FM-II Chapter 03 Complete
FM-II Chapter 03 Complete
TURBOMACHINERY
(Fluid Flow Devices)
Email: [email protected]
Books
Fluid Mechanics; fundamentals and applications by Y. A. Cengel & J.M. Cimbala, 4th Edition
(Chapter 14)
Pumps: Any fluid machine that adds energy to a fluid. Also known as energy absorbing devices
since energy is supplied to them, and they transfer most of that energy to the fluid. The increase in
fluid energy is usually felt as an increase in the pressure of the fluid.
Turbines: are energy producing devices; they extract energy from the fluid and transfer most of that
energy to some form of mechanical energy output. The fluid at the outlet of a turbine suffers an
energy loss, typically in the form of a loss of pressure.
A fan is a gas pump with relatively low pressure rise and high flow rate. Examples include ceiling
fans, house fans, and propellers.
A blower is a gas pump with relatively moderate to high pressure rise and moderate to high flow
rate. Examples include centrifugal blowers and squirrel cage blowers in automobile ventilation
systems, furnaces, and leaf blowers.
A compressor is a gas pump designed to deliver a very high pressure rise, typically at low to
moderate flow rates. Examples include air compressors that run pneumatic tools and inflate tires at
automobile service stations, and refrigerant compressors used in heat pumps, refrigerators, and air
conditioners.
Positive-displacement Machines:
Fluid is directed into a closed volume. Energy transfer to the fluid is accomplished by movement of
the boundary of the closed volume, causing the volume to expand or contract, thereby sucking fluid
in or squeezing fluid out, respectively.
Dynamic Machines:
There is no closed volume; instead, rotating blades supply or extract energy to or from the fluid. For
pumps, these rotating blades are called impeller blades, while for turbines, the rotating blades are
called runner blades or buckets.
Gear Pump Double Screw Pump Flexible-tube Peristaltic Pump Three-lobe Rotary Pump
Dynamic Pumps
Some fundamental parameters are used to analyze the performance of a pump. The mass flow rate of
fluid through the pump is an obvious primary pump performance parameter. For incompressible flow,
it is more common to use volume flow rate rather than mass flow rate. In the turbomachinery
industry, volume flow rate is called capacity and is simply mass flow rate divided by fluid density;
The performance of a pump is characterized additionally by its net head H, defined as the change in
Bernoulli head between the inlet and outlet of the pump
Note:
EGL (Total Fluid Head) = Velocity Head + Pressure Head + Elevation head
Brake horsepower: The external power supplied to the pump is called the brake horsepower.
Pump Efficiency:
All pumps suffer from irreversible losses due to friction, internal leakage, flow separation on blade
surfaces, turbulent dissipation, etc. Therefore, the brake horsepower (bhp) supplied to the pump
must be larger than water horsepower (whp).
Keeping the above statement in mind; we define the pump efficiency to be the ratio of whp to bhp.
There are unfortunate situations where the system curve and the pump performance curve
intersect at more than one operating point. This can occur when a pump that has a dip in its net
head performance curve is mated to a system that has a fairly flat system curve. Although rare,
such situations are possible and should be avoided, because the system may “hunt” for an
operating point, leading to an unsteady-flow situation.
Mechanical Engineering Department 18
Pump Performance Curves and Matching
a Pump to a Piping System
▪ The equation for Hrequired is given is follows: 𝜶 = kinetic energy correction factor
𝑷𝟐 − 𝑷𝟏 𝜶𝟐 𝑽𝟐𝟐 − 𝜶𝟏 𝑽𝟐𝟏
𝑯𝒓𝒆𝒒𝒖𝒊𝒓𝒆𝒅 = + + 𝒛𝟐 − 𝒛𝟏 + 𝒉𝒍𝒐𝒔𝒔𝒆𝒔
𝝆𝒈 𝟐𝒈
▪ The useful pump head delivered to the fluid does 4 things;
▪ It increases the static pressure of the fluid from point 1 to point 2 (first term on the right).
▪ It increases the dynamic pressure (kinetic energy) of the fluid from point 1 to point 2 (second
term on the right).
▪ It raises the elevation (potential energy) of the fluid from point 1 to point 2 (third term on the
right).
▪ It overcomes irreversible head losses in the piping system (last term on the right).
We assume that there is no turbine in the system, although that term can be added back in, if necessary.
Mechanical Engineering Department 19
Pump Performance Curves and Matching
a Pump to a Piping System
𝑷𝟐 − 𝑷𝟏 𝜶𝟐 𝑽𝟐𝟐 − 𝜶𝟏 𝑽𝟐𝟏
𝑯𝒓𝒆𝒒𝒖𝒊𝒓𝒆𝒅 = + + 𝒛𝟐 − 𝒛𝟏 + 𝒉𝒍𝒐𝒔𝒔𝒆𝒔
𝝆𝒈 𝟐𝒈
▪ In a general system, the change in static pressure, dynamic pressure, and elevation may be either
positive or negative, while irreversible head losses are always positive.
▪ In many mechanical and civil engineering problems in which the fluid is a liquid, the elevation
term is important but when the fluid is a gas, such as in ventilation and air pollution control
problems, the elevation term is almost always negligible.
▪ To match a pump to a system, and to determine the operating point, we equate Hrequired of to
Havailable, which is the (typically known) net head of the pump as a function of volume flow rate.
▪ The most common situation is that an engineer selects a pump that is somewhat heftier than
actually required. The volume flow rate through the piping system is then a bit larger than needed,
and a valve or damper is installed in the line so that the flow rate can be decreased as necessary.
A local ventilation system (hood and exhaust duct) is used to remove air and contaminants produced
by a dry-cleaning operation. The duct is round and is constructed of galvanized steel with longitudinal
seams and with joints every 30 in (0.76 m). The inner diameter (ID) of the duct is D = 9.06 in (0.230 m),
and its total length is L = 44.0 ft (13.4 m). There are five CD3-9 elbows along the duct. The equivalent
roughness height of this duct is 0.15 mm, and each elbow has a minor (local) loss coefficient of KL = C0
= 0.21. Note the notation C0 for minor loss coefficient, commonly used in the ventilation industry
(ASHRAE, 2001). To ensure adequate ventilation, the minimum required volume flow rate through the
duct is 𝑽ሶ = 600 cfm (cubic feet per minute), or 0.283 m3 /s at 25°C. Literature from the hood
manufacturer lists the hood entry loss coefficient as 1.3 based on duct velocity. When the damper is
fully open, its loss coefficient is 1.8. A centrifugal fan with 9.0-in inlet and outlet diameters is available.
Its performance data are shown in Table on next slide, as listed by the manufacturer. Predict the
operating point of this local ventilation system, and draw a plot of required and available fan pressure
rise as functions of volume flow rate. Is the chosen fan adequate?
Mechanical Engineering Department 21
Pump Performance Curves and Matching
a Pump to a Piping System
A washing operation at a power plant requires 370 gallons per minute (gpm) of water. The required net
head is about 24 ft at this flow rate. A newly hired engineer looks through some catalogs and decides
to purchase the 8.25-in impeller option of the Taco Model 4013 FI Series centrifugal pump. If the pump
operates at 1160 rpm, as specified in the performance plot, she reasons, its performance curve
intersects 370 gpm at H = 24 ft. The chief engineer, who is very concerned about efficiency, glances at
the performance curves and notes that the efficiency of this pump at this operating point is only 70
percent. He sees that the 12.75-in impeller option achieves a higher efficiency (about 76.5 percent) at
the same flow rate. He notes that a throttle valve can be installed downstream of the pump to increase
the required net head so that the pump operates at this higher efficiency. He asks the junior engineer
to justify her choice of impeller diameter. Namely, he asks her to calculate which impeller option (8.25-
in or 12.75-in) would need the least amount of electricity to operate. Perform the comparison and
discuss.
Mechanical Engineering Department 24
Pump Performance Curves and Matching
a Pump to a Piping System
▪ When pumping liquids, it is possible for the local pressure inside the
pump to fall below the vapor pressure of the liquid, Pv. (Pv is also
called the saturation pressure Psat)
▪ To avoid cavitation, we must ensure that the local pressure everywhere inside the pump stays
above the vapor pressure.
▪ Since pressure is most easily measured (or estimated) at the inlet of the pump, cavitation criteria
are typically specified at the pump inlet.
▪ It is useful to employ a flow parameter called net positive suction head (NPSH), defined as the
difference between the pump’s inlet stagnation pressure head and the vapor pressure head
𝑷 𝑽𝟐 𝑷𝑽
𝑵𝑷𝑺𝑯 = + −
𝝆𝒈 𝟐𝒈 𝒑𝒖𝒎𝒑 𝒊𝒏𝒍𝒆𝒕
𝝆𝒈
▪ The pump manufacturer publishes a performance parameter called the required net positive suction
head (NPSHrequired), defined as the minimum NPSH necessary to avoid cavitation in the pump.
▪ The measured value of NPSHrequired varies with volume flow rate, and therefore NPSHrequired is often
plotted on the same pump performance curve as net head.
▪ Note that since NPSHrequired is usually much smaller than H over the majority of the performance
curve, it is often plotted on a separate expanded vertical axis for clarity.
▪ In order to ensure that a pump does not cavitate, the actual or available NPSH must be greater
than NPSHrequired.
▪ To illustrate the operation of a positive-displacement pump, we sketch four phases of half of a cycle
of a simple rotary pump with two lobes on each rotor
▪ Positive-displacement pumps are a good choice for pumping highly viscous fluids and slurries. They
are used, for example, as automobile engine oil pumps and in the foods industry to pump heavy
liquids like syrup, tomato paste, and chocolate, and slurries like soups.
▪ Positive-displacement pumps have many advantages over dynamic pumps. For example, a positive-
displacement pump is better able to handle shear sensitive liquids since the induced shear is much
less than that of a dynamic pump operating at similar pressure and flow rate. Blood is a shear
sensitive liquid, and this is one reason why positive-displacement pumps are used for artificial
hearts.
▪ A well-sealed positive-displacement pump can create a significant vacuum pressure at its inlet, even
when dry, and is thus able to lift a liquid from several meters below the pump. We refer to this kind of
pump as a self-priming pump
▪ Finally, the rotor(s) of a positive-displacement pump run at lower speeds than the rotor (impeller) of a
dynamic pump at similar loads, extending the useful lifetime of seals
Mechanical Engineering Department 32
Positive Displacement Pumps (Disadvantages)
▪ Their volume flow rate cannot be changed unless the rotation rate is changed. (This is not as simple
as it sounds, since most AC electric motors are designed to operate at one or more fixed rotational
speeds.)
▪ They create very high pressure at the outlet side, and if the outlet becomes blocked, ruptures may
occur or electric motors may overheat
▪ Because of their design, positive-displacement pumps sometimes deliver a pulsating flow, which
may be unacceptable for some applications.
▪ Analysis of positive-displacement pumps is fairly straightforward. From the geometry of the pump,
we calculate the closed volume (𝑽𝒄𝒍𝒐𝒔𝒆𝒅 ) that is filled (and expelled) for every 𝑵 rotations of the shaft.
Volume flow rate is then equal to rotation rate 𝑵ሶ times 𝑽𝒄𝒍𝒐𝒔𝒆𝒅 divided by 𝑵.
𝑽𝒄𝒍𝒐𝒔𝒆𝒅
▪ Example 3-4: 𝑽ሶ = 𝑵ሶ
𝑵
A two-lobe rotary positive-displacement pump, moves 0.45 cm3 of SAE 30 motor oil in each lobe
volume Vlobe. Calculate the volume flow rate of oil for the case where 𝑵ሶ = 900 rpm.
We make the approximation that flow enters the control volume with
uniform absolute velocity 𝑽𝟏 around the entire circumference at
𝟏
→ 𝒏𝒆𝒕 𝒉𝒆𝒂𝒅 = 𝑯 = ∗ (𝝎𝒓𝟐 𝑽𝟐,𝒕 − 𝝎𝒓𝟏 𝑽𝟏,𝒕 )
𝒈
Mechanical Engineering Department 42
Dynamic Pumps
Example 3-5:
triangle in Fig. formed by absolute velocity vector 𝑽𝟐 , relative velocity vector 𝑽𝟐,𝒓𝒆𝒍𝒂𝒕𝒊𝒗𝒆 , and the
tangential velocity of the blade at radius r2 (of magnitude 𝜔r2) we get;
𝟏
𝑯= ∗ 𝑽𝟐𝟐 − 𝑽𝟐𝟏 + 𝝎𝟐 𝒓𝟐𝟐 − 𝝎𝟐 𝒓𝟐𝟏 − 𝑽𝟐𝟐,𝒓𝒆𝒍𝒂𝒕𝒊𝒗𝒆 − 𝑽𝟐𝟏,𝒓𝒆𝒍𝒂𝒕𝒊𝒗𝒆
𝟐𝒈
In words, Eq. 3 states that in the ideal case (no irreversible losses), the net head is proportional to
the change in absolute kinetic energy plus the rotor-tip kinetic energy change minus the change in
relative kinetic energy from inlet to outlet of the impeller. 𝑷 𝑽𝟐 𝑷 𝑽𝟐
𝑯= + +𝒛 − + +𝒛 → 4
Equating Eq. 3 and 4 we get; 𝝆𝒈 𝟐𝒈 𝟐
𝝆𝒈 𝟐𝒈 𝟏
Eq. 5 is commonly called the Bernoulli equation in a rotating reference frame and is valid only for
the ideal case in which there are no irreversible losses through the impeller. Nevertheless, it is
valuable as a first-order approximation for flow through the impeller of a centrifugal pump.
Eq. 6 can be used for preliminary design of the impeller blade shape
1. Angles 𝛽1 and 𝛽2
𝝆𝝎𝑫𝟐
𝑹𝒆 = → As 𝝎𝑫 is a characteristics velocity
𝝁
𝜺
→ Dimensionless roughness parameter [you are already familiar with this term]
𝑫
Mechanical Engineering Department 50
Dynamic Pumps
Pump Scaling Laws:
𝓥ሶ
▪ 𝑪𝑸 = 𝑪𝒂𝒑𝒂𝒄𝒊𝒕𝒚 𝑪𝒐𝒆𝒇𝒇𝒊𝒄𝒊𝒆𝒏𝒕 = 𝝎𝑫𝟑
𝒃𝒉𝒑
▪ 𝑪𝒑 = 𝑷𝒐𝒘𝒆𝒓 𝑪𝒐𝒆𝒇𝒇𝒊𝒄𝒊𝒆𝒏𝒕 =
𝝆𝝎𝟑 𝑫𝟓
𝒈𝑵𝑷𝑺𝑯𝒓𝒆𝒒𝒖𝒊𝒓𝒆𝒅
▪ 𝑪𝑵𝑷𝑺𝑯 = 𝑺𝒖𝒄𝒕𝒊𝒐𝒏 𝑯𝒆𝒂𝒅 𝑪𝒐𝒆𝒇𝒇𝒊𝒄𝒊𝒆𝒏𝒕 = 𝝎𝟐 𝑫𝟐
𝒈𝑯 𝓥ሶ 𝝆𝝎𝑫𝟐 𝜺
=𝒇 , , → 7 Lets understand the interpretation of Eq. 7 and 8 for the sake
𝝎𝟐 𝑫 𝟐 𝝎𝟐 𝑫𝟐 𝝁 𝑫
corelating to similar pumps, in the next slide. Note that these
𝒃𝒉𝒑 𝓥ሶ 𝝆𝝎𝑫𝟐 𝜺 equations contain all the Pi groups so we are actually
=𝒇 , , → 8
𝝆𝝎𝟑 𝑫𝟓 𝝎𝟐 𝑫𝟐 𝝁 𝑫 discussing all those groups.
The requirement of equality of all three of the independent dimensionless parameters can be relaxed
somewhat;
▪ If the Reynolds numbers of both pump A and pump B exceed several thousand, turbulent flow
conditions exist inside the pump. It turns out that for turbulent flow, if the values of 𝑹𝒆, 𝑨 and 𝑹𝒆, 𝑩
are not equal, but not too far apart, dynamic similarity between the two pumps is still a reasonable
approximation. This fortunate condition is due to Reynolds number independence.
▪ In most cases of practical turbomachinery engineering analysis, the effect of differences in the
roughness parameter is also small, unless the roughness differences are large, as when one is
scaling from a very small pump to a very large pump (or vice versa).
Thus, for many practical problems, we may neglect the effect of both Re and 𝜀/D.
Eq. 7 and 8 then reduce to;
𝒈𝑯 𝓥ሶ 𝒃𝒉𝒑 𝓥ሶ
=𝒇 → 𝑪𝑯 ≈ 𝒇(𝑪𝑸 ) =𝒇 → 𝑪𝑷 ≈ 𝒇(𝑪𝑸 ) → 9
𝝎𝟐 𝑫𝟐 𝝎𝑫𝟑 𝝆𝝎𝟑 𝑫𝟓 𝝎𝑫𝟑
Mechanical Engineering Department 53
Pump Scaling Laws: Dynamic Pumps
As always, dimensional analysis cannot predict the shape of the
functional relationships of Eq. 9, but once these relationships are
obtained for a particular pump, they can be generalized for geometrically
similar pumps that are of different diameters, operate at different
rotational speeds and flow rates, and operate even with fluids of different
density and viscosity.
We can also write the pump efficiency equation in terms of these
dimensionless parameters;
ሶ
𝝆𝒈𝓥𝑯 𝝆 𝓥ሶ (𝒈𝑯) 𝝆 𝑪𝑸 𝝎𝑫𝟑 𝑪𝑯 𝝎𝟐 𝑫𝟐 𝑪𝑸 𝑪𝑯
𝜼𝒑𝒖𝒎𝒑 = = = → 𝜼𝒑𝒖𝒎𝒑 = ≈ 𝒇(𝑪𝑸 ) → 10
𝒃𝒉𝒑 𝒃𝒉𝒑 𝑪𝒑 𝝆𝝎𝟑 𝑫𝟓 𝑪𝒑
Since CH, CP, and 𝜂pump are approximated as functions only of CQ, we often plot these three parameters
as functions of CQ on the same plot, generating a set of nondimensional pump performance curves.
The simplified similarity laws of Eq. 9 and 10 break down when the full-scale prototype is significantly
larger than its model; the prototype’s performance is generally better. There are several reasons for this:
▪ The prototype pump often operates at high Reynolds numbers that are not achievable in the
laboratory. We know from the Moody chart that the friction factor decreases with Re, as does
boundary layer thickness. Hence, the influence of viscous boundary layers is less significant as
pump size increases, since the boundary layers occupy a less significant percentage of the flow path
through the impeller.
▪ In addition, the relative roughness (𝜀/D) on the surfaces of the prototype impeller blades may be
significantly smaller than that on the model pump blades unless the model surfaces are micro
polished.
▪ Large full-scale pumps have smaller tip clearances relative to the blade diameter; therefore, tip losses
and leakage are less significant.
Some empirical equations have been developed to account for the increase in efficiency between a
small model and a full-scale prototype. One such equation was suggested by Moody (1926) for turbines,
but it can be used as a first-order correction for pumps as well,
Moody efficiency correction equation for pumps:
𝟏/𝟓
𝑫𝒎𝒐𝒅𝒆𝒍
𝜼𝒑𝒖𝒎𝒑, 𝒑𝒓𝒐𝒕𝒐𝒕𝒚𝒑𝒆 ≅ 𝟏 − 𝟏 − 𝜼𝒑𝒖𝒎𝒑, 𝒎𝒐𝒅𝒆𝒍
𝑫𝒑𝒓𝒐𝒕𝒐𝒕𝒚𝒑𝒆
Note:
𝑫𝒎𝒐𝒅𝒆𝒍
Higher 𝜼𝒑𝒖𝒎𝒑, 𝒎𝒐𝒅𝒆𝒍 and lower 𝑫 will ultimately result
𝒑𝒓𝒐𝒕𝒐𝒕𝒚𝒑𝒆
in higher 𝜼𝒑𝒖𝒎𝒑, 𝒑𝒓𝒐𝒕𝒐𝒕𝒚𝒑𝒆
Note:
Remember than pump specific speed
is a dimensionless parameter
The pump affinity laws are quite useful as a design tool. In particular,
𝓥ሶ 𝑩 𝝎𝑩 𝒏𝑩 suppose the performance curves of an existing pump are known, and
= = 12
ሶ𝓥𝑨 𝝎𝑨 𝒏𝑨 the pump operates with reasonable efficiency and reliability. The
𝟐 𝟐 pump manufacturer decides to design a new, larger pump for other
𝑯𝑩 𝝎𝑩 𝒏𝑩
= = applications, e.g., to pump a much heavier fluid or to deliver a
𝑯𝑨 𝝎𝑨 𝒏𝑨
substantially greater net head. Rather than starting from scratch,
𝟑 𝟑
𝒃𝒉𝒑𝑩 𝝎𝑩 𝒏𝑩 engineers often simply scale up an existing design. The pump affinity
= =
𝒃𝒉𝒑𝑨 𝝎𝑨 𝒏𝑨 laws enable such scaling to be accomplished with a minimal amount
of effort.
After graduation, you work for a pump manufacturing company. One of your
company’s best-selling products is a water pump, which we shall call pump A.
Its impeller diameter is DA = 6.0 cm, and its performance data when operating
at nA = 1725 rpm are shown in Table (next slide). The marketing research
department is recommending that the company design a new product, namely,
a larger pump (which we shall call pump B) that will be used to pump liquid
refrigerant R-134a at room temperature. The pump is to be designed such that
its best efficiency point occurs as close as possible to a volume flow rate of VB
= 2400 cm3/s and at a net head of HB = 450 cm (of R-134a). The chief engineer
(your boss) tells you to perform some preliminary analyses using pump
scaling laws to determine if a geometrically scaled-up pump could be designed
and built to meet the given requirements. Calculate the required pump
diameter DB, rotational speed nB, and brake horsepower bhpB for the
new product.
Fluid Mechanics; fundamentals and applications by Y. A. Cengel & J.M. Cimbala, 4th
Edition (Chapter 14)