designed and built as late as the nineteenth century,
and some of them were fairly efficient, they eventually gave way to more
powerful machines requiring reservoirs.
Modern installations utilize a reservoir of water, which is usually water
collected from a flowing river. The level of the reservoir next to the hydraulic
power station is maintained at a nearly constant elevation by controlling in-
flow from other reservoirs further upstream. The water flows into the hydrau-
lic turbine through a large pipe, known as a penstock. It leaves the turbine
through a diverging duct, known as a draft tube, and enters a downstream
reservoir, known as a tail race. The available head, which ranges from several
feet to several thousand feet in existing plants, is the vertical distance between
the free surfaces of the water in the reservoir and the tail race. A schematic
diagram is shown in Figure 10.1.
The designer is generally presented with an available flow rate Q, based
on runoff records, and an available head H. The turbine speed N will also be
given, since it wili usually be required to drive a generator at a prescribed rate.
a7
98 HYDRAULIC TURBINES
Reservoir
level
vie
Turbine
Draft tube
Tailrace
The choice of the type of turbine will then follow naturally after a calculation
of specific speed, since we can observe in Table 3.2 that the ranges of specific
speeds corresponding to peak efficiency are quite different for Pelton, Francis,
and Kaplan Turbines. Thus the Pelton, or tangential-flow, turbine is most
efficient for N, < 0.3, the Francis, or radial-flow, turbine is best for N, be-
tween 0.3 and 2.0, and the Kaplan, or axial-flow, turbine is best for Nea2u:
The Pelton-type turbine is illustrated in Figure 1.3. For this turbine the pen-
stock ends in a nozzle which creates a high-speed waterjet. The latter impinges
on vanes in the form of hemispheres or half-ellipsoids. The force on the vanes,
created by deflection of the water through just less than 180°, drives the
wheel around against the resistance of a load.
As indicated in Figure 10.2, the absolute velocity V, of the jet is the
arithmetic sum of its relative velocity W, and the vane speed U. Leaving the
vane at some small angle 8, , the tangential component of the absolute velocity
is given by
Vio
= W, cos, -U (10.1)
4._______________.
Vane TB Vv,
We
W,=W,=V, =U (10.3)
we have
Differentiation with respect to U/V, leads to the result that maximum hy-
draulic efficiency corresponds to
zl (10.7)
Vi opt 2
The shaft power P of the Pelton turbine may be estimated from the
foregoing relations if leakage and mechanical losses are also considered. The
latter losses are usually small, e.g., 3 to 5 percent, and are accounted for in
the volumetric and mechanical efficiencies n, and n,,, as with pumps. Thus
we can write
Pe Oy wae ues)
or, in terms of volume flow rate and head,
P= Nm NyNyPQsH (10.9)
in which the product of three efficiencies is usually denoted by n, the over-
all efficiency. It should be noted that the flow rate also depends on the head,
ee
o=(7)arccasiy"” (10.10)
100 HYDRAULIC TURBINES
V2
where C is a coefficient which accounts for head loss due to friction in the
nozzle, control valve, and penstock, and d is the jet diameter. Of course,
several jets may be used around the periphery of the wheel, in which case the
flows would be additive.
The ratio of wheel diameter D to jet diameter d varies from 6 to 25.
The buckets are positioned close together to avoid spillage, the number of
buckets varying from 20 to 30 per wheel. Such designs result in overall
efficiencies of 80 to 90 percent.
The shaft power P will again be calculated from (10.9) with hydraulic
efficiency defined by
e. U, Vi COs a,
Ny Sener (O12)
CbrD
= (10.13)
~ (2gH)'2
FRANCIS TURBINE 101
where b denotes axial width at the inlet and C denotes a velocity coefficient
which depends on frictional resistance along the entire flow path from
reservoir to tail race. Typically, the coefficient C is in the range 0.6 to 0.7.
A Cordier diagram, such as that appearing in Figure 3.3, can be used to
determine a suitable wheel diameter D. This is obtained from the specific
diameter after the Cordier diagram is entered for a known specific speed. The
desired flow rate is assured by determining the appropriate axial width b, of
the runner from (10.13).
Figure 10.4 illustrates differences in vane design arising from differences
of the specific speeds of Francis runners (rotors). Low specific speeds imply
low flow rates, which result in a smaller axial width b, and angle a, , as well
as a smaller exit diameter at the base of the rotor. Water leaves the high-
specific-speed vane with a large axial component, in contrast to the large
radial component present at the vane exit of a low-specific-speed runner. The
inlet vane angle 6, is also increased and can be as much as 135°. The lower
edge of the high-N, vane must be twisted to provide axial discharge velocities
are Lea
_— Yeap
(a)
Figure 10.4 Francis turbine runner designs: (a) low specific speed; (b) high
specific speed.
102 HYDRAULIC TURBINES
at varying blade speed. The rotor exit diameter is as large as, or larger than,
the mean inlet diameter, as shown in Figure 10.4, to accommodate the higher
flow rate associated with larger values of N,.
As indicated in Figure 1.3, the connection between the turbine casing
and the tail race, called a draft tube, is installed to conserve the available head
between these two levels. The draft tube accomplishes a diffusion as well,
and the exit velocity from the system is thus reduced, improving overall
efficiency.
E=UV,, (10.14)
BeUVuy
NH (10.15)
gH
The flow rate can be expressed in terms of a hub diameter D,, a tip diameter
D,, and a coefficient C,, which accounts for head loss due to friction. Thus
we may write
According to Kadambi and Prasad (1977), the coefficient C, can vary from
0.35 to 0.75. Power to the turbine shaft is calculated from (10.9), as before.
10.5 Cavitation
gaueae
H
(10.18)
Critical values of o, corresponding to incipient cavitation, have been correlated
with specific speed N,, defined as NP1/2/H*/4. For the Francis turbine Shep-
herd (1956) gives the critical Thoma parameter correlation as
N g2
o= 0.625(™*) (10.19)
in which specific speed is calculated using rpm, hp, and ft units for N, P, and
H, respectively. Similarly, the critical value of o for Kaplan turbines is given as
104 HYDRAULIC TURBINES
he PNY?
o = 0.28
+ —— |—— (10.20)
7.5 \100
References
Problems
10.1 Develop an expression for design torque (maximum efficiency) for the
Pelton turbine in terms of wheel diameter and jet characteristics.
10.2 For the turbine of Problem 10.1, H = 100 ft of water, By =0,K =1,
jet diameter = 6 in., C = 0.94, n = 0.9, and N = 120 rpm. Find the following
for maximum efficiency:
a. Flow rate
b. Power
c. Torque
d. Wheel diameter
11.1 Introduction
105
106 WIND TURBINES
Rotor
the vertical-axis type differs from the radial-flow turbine in that the flow is
across the rotor rather than radially inward. Figures 11.1 and 11.2 show typi-
cal arrangements for these two types of wind turbines.
In what follows an attempt will be made to discuss briefly the aerody-
namic design of the basic horizontal and vertical types of wind turbines. There
are many considerations which are not discussed in this text, such as speed
control, energy storage, and structural design. Generally, the wind machine
must be designed to operate continuously at constant speed with various
wind and weather conditions and to drive a generator or pump to supply
immediate and future energy needs. Schemes for meeting these and other
design requirements are discussed in books by De Renzo (1979), Bossel (1977),
and Cheremisinoff (1978).
If one imagines the turbine rotor to be replaced by a disk of area A which can
extract energy from the stream, as shown in Figure 11.3, then a useful ana-
lytical result follows through application of the basic equations. The theory
has been developed by Glauert (1959) for propellers and by Betz (1966) for
windmills. As indicated in Figure 11.3, the actuator causes a divergence of
streamlines and a deceleration of the fluid from an upstream speed V to a
ACTUATOR THEORY 107
p(V -v,)?
————————— (11.3)
2
we can see that the drop in total pressure across the actuator is simply
pv 2
Po - Po = PW ~~ (11.4)
Vv vay V-y,
—_ SEE phen —_~
|
|
|
Disk of
area A
Since the static pressure drop at the actuator is equal to the total pressure
drop, we have
Po - Po =P -P (11.5)
Thus the pressure difference on the two sides of the actuator is given by
; V r= vy
D=(p-p’)A CLE)
or
D = pAy, (11.8)
V = Vy
Viucaee ¥ (11.9)
D = 2pAVv(V - v) (11.10)
n - DV
12 CG ea
ACTUATOR THEORY 109
where T is the torque and 22 is rotational speed, then the efficiency compares
the actual shaft power produced to the “thrust power” available from the
stream. The actual retardation of the stream is less than DV, and it is derived
as the rate of decrease of kinetic energy:
v2 (V-vy, )
KEKES o(V OWA
va(® ; Le
(11.12)
In simplest form (11.12) reduces to D(V - v), which is somewhat less than
DV. Since this energy ideally is transferred to the rotor, we have
n sy (11.14 )
Vv
7
To determine the efficiency corresponding to maximum power, (11.14)
is used to eliminate v in (11.10). The torque-speed product is replaced by
power P; thus
P=TQ Cit215)
P = 2pAV3n7(1 - 7) (11.16)
dP
= 2pAV3[2n(1 - n) - 7] (11.17)
dn
Setting the derivative equal to zero gives one the optimum value of 7, namely,
or (11.18)
Putting (11.18) into (11.16) gives the maximum wind turbine power as
8 CIIETS)
Pax i 97 p AN?
Since the rotor imparts angular momentum to the air as it passes between the
blades, a change in whirl velocity AV,, must be indicated by the theory. The
change AV,, is expected to be proportional to the blade speed U, which is the
product of rotational speed Q and radial position r. This may be expressed as
u=(1-a)V Cir2
The actuator theory has shown that a = 1/3 for maximum power, since a is
identical with v/V in (11.18). Wilson and Lissaman (1974) show that the tur-
bine delivers power for 0 <a< 0.5. The theoretical value is helpful in that a =
1/3 can be chosen as a first approximation in design calculations.
Blade
(l-a)V
Referring to Figure 11.4, we see that the velocity triangle is easily con-
structed from a knowledge of radial position r, blade rotational speed Q,
wind speed V, and the coefficient a. Normally, the designer would be given
values for wind speed V and rotor speed Q. The solution of Glauert for the
optimum actuator disk given by Wilson and Lissaman (1974) is useful in ob-
taining approximate values of the parameter a at each radial position along
the rotor blade. Glauert’s solution is
(oe (ale
pe clamese | hCaD) (11.22)
from which the dimensionless group rQ/V can be determined for any value of
the parameter a. The relation can be used to make tables or graphs for practi-
cal use. Figure 11.5, based on (11.22), is such a plot. It is clear from Figure
11.5 that a © 1/3 over most of a practical blade. The axial component of the
absolute velocity at the rotor is then approximately two-thirds of the wind
speed.
The Glauert theory also relates the parameter a’, defined in (11.20),
to the parameter a by the relation
lis 3a
,
Cir 223)
4a-1
.34
52S)
s2
sell|e
Factor
a
0 | ee 3 4 S) 6 7
which follows from the Glauert theory presented by Wilson and Lissaman
(1974). Using (11.26), a value of the chord of the airfoil section for the tur-
bine blade can be determined at each radial position in terms of the airfoil
lift coefficient and the velocity diagram angle ¢ applicable at that position r.
The number of blades nz would have to have been previously determined
using the guidance of correlations such as those presented in Figure 11.6.
Lift and drag coefficients are utilized to compute torque for the design
obtained. The tangential component of the blade force per unit length is
given by
HORIZONTAL-AXIS MACHINES 113
24 '
i)
\
{
x
'
\
\
\
\
‘
\
Number
Blades,
of
MB
P=TQ (Cine29))
be satisfied. An iterative procedure can be carried out using the initially chosen
values of a and a’ to start the calculation of ¢, a, Cy, Cp, and new values of a
114 WIND TURBINES
and a’ from (11.30) and (11.31). After convergence is obtained the set of
values may be used to calculate blade force and torque.
where a is given by
(l-a)V Ww
NgCRAlsin 6
a= ——_______
IRV (11.33)
and R is the maximum radius of the blade with respect to the axis of rotation, as
shown in Figure 11.2. The airfoil chord is denoted by C, and the angular posi-
tion @ is as shown in Figure 11.7. The power is obtained from (11.29).
Blackwell (1974) finds that maximum efficiency is obtained when the
tip-speed ratio RQ/V is around 6. The velocity diagrams of Figure 11.7 indi-
cate a tip-speed ratio of about 4. Blackwell finds that vertical-axis turbines of
this type fail to deliver power for RQ/V <3.
Two other types of vertical-axis machines which have achieved fame are
the Savonius (S-shaped blade) and the Madaras (cylinders), For a discussion
of these and other types the reader is referred to a book by Simmons (1975).
References
Abbott, I.H., and A.E. von Doenhoff. 1959. Theory of Wing Sections. Dover,
New York.
Betz, A. 1966. Introduction to the Theory of Flow Machines. Pergamon,
Oxford.
Blackwell, B.F. 1974. The Vertical-Axis Wind Turbine, “How it Works”. Sandia
Albuquerque Laboratories Report No., SLA-74-0160, Albuquerque,
New Mexico.
Bossel, V. 1977. Energie vom Wind. Deutsche Gesellschaft fur Sonnenenergie,
Munich.
Cheremisinoff, N.P. 1978. Fundamentals of Wind Energy. Ann Arbor Science,
Ann Arbor, Michigan.
Csanady, G.T. 1964. Theory of Turbomachines. McGraw-Hill, New York.
De Renzo, D.J. 1979. Wind Power. Noyes Data Corporation, Park Ridge,
California.
Glauert, H. 1959. The Elements of Aerofoil and Airscrew Theory. Cambridge
Press, Cambridge, Massachusetts.
Simmons, D.M. 1975. Wind Power. Noyes Data Corporation, Park Ridge,
California.
Wilson, R.E., and P.B.S. Lissaman. 1974. Applied Aerodynamics of Wind
Power Machines. Oregon State University. Report No. NSF-RA-N-74-113.
Problems
11.1 The Boeing MOD-2 wind turbine has a blade diameter of 300 ft and
reaches a rated power of 2.5 MW at 27.5 mph. What is its power coefficient
under these conditions?
116 WIND TURBINES
11.2 For a horizontal-axis wind turbine operating at a = 1/3, show that the
entropy increases through the actuator disk is given by
Me ERO)2)
9 Iss)
Sketch the flow process from upstream, through the actuator disk, and into
the wake on a T-S diagram.
117
118 INDEX
[Efficiency | Performance, 27
volumetric, 34, 99 Polytropic, 16
wind-turbine, 108 Power
Energy balance, 11 brake, 38
Energy transfer, 12, 14 turbine, 109
Enthalpy turbomachine, 13
static, 11 Pressure rise
total, 12 compressor, 47, 71
Euler, 14 pump, 33,57
Pump
Fan axial-flow, 61
axial, 65 centrifugal, 17
centrifugal, 41 laws, 26, 37
laws, 37 similitude, 26
Force, blade, 56, 113
Ratio
Head hub-tip, 62 .
cavitation, 38 pitch-chord, 57
ideal, 18, 32 tip-speed, 114
total, 35 Reaction, 15, 82, 93
Reynolds number, 24
Impeller
compressor, 47 Slip, 34
pump, 38 Solidity, 58, 62, 69, 84
Incidence, 83 Speed
acoustic, 51
Loss specific, 24, 28, 61
boundary-layer, 59 Stage
leakage, 60 Curtis, 92
pump, 34 number, 81
secondary-flow, 60 Surge, 51
[Turbine] Volute, 40
steam, 87 Vortex, 64
wind, 105
Work, 75
Velocity, 9 Work-done, 72
Volume flow, 17
]
oo0728—ee05
“IQ ier
nd
,
about the textbook... ae aes x
Based on the courses Professor Logan has taught, the material in this textbook is d
arranged sequentially in order to provide a clear, unified exposition of the topic, = 9
Beginning with the first principles of physics—conservation of mass, momentum, ~
and energy—the text proceeds to centrifugal pumps. More complex machines are
then considered, concluding with the basic types of hydraulic turbines. Each
chapter addresses the questions of how the ee of turbomachine theory _—
may be applied in design and how they may be used to predict the performance
of the turbomachine under consideration. littsarathve to accompany the
text and a solutions manual is available. :
i
i
4
EARL LOGAN, JR. is a member of the faculty of the College of Engineering and
Applied Science at Arizona State University, Tempe, Arizona. He received the ;
B.S. degree (1949) and M.S. degree (1958) from Texas A&M University and the :
Ph.D. degree from Purdue University (1961). Dr. Logan’s research focuses on
turbulent boundary layers and he has published various technical papers on fluid |
mechanics, heat transfer, and flow measurement. He is a member of the Ameri- ,
can Society of Mechanical Engineers, American Society for Engineering Educa-
tion, and the American Physical Society. |