Sie sind auf Seite 1von 10

Fluids Handling

Preventing Pulsation
Problems in
Piping Systems
Eliminate unnecessary damage in piping
systems by following these guidelines to
avoid pulsations in pumps and compressors.

Mark Corbo
No Bull Engineering PLLC

ne of the foremost concerns facing pump and compressor users today is pulsation problems in their
piping systems and manifolds. In cases where a fluid
excitation is coincident with both an acoustic resonance and
a mechanical resonance of the piping system, large piping
vibrations, noise and failures of pipes, supports and attachments can occur. Other problems that uncontrolled pulsations may generate include cavitation in pump suction lines,
valve chatter and failures, and degradation of pump/compressor performance.
The aim of this article is to provide users with a basic
understanding of pulsations and to provide tips for prevention of field problems. Accordingly, the article foregoes the
high-level mathematics in the interest of presenting fundamental concepts in physically meaningful ways. A more indepth treatment of all subjects presented herein is provided
in Ref. 1.

Fundamentals
A pulsation is simply a fluctuation in pressure that travels
through the fluid in a pump or compressors piping system.
At any point in the flowing fluid, a distinction must be made
between the fluctuating parameters and the steady-state parameters. A fluids steady-state parameters are its flowrate
and its static pressure distribution along the length of the
pipe. When pulsation occurs, the fluctuating pressures and
flows are superimposed on the steady-state values. In this
case, the pressure at a given point can be expressed as:
P(t) = Pss + Pcyclic sin (t)
22

www.cepmagazine.org

March 2006

(1)

CEP

where: P(t) = pressure at a given point as a function of time,


lbf /in.2; Pss = steady-state pressure, lbf /in.2; Pcyclic = fluctuating pressure or pulsation, lbf /in.2; = angular frequency,
rad/s; and t = time, s. Since the main focus of this article is
on pulsations, the discussion will mainly focus on the fluctuating value.
When pulsations occur in a piping system, they are propgated through the fluid as acoustic waves (or longitudinal
mechanical waves). Acoustic waves moving through a fluid
are known as traveling waves and have the familiar wave
properties of frequency, period, wavelength and velocity. For
any given fluid at a given temperature and pressure, all disturbances travel at the same speed, known as the acoustic
velocity. Any acoustic wave actually consists of two waves
a pressure wave and a velocity (or displacement) wave,
which are always 90 deg out of phase with one another. Although a pulsation can be described by either wave, it is
usually more practical to focus on the pressure wave.
Whenever an acoustic traveling wave moving through a
pipe encounters a change in flow area, it is reflected. Some
physical elements that generate reflections include closed
ends, open ends, changes in pipe diameter (expansions and
contractions), branches, tees and flow restrictions (orifices,
valves, etc.). At such locations, two new traveling waves are
generated a reflected wave and a transmitted wave. The
reflected waves frequency, propagation speed and wavelength are always the same as those of the incident wave.
The transmitted wave proceeds in the second pipe in the
same direction as the incident wave. The transmitted wave
will always have the same frequency as the incident wave

Nomenclature

A
c
crigid
d
D
D1
D2
E
fN
fs
Fdyn
Fnet
g
H
Kbulk
L
Lf
N
P
Pcyclic
Pi
Pmean
Pss
Ptank
Pvapor
P(t)
Pr
P
P*
Q
Qin
RF
St
t
tw
u2
U
v0
v
V`

= pipe cross-sectional area, in.2


= acoustic velocity, ft/s
= acoustic velocity using rigid pipe assumption, ft/s
= pipe diameter, in.
= characteristic dimension of obstruction, ft
= larger diameter in a pipe with an expansion, in.
= smaller diameter in a pipe with an expansion, in.
= pipe elastic modulus, lbf/in.2
= Ntth natural frequency, where N is an integer, Hz
= frequency that vortices shed from one side of body, Hz
= shaking force at contraction or expansion, lbf
= net shaking force acting on elbow, lbf
= gravitational acceleration, ft/s2
= magnitude of pressure wave in terms of head, ft
= fluid bulk modulus, lbf/in.2
= effective length of pipe, ft
= effective length of fluid inertia element, in.
= all positive integers, i.e., 1, 2, 3, etc.
= pulsating pressure at bend, lbf/in2
= fluctuating pressure amplitude or pulsation amplitude, lbf/in.2
= pressure drop across fluid inertial element, lbf/in.2
= mean static pressure, lbf/in2
= steady-state pressure, lbf/in2
= tank pressure, assumed constant throughout tank, lbf/in.2
= liquid vapor pressure, lbf/in.2
= pressure at a given point as a function of time, lbf/in.2
= pressure drop across fluid resistance element, lbf/in.2
= pressure pulsation generated by impeller, lbf/in.2
= normalized pressure pulsation coefficient
= volumetric flowrate, in.3/s
= volumetric flowrate into tank, in.3/s
= acoustic resistance, lbfs/in.5
= Strouhal number, dimensionless
= time, s
= pipe wall thickness, in.
= circumferential velocity at impeller, in./s
= characteristic flow velocity, ft/s
= fluid velocity prior to valve closure, ft/s
= change in flow velocity, ft/s
= volume of the tank, in.3

Greek Symbols

= fluid density, lbfs2/in.4


L
= density of fluid subjected to water hammer, lbfs2/ft4

= bend angle, deg

= angular frequency, rad/s


Parameters defined in units of inches or feet may be assigned either unit
in an equation for dimensional consistency.

but, depending on the fluid properties in the second pipe, its


acoustic velocity and wavelength could be different from
those of the incident wave.
Since the incident and reflected waves always have the
same frequency, the pressure and velocity at any point in the
fluid can be obtained by adding the individual contributions
from the two waves. The superposition of two or more traveling waves that have the same frequency generates another
wave known as a standing wave, which assumes the same
frequency as its constituents. The standing wave is so named
because it looks as if it is standing still.
In a standing wave, all particles execute simple harmonic motion at the same frequency. However, in direct contrast to the traveling wave, the vibration (or pulsation) am-

plitudes for all particles are not the same. Instead, they
vary with the particles location. The points of primary interest in a standing wave are the antinodes and the nodes.
Antinodes are points that have the maximum amplitude,
while nodes are points that have zero amplitude. Since the
pressure and velocity waves are always 90 deg out of
phase with one another, pressure nodes are always velocity
antinodes and vice versa. Closed ends of pipes are always
pressure antinodes (and velocity nodes), while open ends
are pressure nodes.

Acoustic velocity
The technical definition of acoustic velocity is the rate
at which a pressure disturbance travels within a fluid. Typical acoustic velocities for air and water are 1,087 ft/s and
4,760 ft/s, respectively (2). The baseline equation for
acoustic velocity is:
c = (1/12)(Kbulk /)1/2

(2)

where: c = acoustic velocity, ft/s; Kbulk= fluid bulk modulus,


lbf /in.2 and = fluid density (lbf s2/in.4). The presence of
Kbulk in Eq. 2 indicates that the fluids compressibility is important. In fact, c would be infinite for a perfectly incompressible fluid.
Eq. 2 is only valid for flow through rigid pipes. Any
flexibility in the pipe walls acts as another compressibility factor in the system, which further lowers the value of
c. Although this effect is relatively unimportant when the
fluid is a gas, it can be significant for liquids. The compressibility imparted by pipe walls is normally accounted
for using the Korteweg correction, shown below:
c = crigid /( 1 + (Kbulk d / (Etw)))1/2

(3)

where: crigid = acoustic velocity calculated using Eq. 2 in


accordance with the rigid pipe assumption, ft/s; d = pipe
diameter, in.; E = pipe elastic modulus, lbf /in.2; and tw =
pipe wall thickness, in. If a liquid contains even minute
amounts of entrained gas, c is drastically reduced (3). It is
noteworthy to mention that the introduction of an amount
of air as small as 0.1% by volume can essentially cut the
value of c in half.

Organ-pipe resonances
If a traveling acoustic wave is introduced into one end of
a finite length pipe, it will be reflected back from the other
end and form a standing wave. If the wave frequency and
the pipe length are such that the reflected wave arrives at the
source at exactly the same time as a new incident wave is
being generated, the amplitude of the standing wave can
quickly grow to an extremely large value. This condition is
known as acoustic resonance, and the frequencies of the resonating waves are known as resonant frequencies.
Organ-pipe resonances are the best known ones. They
refer to the half-wavelength resonance and the quarter-wave-

CEP

March 2006

www.cepmagazine.org

23

Fluids Handling

L
+P

+P

+v

Pressure Mode
Shape

Pressure
Amplitude
vs. Time

Figure 1. Pressure- and velocity-mode shapes.

length resonance. A half-wavelength resonance occurs in a


pipe of constant diameter that has the same condition at both
ends either both ends are open or both ends are closed.
For a pipe having two closed ends (i.e., a closed-closed
pipe), any frequency that yields pressure antinodes at both
ends is a resonant frequency. There may be multiple antinodes between the ends or no antinodes at all. Stated another
way, the natural frequencies are those that yield an integral
number of half wavelength resonances over the designated
stretch of the pipe length. The natural frequencies for such a
pipe are given by the following equation:
fN = Nc /(2L)

(4)

where: fN = the Nth natural frequency, Hz; N = all positive integers; and L = the effective length of the pipe, in. In order
to illustrate the systems physical behavior, the modes
shapes are usually plotted for acoustic standing waves in the
same fashion as they are for mechanical vibrations. The only
difference is that each acoustic mode has two distinct shapes
one for pressure and one for velocity. The pressure-mode
shape provides the amplitude of pressure fluctuation at each
point, while the velocity-mode shape shows the velocity sinusoidal amplitude at each point.
Figure 1 illustrates the pressure- and velocity-mode
shapes for the first (or fundamental) mode of the closedclosed pipe. The pressure-mode shape consists of a single
half-wave. It has antinodes at both closed ends and a node at
the mid-length, while the velocity-mode shape is exactly the
opposite. The variation of pressure with time is also depicted
at several points along the pipe.
The second-, third- and other higher-order modes are
known as overtones. Higher-order frequencies that are integer multiples of the fundamental mode are called harmonics.
For an open-open pipe, the situation (when resonances
occur) is very similar except the boundary conditions that
must be satisfied are displacement antinodes at both ends. The
natural frequencies for this case are again given by Eq. 4. Although the natural frequencies are identical to those for a
closed-closed pipe, the mode shapes are different.
Another major organ-pipe resonator is a constant-diameter
pipe featuring one open end and one closed end, or an openclosed pipe. The boundary conditions that must be satisfied
for this pipe are a pressure antinode at the closed end and a

24

www.cepmagazine.org

Velocity Mode
Shape

March 2006

CEP

pressure node at the open end. The natural frequencies for


such a pipe are given by:
fN = c (2N 1)/(4L)

(5)

The fundamental mode of an open-closed pipe consists of


a single quarter-wave. For this reason, the open-closed pipe
is often referred to as a quarter-wave stub. However, in direct contrast to the open-open and closed-closed pipes, not
all of the harmonics are present. Specifically, all of the even
harmonics are missing. This is because the even harmonics
of the fundamental mode all consist of integral multiples of
half-wavelengths that have already been shown to yield the
same condition at both ends (i.e., either both are nodes or
both are antinodes). Since the open-closed pipe requires opposite boundary conditions at the two ends, only the odd
harmonics are present.
An important point to remember is that the above simplified relations are only applicable to segments of pipe where
the diameter is constant. If a segment contains a change in
diameter, even a small one, reflections can occur, rendering
the previous equations invalid.

Acoustic resonance
It should be emphasized that the existence of quarterwave and half-wave resonant modes in a piping system
does not necessarily mean that the system will encounter
pulsation problems. In fact, all piping systems will have
such modes, along with other resonant modes that are more
complex. In order for a potentially problematic mode to
exist, there must be a means for exciting it. In most cases,
the excitation is in the form of a flow or pressure variation
generated by a pump or compressor. In general, resonance
will occur when the total time it takes for the excitation to
travel up the pipe, be reflected, and return to the source, is
such that the reflected wave is perfectly in phase with a subsequent perturbation being generated by the source. In other
words, the frequency of the excitations must match one of
the resonant frequencies of the system. When this happens,
standing waves are formed in the piping system, and the
pressure and displacement amplitudes can build up to very
large, potentially destructive values. Since the large pressure
amplitudes are the element of most concern, the pressure
antinodes are the areas of concern. It should be remembered

Constant Area Pipe:


D = 1 in., L = 240 in.

Open End

Closed End

Variable Area Pipe:


D1 = 2 in., L1 = 80 in.
D2 = 1 in., L 2 = 160 in.

Open End
Closed End

Figure 2. Constant- and variable-area pipes.


that, even at resonance, the fluctuating pressures at a pressure node are zero.
If the system contained absolutely no damping, the pressure fluctuations at the antinodes would theoretically be infinite at resonance. However, all real piping systems contain
acoustic damping due to several phenomena, with the most
prominent being parasitic losses within the piping due to
pipe friction, restrictions, bends, etc. Since these losses vary
with the square of the flowrate, the system damping is much
larger for higher steady-state flows than for low flowrates.
For this reason, acoustic resonances are more prominent at
low flows, where the system is very lightly damped (4).
Even when the flows are not low, most pump and compressor piping systems are relatively lightly damped from
an acoustic standpoint. A measure of the amount of damping is the amplification factor at resonance, or Q factor,
which is analogous to that associated with mechanical vibrations. The larger the amplification factor, the sharper the
peak (corresponding to lighter damping) and the more dangerous the mode is. Typical piping systems have been found
to have amplification factors around 1530, with 80 being
the highest ever observed by the author. In general, the amplification factor is dependent on flow, line size and resonance frequency.

Acoustic behavior of real piping


The two organ-pipe resonances most likely to be excited
in turbomachinery are the open-open and open-closed types.
Since in most installations, the pump or compressor discharge line ends in a manifold connection or connections of
larger cross-section, the discharge line almost always has an
open end at its terminal end. Likewise, since the pump or
compressor suction line usually begins at a reservoir or
tank, the suction pipe almost always begins with an open

end. Additionally, in a reciprocating pump or a compressor,


the pump/compressor ends of the discharge and suction
lines normally behave as closed ends, since a pressure wave
traveling toward the pump or compressor will either hit a
closed valve or a piston (a piston behaves as a moving
closed end).
Other configurations that are not perfectly open or
closed can be treated as such if they obey the 2:1 rule.
That is, if the diameter increase or decrease at a valve or enlargement/contraction is 2:1 or more, the configuration can
respectively be treated as either an open-end or closed-end
pipe. Then, one can use Eqs. 4 or 5 to obtain the natural
frequency of the configuration.
Every piping system arranged in series has two different
acoustic natural frequencies (5). The first is the theoretical
frequency one obtains by using the organ-pipe relationships
and making reasonable approximations in areas where there
is a change in diameter. The second is the actual frequency
observed in the field. Unfortunately, due to partial reflections
occurring at each area change, except in the simple case of a
constant diameter pipe, the actual resonant frequency is usually quite different from the theoretical frequency.
These differences are significant even in relatively simple configurations. The author demonstrated this by analyzing the two configurations shown in Figure 2 (assumed to
be flowing water) using an acoustic simulation code developed by the authors company. The top configuration, referred to as the constant area pipe, is a classic quarter-wave
stub whose fundamental frequency was found to be 58.7
Hz, which is very close to that obtained using Eq. 5.
Unfortunately, reasonable approximations are not very
accurate for the variable area pipe (Figure 2, bottom). One
methodology might be to analyze the 1-in.-dia. pipe as an
open-open pipe using Eq. 4, which yields a fundamental
frequency of 176 Hz. Another reasonable approach might
be to treat the 2-in. pipe as an open-closed pipe using Eq. 5,
which also yields a fundamental frequency of 176 Hz. A
third approach might be to ignore the change in diameter
and simply look at the entire pipe as an open-closed pipe,
which yields the same value obtained for the constant area
pipe, or 58.7 Hz. However, the actual fundamental frequency determined by the code is 36.8 Hz, which is far from all
of the theoretical values.
In more complex systems, such as those containing
branches or parallel paths, the theoretical frequency does
not have any meaning, since there is no simple manner of
calculating it that can be reasonably justified. However, as
in the preceding example, the actual resonant frequency can
be calculated with the aid of a good acoustic-analysis computer code. As is the case in simple pipes, higher-order natural frequencies are also present in complex systems, but they
are seldom harmonics of the fundamental.
When a piping system is subjected to an excitation
whose frequency is not in resonance with any of the systems acoustic natural frequencies, the excitation wave sim-

CEP

March 2006

www.cepmagazine.org

25

Fluids Handling

ply passes through the system as a traveling wave without


the formation of any appreciable standing waves. In this
condition, the pulsation amplitudes everywhere in the system are essentially the same as those of the source.

Lumped-acoustic systems
In general, acoustic systems, like mechanical systems,
can be broken down into two basic types distributed
systems and lumped systems. A distributed acoustic system
is one in which all of the properties associated with flow,
velocity, and pressure are continuously distributed along
the length of the system. A simple example of a distributed
system is a cantilever beam. On the other hand, a lumpedacoustic system is one in which the acoustic properties can
be lumped together at various points in the system. A
prominent example is the mass-spring-damper system.
There are three basic elements in lumped-acoustic systems. The first is fluid inertia, which is sometimes referred to as fluid inductance. This type of element is represented by a narrow or constricted passage that is short
enough to allow its contained fluid to behave as a rigid
body, the inertia of which must be overcome by the
acoustic pressure difference across it. The governing
equation for this type of element is:
Pi = (Lf /A)dQ/dt

(6)

where: Pi = pressure difference needed to overcome inertial effects of the fluid flowing through the pipe, lbf /in.2; Lf
= effective length of the fluids inertial element, in.; A =
cross sectional area of the pipe, in.2; and dQ/dt = derivative of volumetric flowrate with respect to time (in.3/s)/s.
The second basic element is the fluid compliance (i.e.,
fluid capacitance) element. This type of element is represented by a chamber or tank and represents the ability of a
fluid to store some of the volume flow entering it by
converting it to pressure energy by virtue of its compressibility. For long wavelengths, the acoustic pressure is consistent throughout the entire volume of the tank. The governing equation for this element is:
Qin = (V /Kbulk)dPtank /dt

(7)

where: Qin = the volumetric flow into the tank, in.3/s; V is


the volume of the tank and dPtank /dt = derivative of tank
pressure with respect to time, (lbf /in.2)/s.
The third basic element is the fluid resistance element.
This type of element is represented by an orifice, valve, restriction, etc., having a fairly small flow area, such that a
relatively large pressure drop is required to generate fluid
flow through the element. The governing equation for this
element is:
Rf = Pr / Q

26

www.cepmagazine.org

(8)

March 2006

CEP

where: Rf = acoustic resistance, lbf s/in.5; and Pr = pressure drop across the element, lbf /in.2.

Reciprocating pump/compressor excitations


In order for resonant pulsation problems to occur, an
excitation source is needed to excite the acoustic system at
one of its resonant frequencies. In general, reciprocating
pumps and compressors are probably the most notorious
sources of these types of excitations. Although these are
not the only positive-displacement devices that can cause
pulsation problems (gear and vane pumps also can), they
are the only ones discussed in detail. In general, the pulsations generated by positive-displacement pumps and compressors are low frequency and high amplitude, while those
generated by centrifugal pumps/compressors are high frequency and low amplitude.
Most reciprocating pumps and compressors are simple
slider-crank mechanisms that consist of a rotating
crankshaft, connecting rods and pistons. They are essentially like an automobile engine running with the power
transmission in the opposite direction. Specifically, the
crankshaft, which is normally driven by an electric motor,
transmits torque to the connecting rods, which convert the
crankshafts rotary motion into the reciprocating motion of
the piston assemblies.
It should be noted that because the suction and discharge valves for any given cylinder in a reciprocating
pump or compressor are never simultaneously open, the
suction and discharge systems in a reciprocating application are completely isolated from each other from an
acoustics standpoint. They represent two distinct acoustic
systems that must be analyzed separately.
Reciprocating pumps and compressors generate two
distinct types of excitations that can generate pulsations in
the suction and discharge piping flow excitations and
acceleration excitations. Flow excitations are almost always predominant in the discharge piping. However, in
some pump suction systems, the acceleration excitations
are the more dangerous of the two.
For example, flow excitations arise because the action
of the pistons, along with the opening and closing of the
valves, generates a pulsatile flow in both the suction and
discharge systems. Since the same effects are occurring at
all cylinders, these excitations normally vary inversely
with the number of plungers, since the plungers can be
phased to smooth out these effects. If the plunger connecting-rods were infinitely long, it can be shown that the flow
pulsations are perfectly sinusoidal and that each cylinder
would only generate a 1X excitation. However, the finite
length of the connecting rod also gives rise to a 2X excitation at each cylinder. Thus, the primary flow excitations
occur at the plunger frequency and at twice the plunger
frequency, which would be 3X and 6X for a triplex pump.
The flow pulsations convert to pressure pulsations
based on the nature of the discharge and suction piping.

Since the discharge pressure is normally set by resistance


losses in the discharge piping (e.g., pipe friction, bends,
expansions, etc.) and since these losses are normally a
function of the square of the flowrate, the squared law
causes the pressure pulsations to be larger (percentagewise) than the flow pulsations. For example, consider a
hypothetical system that has a discharge pressure of 500
psi that is totally generated by pipe friction loss at nominal
flowrate. A flow pulsation of 20% would translate into a
pressure variation of 44% (1.202 = 1.44), or 220 lbf /in2,
which is substantial. That is why flow pulsations are normally more dangerous in discharge piping.
The other type of excitation generated by reciprocating
pumps is the acceleration excitation. Although these excitations are also theoretically present in reciprocating compressors, they are normally inconsequential. Acceleration excitations can be understood by visualizing a fluid completely at
rest in the suction piping during the discharge stroke. Once the
suction stroke begins and the suction valve opens, the pump
accelerates the fluid in the suction line from rest to maximum
piston velocity. At the end of the suction stroke, the suction
valve closes and the fluid decelerates to rest. The same phenomenon occurs with the fluid in the discharge line.
All of these sudden accelerations generate pressure pulses
in both the suction and discharge lines. Since typical reciprocating pump acceleration excitations are around 25 lbf /in.2
(6), which is significantly smaller than the typical flow excitations (220 lbf /in.2 for discharge piping in the example
given above), acceleration excitations usually have little impact on discharge piping. However, since pump-suction systems normally employ short pipes of large diameters to meet
net-positive-suction head (NPSH) requirements, the frictional pressure drops, and flow-based excitations in suction lines
are usually quite low. (The NPSH is simply the difference
between the absolute static pressure and the liquids true
vapor pressure at the inlet temperature.) Thus, acceleration
excitations are often the ones of concern in pump-suction
piping. Unfortunately, unlike flow excitations, increasing the
number of cylinders in a reciprocating pump has limited
benefit regarding acceleration excitations.

Centrifugal pump and compressor excitations


Although the reciprocating pump and compressor are
the most notorious sources of pulsation problems in the
chemical process industries, centrifugal pumping and compression systems are also susceptible. Centrifugal pump
and compressor systems suffer from pulsation problems
due to three very different phenomena. First, the
pump/compressor can generate high-frequency pulsations
at vane-passing frequency that can excite acoustic resonances in the manner just described for reciprocating machines. Second, vortex shedding that occurs at discontinuities in the piping system can also excite acoustic resonances. Third, complex interactions between the dynamic
characteristics of the pump/compressor and those of the
piping system can generate instabilities (7).

Centrifugal pump/compressor impeller excitations occur at


vane-passing frequency and its harmonics (particularly the
second). These vane-passing excitations are generated by the
action of the impeller vane passing stationary objects that are
in close quarters to it. In direct contrast to the pulsations generated by positive-displacement machines, centrifugal impeller pulsations normally occur at high frequencies (typically
> 500 Hz) and have relatively small amplitudes. According to
Ref. 8, vane-passing frequency pulsations are approximately
0.33% of the pressure rise generated by the impeller. The following equation yields the normalized pressure pulsations
(P*) generated by the wake in an impeller (9):
P* = P / (u22 / 2)

(9)

where: P* = normalized pressure pulsation coefficient; P


= amplitude of the pressure pulsation generated by the impeller, lbf /in2; and u2 = circumferential velocity at the impeller exit, in./s. Data taken on centrifugal pumps revealed
that virtually all pressure pulsations possess normalized coefficients below 0.015 at the best efficiency point (BEP) flow, and
below 0.020 at 25% of BEP flow (9).
Although the amplitudes of vane-passing pulsations are
very difficult to predict analytically, factors that
are known to have an impact on their magnitudes include:
Radial clearance between the impeller tip and the stationary collector (e.g., diffuser vanes or volute cutwaters). In
general, increasing this clearance is the surest way to reduce
pulsation amplitudes. Pulsations have been empirically found
to vary inversely with this clearance to the 0.77 power (9).
Flow. Pulsations are normally minimized at BEP.
Number of vanes. Increasing the number of vanes normally reduces pulsations.
Type of impeller. Pulsations tend to increase with the increasing specific speed of the impeller. Thus, the pure radial
impeller has the lowest pulsations, followed by the Francis impeller and axial impeller.
Staggering of impellers in multi-stage machines. Staggering acts to reduce pulsations.
Impeller tip speed. Higher tip speeds generally mean
larger pulsations.
Unlike the pulsations generated by a reciprocating machine which are sometimes large enough to cause problems even in the absence of resonance centrifugal-impeller excitations are so small that they can only do damage if they are amplified via acoustic resonance. Additionally, the impeller must be located at or near an antinode in
the velocity-mode shape in order for amplification to
occur. Finally, unlike reciprocating machines, a centrifugal
impeller does not isolate the discharge and suction systems
from each other so all of the piping must be analyzed as a
single system. When modeling these systems, the impellers acoustic characteristics must be accounted for.
This is usually done through a specialized transfer matrix
that captures how the impeller behaves when exposed to
flow and pressure perturbations.
CEP

March 2006

www.cepmagazine.org

27

Fluids Handling

Table 2 of Ref. 9 provides a list of design guidelines for


keeping impeller pressure excitations in check. The most
basic rule of thumb is that the ratio of the collector diameter
to the impeller diameter should exceed 1.04 and that the impeller and collector should never have the same numbers of
vanes. Table 3 of the same reference provides design features that can be incorporated into centrifugal pumps to further reduce pulsation amplitudes.

Vortex shedding
The next pulsation excitation mechanism often present in
centrifugal machines is vortex shedding. In contrast to the
excitations previously described, vortex-shedding excitations
occur within the piping system, not at the pump or compressor. Vortex shedding occurs when flow passing over an obstruction, side branch, or piping discontinuity generates fluid
vortices at a regular interval. Under certain conditions, these
vortices can excite acoustic resonances and generate pulsation problems.
When a fluid passes over any body having a bluff
(broad) trailing edge, such as a cylinder, vortices are shed
from both sides of the body. The vortex shedding is not simultaneous; instead, vortices are alternately shed from one
side and then the other. This alternating shedding causes
pressure fluctuations to occur on both sides of the body.
The frequency at which vortices shed from each side of a
body has been empirically determined to be a function of a
dimensionless parameter known as the Strouhal number
(St), which is defined as follows:
fS = St U / D

(10)

For internal flows, such as those in pipes, the characteristic


flow velocity, U, is simply the mean velocity in ft/s. For
external flows, such as those over a cylinder, the characteristic flow velocity is the free stream velocity. For all
practical purposes, the characteristic dimension, D, is assumed to be the maximum width of the structure perpendicular to the flow. For example, if the body is a cylinder,
D is simply the diameter. For the case of flow past a tee or
branch, D is the inside diameter of the branch. For abrupt
expansions, the characteristic dimension is simply the difference between the two pipe diameters.
It is customary to refer to the frequency obtained from
Eq. 10, that associated with only one side of the body, as
the vortex shedding frequency. Accordingly, it is easily
seen that the fluctuating pressure force generated by the
vortex-shedding acting in the direction perpendicular to
the flow (i.e., lift direction) acts at the vortex shedding frequency. On the other hand, the fluctuating force acting in
the direction of flow (i.e., drag direction) acts at twice the
shedding frequency.
Plots of St vs.Reynolds number (Re) obtained from
tests for flow around circular cylinders are provided in
Ref. 10. Most of the measured St values are in the vicinity

28

www.cepmagazine.org

March 2006

CEP

of 0.20. Additionally, Figures 36 of Ref. 11 give representative values of St as functions of Re for various geometries. A formula for St for flow past a sidebranch that
is valid for Re > 1.6 107 is available in Ref. 12.
In centrifugal pumping and compression systems, vortex
shedding primarily occurs at valves and other restrictions
and at flow past tees and side branches. The types of valves
where significant vortex shedding can occur include relief
valves, throttling valves and pressure regulators. Since the
pulsation amplitudes of the vortices are always very small,
they can only create problems if they excite an acoustic resonance. Additionally, vortices can only excite a resonance if
they are located at or near a velocity antinode. One of the
reasons that vortex-related problems frequently occur in side
branches is that most side branches act as quarter-wave stubs
that have a velocity antinode (open end) right at the location
of vortex generation. In order to avoid side branch problems,
the side branch should be designed to minimize its length-todiameter ratio and to eliminate all sharp edges, which generate vortices (13).
A case study in Ref. 8 offers a good method for troubleshooting pulsation problems if vortex shedding is suspected to be the excitation source. Using the observed pulsation frequency, the known flow velocity, and the expected Strouhal number, the authors back-calculated the characteristic dimension, D, from Eq. 10. They then searched
their system for a flow obstruction or gap of about that
size and pinpointed the stiffening rings on the strainer as
the cause of their problem.
One of the best ways to minimize or eliminate vortex
shedding is via streamlining of the downstream side of the
flow obstruction. In order for this path to be effective, the
included angle between the two downstream surfaces cannot be more than eight to ten degrees (11). Vortex shedding can also be reduced by adding a vortex suppression
device to the obstruction. Figure 3-23 of Ref. 11 shows
several common configurations.
If prevention of vortex shedding is not practical, the
vortex shedding frequencies can be shifted away from
acoustical and/or mechanical natural frequencies by making the effective width of the obstruction larger or smaller.
Alternatively, the flow path can be modified to change the
velocities across the obstruction.

Water hammer
All of the pulsation excitations discussed so far can be
classified as steady-state excitations in that they are capable of generating pulsations that can last an indefinite
amount of time. However, transient pulsations, which can
be grouped together under the term water hammer, can
also cause significant problems.
Water hammer is normally associated with the rapid
changes in internal fluid pressure that occur in a pipe when
the flow is suddenly interrupted via the closing of a valve
and the hammering sound that may accompany those

changes. A common, everyday example is the noise that


sometimes occurs in the pipes of an old building when a
water faucet is shut off. However, any transient pulsations
that occur as the result of a rapid change in flow conditions can be referred to as water hammer. These phenomena are sometimes also referred to as surge or fluid shock.
When a valve in a flowing pipe is suddenly closed, a
large pressure spike occurs at the valve. If the pressure
spike is not sufficient to rupture the pipe or some other
component, the compression wave created will reverse and
travel back down the pipe toward the pump. When the
wave hits the check valve or pump, it will again reverse
and flow back toward the valve that triggered the entire
scenario. It will continue to reverberate until something finally breaks or the energy completely dissipates due to
system damping. Even if nothing fails initially, this scenario subjects the system to repeated stresses, which could
ultimately lead to fatigue failures.
It should be noted that the compressibility of the liquid
and elasticity of the pipe serve to cushion the shock generated in the above scenario. In fact, if the pipe were rigid
and the liquid were truly incompressible, the pressure spike
would be infinite in the case of instantaneous valve closure.
The amplitude of the pressure spike H generated by the
sudden valve closure is calculated as:
H = (c /g) v

(11)

where: H is the change in head, ft; v = change in fluid


velocity, ft/s; c = acoustic velocity, ft/s; and g = gravitational acceleration, ft/s2. For the simple case where a fluid
is initially flowing at a velocity v0 and the valve is suddenly closed, the amplitude of the pressure wave that travels
upstream is Lcv0.
The induced pressure pulsation is strongly dependent on
the acoustic velocity. However, the impact of the acoustic velocity goes even further than that, since it determines the
length of time that corresponds to an instantaneous change
in valve position. This is because as long as the valve closes
in a shorter time than the time it takes for the pressure wave
to travel down the pipe to the other end (having a length of
L), reflect, and then travel back up the pipe to the valve, the
impact on the system is exactly the same as if the valve had
closed instantaneously. This critical time interval is 2L/c. As
long as the valve closes completely in less than this time, the
impact on the system will be the same, regardless of how
quickly the valve actually closes.
Although all systems are different, a valve that closes in
less than 1.5 s will normally behave as if the closure is instantaneous (14). In cases where the valve closure cannot be
considered instantaneous, the head rise can be determined
from a modified form of Eq. 11 (15). To calculate the equivalent initial velocity that generates the head rise for the case
where the pipe is made up of a series of pipes of different diameters, see Ref. 16.

In a typical pumping system, the events that may lead to


water hammer include rapid opening or closing of a valve,
abrupt startup of a pump against a pipeline full of static fluid;
and abrupt shutdown of a pump. Water hammer can have potentially harmful effects on the system. If the pressure generated by the transient is high enough, pipes could rupture or
valves and other components could fail. In general, since suction pipes tend to have short lengths and low flow velocities,
water hammer seldom originates in the suction piping. Additionally, since elbows and bends provide damping, complex
circuits dont normally suffer from water hammer, either.
There are several methods for reducing the effects of
water hammer. The valve or other quick-acting component
can be modified to smooth out its flow and/or lengthen its
opening or closing time. Several practical means for doing
this are provided in Ref. 17. Also, since the severity of
water hammer is highly dependent on the acoustic velocity,
a reduction in acoustic velocity will always be beneficial.
This can be accomplished by adding a small amount of air
to a liquid system or by switching to non-circular piping.
Further, since the pressure spike is proportional to the initial
flow velocity, doubling the pipe diameter can reduce the
spike by a factor of four.

Piping vibration
Thus far, the article has focused on the main sources of
pulsation problems in piping systems. Equally important are
the adverse effects that pulsation can render. Although there
are many of these, the biggest problem occurring in discharge piping is vibration, and the biggest headache associated with pump-suction systems is cavitation. Accordingly,
these two areas are the subjects of this and the next section.
On their own, pressure pulsations are not capable of producing piping vibration. For instance, if an infinitely long
straight pipe were subjected to traveling pressure waves of
extremely large amplitudes, no vibration would occur. In
order for vibration to occur, there must be some way for the
pressure pulsations to translate into an oscillatory force that
is applied to the piping. For instance, if the straight pipe
were terminated in an elbow, the pulsations acting on the
elbows cross-section would generate the needed oscillating
force and the pipe would probably vibrate excessively. The
elbow is, thereby, referred to as a point of acoustical-mechanical coupling. In general, such points occur when the
pulsating pressures act on unbalanced areas. Components in
typical piping systems where this type of coupling occurs
include elbows, sudden expansions and contractions, capped
ends of bottles and manifolds, orifices, and valves.
In typical piping systems, the most common coupling
point is at elbows or bends. If the pulsating pressure at the
bend is P, the amplitude of the resultant dynamic force is:
Fnet = 2PA cos(/2)

(12)

where: A is the cross sectional area of the pipe. Looking

CEP

March 2006

www.cepmagazine.org

29

at Eq. 12, as the bend angle, , is made greater than 90


deg, the magnitude of the shaking force is reduced. Thus,
the worst case is also the most prevalent the 90-deg,
right-angle elbow. Accordingly, if the piping must make a
90-deg change in direction, it is better to employ two 45deg elbows than a single 90-deg bend.
Another source of dynamic coupling in a piping system is
a change in piping diameter, either a contraction or expansion. It is easily shown that for either case, the shaking or
dynamic force is calculated as:
Fdyn = P( / 4)(D12 D22)

(13)

where: D1 and D2 are the respective larger and smaller diameters of a pipe. Another source of coupling is any enclosed
vessel such as a surge volume, bottle, manifold, etc. In such
devices, the pulsation forces present at each end must be assessed, along with those acting on any vessel internals present such as baffles and supports, in order to determine the
total unbalanced force. There are some acoustic modes that
generate large shaking forces on a vessel and others that
generate no force, regardless of the amplitude of the pulsations (1). It all depends on the phasing of the forces with respect to one another.
A piping system behaves in the manner of any elastic mechanical system it has natural frequencies that are characteristics of the system. If the piping system is excited by an
oscillatory force having a frequency equal to or close to one
of its natural frequencies, mechanical resonance occurs and
the vibration amplitudes, forces, and stresses can become excessive. In typical piping systems, mechanical resonance can
be associated with amplification factors as high as 20 (18).
The worst case scenario is when a piping system has an
acoustic resonance that is coincident with or close to a piping mechanical resonance. In this case, the acoustic amplification factor, which can be as high as 15, combines with the
mechanical amplification factor (20, from above) to generate
an overall amplification factor that can be as high as 300
(18). It is not difficult to visualize how such a situation could
lead to problems.
The most important task in piping system design is, therefore, to ensure that acoustic resonances are not capable of
exciting any piping mechanical resonances. In order to ensure that a piping system will operate troublefree in the field,
a thorough acoustic analysis should be combined with a
good piping-vibration analysis (for details, see Ref. 19). Additionally, the pulsation-induced shaking forces can also be
calculated and compared to the guidelines given in Ref. 20.
General rules that can be employed during design to minimize the chances of piping vibration problems include:
the number of elbows should be minimized
in the areas where elbows must be employed, 45-deg
bends are preferential to 90-deg versions
all elbows and concentrated masses, such as valves,
should be rigidly supported with clamps

30

www.cepmagazine.org

March 2006

CEP

Pressure lbf / in.2

Fluids Handling

Pcyclic
Pmean
Pvapor
P=0
Time

Figure 3. Pulsation-generated cavitation.


avoid using threaded connections for instrumentation
(they are susceptible to vibration-related failures at thread roots)
use pipe braces, not hangers, to support the piping
attach pipe supports to rigid anchor structures
the stiffness of piping clamps and supports should be at
least twice the stiffness of the basic pipe span.

Cavitation
Although piping vibration is often a troublesome problem
in pump discharge piping, in suction piping, the pressure levels are normally too low for significant vibration amplitudes to
arise. Instead, the most likely adverse consequence of large
pulsation levels in pump suction systems is cavitation.
Cavitation occurs when the local static pressure in a liquid falls below its vapor pressure. In the initial stages of cavitation, microscopic gas nuclei within the liquid give rise to
the formation of gas bubbles in regions of low static pressure. The pumps flow then carries many of these bubbles to
regions where the static pressure is above the vapor pressure.
This causes the bubbles to collapse and produce high-intensity pressure waves. The localized pressures and temperatures can be as high as 60,000 lbf /in.2 and 1,500F (21).
Cavitation can occur at the pump inlet, inside the cylinders
of a reciprocating pump, in the suction manifold, and in other
places where changes in fluid velocity occur. The repeated
pulses generated during bubble collapse can generate pulsations, noise, and vibration of pump components or piping.
They also can do significant erosion damage to the pump internals. Cavitation also can degrade pump performance,
sometimes drastically, if enough bubbles are present.
Figure 3 shows how pressure pulsations can cause cavitation. It is seen from the figure that the mean static pressure
for this hypothetical pump, Pmean, is above the liquids vapor
pressure, Pvapor. Accordingly, if there were no pulsations in
this suction system, there would be no cavitation. However,
the presence of a pulsating pressure, Pcyclic, means that the
minimum static pressure in the system is now Pmean P cyclic.
If, as is shown in the figure, this minimum pressure is
below the vapor pressure, the liquid will theoretically cavitate. Thus, higher pulsation amplitudes increase the likelihood of cavitation.
If pressure pulsations are successful at generating signifi-

Literature Cited
1.

Corbo, M. A., and C. F., Stearns, Practical Design Against


Pump Pulsations, Proceedings of the Twenty-Second International Pump Users Symposium, Turbomachinery Laboratory, Texas
A&M Univ., College Station, TX., pp. 137177 (2005).
2. Resnick, R., and D. Halliday, Physics Part One, 3rd ed.,
John Wiley and Sons, Hoboken, NJ (1977).
3. Singh, P. J., and N. K. Madavan, Complete Analysis and Simulation of Reciprocating Pumps Including System Piping, Proceedings of the Fourth International Pump Symposium, Turbomachinery Symposium, Texas A&M Univ., College Station, TX, pp.
5573 (1987).
4. Sparks, C. R., On the Transient Interaction of Centrifugal Compressors and Their Piping Systems, ASME Journal of Engineering for Power, 105 (4) pp. 891901 (1983).
5. Wylie, E. B., and V. L. Streeter, Fluid Transients in Systems,
Prentice Hall, Upper Saddle River, NJ (1993).
6. Miller, J. E., Characteristics of the Reciprocating Pump, Proceedings of the Fifth International Pump Users Symposium, Turbomachinery Laboratory, Texas A&M Univ., College Station, TX,
pp. 163173 (1988).
7. Greitzer, E. M., The Stability of Pumping Systems - The 1980
Freeman Scholar Lecture, ASME Journal of Fluids Engineering,
103 (2), pp. 193242 (1983).
8. Lewis, A. L., et al., Control Valve Induced Pipeline Vibrations in
a Paper Pulp Pumping System, Proceedings of the Fourteenth International Pump Users Symposium, Turbomachinery Laboratory, Texas A&M Univ., College Station, TX, pp. 4959 (1997).
9. Guelich, J. F., and U. Bolleter, Pressure Pulsations in Centrifugal Pumps, ASME Journal of Vibration and Acoustics, 14 (2).
pp. 272279 (1992).
10. Au-Yang, M. K., Flow-Induced Vibration of Power and Process
Plant Components: A Practical Workbook, ASME, New York (2001).
11. Blevins, R. D., Flow-Induced Vibration, Krieger Publishing
Co., Malabar, FL (2001).

cant cavitation, it can usually be easily seen in a pressure vs.


time plot obtained from the field. Whereas normal pulsation
generates a sinusoidal wave, the presence of cavitation will
cause a squaring off of the bottom portions of each of the
waves. After the initial signs of cavitation, the pressure signature also often shows a series of sharp pressure spikes, indicative of cavitation bubble collapse.

MARK A. CORBO is the founder and chief engineer of No Bull Engineering


(46393 Vianne Court, Temecula, CA 92592; Phone/Fax: (951) 694-3807;
E-mail: markatnobull@earthlink.net), a technology-focused
engineering and consulting firm. His primary responsibilities include
rotating equipment engineering design and analysis, troubleshooting
and third-party design audits for clients in the turbomachinery and
aerospace industries. Prior to founding No Bull eight years ago, Corbo
was a consulting engineer at Mechanical Technology, Inc. (MTI) and
spent 12 years in the aerospace industry designing pumps, valves, and
controls for gas turbine engines. His fields of expertise include
rotordynamics, fluid-film journal and thrust bearings, acoustic
simulations, Simulink dynamic simulations, hydraulic and pneumatic
flow analysis, and mechanical design. He has authored more than a
dozen technical publications, including one that won the Best Case
Study award at Bently Nevadas Iscorma 2001 rotordynamics
conference. Corbo holds BS and MS degrees in mechanical engineering
from Rensselaer Polytechnic Institute, is a registered professional
engineer in New York, and is a member of ASME, STLE and The
Vibration Institute.

12. Rogers, L. E., Design Stage Acoustic Analysis of Natural Gas


Piping Systems in Centrifugal Compressor Stations, ASME J. of
Eng. for Gas Turbines and Power, 114 (4). pp. 727736 (1992).
13. Baldwin, R. M., and H. R. Simmons, Flow-Induced Vibration
in Safety Relief Valves, ASME Journal of Pressure Vessel Technology,108 (3), pp. 267272 (1986).
14. Cornell, G., Controlling Surge and Pulsation Problems, Pumps
and Systems, 6(9), pp. 1218 (1998).
15. Graf, E., and J. Marchi, The Application of Computer Simulation and Real Time Monitoring to Minimize the Pressure Pulsations in Complex Pumping Systems, Proceedings of the Fourteenth International Pump Users Symposium, Turbomachinery
Laboratory, Texas A&M Univ. College Station, TX,
pp. 6168 (1997).
16. Stepanoff, A. J., Elements of Graphical Solution of Water-Hammer Problems in Centrifugal-Pump Systems, ASME Transactions, 71, pp. 515534 (1949).
17. Dodge, L., Reduce Fluid Hammer, Product Engineering,
pp. 276283 (1960).
18. Parry, W. W., System Problem Experience in Multiple Reciprocating Pump Installations, Proceedings of the Third International
Pump Symposium, Turbomachinery Laboratory, Texas A&M
Univ., College Station, TX, pp. 2125 (1986).
19. Wachel, J. C., et al.,Piping Vibration Analysis, Proceedings of
the Nineteenth Turbomachinery Symposium, Turbomachinery
Laboratory, Texas A&M Univ., College Station, TX,
pp. 119134 (1990).
20. Tison, J. D., and K. E. Atkins,, The New Fifth Edition of API
618 for Reciprocating Compressors Which Pulsation and Vibration
Control Philosophy Should You Use? Proceedings of the Thirtieth
Turbomachinery Symposium, Turbomachinery Laboratory, Texas
A&M Univ., College Station, TX, pp. 183195 (2001).
21. Beynart, V. L., Pulsation Control for Reciprocating Pumps,
Pumps and Systems, 7 (11), pp. 2436 (1999).

The manner in which pump designers traditionally attempt to avoid cavitation is via specification of NPSH at the
pump inlet flange. The standard Hydraulic Institute equation
for determining the required NPSH includes an acceleration
head term that is supposed to account for dynamic effects.
However, inspection of the equations used for its calculation
reveals that its basically an attempt to account for a dynamic phenomenon with a static calculation. Not surprisingly,
there are many situations, especially if acoustic resonances
occur in the suction system, where this method underpredicts
the pulsation amplitudes, often by a substantial amount. In
such cases, the calculated NPSH could very well meet the
required NPSH, but the pump may still suffer from severe
cavitation. Naturally, performance of a good acoustic analysis should rectify this problem.

Conclusion
Although guidelines for preventing pulsation problems
have been provided, there are situations where these guidelines alone are not enough, and pulsation-control devices
(surge volumes, accumulators, acoustic filters, etc.) are necessary. A follow-up article slated for publication this spring
in CEP will discuss these devices in detail.
CEP

CEP

March 2006

www.cepmagazine.org

31

Das könnte Ihnen auch gefallen