Sie sind auf Seite 1von 155

Article #1: How Does Pump Suction Limit the Flow?

One of the claimed advantages of the centrifugal pumps over positive displacement pumps is their ability to operate over a wide range of flow. Since a centrifugal pump operates at the intersection of a pump curve and a system curve, by varying the system curve the operating point of the pump is easily changed:

Figure 1-1 Flow control of the centrifugal pump by the discharge valve The convenience and simplicity of such flow control by the discharge valve throttling comes at a price, because a pump is thus forced to run either to the left, or to the right, of it's best efficiency point (BEP). However, the real danger of operating the pump too far off-peak comes from the suction side considerations. Too far to the right - and you are easily risking to run out of the available NPSHA, causing cavitation problems. Too far to the left - flow recirculation at the impeller eye will let itself known through the noise, vibration, and damage. Thus, the flow must be limited on both sides of the BEP:

Figure 1-2 Pump operating range has limits

Consider the first limitation - high flow. Centrifugal pump stops pumping when liquid turns to vapor. This happens when the pressure somewhere inside the pump drops below liquid vapor pressure. Vapor pressure depends on the temperature, and a few other things. As we know, water turns to vapor at 212 oF at atmospheric pressure, when we boil water in the open pot. If the pot were closed, the water would reach higher pressure before it boils. Conversely, if the pressure were reduced (vacuum), water would boil at lower temperature. It will boil at room temperature, if the absolute pressure is less then about 0.4 psia. Water has low vapor pressure, but other substances may have very high value. Freon, for example, has vapor pressure of about 90 psia, and ethane value of vapor pressure is about 700 psi, - at 80 0F. Knowing vapor temperature without relating it to a corresponding temperature is meaningless. Sometimes it is good to have a tabulation, or a graph, showing the relationship between the vapor pressure and temperature. The higher the temperature - the higher the vapor pressure is. Centrifugal pump is a "pressure generator", produced by the centrifugal force of its rotation impeller. The pressure gets higher as flow progresses from the suction to discharge. This is why vaporization of liquid is most likely to happen in the inlet (suction) region, where the pressure is lowest. In practice, it is difficult to know exactly when vaporization (cavitation) happens, so it is wise to keep some margin of available pressure over vapor pressure. Pressure is expressed in "psi", but can also be expressed in feet of water, and the conversion formula is: FT = PSI x 2.31 / SG, where SG is specific gravity. This pressure, expressed in feet of water, is called discharge head at the pump exit side, or suction head on the inlet side. The difference is a pump developed head, also called a total dynamic head (TDH). These heads must

include both static and dynamic components. Static part is what we measure by the gage in front of a pump, and dynamic, according to Bernoulli, is velocity head V2/2g. For example, suppose an inlet pressure gage installed in a 2" pipe directly in front of a pump delivering 100 gpm oil with specific gravity SG = 0.9, reads 10 psig. To calculate velocity head, find the pipe net area, which is A = 3.14 x d2 / 4 = 3.14 x 22 / 4 = 3.1 in2. The velocity can be calculated by the formula: V = (Q x 0.321) / A = (100 x 0.321) / 3.1 = 10.4 ft / sec Then, the velocity head is: V2 / 2g = 10.42 / (2 x 32.2) = 1.7 ft, or, converted to psi is = 1.7 x 0.9 / 2.31 = 0.7 psi The total suction pressure is then 10 + 0.7 = 10.7 psi, or, if expressed in feet of water, = 10.7 x 2.31 / 0.9 = 27.5 feet It is best to have gages as close as possible to the pump, on the suction and discharge sides. Unfortunately, often these gages are not installed, (which somehow happens more often on the suction side), and suction head in front of the pump is estimated by calculations, by subtracting the pressure (head) losses from the known value of head upstream, and adjusting by elevation correction, according to Bernoulli. In many cases, the upstream datum is a known liquid level in a suction tank. Examples: a) Tank open to atmosphere:

Figure 1-3a: Open tank

Figure 1-3b: Pressurized tank

hsuction = 2.7x2.31/0.9 + 10 7 = 9.9 Figure 1-3c: Tank under vacuum For water and similarly low viscosity liquids, suction losses are usually low, and often are disregarded. However, for more viscous substances, such as oils, these losses can be substantial, and may cause the pressure in front of the pump drop below the vapor pressure, causing cavitation. This is why the inlet velocity must be minimized, as the losses depend on velocity squared.

Longer pipe runs, bends, turns and other restrictions, add to inlet losses, leading to further pressure reduction in front of a pump. As a quiz, using the examples above, see if you can figure out what happens to inlet pressure if the pipe diameter is doubled? Or made half the diameter? (If you do send the answer to us, and will publish it the Pump Magazine). To avoid cavitation, what matters is not the suction pressure, but how much higher it is then the vapor pressure of the liquid being pumped. This is where a concept of NPSH comes handy. The available NPSHA thus is simply the difference between this total suction head, as discussed above, and vapor pressure, expressed as head, in feet. Pump manufacturers conduct tests by gradually lowering suction pressure, and observing when things begin to get out of hands. For a while, as pressure decreases (i.e. NPSHA gets smaller), nothing happens, at least nothing obvious. A pump, operating at a set flow, keeps on pumping, and develops constant head. At some point, when the value of suction pressure (and corresponding NPSHA), reaches a certain value, a pump head begins to drop, which typically happens rather suddenly:

Figure 1-4: Development of Cavitation Actually, things are happening inside the pump well before the sudden drop of head, but they are not as obvious. First, at still substantial suction pressure, small bubbles begin to form. This is called incipient cavitation - sort of tiny bubbles in your water cattle that begins to percolate before water is fully boiling. These small bubbles are formed and collapse, at very high

frequency, and can only be detected by the special instrumentation. As pressure is decreased further, more bubbles are formed, and eventually there are so many of them, that the pump inlet becomes "vapor-locked", so that no fluid goes through, and the pump stops pumping - the head drops and disappears quickly. It would be nice if enough pressure was always available at the suction so that no bubbles were formed whatsoever. However, this is not practical, and some compromise must be reached. The Hydraulic Institute (HI) has established a special significance to a particular value of NPSHA, at which the pump total developed head drops by 3%. The value of this NPSHA, at which a pump losses 3% TDH, over (i.e. in access of) vapor pressure is called net positive suction head required (NPSHr) in order to maintain 3% TDH loss. NPSHr = (Hsuction - Hvapor), required to maintain 3% TDH loss NPSHr is, therefore, established by actual test, and may vary from one pump design to another. In contrast, the available NPSHa, has nothing to do with a pump, but is strictly a calculated value of total suction head over vapor pressure. Clearly, NPSHA must be greater then NPSHR, in order for a pump to make its performance, i.e. to deliver a TDH, at a given flow. It is easy to know when a NPSH problem is obvious - a pump just stops pumping, but the vapor bubbles do not need to be so dramatically developed to cause TDH drop, even smaller bubbles can cause problems. The evolved bubbles get carried on through the impeller passage, at which pressure is rising from inlet to exit of the blade cascade. This increased pressure causes the reverse to what happened to a bubble "awhile back", when it first became a bubble formed from a liquid particle during phase transformation (boiling). Now, the bubble is at the somewhat higher pressure, which tries to squeeze it, against the vapor surface tension that keeps the bubble a bubble. The bubble collapses (implodes), with a sudden in-rush of surrounding liquid into a vacuum space previously occupied by the bubble. The inrush is

accompanied by a tremendous, but a very localized, pressure shock, which, if imploded in the vicinity of the metal (impeller blade), would cause a microscopic hammerlike impact, eroding a small particle of metal. With enough bubbles and enough time, the impeller vanes can be eroded away quickly, a phenomenon known as cavitation (hence the word) damage. This is why an NPSHA margin (M=NPSHA-NPSHR) is important, which is typically at least 3-5 feet, and preferably should be even more, if possible. The NPSHR, discussed above, was so far limited to a particular flow on a pump performance curve. At higher flow, the internal fluid velocities are higher, and, according to Bernoulli, the static pressure (or static head) part becomes less, i.e. closer to vapor pressure. The static pressure, therefore, must be increased externally, i.e. a higher value of NPSHR is needed for higher flows. This is why the NPSHR curve shape looks like this:

Figure 1-5: Ample margin of NPSHA is important It is important to specify an ample margin of NPSHA over the pump NPSHR for a complete range of operation, and not just at a single rated flow point. The following example illustrates a common mistake, leading to the NPSHproblem. The pump was procured with the intend to deliver between 350-500 gpm, and the manufacturer quotation indicated 16 feet required NPSHR at 500 gpm. As a process later changed, more flow was required, and the discharge valve was opened to allow pump to deliver more flow, 750 gpm. However, as can be seen from Figure 1-5, at about

700 gpm, the NPSHR exceeded the NPSHA available at the installation, and pump started to experience typical NPSH problems - noise, loss of performance, and impeller cavitation damage. An instinctive thought to address the issue of cavitation due to flow-run out operation is to "overkill" on a pump size, and therefore always stay to the left of the BEP. In the example above, a larger pump, having same 16 feet NPSHR, but at 750 - 800 gpm, would never run out of the NPSHA. That is true, and, in fact, this is exactly what has been a common practice in the past, where an oversized (and, by the way, more expensive) pump would be specified "to make sure", - just to discover other, just as severe problems. When a centrifugal pump operates below certain flow point, a phenomenon known as flow recirculation in the impeller eye starts. This depends on several design factors, such as suction specific speed (see in other article of Pump Magazine), but generally recirculation begins below 80-60% flow, and becomes quite sever below 40-20%. At even lower flows, recirculation may become especially severe, and is known as surge - violent, lowfrequency sound, accompanied by strong low-frequency vibration of the pump and piping:

Figure 1-6 Problems come up when pump operates at too low flow In addition to obvious mechanical problems with recirculation, the flow undergoes a complex vortexing motion at the impeller inlet (eye), with localized high velocities of the vortex causing horse-shoe looking cavitation damage, usually on the "blind" side of the

blade, as compared to high-flow cavitation. Other problems add oil to the fire - radial thrust, which skyrockets at low flow, causes deflections of the shaft, leading to seal leaks, bearings life reduction, and even shaft breakage (see other articles of the Pump Magazine on these subjects). Troubleshooting methods and failure analysis techniques help to pinpoint a cavitation problem with a particular pump. The indications of the high flow cavitation are different from the low flow recirculation damage. Side of the blades, the extend and shape of the cavitation trough, can be helpful in determining the causes of each individual problem. To learn more about this topic, e-mail your comments to us at:
Below are some comments from a reader: Dear Dr. Nelik,

How does a change in pump suction configuration affect the system curve? I've read the Article #1 "How Does Pump Suction Limit the Flow" and I've also read reader question #58 that deals with changes to the inlet and outlet piping to a pump. After carefully reading Article #1, I can conclude that suction side considerations do not truly limit flow, but flow must be adequate to prevent cavitation (too high flow - low pressure - cavitation) or flow recirculation at too low flow. To stay out of these problem areas, one must design the system (downstream of the pump) and select a pump so that the operating point flow rate does not cause upstream problems. My question revolves around how the upstream configuration affects the system curve - if at all. There appears to be a degree of independence between the system configuration downstream of the pump and the upstream configuration, since the only influence on where the system curve moves to is dependant on the total static head (difference in elevations in the fluid levels for the suction and discharge, and any additional pressurization if present). Friction in the suction line does not appear to influence the system curve but it does influence the NPSHA. Reader question #58 describes changes to inlet and outlet piping to support a ship in dry-dock. The first step is to determine where the system curve for the new outlet piping intersects the pump curve and you get an operating point. Then for that flow rate, check the NPSHA for the new suction piping configuration against the NPSHR required for the pump. If NPSHA is greater (with some margin) than NPSHR, you should be OK. This leads me to my question. We have two identical generators in two separate machinery rooms on a ship. For simplicity, let's say that the inlet and outlet piping for the Seawater cooling is the same length for both generators - 20 feet of inlet piping and 20 feet of outlet piping. The only difference is that because of space constraints, one generator has 4" inlet piping to the generator skid, and the other has 6" piping to the skid. When the suction line reaches the skid, the piping is again identical for both - a one-foot long 5" line that enters the pump. The size of the outlet piping from the skid on both is the same. The system curve (dependent on the downstream pipe configuration) should be identical for both. The total

static head is identical for both. Therefore based on the methods described above, the pump should operate at the exact same point for both. Since the suction line size is different, the NPSHA to the two pumps is different and the pressure at the pump inlet is different (there are more losses through the 4" suction). Provided there is no cavitation in either case, would this influence the system curve, and therefore the operating point, in any way? Based on what I've read, I think the answer in no. That's why I said before there seems to be a degree of independence between the piping upstream and the piping downstream as far as operating point is concerned. When the units were running, both pumps should be operating at the same outlet pressure and flow, but upstream of the pump, the water would obviously be going faster through the 4" suction pipe. This is what I think will happen, and I want to verify. Regards, Lionel A. Sequeira Mechanical Engineer, Dept of the Navy 9/30/04
Editor comments: Lionel: You got the concept pretty well, with just a small correction. Both pumps will impart the same amount of energy to the fluid. (And, technically, pump head is exactly that: energy per unit of mass). Thus, both pumps will show the same differential pressure at a given operating flow (assuming, as you said, no cavitation). However, the first pump will start off at a lower inlet pressure, because it will loose more pressure on friction between the supply tank and pump inlet pipe due to smaller (4) pipe. Thus, the discharge gage will show lower pressure for the first pump, - by the amount of difference in losses between the first pump suction losses and the second pump suction line losses. Remember, pump head is not discharge pressure expressed in feet of water, but the differential pressure, expressed in feet of water. Another way of thinking this is that the pump head is the difference between total energies between discharge side (near the pump) and the suction side (also right at the pump), corrected by gage elevation differences (which is usually minor). The discharge head is a sum of pressure head, velocity head and elevation head. Same goes for the suction. Assuming the gages are at the same elevation, this cancels out. The velocities are the same, as in the very immediate proximity (discharge flange at exit, or a 5 short pipe at suction) pipe sizes are the same, thus velocity heads are the same for a given flow. Since the total differential is the same, then the pressure difference is the same as well. But, the suction inlet pressure is lower, and, adding the constant differential amount to it makes the discharge pressure lower for the first pump. I am impressed that you are getting really deep into pump hydraulics! As a suggestion also consider to attend our Pump School, - and bring-in your colleagues as well ! We go over these sorts of things, plus much more, at our regular Pump School sessions in Atlanta. It is intense, informative, and fun. In fact, many Navy engineers have attended in the past, as well as maintenance folks from chemical plants, refineries, power plants, paper mills, and other industries. Keep on pumping! Dr. Lev Nelik, P.E. Editor, Pump Magazine

DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article #2: Specific Speed (NS) - Why Is It Dimensionless, or - Is It?


Pump Specific Speed is one of the several dimensionless parameters used in pumps. Pump designers like to use dimensionless numbers, because his allows them to analyze and compare pump performance regardless of size. Otherwise, how would you compare a 6" pump performance with a 10" pump? What is a typical maximum efficiency is to be expected from a given pump? There is an infinite variation of pump sizes, flows, generated pressures, speeds, etc.- so, it would be impossible to compare a performance of the 6" pump running at 3200 rpm at 200 gpm, with a 10 pump running at 1400 rpm at, say, 310 gpm. Besides, the same performance can be achieved by a wider impeller with a narrower volute, as by the narrower impeller with a wider volute, if their outside diameters are different. In other words, there is an infinite variation of operating parameters (flows, heads, speeds) that depend on an infinite variation of pump geometry internals. A specific speed NS reduces this infinite variation of performance parameters into a single number, and an infinite variation of geometries into a single characteristic non-dimensional impeller profile shape. Then, a given impeller shape would point at the unique performance factor, i.e. specific speed. Obviously, to eliminate dependency of the operating parameters specific speed must be dimensionless, otherwise the remaining dimension (gallons, pounds, feet, etc.) would affect the comparison. Originally, specific speed was defined as:

OmegaS = Rotating Speed x Flow0.5 / Head0.75 , The units had to be in any system of "consistent" units, for example: Rotational Speed - 1/sec Flow - ft3/sec, Head in feet times the gravitational constant (which has units of ft/sec2):

This is how the initial definition of "specific speed" was meant to be "dimensionless". But, such set of units is awkward, and the gravitational constant is simply dropped, and flow is entered in gpm, and rotational speed in RPM. Specific speed thus defined is denoted NS. Strictly speaking, such modified definition is not dimensionless, but the proportionality terms are constants (such as ft3/sec x 448.8 = gpm), which makes the NS directly proportional to "OmegaS". The non-dimensional nature of the NS (disregarding the dropped constant) makes it independent of rotational speed RPM. This is because of the affinity laws governing the operation of the centrifugal pump. These state that flow varies linear with speed, and head as a square: Q ~ RPM H ~ RPM2 If RPM is changed n-times, flow changes n-times also, and head changes n2 times.

(Note that NS has not changed when RPM changed). Incidentally, similar rule happen to apply to impeller OD cut: if impeller diameter changes n-times: flow changes linearly, and head as a square. The net result: no change in NS. Also, if all geometric linear dimensions of a pump are changed at the same ratio, its specific speed still remains constant. (This part, however, is more specialized, - usually of the realm of the pump hydraulics designers - see special discussion on that):

Same shape, but different size i.e. Specific Speed is a measure of the impeller geometry similarity (i.e. affinity), - not the size of it. In other words, if you are looking at a cross-sectional drawing of a pump, there would be a certain specific speed NS, unique to the impeller shape regardless whether you are looking at a cross-section of a 6" design, or a 60" design. It was discovered that a pump efficiency at the best efficiency point (BEP) depends mainly on the Specific Speed, and a pump with Specific Speed of 1500 is more efficient then the one with specific speed of 1000. Charts and publications were developed, and are available as an estimate of a "reasonably achievable" efficiency for a given pump, versus actual. Similarly to pump specific speed (NS), a concept of a suction specific speed (NSS) has also been introduced. Here, the head at the denominator is substituted by the NPSHR at the BEP. But this is a separate discussion (see among other Articles).

To learn more about this topic, e-mail your comments to us at: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article #3: Suction Specific Speed (NSS)


Suction Specific Speed is another dimensionless parameter used in centrifugal pumps. Much of the discussion on specific speed (NS) applies to suction specific speed (NSS). (See topic "Specific Speed (NS)why is it dimensionless, or - is it?"). While specific speed (NS) is mostly related to the discharge side of the pump, the suction specific speed deals primarily with its suction (inlet) side. The head (H) term in the denominator of the defining formula for the NS is substituted by the NPSHR: NSS = RPM x Q0.5 / NPSHR0.75, where flow is in gpm, and NPSHR is in feet. Values of NSS vary from about 6,000 to 15,000, and sometimes even higher for the specialized designs. From the discussion of a pump suction performance (see topic "How does pump suction limit the flow?"), we know that conflicting demands are imposed on a pump system by the pump user and a pump manufacturer. A user would prefer to provide as low NPSHA as possible, as it often relates to a system cost: for example, higher level of liquid in the basin of the cooling water pumps requires taller basin walls, or deeper excavation to lower a pump centerline below the liquid level. A pump manufacturer, on the other hand, wants to have more NPSHA, with an ample margin above the pump NPSHR, to avoid cavitation, damage, and similar problems.

In other words, a wider margin (M) can be achieved either by increasing the NPSHA, or decreasing the NPSHR, since M = NPSHA - NPSHR Thus it may appear that a lower NPSHR design is preferential, and a competing pump manufacturer might present a lower NPSHR design as automatically translated into construction cost savings - because of not having to increase the NPSHA. Since a lower NPSHR design means a higher value of NSS (according to the definition above), the highest NSS design might seem to look best. In reality, however, this is not so. In a topic "How does pump suction limit the flow?" it was explained that higher flow velocities result in reduction of the static pressure, which may then become dangerously close to the fluid vapor pressure and cavitation. Thus, lower velocities result in higher localized static pressure, i.e. a safer margin from the cavitating (i.e. vaporization) regime. Since the velocity is equal to flow divided by the area, a larger area (for the given flow) reduces the velocity, - a desirable trend. This is why a larger suction pipe is beneficial at the pump inlet. Cavitation usually occurs in the eye region of the impeller, and if the eye area is increased - velocities are decreased, and the resulting higher static pressure provides a better safeguard against vaporization (cavitation). So, a larger impeller eye seems like a way to lower the NPSHR:

Figure 3-1 Larger impeller eye results in lower NPSHR at BEP, but certain problems arise at off-peak operation Unfortunately, the flow of liquid at the impeller eye region is not as simple and uniform as it is in a straight run of a suction pipe. Impeller eye has a curvature, which guides the turning fluid, like a car along the

sharp curves of the road, into the blades and towards the discharge. If a pump operates very close to its BEP, the inlet velocity profile becomes proportionally smaller, but the fluid particles stay within the same paths: If, however, a pump operates below its BEP, the velocity profile changes, and no longer can maintained its uniformity and order. Fluid particles then begin to separate from the path of the sharpest curvature (which is the impeller shroud area), and the resulting mixing and wakes produce a turbulent, disorderly flow regime, which makes matters difficult from the NPSHR standpoint.

Figure 3-2: Even thought large eye impeller has better NPSHR at BEP, it has flow separation problems at low flow The upshot of all this is that a larger impeller eye does decrease the NPSHR at the BEP point, but causes flow separation problems at the offpeak low-flow conditions. In other words, a high Suction Specific Speed (NSS) design is better only if a pump does not operate significantly below its BEP point. Interestingly, with very few exceptions, there is hardly a case where a centrifugal pump operates strictly at the BEP. The flow demands at the plants change constantly, and operators throttle the pump flow via the discharge valve. High NSS designs are known to result in reliability problems because of such frequent operation in the undesirable low flow region. Actual plant studies have shown, that above NSS of 8500 9000, pumps reliability begins to suffer - exponentially:

Figure 3-3 Plant experience shows that impellers designed with NSS greater than 9000 have poor reliability record Realizing this, around mid-80s, users started to limit the value of the NSS, and a Hydraulic Institute uses NSS = 8500 as a typical guiding value. It might be of interest for you to calculate the values of NSS of your pumps, and find out if a correlation between those and reliability exists at your plant. To learn more about this topic, e-mail your comments to us at: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article 4: What Is L3/D4 and Why Is It Good If It Is low?


L3/D4 ("L-cubed over D-to-the-fourth") is a measure of pump rotor stiffness, it's ability to resist radial load and to minimize deflection. It comes from the basic cantilevered beam deflection formula, which you can find in any book on mechanics: y = F x L3 / (3 x E X I), where F is radial load, L is cantilevered length, E modulus of the elasticity of the material, and I is moment of inertia.

Figure 4-1 Pump shaft deflects under load, as a structural beam

Load F could be a cantilevered weight of the overhung load, a centrifugal force created by the end load unbalance, a hydraulic radial thrust of a centrifugal pump, or a combination of forces. These forces can be static and not changing direction (such as weight), or dynamic (such as rotating unbalance). For a circular shafts I = 3.14 x D4 / 64, and thus a deflection at a given force is proportional to: y ~ L3/D4, or abbreviated it is often written as L3D4 Thus L3D4 becomes a criterion for an indirect assessment, or a comparison, of a rotor deflection under load. Mechanical seals can not tolerate much deflection, and are prone to leakage if their faces are displaced by more then 0.001" - 0.002".

Figure 4-2: Seals will leak as excessive loads causes shaft deflection and seal faces misalignment The lower L3D4, the less is shaft deflection, which is better for the seals. If L3D4 becomes too large, a pump shaft can snap, especially if operated close to shut-off, where hydraulic radial loads are excessive (see other related articles in the Pump Magazine on this subject). ANSI pumps have L3D4 ratios range from 20 to 120, but new designs have been introduced with this ratio below 10.0 (see related articles,

such as "Barrier" design, that combines mag-drive and gas seal technologies). You can easily determine the L3D4 ratio of your pump by measuring the length of the shaft from the center of the bearing closest to the impeller and impeller centerline, and the diameter of the shaft under the bearing. Then, cube the length, raise the diameter in to fourth power, and obtain the ratio. A shaft diameter changes from the bearing towards the impeller, but its value under the bearing is taken nominally. Tabulating these ratios for different pumps, you can make you own plant database of L3D4 of different designs.

Figure 4-3: L and D parameters for a pump rotor As a word of caution, the ultimate manifestation of pump reliability is its operating history, which could at times conflict with what a L3D4 number would indicate. Nevertheless, it is a good guide, and helpful as one of several design factors that may have an effect on the pump reliability. To learn more about this topic, e-mail your comments or questions to: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article #5: Relief Valves for PD Pumps - Internal or External?


Positive Displacement pumps are knows as "flow generators", while centrifugal pumps can be thought as pressure (or head) generators. It is easy and convenient to control the centrifugal pump with a discharge valve: it changes a system curve, resulting in pump operating at different flows (see Article "How does pump suction limit the flow?"). Not so for the positive displacement pumps. All types of a PD pump (piston, rotary gear, lobe, vane, etc.) have a major similarity, which distinguishes them from the centrifugal pumps: PD pump curves show that they produce almost (depending on viscosity and internal clearances) the same flow, regardless of the differential pressure:

Figure 5-1: Hydraulic performance pump curves - centrifugal (head versus flow) and positive displacement (flow versus differential pressure) A PD pump "does not know and does not care" what happens on the system side - with each rotation of a shaft, a rotary gear pump, for example, moves the same amount of fluid from its inlet, to discharge regardless of the differential pressure (second order dependency actually does exist, as function of viscosity and clearances, but this is covered in another article). The pump does, however, require more torque, more power, as pressure increases, but as far as flow is concerned - it stays constant. As a result, a PD pump will try to move the same amount of fluid, regardless of the position of a discharge valve (or other system-related conditions downstream) - resulting in pump discharge pressure building up very rapidly if valve begins to close - sometimes with almost

negligible incremental turn of the valve. Eventually, something is bound to fail - a burst hose, a sheared off shaft, a seal, and so forth. Such failures are catastrophic and dangerous. To prevent them, a relief valve should always be employed with a PD pump. Some pump designs incorporate the internal in-built relief valves, and others do not and require the external relief valve. The advantage of the internal relief valve is obvious - it is an integral pump of the design, and comes with a pump. However, internal valves are usually limited to application of relatively "benign" substances, such as, for example, oils. The reason for that is as follows. Say a gear pump is pumping against a 60 psi differential pressure, and its internal relief valve spring is set to open the valve to by-pass at a 100 psi (actually it would crack somewhat earlier, say at 90 psi, and reach a full by-pass at 100 psi). When the valve opens, 100 psi drives the liquid from the discharge port immediately to a suction port, in a very close, "short-circuited" manner. The energy delivered to the liquid at 100 psi differential has to go somewhere - and it goes to heat, with vaporization (flashing) of the liquid possible. Depending on the specific heat of the fluid, and the rate of the heat rejection through the casing walls, flashing may be more, or less, likely. The nature of the pumped fluids is also important: overheating of oils, as bad as it is, may not be as dangerous as overheating of a sulfuric acid, caustic, or other similar chemical. It is better in such cases to use the external relief valve, where fluid would bypass from the discharge back to the tank, eliminating pump short-circuiting altogether. For this reason, pumps designed for chemicals and other similar substances should not use internal relief valves, but the external ones. To learn more about this topic, e-mail your comments to us at: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article #6: Metallurgy - is 316ss Soft or Hard?


What sort of question is that?! Everyone "knows" that 316ss is very soft - as any austenitic stainless steel would be! Right? Mmmmm... Why is stainless steel so good against cavitation? - if it were soft, wouldn't the imploding bubbles (see Pump Magazine Article on that subject) erode the material away at no time? The reason for such excellent resistance of stainless steel to cavitation is it work-hardening property. You can easily make an indentation on a surface of an unworked stainless steel strip. However, after hitting the strip with a hammer for some time, it becomes much difficult to make such indentation - the surface work-hardens. (Ironically, this is why stainless steel, as soft as it is, is difficult to machine - it work-hardens as a cutting tool goes over it, resulting in a chunky, chisel-like chip, instead of a smooth clean cut). The same happens when cavitation bubbles bombard the surface of a stainless steel - it work-hardens, and begins to resist any further cavitation very quickly. 316ss (CF8M) resists cavitation about 10-15 times better than cast iron, because of this work-hardening characteristics. Of course, further metallurgical modifications can make stainless steel even more resistant to cavitation - for example, CA6NM is roughly 2-3 times more resistant to cavitation as compared to 316ss. Things are not always as they first appear! To learn more about this topic, e-mail your comments to us at: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article #7: Lubrication Practices: How GreaseLubricated Bearings Function


A shielded, grease-lubricated ball bearing can be compared to a centrifugal pump having the ball and-cage assembly as its impeller and having the annulus between the stationary shield and the rotating inner race as the eye of the pump. Shielded bearings are not sealed bearings. With the shielded type of bearing, grease may readily enter the bearing, but dirt is restricted by the close fitting shields. Bearings of the sealed design will not permit entry of new grease, whereas with shielded bearings grease will be drawn in as the bearing cage assembly rotates. The grease will then be discharged by centrifugal force into the ball track of the outer race. If there is no shield on the backside of this bearing, the excess grease can escape into the inner bearing cap of the equipment bearing housing. Single-shield bearings Plants applying best-in-class-practices today consider the regular single shield bearing with the shield facing the grease supply (Figure 7-1) to be the best arrangement. Their experience indicates this simple arrangement will extend bearing life. It will also permit an extremely simple lubrication and relubrication technique if so installed. This technique makes it unnecessary to know the volume of grease already in the bearing cartridge. The shield serves as a baffle against agitation. The shield-toinner-race annulus serves as a metering device to control grease flow. These features prevent premature ball bearing failures caused by contaminated grease and heat buildup due to excess grease. Further, warehouse inventories of ball bearings can be reduced to one type of bearing for the great bulk of existing greaselubricated ball bearing requirements. For other

services, where an open bearing is a "must," as in some flush-through arrangements, the shield can be removed in the field.

Figure 7-1 Single-shielded equipment bearing, with the seal facing the grease cavity Double-shielded bearings Some manufacturers still subscribe to a different approach, having decided in favor of double-shielded bearings. These are usually arranged as shown in Figure 7-2. The housings serve as a lubricant reservoir and are filled with grease. By regulating the flow of grease into the bearing, the shields act to prevent excessive amounts from being forced into the bearing. A grease retainer labyrinth is designed to prevent grease from reaching the inner side of the bearing.

Figure 7-2 Double Shielded bearing with grease. Metering plate facing the grease reservoir On equipment furnished with this bearing configuration and mounting arrangement, it is not necessary to pack

the housing next to the bearing full of grease for proper bearing lubrication. However, packing with grease helps to prevent dirt and moisture from entering. Oil from this grease reservoir can and does, over a long period, enter the bearing to revitalize the grease within the shields. Grease in the housing outside the stationary shields is not agitated or churned by the rotation of the bearings and consequently, is less subject to oxidation. Furthermore, if foreign matter is present, the fact that the grease in the chamber is not being churned reduces the probability of the debris contacting the rolling elements of the bearing. On many pieces of equipment furnished with greaselubricated double-shielded bearings, the bearing housings are not usually provided with a drain plug. When grease is added and the housing becomes filled, some grease will be forced into the bearing. At this point any surplus grease will be squeezed out along the close clearance between the shaft and the outer cap because the resistance of this path is less than the resistance presented by the bearing shields, metering plate, and the labyrinth seal. Open bearings High-load and/or high-speed bearings are often supplied without shields to allow cooler operating temperature and longer life. One such bearing is illustrated in Figure 7-3.

Figure 7-3 High load/ high supplied without shield

speed

bearings

are

often

If grease inlet and outlet ports are located on the same side, this bearing is commonly referred to as "conventionally grease lubricated." If grease inlet and outlet ports are located at opposite sides, we refer to "cross-flow lubrication." Figure 7-4 shows a cross-flow lubricated bearing.

Figure 7-4 Open lubrication

bearing

with

cross

flow

grease

Lifetime lubricated, "sealed" bearings Lubed for-life bearings incorporate close-fitting seals in place of, or in addition to shields. These bearings are customarily found on low horsepower equipment or on appliances, which operate intermittently.

Although a large petrochemical company in West Virginia has expressed satisfaction with sealed ball bearings in certain equipment applications as long as bearing operating temperatures remained below 150o C. (300o F) and speed factors DN (mm bearing bore times revolutions per minute) did not exceed 300,000. Close-fitting seals can cause frictional heat and that loose fitting seals cannot effectively exclude atmospheric air and moisture, which will cause grease deterioration. Procedures for Re-Greasing Equipment Bearings Rotating equipment bearings should be re-greased with grease, which is compatible with the original charge. It should be noted that the polyurea greases often used by the equipment manufacturers may be incompatible with lithium base greases. Single-Shielded Bearings To take advantage of single-shield arrangements, Phillips Petroleum developed three simple recommendations: 1. Install a single-shield ball bearing with the shield facing the grease supply on equipment having the grease fill and-drain ports on that same side of the bearing. Add a finger full of grease to the ball track of the backside of the bearing, during assembly. 2. After assembly, the balance of the initial lubrication of this single-shielded bearing should be done with the equipment idle. Remove the drain plug and pipe. With a grease gun or high volume grease pump, fill the grease reservoir until fresh grease emerges from the drain. The fill and drain plugs should then be reinstalled and the equipment is ready for service. It is essential that this initial lubrication not be attempted while the equipment is running. It was observed that to do so would cause, by pumping action, a

continuing flow of grease through the shield annulus until the overflow space in the inner cartridge cap is full. Grease will then flow down the shaft and into areas where it is not wanted. This will take place before the grease can emerge at the drain. 3. Relubrication may be done while the equipment is either running or idle. (It should be limited in quantity to a volume approximating one-fourth the bearing bore volume.) Test results showed that fresh grease takes a wedge-like path straight through the old grease, around the shaft, and into the ball track. Thus, the overflow of grease into the inner reservoir space is quite small even after several relubrications. Potentially damaging grease is thus kept from the stator winding, in motors. Further, since the ball and cage assembly of this arrangement does not have to force its way through a solid fill of grease, bearing heating is kept to a minimum. In fact, it was observed that a maximum temperature rise of only 20 o F. occurred 20 minutes after the grease reservoir was filled. It returned to 5 o F rise two hours later. In contrast, the double-shield arrangement caused a temperature rise of over 100 o F (at 90 o F ambient temperature the resulting temperature was 190 o F.) and maintained this 100 o F rise for over a week. Double-Shielded Bearings 1. Ball Bearings A. Pack (completely fill) the cavity adjacent to the bearing. Use necessary precautions to prevent contaminating this grease before equipment is assembled. B. After assembly, lubricate stationary equipment until a full ring of grease appears around the shaft at the relief opening in the bracket. 2. Cylindrical Roller Bearings A. Hand pack bearing before assembly.

B. Proceed as outlined shielded ball bearings.

in

(1)

and

(2)

for

double

If under-lubricated after installation, the double shielded bearing is thought to last longer than an open (non-shielded) bearing given the same treatment because of grease retained within the shields (plus grease remaining in the housing from its initial filling). If over-greased after installation, the double-shielded bearing can be expected to operate satisfactorily without overheating. This will be as long as the excess grease is allowed to escape through the clearance between the shield and inner race, and the grease in the housing adjacent to the bearing is not churned, agitated and caused to overheat. It is not necessary to disassemble equipment at the end of fixed periods to grease bearings. Bearing shields do not require replacement. Double-shielded ball bearings should not be flushed for cleaning. If water and dirt are known to be present inside the shields of a bearing because of a flood or other circumstances, the bearing should be removed from service. All leading ball-bearing manufacturers are providing reconditioning service at a nominal cost when bearings are returned to their factories. As an aside, reconditioned ball bearings are generally less prone to fail than are brand new bearings. This is because grinding marks and other asperities are now burnished to the point where smoother running and less heat generation are likely. Open Bearings Equipment with open, conventionally greased bearings is generally lubricated with slightly different procedures for drive-end and opposite end bearings.

Figure 7-5 Double-Shielded Bearings 1. Lubrication procedures for drive-end bearings A. Relubrication with the shaft stationary is recommended. If possible, the equipment should be warm. B. Remove plugs and replace with grease fitting. C. Remove large drain plug when furnished with the equipment. D. Using a low pressure, hand operated grease gun, pump in the recommended amount of grease, or use 1/4 of bore volume. E. If purging of system is desired, continue pumping until new grease appears either around the shaft or at the drain opening. Stop after new grease appears. F. On large equipment provisions have usually been made to remove the outer cap for inspection and cleaning. Remove both rows of cap bolts. Remove, inspect and clean cap. Replace cap, being careful to prevent dirt from getting into bearing cavity. G. After lubrication allow the equipment to run for fifteen minutes before replacing plugs. H. If the equipment has a special grease relief fitting, pump in the recommended volume of grease or until a one-inch long string of grease appears in any one of the relief holes. Replace plugs. I. Wipe away any excess grease which has appeared at the grease relief port. 2. Lubrication procedure for bearing opposite drive end

If bearing hub is accessible, as in smaller equipment with large couplings or drip-proof equipment, follow the same procedure as for the drive end bearing. For fancooled equipment note the amount of grease used to lubricate shaft end bearing and use the same amount for commutator-end bearing. Motor bearings with grease inlet and outlet ports on opposite sides, are called cross-flow lubricated. Regreasing is accomplished with the equipment running. The following procedure should be observed: 2.1. Start equipment and allow to operate until normal equipment temperature is obtained. 2.2 Inboard bearing (coupling end): A. Remove grease inlet plug or fitting. B. Remove outlet plug. Some equipment designs are equipped with excess grease cups located directly below the bearing. Remove the cups and clean out the old grease. C. Remove hardened grease from the inlet and outlet ports with a clean probe. D. Inspect the grease removed from the inlet port. If rust or other abrasives are observed, do not grease the bearing. Tag equipment for overhaul. E. Bearing housings with outlet ports: (1) insert probe in the outlet port to a depth equivalent to the bottom balls of the bearing; (2) replace grease fitting and add grease slowly with a hand gun. Count strokes of gun as grease is added; and (3) stop pumping when the probe in the outlet port begins to move. This indicates that the grease cavity is full. F. Bearing housings with excess grease cups: (1) replace grease fitting and add grease slowly with a handgun. Count strokes of gun as grease is added and (2) stop pumping when grease cavity is full. 3. Outboard bearing (fan end):

a. Follow inboard bearing procedure provided the outlet grease ports or excess grease cups are accessible, b. If grease outlet port or excess cup is not accessible, add 2/3 of the amount of grease required for the inboard bearing. c. Leave grease outlet ports open-do not replace the plugs. Excess grease will be expelled through the port. d. If bearings are equipped with excess grease cups, replace the cups. Excess grease will expel into the cups. Lev Nelik Pumping Machinery, LLC

Some additional feedback from the professionals in the field:


Roy A. Forson, CLS Maintenance Manager Imerys Performance Minerals : Type A: The proper bearing re-lubrication with a drain plug is: 1. Check Temperature 2. Wipe grease fitting 3. Remove purge plug and clean out hard grease (unit must be down. If you cannot take the top cap off, then insert small spoon-type tool and feel for hard grease. If no hard grease is found, then continue below. 4. Check seals for excessive grease (blown seals, cracked seals. These should be replaced) 5. Pump new grease until new grease exits out of the purge plug drain. 6. Run unit with purge plug removed for 10-30 minutes. (This will allow the grease to equalize) 7. Clean and replace the purge plug. Type B: The proper bearing re-lubrication without a drain plug is: 1. Check temperature 2. Shut down & immediately remove grease fitting to allow the bearing to self-purge 3. Clean and replace the fitting 4. Pump a limited quantity of grease to avoid rupturing the grease seal 5. Remove the fitting and allow for equalization 6. Repeat until grease comes out

7. Replace grease fitting


Type C: The proper bearing re-lubrication with a relief fitting is: 1. Check temperature 2. Clean fitting and pump grease until it relieves 3. If it doesn't relieve, fitting may be plugged. Remove and replace 4. Run and check for relief/temperature

Additional general considerations for re-lubrication of bearings other than electric motor bearings:
(Remember, if you are not going to use any calculation for re-greasing, then you are guessing) 1. If unit is running, check temperature of all bearing surfaces. It is important to know if it's running hot. 2. Using all safety precautions, locate grease fitting and clean the fitting of debris by wiping it off with a clean rag. 3. Verify the type of grease in the grease gun - do not take for granted that what is marked on the pump is what's in the pump. 4. Pump a little bit of grease out of the gun and wipe the end. This will remove any debris trapped in the nozzle 5. 6. Install the end of the grease nozzle onto the grease fitting and verify it is on properly Pump grease into the bearing based on the plant "grease amount" recommendations.
Caution - if grease will not pump into the bearing, the grease fitting is probably bad (check valve probably stuck)

7. Observe the seal of the bearing and be sure not to over-grease. Over-greasing will cause the seal to open, consequently breaking the seal. 8. After greasing, listen for any unusual sounds. There are more technical instructions for the serious user like calculating re-lubrication amounts based on bearings size, re-greasing frequencies based on bearing size, speed, conditions, etc.

Know the thickener type. If grease tube mentions polymers - these are additives that can improve things like increase base oil viscosity and increase adhesiveness and really have little to do with compatibility. The thickener type is most important. Some thickeners are not compatible with others. For example, aluminum is not compatible with lithium, polyurea is not compatible with most thickeners, barium, clay, calcium thickeners are not compatible with other types. Some of the complex mixes are compatible, but you really have to know what it is your are commingling. My advice is to conduct a compatibility test in a lab. Also, when mixing two incompatible thickeners, you may get one of two typical reactions - a soft and soupy solution where the grease does not stay put, or you can get a rock hard material in the bearing housing. This occurs when the oil leeches from the grease and all that remains is the thickener. Additionally, remember grease is comprised of three components - base oil, thickener, and additives. 85% of the three components is oil so it's important to know what type of oil is being used, mainly synthetic. Sorry to be long winded, but most places I use to go to have no idea of the differences in greases. In many cases, a salesman will try to sell you something and tell you it's compatible, but he really may not

know. Roy A. Forson, CLS Maintenance Manager Imerys Performance Minerals

To learn more about this topic, e-mail your comments to us at: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article #8: Unstable Curves Tell Me More!


This Article was initiated as a response to a question from one of our readers: Can you tell me more about unstable curves? Optimal designs of centrifugal pumps result in unstable curves (near to dead head), but in which cases is this creating a problem, and what can you do to resolve other than redesigning the pump? Walter A. Koch VERDER Group, Liquids Division, Europe Dear Walter: A stable curve is very important for a pump operation, - especially for pumps operating in parallel. The higher the energy level, and the more critical an installation is the more this could become an issue. API 610 even states that pumps that have stable head/capacity curves (continuous rise to shutoff) are preffered for all applications and are required when parallel operation is specified. When parallel operation is specified, the head rise shall be at least 10 percent of the head at rated capacity

Fig. 1 Stable (left) and Unstable (right) H-Q curves A centrifugal pump operates at the intersection point of a pump curve and a system curve. A system curve is a parabola starting from zero in case of mainly friction losses (long pipe with restrictions, such as valves, fittings, etc.), or a parallel line in case of mainly static head (pumping up to a vessel). Or, it can also be a combination of both:

Fig. 2 A pump operates at the intersection of a pump H-Q curve and a system curve If the pump curve is stable, there is always a unique point (A) an intersection of a pump curve and a system curve. If the pump curve is unstable, the region between B and F has two possibilities at either flow Qb, or Qf :

Fig. 3 Stable curve (left) has a single definition of an intersection between a pump curve and a system curve. Unstable curve (right) has two flows where a pump can operate, at the same head. Imagine a parallel operation, with two pumps piped to a common header. Suppose Pump-1 is running and Pump-2 is idle, ready to be brought online. Starting of a pump is usually done near the shut-off (valve just slightly cranked open), in order to minimize motor load. If Pump-1 is running in a funny region, say at point C (where curve is unstable), the system head is Hc, - i.e. higher then the shutoff head Hf, which is

what Pump-2 will generate at first, when it starts. Therefore Pump-2 can not open the check valve, which is held closed by the higher pressure Hc imposed by the already-running Pump-2.

Fig. 4 Two pumps in parallel pump P2 is having trouble starting Imagine next that several pumps are already running in parallel. Since they discharge to a common header, their discharge head must be the same. However, each pump may have different flows either Qc, or Qe. If a plant operator wants to increase the total flow and opens the discharge valve more, Pump-1 will increase its flow, and its head will decrease (Qcc, Hcc). The new system head Hcc will now push the Pump-2 to lower flow (Qee). Eventually, a stronger pump may completely take out the weaker pump to near, or at the shut-off head. Operator, only noticing a total increase in flow, would not even know of this happening, - while the Pump-2 unexplainably begins to vibrate, shake, and possibly fail:

Fig. 5 Pump P1 is forcing pump P2 out Not a nice thing An excellent source of reference on this, with a more detailed explanation of unstable curves, is in a book (a classic!) by A.F. Stepanoff, Centrifugal and axial Flow Pumps, John Wiley publication, 2nd Edition, page 293. There, the author also addresses the system conditions that would contribute to, or further aggravate, the situation: The mass of water must be free to oscillate, - a typical scenario in boiler feed applications There must be a member in the system which can store and give back the pressure energy or act as a spring in a water system. In a boiler feed pump cycle, the elastic steam cushion in the boiler also serves same purpose. Long piping can also do the same. Stepanoff also provides recommendations: By-passing part of the capacity to the suction supply tank. Automatic capacity governor near the boiler, with very slight throttling at the pump to stop H-Q swings. Braced piping, except for the provision for heat expansion. Avoid operation near critical point. For higher Specific Speed pumps (see separate article on Ns and Nss definitions), such as axial flow pumps, the instability happens at flows substantially closer to BEP, as compared to lower specific speed pumps (such as boiler feed). However, this instability is local, and the curve continues to rise again, after the local region of instability:

Fig. 6 High specific-speed (Ns) pumps have instability starting much closer to BEP, as compared to lower specific-speed units

As the reader has pointed out, and it is true - that the best designs, with regard to efficiency, often end up with unstable curves. Such is the nature of hydraulics! - of the pumping machinery. A better and more practical compromise is not to push the efficiency overly high at a single BEP (best efficiency) point, but to have a more balanced design, where efficiency may still be good overall, but the curve is stable. This is because, in practice, it is almost impossible to limit a pump operation to a very near region of flow. All pumps, regardless of their energy level, may experience these difficulties with curve instability. As a practical matter, small pump, however, such as inexpensive commercial units, or even small pumps for chemicals, such as ANSI, usually either do not operate in parallel too often, or, if they do, they can be started at more open valve. Since the energy level is small enough, and the duration is short, as compared to a much more powerful units (such as boiler feed pump, circulating vertical pumps, etc.), there is seldom (but could happen!) a problem with these pumps. By note to the readers we welcome any additional comments on the subject. If anyone would like to expand or add on a personal experience regarding curves instability and practical ways of how it was handled please let us know, so that we can add your contribution in the next edition. DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Pump Magazine Publications

The Essence of Equipment Failure Analysis Theory, Approach, and a Case Study
by Sourav Kumar Chatterjee

Manager, Rotating Equipment Hindustan Petroleum Corporation, Ltd. Mumbai Refinery, India July 9, 2002
Abstract Failure is incapability of an item to deliver desired level of service as specified by design/expected by user, under specified condition. A thorough analysis of Root Cause of Failure is followed by the detailed Field Case History of a seal failure of a pump at a refinery. Human factor, logistics and team assignment is analyzed, along with tracking technical aspects of a problem. Actual data for a pump operation around the failure period is related to mean time between failures (MTBF) and a follow-up monitoring plan, after problem evaluation and correction, is established. An interesting and informative case for practicing plant engineers, maintenance and operating personnel, to compare notes and learn. Analysis Analysis is a technique where a set of useful information on an event under consideration is compared with a set of design information of same areas and pertaining to same item involved in the event, - to find out the deviations, followed by a logical conclusion on cause of eventuality, using expert system which possess wide database on similar type of events or using human expertise. Failure analysis Failure analysis is an analytical technique used by professionals of all field at various functions to protect against potential problems in process & products. What is a Potential Failure? The identifiable & measurable physical condition of an item, which may be equipment / person / system, and which indicate that the functional failure is about to occur or

in the process of occurring is known as potential failure. The term potential implies strong probability of occurrence. EXAMPLES: Temperature of running equipment parts (bearing housing casing lube oil, etc) Visible leaks and wear Vibration level indicating potential bearing failure Wear particles in gearbox oil showing imminent gear failure What is a Failure Mode? It is an event, likely the cause for the condition of each failure state. In other words, it is the manner in which an item could potentially fail to meet the functional requirement, or design intent, or both, as defined and/or acceptable to the end user. Some typical failure modes: * Bent * Incorrect adjustment * Broken * Internal leak * Contaminated * Jamming What is a Failure Effect? Failure Effect indicates the result of failure, and makes us realize the following: Evidence that the failure has occurred. Safety, environmental & social consequences The way in which the production or operation or system is getting affected The physical damage caused by the failure Action to be taken to repair/revive /cure the system and arrest further deterioration Some typical failure effects: * Leakage of pumpage * Low pressure

* Low flow * No production * Erratic operation * No control * High vibration * Poor performance * Rough finish * Unstable operation * Operating parameter fluctuation * Intermittent operation * Deterioration of product quality Objectives of Failure Analysis To find out Root Cause of failure and remedial actions Recognize and evaluate the potential failure modes Higher organizational, environmental, social and human security and safety Identify actions, which could eliminate or reduce the chance of potential failure from occurring Cost control Higher productivity Documentation of the process for future reference and monitoring Core View of Failure Event In an apparent assessment, though failure event leads to losses, hazard, despair, discrimination and all similarly negative notions, - but, ironically, there are few positive features also are inherent in it if followed by Root Cause analysis.

H Loss A Z A R D

Despair

A formalized approach is of utmost necessity to carry out effective and successful failure analysis. Such concept generally comprises of five main activities. Data collection Formulation of probable cause areas Analysis Remedial measure Documentation and corrective actions Data Collection The success of a failure analysis greatly depends on data collection. Out of so much data, the technique of picking up relevant data accurately is a highly skilled job. Many times the analysis sets back as concerned personnel become at a loss to understand what data are required, and how to get it. For an equipment failure following steps may be followed: 1- Identify the equipment & component 2- Find out potential failure mode or failure effect 3- Find out designed parameters (constructional & operational)

4- Note observations on operating parameters (during failure) & constructional parameter on dismantling Formulation of Probable Causes
Type of equipment and accessories Constructional features Service condition Type of component failed Nature of failure Potential failure modes observed before failure Last maintenance details and MTBF

Probable causes 1) 2)--3)---

0---

Input Analysis

Remedial Measures Remedial measures are adopted based on area of root cause and feasibility study for implementation. Design problem Installation problem Assembly problem Mal-operation Raw material/spare part problem

Documentation and Corrective Actions Documentation is an arrangement / system to keep useful information in meaningful manner which can be retrieved easily whenever required. It is also information for concerned in relation to pertinent item or event, which is basic requirement for further development, and progress. Documentation of entire failure analysis event must be done in designated item field and in prescribed format highlighting details of event and total observations, analysis considerations, justification for selecting appropriate measures, implementation details, effect and observations after implementation, update of P&ID / Datasheet / Drawings, indicating cause, date and agency involved.

____________________________________________________ _________ Plant, site, or an enterprise modernization means continuous alteration of policies & approaches with the goal of making positive response to the needs, in terms of quality of product, its cost effectiveness, time, availability, and safety. In course of doing that, a careful study must becarried out to select appropriate measures and to identify its key aspects for successful operation. Appropriate monitoring of performance of such new systems also has immense importance. Absence of mandatory accessories for operation and monitoring can lead to trouble and hazardous situation. This case study presents a situation where an ECS seal for emission control has failed creating hazard due to improper monitoring system and supporting accessories.

1A- Equipment type: centrifugal pump, back pull out design Tag no.- 14P19 Location Cr. LPG Service C3+C4 (Propane+Butane+propylene)

1B- Mechanical parameters: Bearing type NU310/7310*2 Seal type ECS seal Double tandem Flushing Plan 02,62 water quenching Seal box venting to closed flare Cooling Plan- Plan G Lubrication type- Oil splash lubrication Lube oil grade-Turbinol-68 Suction and discharge nozzle size-6"*300 & 4"*300 MOC of Major parts- SS-410, SS-316, CS 1C- Operation parameters: Service fluid: cracked LPG Temp: 45 C Flow: 115 M3/hr Sp.gr: 0.49 head: Diff. 75M 1.0M

NPSHr PSI

Suct.pr: 210 Disch. pr: 263.5 PSI RPM: 1450 Min. flow: 28 M3/hr Vapor pressure at p.t.: 200 PSI 1D- monitoring facility: Online primary seal failure detection facility Alarm / trip connected seal failure alarm Failure detection probability: Fair Incident Failure mode and effect: both primary & secondary seal leaked (ECS)

Time of failure: 1st May 2002 @ 3 AM Detected by: Operation personnel Immediate action taken: Pump stopped and isolated immediately Safety hazard: Yes Environmental Hazard: Continuous leakage of LPG through seal. Failure Reporting time: 1st May 2002 @ 10 AM Input process conditions: Suction condition trend: O.K. Temperature trend: constant Suction flow trend: N/A Suction source level / pressure: suction drum pressure & level trend constant Output process conditions: Discharge flow trend: though the reflux flow trend found constant, heavy fluctuations observed in LPG run down flow and back pressure. Discharge Temperature trend: N/A Discharge pressure trend: N/A Observations at site: Cooling/flush Line and jacket condition: cooling water lines found through and clear. Scaling found inside stuffing box jacket. L.O Condition: Good No contamination observed. Coupling condition: Good and intact Foundation condition: OK Alignment readings on decouple: within limit Suction and discharge piping alignment: no piping stress Piping Foundation condition: in order ECS Seal system: flare vent line found plugged Observations: Bearing condition: bearings found good and intact, no radial and axial play observed. Bearing housing condition: OK

Seal parts condition: heavy pitting on seal ring mating face. Seal ring packing ( "O" ring) totally burnt. Heat marks on Insert mounting burnt and damaged. Rotary unit springs found broken in pieces. Dust of carbon found around seal parts. Observations on secondary seal: wave springs broken, bellows found punctured. Rotary face and packing good and intact. Heat marks on shaft at sleeve sitting portion. Shaft condition/runout: OK, runout 0.001" Impeller / lock nut condition : lock nut intact, impeller found cracked at back shroud sleeve/ bush Clearances: wear marks on sleeve at steam purge bush position Wearing conditions & clearances: rubbing marks on both wear rings. Clearances found: 0. 045" and 0.050" as compared with designed 0.026" & 0.030" front & back respectively (suction and discharge) Condition of other related parts: coupling teethes well. Throat bush clearance also found increased by 0.015 MTBF: 12 months Last PM & observations: 11th April, 2001 BCW lines were clear, coupling condition was good, bearing good, foundation bolts OK, alignment was off realignment was carried out, coupling run out OK. Last failure details and cause: pump was removed for seal leak on 14/03/2001. Subsequently single seal was replaced with ECS seal . Last overhauling details with activities: Bearings were changed, ECS seal was installed. Parts used from OEM/local: OEM Vibration trend since last O/H:

Starvation (loss of flow?) Bend shaft Bearing failure Misalignment Loose rotor assembly Sealing system problem Observations and conclusions were based on type of equipment and accessories, constructional features, service condition, type of component failed, nature of

failure. Potential failure modes observed before failure are depicted in the chart:

Detail Analysis and Discussion The heat mark on seal parts, sleeve and fatigue failure of wave spring bellow of 2nd seal and spring of primary seal, eventually reveal the parts were exposed to high temperature and high stress causing catastrophic failure. Moreover, the alarm on failure didn't activate which is major flaw in ECS seal system and calls for immediate rectification. It may be noted that this seal was installed during March 2001 and the vapor recovery line has been connected to flare system only on April 2002. During the operation of seal this was kept plugged, as LPG is prohibited item or releasing to atmosphere. It is evident from observations of failed parts that primary seal failed first which could not be noticed, as alarm system didnt work. The seal kept on running on ECS seal and only on failure it both failures got exposed leading to hazardous situation. Failure of Primary Seal The flushing plan 02.62 (water quench) for this service, always has a tendency towards getting inadequate seal flush. This is because the pump design, which has back wearing and throat bush restriction to stuffing box along with impeller balancing holes. Due to this design, the stuffing box pressure always equals to suction pressure, which is very close to vapor pressure at process temperature. Hence rise in temperature at seal box can create vaporization at seal box and faces leading to loss of seal face lubrication. More dead end vapor recovery system also didn't allow the vapors at primary seal face and got accumulated at ECS seal box, pressurizing ECS seal box and increasing face loading on ECS seal. After some time the heat generated due to seal friction would add more heat to entrapped vapor causing the rise of pressure due to constant volume. This enhanced pressure will act on secondary seal box at O.D and on inner diameter of primary seal insert squeezing off any possible lubrication film, which was already constrained due to type of flushing plan.

Thus the compression units were subjected to abnormal stress due to increased pressure along with high heat due to lubrication less rubbing of seal faces. In this case the primary seal leak took place due to inductee seal flush (evident from heat mark and carbon dust) followed by reversed pressure, causing damage of o-rings and compression of unit springs. Pitting on the seal face appears to be due to blistering as a result of heat concentration The hairline crack on impeller surface across the radius is also due to corrosion fatigue as it was subjected to cyclic stress due to flow variation within corrosive environment as the H2S, which is present in LPG (15000 PPM). Failure of ECS seal This failure was the result of high load on wave spring due to vapor concentration at seal box and rapid wear due to high face loading and lack of lubrication. Actually this seal face has less contact area so that heat generation be less and designed for operating under minimum box pressure. Once first seal is failed, this seal provided service for short period, allowing planned (although urgent) shutdown for seal repair. The wearing clearances increased due to temporary rotor bow at impeller end while operating under fluctuating load condition away from BEP. Scale formation in stuffing box jacket further caused poor cooling effect and heat dissipation. Calculation of heat generation at seal faces: Pressure-velocity facto (PV) Heat Generation at seal: Q=C1 x PV x f(Ao), B.T.H/Watt b= seal balancing ratio, 0.7 K=Pr. gradient factor, 0.3 for light liquid Psp= spring pressure = 0.45 bar Vm=velocity at mean diameter 3.14x65x3000/1000 x 60 = 10.5 m/sec

f= coefficient of friction, 0.07 for C/TC Combination Ao=Seal face area of seal ring = .001 sq. m C1= 1 for SI unit PV = {(12-0.45) x (0.7-.0.3) + 0.45} x 10.5 = 53.025 bar m/sec Q= 1x 53.025 x0.07 x0.001 =0.037 watt/sec=0,037/4.18=0.009 cal/sec * Or Q= 0.009 x60 x60=31 cal/hr * Cal= watt/J (J=4.18 Jules/sec) This undissipated heat will cause rise in temperature of LPG vapor at constant volume and the rise per hour could be calculated by using gas law: P1 x V1 / T1 = P2 x V2 / T2 Root Cause of Failure The improper flushing plan and lack of vapor escape feature is the Root Cause of failure of primary seal. The non-function of alarm system and absence of vapor recovery connection are the root causes for ECS seal failure. The lack of cooling due to jacket scale also a cause to accelerate the failure.

1. Seal flush system modification to API Plan 11 that will maintain higher stuffing box pressure and enough flush. 2. De-scaling of stuffing box jacket and thorough inspection during preventive maintenance to be carried out. 3. The diff. temperature of cooling water to be monitored for effective heat dissipation. 4. The vapor recovery line to be connected properly to flare header.

5. Rectification of alarm annunciation system for seal failure. Timing Schedule and Team Assignment Activity no 1,2,4 - by maintenance Item 3 - by operation / PAD Item 1 - in consultation with seal manufacturer during next available opportunity Item 5 by Instrument section Remaining items to be implemented immediately Follow-up: Six months observation Document and update of records and history log to be maintained after corrective measures are implemented and continued during the following satisfactory operation of period of one year. Readers Feedback, Questions, Discussion and the Author Comments: There are several items that are a bit unclear, and this maybe due to nomenclature usage. The author assumes the reader knows what a double/tandem seal is If double tandem, then why a disaster bushing (plan 62)? (Here again the nomenclature leads us to some confusion that could have been enhanced by including a simple schematic of the seal The cracking of the impeller suction end ring is not covered sufficiently to dismiss total doubts of a failure initiation there. Did this cause the failure of the throat bushing, was there one? And consequently - the loss of pressure in the seal cavity? A great understanding of a seal design is shown in the analysis by Mr. Sourav Kumar Chatterjee. I am curious of what was the cross-sectional face width of the seal? In refinery applications for LPG one usually uses 0.125

inch, as a rule of thumb, keeping the balance ratio intact in the seal as manufactured (in this case 70%). Face cross-section greater than 0.125 in., in extreme cases with stuffing box pressures 150 to 200 lbs, can generate sufficient face-to-face loads to turn the seals on the sleeve/shaft and begin machining themselves loose. This can take place in a shorter time frame than the dead ended box can affect the seal operations. Would appreciate if Mr. Chatterjee could go over some of these to clear up

Author Response:
I am thankful for valuable comments. The seal is Emission Containment Seal (ECS) which works on nearly zero (particularly if N2 buffer provision is made) emission. The secondary ECS seal runs dry due to minimized contact area and a special grade material (C/SC) for face combination.

To flare through NRV

To minimize seal box pressure (the vapor from primary seal to be vented to flare) the general arrangement is as follows:

The original seal for this pump was a single seal and having plan 02.62. The 62 is a quenching provision and for LPG service water quenching is given to avoid ice formation in seal area due throttling expansion at close clearance areas when pump remains in stand-by condition. Initially, conversion to ECS seal retained the same plan 02.62 due to following reasons: The same plan had been working satisfactorily for single seal since commissioning in 1994 In corporation of plan 11 needs major modification in stuffing box, which is time consuming affair. Unfortunately, the detail analysis of the effectiveness of 02.62 for ECS primary seal had not been done to pinpoint the differences between the earlier single seal and ECS seal. The cracks on the impeller are due to corrosion fatigue. Similar cracks were noticed and on micro-structure examination it was determined they had developed because of degradation of bonding due to chemical corrosion and subjected to variable pressure, as there is frequent pressure variation of rundown header. For other pump the impeller material was changed to CF8M from 410ss and for this pump also same modification is done.

Hairline cracks on impeller face:

D1 D2 D3

Se al Box area side Atmospheric

Contact Area = 4/x D1 2 Exposed area = 4/x D1 2

- 2D x 4/2 - 3D x 4/2

Closing force = Exposed area x Liquid pressure at stuffing box Opening Force= contact area x mean face pressure due to hydrodynamic force Normally, the closing force is higher than opening force so as to keep face contact stable during adverse operating condition like cavitations, high vibration etc. Balancing of Seal faces= Opening area/ exposed area. Hence for same seal size an increase in face width will cause increase in opening force-making seal unstable in adverse operating condition. If for same seal size the seal face width is too small also the unbalanced closing force will squeeze the film thus loss of lubrication will take place leading to face damage. Seal face width is selected based on pressure-velocity (PV) factor, service liquid, and size of the seal. The stick lip condition, when rotary face tends to transmit torque to stationary face, can take place, if there is too much pre-load spring compression applied, or in case there exists congealing substances (VTB, LSHS etc.) due to improper heating or absence of purging steam) between two faces.

Distortion of carbon face can take place due to heat or impurities causing higher frictional force or higher frictional bonding with mating face. Increase in throat bush clearance is not very high. Moreover, as the impeller is having balancing holes, the in throat bush area will not affect the seal box pressure remarkably as it is already close to suction pressure. Only the circulation would increase and the liquid film support to the shaft would be absent, leading to lowering of critical speed nominally and the rubbing of wearing surfaces could also take place. Hope this helps to answer questions mentioned by the readers in the Editorial feedback. I sincerely thank the Pump Magazine publisher and the readers for their comments. Sourav Kumar Chatterjee Manager Rotary Equipment HPCL Mumbai India
We received a comment on this article form our reader from Spain:
Dear Sir I enjoyed reading the two cases Mr. Chatterjee presented on mechanical seal troubles and solutions for a pump for the reduced crude at the bottom of a column at a Primary Distiller Unit. I work in a refinery on a problem of continuous failures of the mechanical seal. This mechanical seal began to fail shortly after the installation in 2003. I think there are several factors: one related to the cavitation due to high temperature of reduced crude (624 C measured by a temperature transmitter) and low suction pressure (8 psig measured by a pressure transmitter). The mechanical seals for the petroleum industry are governed by API 682, and the supplier used Plan I with Plan 11 and Plan 53 with a barrier fluid cooled in a convective tank (not forced). We experienced the leakage when using Plan 54 (with a forced barrier fluid through a heat exchanger) after a short time. The current mechanical seal have bellows in a dual tandem pressurized arrangement with Plan 54 with a synthetic oil (Royal Purple Barrier Fluid GT 910) and with a recently installed steam quench extracted from a saturated steam line and demoisturized through a mechanical separator with a steam trap at the bottom of the separator, although I noticed that the condensate is leaving the mechanical seal directly to the drain (sewage) without a steam trap. Is trap needed to avoid condensate on the mechanical seal chamber? I am concerned if the steam entering mechanical seal chamber is dry, as wet steam would cause problems of coking at seal bellows. How can I make sure steam to the seal chamber is dry?

Considering low NPSH due to vapor pressure of reduced crude at the operating temperature of 624 C, operators said that the cavitation of pump occurs mainly when starting the pump. Do you know what is vapor pressure of a typical reduced crude at the temperature of 624 C ? I think it varies according to the crude composition, but an approximate value would help.. I also think that a probable problem could be the diameter of suction pump of 10 inches and has not been increased when we installed a new pump from the old pump of lower capacity. Thus a new pump has more friction loss leading to less suction pressure and risk of cavitation. The old pump of less capacity in gpm did not have problems with mechanical seal failures. Also, operating at low flow, the valve at the outlet of pump is throttled to less than 50 % of nominal capacity. Vibration of pump after major repair of mechanical seals and bearings went down from 6.4 to 2.5 mm/sec and acceleration went down from 12.4 to 0.74 G. Are these ok? There are three pumps in parallel pumping reduce crude from bottom of column. Two of them operate together, and have no seal problems. The third pump operates alone at reduced capacity has mechanical seal failures and control valve pinched. Initially, the pump had the same problem experienced in Case 1 you presented with the barrier fluid circulating only when the pump was in operation and a failure occurred when there was no barrier fluid at the fluid reservoir and coke was found under the bellows of the secondary seal, which is why the seals faces opened. Plan 54 users synthetic oil (Royal Purple Barrier Fluid GT 910) as a barrier fluid, which the supplier claims to be compatible with the reduced crude. I would appreciate any comments about the probable causes of our continuous mechanical seal failures, and any suggestions you may have to solve the problem. Best regards Manuel Luque Casanave August 14, 2006 Dear Mr. Casanave, - thank you for your comments, and it is good to see that our readers keep in touch with the publications long after they first appear at Pump Magazine. We will let Mr. Chaterjee know your input, as well as are asking our readers to provide any additional comments, thoughts, and ideas. In my view, suction pressure was an issue, as you noted, and maintaining proper level of buffer fluid was an issue. Perhaps an automated level detection in the buffer tank would keep it from running dry. Regarding vapor pressure of oil, it may vary from one site to another, depending greatly on specific composition. I would recommitment you involve your local laboratory to test the oil, and determine its characteristics, including vapor pressure. Thanks again, for your valuable contribution and an interesting case. Dr. Lev Nelik, P.E. Editor Pump Magazine

DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article #10: PUMP-OUT VANES


This article came out in response to a question from our reader: Could you explain how an ANSI pump impeller clearance (both open vane impeller with back pump out vane and reverse vane impeller with balancing holes) setting affect pump performance? What are the advantages and disadvantages of these impellers?

The main reason to use pump-out vanes (POV) is to change the pump axial hydraulic thrust. Take a look at the picture:

The rotation of the impeller results in dragging into rotation of the fluid in the gap between the impeller and casing walls. This is similar to a motion of a teaspoon in a cup, or a disk spinning inside containment. The resulting motion is referred to as forced vortex. Such vortex setsin in the front and back gaps between the casing walls and the impeller front (shown on the right) and a back hub (shown on the left). The

pressure distribution in the gap is parabolic higher at the impeller OD, and gradually reducing towards the shaft centerline. Pressure time area is force which is exerted on the impeller from both sides (FR and FL). The difference between these forces is hydraulic axial thrust, which is ultimately transmitted to the bearings, and thus is desirable to be small. This pressure at a given position in a gap depends on the radius, rotational speed of the fluid (divided by the impeller rotational speed), and the gap. Curve (1) shows the static pressure distribution behind the impeller hub without the pump-out vanes. As we know from basic hydraulics, - the faster the fluid moves, the lower the static pressure is. So, if we could make the fluid in the gap to rotate faster, the static pressure would be reduced, and the force FL to become smaller closer, and hopeful equal to the FR to reduce, or eliminate the net thrust. Without the POV, the fluid in the gap is rotated only by the friction (drag) of the impeller hub wall. It rotates with the same speed as the impeller right at the impeller wall surface, but (so called no-slip-condition) is not rotating at the casing wall, since that wall is stationary. Thus, on the average, the bulk of the fluid in the gap is spinning at the angular velocity equal to half of the angular velocity of the impeller. But, if we add the POV, the fluid becomes trapped within the POV space, and thus rotates with the same speed as the impeller i.e. double of what the fluid does in the absence of the POV. This, of course, assuming the gap between the POV and the casing wall is (theoretically) zero (x=0). The number of the POVs actually does not have to be equal to the number of the impeller main blades, but often is, due to casting production process, for simplicity. Obviously, the gap x can not be zero, so the actual reduction of the pressure profile (curve marked as 2, is less, depending on the gap x. And, if this gap becomes too large, the effect of the POV diminishes, and eventually disappears. It turns out that, the POV are most effective at x=0, and become completely non-effective when x=t, i.e. when the liquid gap (x) becomes equal to theheight of the pump-out vanes (t). (Papers are written on this subject, such as Zankers, explaining whys and whats, and those of you who have a couple of sleepless nights to study them let us know, and we will get you the material).

The balancing holes are also used to reduce pressure distribution, i.e. a similar idea as the pump-out vanes. In fact, this is why they are called balancing. To be effective, the impeller must have a tight clearance between it and a casing (not shown on a picture above), to separate the higher pressure zone, from a lower pressure zone. The balancing holes thus connect the back of the impeller with the inlet area, where pressure is low (close to suction). Some leakage will occur, reducing the efficiency. If there is no clearance, such as shown above, the leakage will be greater, reducing the efficiency even more. Another reason for the POV is to reduce the pressure at the mechanical seal area. The effect of these vanes can be very strong, and, sometimes even result in creating vacuum, and boiling out of the liquid. This could be trouble, as mechanical seals do not like to operate in a vapor environment. Regarding the performance there is price to pay for the thrust balance, as the additional power required to spin the liquid faster takes away the efficiency. This is why, the higher energy pumps, such as API, or boiler feed, rarely have the pump-out vanes, while the ANSI pumps, which are relatively lower horsepower units, have these. Note the picture above shows an open impeller. The front gap, between the impeller and the casing, must be tight (typically 0.015 0.030, depending on a pump size). The challenge is thus to maintain both front and back gaps small from the efficiency and thrust standpoint. Close impellers solve this problem, but at the expense of reduced ability to handle stringy, such as fibrous, solids. And finally, regarding the reverse vane impeller. The impeller shown above is a standard design, such as, for example, manufactured by Goulds, model 3196. To adjust the front clearance, the rotor must be pushed against the casing, and then backed-off by the amount of design clearance. It is sometimes desirable to keep the casing piped-up, as it is somewhat a chore to re-pipe it. This requires re-setting of the front gap, during maintenance, on site and, if it rains or snows you freeze and catch a cold! If the impeller is turned around, such that the clearance gap is between the impeller and the stuffing box (or a sealing chamber), then

this clearance can be set at the shop, and the rotor can then be simply brought to a casing and bolted on quickly. Example of such design is former Durco Mark III design. As you can see, each design option has advantages, and disadvantages, and the ultimate decision remains with a specific application requirements. We welcome your comments to: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

ARTICLE #13: PARALLEL OPERATION OF CENTRIFUGAL PUMPS and the Importance of Curve Stability
Some time ago, one of our readers asked a question about a parallel operation of pumps. This prompted a discussion which was posted under Q&A Section, Question #25. Later on, another reader inquired about the unstable curves and their effect on pumps parallel operation, which resulted in another Pump Magazine Article (#8). This generated interest from our friends at Pump-Flo company, who asked our permission to reprint the article on their web site. Pump-Flo is a maker of pump hydraulic selection curves program, for the pumps made by several pump manufacturers. Their readers also liked the topic on parallel operation, and more discussion followed, which we posted in Q&A Section, under Question #48. We have been our readers, convinced us (#13), which receiving more questions and comments from and the latest one (which we show below) of a need to respond with a formal article is what you are reading.

Question:

A chemical pump supplier engineer was asked why two of his pumps in parallel appeared to be unstable at certain points on the curve, notwithstanding that the curve was an inherently stable curve with continuous rising gradient to shut off. Both pumps are identical, same diameter impeller, end suction type. This instability was experienced in a mining plant on a cooling tower system with acidic water at 50 deg C. The instability could be corrected by altering the settings of the discharge gate valve, i.e. changing the head and flow on the pump. The suction head was more than adequate at 10 psig. The NPSHr was less than 0 psig. His explanation was linked to the suction manifold, which directed liquid in a straight feed to one pump [A] and liquid via 2 x 90 degree elbows to the second pump [B]. There was a very short straight portion of pipe between the 2nd elbow and the suction flange on pump B, approximately 3-4 pipe diameters [pipe dia 10"] This arrangement it was argued, caused the suction head on the 2nd pump to fluctuate rapidly, both increasing and decreasing the suction head to pump B beyond the smooth flow suction head experienced by pump A. This fluctuation or vortexing could be substantial enough to cause the 2 pumps to fight themselves and result in the instability experienced. While the argument is plausible and credible, is this a recognized phenomenon in the industry that can and has occurred on a number of other occasions? The solution, if this is the cause of the instability, would be of course to change the manifold to more smoothly direct flow to the 2nd pump, eliminating the 90 degree elbows. Hugh Lloyd National Process Equipment, Canada Answer: Dear Hugh, As we understand it, this is a setup you have:

V3

V1 EL2 EL1

V2

Note: elbows EL1 and EL2 are shown within the same plane for sketch simplicity, but in reality most likely are in different planes. The stable curve the supplier engineer pointed to is probably a manufacturer curve from the catalog, which probably looks something like this:

Head

Flow The H-Q curve is continuously rising, i.e. deemed stable. Pump Magazine Article #8 talked about unstable (drooping) curves:

H Q As was explained in Article #8, unstable curves cause problem for pumps operating in parallel, which would explain Hughs problem. However, what is puzzling is why the two pumps, having stable curves, behave as if their curves are unstable? However, the pumps are not identical. Clearly, the suction approach to Pump B is different from Pump A. While it is generally known that sharp turns, multiple elbows, short pipes, and other obstructions are not a good thing for pumps, and numerous Case Histories have been published to illustrate and warn the pump users not to do that (Hydraulic Institute even issued guidelines on suction approach dimensions, sump configurations, etc. all to help avoid problems) questions nevertheless come up each time another problematic parallel operation surfaces. There are many reasons why tipsy-turvey suction could cause instability. Lets take a look how an apparently stable H-Q curve, even as tested by the manufacturer, can become unstable when a pump ends up with a curvey suction. There is no other way to understand this without

having to delve into the heart of the impeller hydraulics. The head generated by the impeller is the difference between the velocity angular momentum between the impeller inlet and exit:

H = [ (U x Vtheta)OD - (U x Vtheta)eye ] / g

Impeller OD

Impelle r Eye

The g is gravitational constant, U is peripheral rotational velocity, subscript theta means peripheral projection of the absolute flow velocity vector, and H is impeller head. If fluid velocities are taken in feet per second, and g=32.17 ft/sec2, then head H comes out in feet. There is also hydraulic efficiency involved, but we skip that for simplicity. Hydraulic designers refer to this in terms of velocity triangles, which essentially is a set of fluid velocity vectors are the inlet and the exit of the impeller:

W1 V1 U1 Alpha1 = 90 degrees U2 Alpha2 = 20 degrees

V2

W2

The projection of the absolute velocity V onto peripheral velocity U is called Vtheta . For most end-suction pumps, the absolute flow angle at

the inlet is 90 degrees, which is also referred to as no-prewhirl condition. Thus the projection of the V1 velocity onto U1 is zero, and thus the product (U x Vtheta)eye = 0 For example, if a Pump A develops 100 feet of head at the BEP point of 1000 gpm, the impeller exit component contributes all 100 feet if there is no pre-whirl, and there is no effect of the impeller inlet no addition and no subtraction. At lower flow, the exit velocity triangle changes (Vtheta)OD component gets larger, and thus more head is generated. At the shut-off, there is no through-flow, and the (Vtheta)OD essentially overlays directly over the vector U2. Thus, at zero flow (shut valve), the head is the highest, - for typical end-suction pumps, such as ANSI, for example, the rise to shut-off is approximately 110 - 130%. And in order for the H-Q curve to be stable this rise must be continuous in order to avoid instabilities if operated in parallel. Double-suction pumps, for example, are different. Their casing forms a toilet-bowl-shaped suction, which just as a toilet! creates a prewhirl:

U1 Alpha1 = 45-60 degrees Now the product (U x Vtheta)eye is not zero! If this product amounts to, say, 30 feet, the impeller would generate less head: 100-30 = 70 feet. However, hydraulic designers would compensate for that by altering the impeller exit geometry width, blade angles, etc. so that the exit portion would produce extra head, say 130 feet and the net result would remain the same 130-30 = 100 feet. Therefore, by changing impeller geometry, it possible to change the curve shape, including the rise to shut-off. Likewise, by changing the pre-rotation (pre-whirl), it is also possible to change the rise-to-shut-off. In fact, the change could go either way the head at the BEP can either increase, or decrease depending whether the pre-rotation is positive (as shown above), or negative. The shut-off head, however, does not depend on prerotation component.

When a designer accounts for a known pre-rotation in the casing (such as for double suction pumps example), he or she does it in a way to preserve the flow dynamics, and maintain smooth flow and good efficiency. The designer has not control, however, if the pre-rotation is created artificially, such as bad suction piping, multiple elbows, and so on, - including the direction of the prerotation. If the deck gets stacked the wrong way, the flow in the suction piping may end up with negative rotation, - and double-turning elbows are known to be nasty on flow patterns.

The resulting velocity inlet triangle may thus look like this:

Alpha1 = 140 degrees, i.e. negative Vtheta results. This could create a negative inlet term - for example, (U x Vtheta)eye = - 40 feet (minus). Since the exit is fixed, the net head would become 100 (40) = +140 feet, - which is obviously greater then the shut off head, - i.e. an unstable curve got created:

Hpeak=140 Hso= 120 Q As Hugh rightly noted, the solution, if this is the cause of the instability, would indeed be to change the manifold to a more conventional and appropriate inlet piping configuration. Even though there could always be

other reasons involved, a good piping is not a guarantee, but a beginning to a trouble-free pump operation. Next, lets look at the parallel operation in more detail, starting with stable curves. Situation 1: Single Pump

The pump must generate pressure (head) to overcome hydraulic losses in the system. These losses consist of losses in a valves, bends, elbows, heat exchanger and pipe friction. From basic hydraulics, a friction loss is proportional to the square of flow: HLOSS = k x Q2 The summation of individual losses is equal to a total system hydraulic loss: HLOSS = KVALVE X Q2 + KELBOW X Q2 + KHE x Q2 + = (KVALVE + KELBOW + KHE +) x Q2 = kSYSTEM x Q2 We were able to do the math above and add the individual hydraulic coefficient due to the fact that the flow is the same throughout the system. We could not do that so easily if the system had branches, with flows branching out all over. This is analogous to the electric circuit, where the same current flowing from one component to the next result in different voltage drops across the individual resistors. In hydraulics, the pressure drop due to fluid flow is similar to the voltage drop due to electric current.

Once the kSYSTEM is calculated, a system curve can be plotted: hLOSS

Flow If the pump curve and a system curve are plotted together, an intersection thus determines the pump operating point, because it belongs to both curves: Head

Flow (Q) Situation 2: (2) pumps in parallel, but with a common resistance element (valve)

Flows Q1 and Q2 combine, and that is what flows through the valve: Qvalve = Q1+Q2 The valve curve is similar parabola as shown in the previous example: h = k x Qvalve2 = k x (Q1+Q2)2 Mathematically, a pump curve is a Head as a function of Flow. Or, it can be also expressed as Flow as a function of Head, i.e.: Q1 = f(H1), and Q2 = f(H2) In general, the pumps do not have to be identical. Then, Qtotal = Q1+Q2 = f(H1) + f(H2) Since Qtotal = Qvalve, the combined pump curve and a valve (i.e. system) curves can be plotted and the intersection would determine the head (pressure) where the pump/system will operate:

Head 1 pump (Qtotal = Q1 x 2) 500 600 Flow, gpm 2 pumps

(for simplicity both pumps here were assumed to be identical)

Note an interesting thing: when only one pump was operating (point 500 gpm), the flow was 500 gpm. But when the second pump also came on line, they together, did not produce 500x2 = 1000 gpm! Instead, only 600 gpm flows through the system! Clearly, the answer to that is in the shape of the system resistance curve. However, if the pumps were pumping against mainly a static head (such as pressurized tank) then the flows would be additive! Why? Because the static head curve is not a parabola, but a straight line, independent of flow, and thus parallel to the flow axis:

Head 100 psi (system)

500

1000 Flow, gpm

500 gpm 500 gpm

1000 gpm 100 psi

Note: in general, a combination of friction as well as static heads may be present, although usually either one or the other scenario is more predominant. Situation 3: (2) pumps in parallel, but NOT with common resistance Now, that is getting interesting. In the previous example, the change of the valve setting (changing valve k coefficient) affects both pumps the same way. But what if we have a system like this:

Common valve, V0 Valve V2 Lets assume, for simplicity, two identical pumps, and the valves are identical also. Assume also that each valve is throttled the same percentage (i.e. having the same coefficient k):

Head,ft (Hso= 120 ft) CURVE

PUMP CURVE HBEP = 100 ft

hsys VALVE

Flow, gpm Flow, gpm 250 300 500 (BEP) 400 Digitize several points form the valve curve into a Table below to make it easier to work with, as we go:
Flow, gpm hsys, ft 100 200 300 400 4 16 36 64 500 100 600 144

We know the following: Assume there are no other losses except for the (2) valves. In other words, all other losses have been combined into the (2) equivalent valves V0 and V2 Valve V0 sees the combined flows: from Pump 1 plus Pump 2 The head generated by the Pump 1 is equal to the pressure drop (head loss) across the Valve V0, because of the pressure continuity the nodal point at the exit of the Pump 1 is the same as the entry point into the Valve V0 Valve V2 sees flow from Pump 2 only System must be in balance, to satisfy individual pump H-Q curves, and individual valve h-Q curves. The solution is iterative: a) Guess at the flow from Pump 1, Q1=250 gpm. b) Per H-Q curve of Pump 1, head H1=118 feet c) Head loss across valve V0 is equal to head generated by Pump 1, i.e. valve h1=H1=118 ft d) From the valve V0 curve, Q0 = 543 (we actually cheated a bit here got the equation for the parabola, h0 = Q02/2500, and then just plugged-in the numbers)

e) Total pumps flow is equal to what is flowing through valve V0, i.e. Qtotal=Q0=543 f) Flow through Pump 2 is the difference: Q2 = Q0 Q1 = 543-250 = 293 gpm g) From the Pump 2 curve (we used identical pumps in this case), H2 = 116 ft (approximately, as it is more difficult to derive the polynomial equation for the pumps. For the valve it is easy - parabola) h) A pressure drop across the valve V2 is the difference between the pressures at its exit and inlet. The exit pressure is equal to what Pump 1 generated, and the inlet is what Pump 2 generated, i.e. h2 = 116-118 = 2 ft. This is impossible, i.e. our first guessed flow from Pump 1 was wrong. Lets re-guess and go around again: a) New guess, Q1=300 gpm b) H1 = 115 ft c) h0 = H1 = 115 ft d) Q0 = 536 gpm e) Qtotal = Q0 = 536 gpm f) Q2 = Q0 Q1 = 536 300 = 236 gpm g) H2 = 119 ft from pump curve h) h2 = H2 H1 = 119-115 = 4 ft i) From valve curve (or interpolating from the Table), Qvalve2=100 gpm j) Q2 = Qvalve2 = 100 gpm (from the flow continuity what flows to Valve 2 has to come out from Pump 2). This is not equal to Pump 2 flow as derived at step (f), so we need to guess again. (But we are getting closer). a) Next guess, Q1 = 400 gpm b) H1 = 113 ft c) h0 = H1 = 113 ft d) Q0 = 532 gpm e) Qtotal = Q0 = 532 gpm f) Q2 = Q0 Q1 = 532-400 = 132 gpm g) H2 = 119.5 ft from pump curve h) h2 = H2 H1 = 119.5-113 = 6.5 ft i) Qvalve2 = 127 gpm j) Q2 = Qvalve2 = 127 gpm

We now reached a reasonable convergence. Thus the final answer is: Pump 1: flow = 400 gpm, head = 113 ft Pump 2: flow = 132 gpm, head = 119.5 ft The above solution has a numeric (analytical) counterpart, and could be solved by writing a set of simultaneous equations and solving them: h2 = k2 x Q22 h0 = k0 x (Q1+Q2)2 we could even have different valves or valve settings (different k0 and k2) H1 = f(Q1) or Q1 = f(H1) H2 = f(Q2) or Q2 = f(H2) pumps also could be different H1 = h0 h2 = H1 H2 (6) equations and (6) unknowns: Q1, Q2, H1, H2, h1, h2 Clearly, even a simple case of just two pumps and two valves can be a rather time consuming task to solve. Clearly, for a real pumping system, with many pumps and components, a manual method is impractical. This is when computers come handy, asRay Hardee P.E., Engineered Software, Inc, explained in the discussion for Question #48. Obviously a more refined math and methods can be employed but lets leave that for the programmers and mathematicians! For now, and from the practical point of view, - if we have a piping network program like that and a clear manual of how to use it a solution of even a complex system would become an easy job. One of such companies is Sunrise Systems, and you can get more information about them and their product via their web site at www.sunrise-sys.com Note that we have only touched on systems with stable curves. What happens if the pump curves are not stable? Well, mathematically, we get multiple solutions. In other words, just as was explained in Article #8, a pump would be technically able to operate at two flows for the same given head. How do programs like the ones mentioned solve this?! That,

however, is another issue. In the meantime stay away from the unstable curves, and, for that matter from bad piping! Keep on pumping! Dr. Lev Nelik, P.E., Apics Pumping Machinery Note: if you would like to learn more about pumps, systems, theory, as well as hands-on pump reassembly work come to our next Pump School learn the details on the schedule and the agenda in Section PUMP SCHOOL of the PUMP MAGAZINE. We welcome your comments to: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

The following article by Bob Hart, a former Principal Consultant at DuPont Company, is presented here in response to a question posed by a reader of Pump Magazine, on the important subject of Net Positive Suction Head: required versus available. We have posted the readers question, and Bob Harts response, followed by this insightful technical article on the subject, for the benefits of the pump users. Additional related subjects can be found within the Pump Magazine site, via a Search function from the main site entry. First, a question by a pump user: Dear Pump Magazine, How many times NPSH-A can be higher than NPSH-R? As per CEP article (March 1993 page-81), and Chemical Processing (December 1994, page 49), a higher NPSH-Available is not good for the pump (?!). Is it true? We have two situations in the company I work:

(a) pump operates (Ammonia) with 18.6 ksc(g) suction pressure, temperature 37.8C. It runs continuously without problems for over 25 years. The pump data sheet reads NPSHA = 20, and NPSHR = 11 feet. (b) another pump operates with 8.46 Ksc suction pressure, temperature 65.8C(weak ammonium carbonate solution). This pump is also running without any problems for about 10 years. The pump data sheet reads NPSHA is only 4.9 feet, though the operating pressure is 8.46 Ksc. If I am correct, the Head = P/Density = 8.46x14.22*/1.09/62.4 = 259 feet (1.09 sg) Could you comment? Thanks in advance, Ramasami Anbazhagan, a Pump User Bob Hart comments: For the reader to fully understand the following comments, it must be understood the term NPSH-Required, as defined by the pump industry, places the pump in a degree of cavitation that is an unacceptable operating condition. The more descriptive and appropriate term for this very precise hydraulic condition that is now defined as the NPSHRequired of a pump would be to define it as the NPSH-Instability (NPSHI) characteristic of the pump. All pumps will experience a loss of head and many will become totally unstable (surge) when inlet hydraulic conditions (NPSH-Available) approaches the NPSH-Required value. The CEP 1993 and Chemical Processing 1994 articles referenced are not readily available to this writer and hence the exact basis for the remarks may not be addressed in the following comments. However, there are various articles that have been written in the past few years that describe the cavitation phenomena that point out, quite correctly, the rate of deterioration of the surface on which the vapor bubbles implodes, is greater when the bubbles are small than when they are large. The size of the vapor bubbles that will exist within a pump is dependent on the margin between the terms NPSH-Available and NPSHRequired. Extensive testing has been done to identify these phenomena on specific pumps.

When developing a pumping system, there are numerous decisions that must be made, frequently without the benefit of knowing the exact pump that will be applied to the system. These decisions frequently use the best information and judgment available to the individual developing the system at the time with the end result not having the advantage of laboratory test precision. The remarks in the inquiry would lead me to believe there may still be a misunderstanding about the definition of NPSH-A as it relates to the Suction Pressure of a pump. If the liquid being pumped has a very high Vapor Pressure, the pump can have a high Suction Pressure but a very low NPSH-Available. The NPSH-A may be less than the Submergence Level of the pump when considering the friction losses in the piping system. In response to the specific question: Is there a practical upper limit to the Ratio of NPSH-A/NPSH-R?, I would like to offer the following: My industrial application experience has not identified maintenance, operating or reliability problems that have limited plant production that could be directly attributed to excessive NPSH-Available compared to the NPSH-Required. On the other hand, inadequate NPSH margin is frequently a major contributing factor to these three parameters (maintenance, operation, reliability) that limits plant production. There are a number of practical reasons that an excessive NPSH margin is normally not applied to pump installations. Four major factors are: Installation Costs Increase with larger NPSH Margins The Actual NPSH-Available is typically less than calculated due to Fluid and Piping conditions. Manufacturers NPSH-R Test Conditions are for new pumps Hydraulically stable fluids and piping systems seldom encountered in actual installations Actual Operating Conditions will typically vary significantly from the single point Data Sheet information. Robert J. Hart, P.E. April 15, 2003

*************************************************************************************** *** Technical Article CAVITATION A DANGEROUS ENEMY TO PUMPS By Robert J. Hart, P.E. Robert J. Hart Enterprises, LLC The term Net Positive Suction Head Required applies to all pumps. It is a term that has been and continues to be misunderstood by many of those selecting pumps and designing the piping systems in which they are installed. As a result, a number of pumps operate in varying degrees of cavitation, an underlying cause of the high maintenance costs often associated with them. Cavitation, at its worst, sounds like loose gravel passing through the pump. Damage to seals, bearings and impellers will usually be experienced well before the noise of cavitation can be detected by the human ear. The values of NPSH-Required, published in most manufacturers Generalized Performance Curves, are values that place the pump in controlled, but heavy cavitation. The NPSH-Required is defined as the NPSH applied to the pump at a given flow rate which causes sufficient cavitation to reduce the Total Dynamic Head (TDH) by 3%. This is an established pump industry standard procedure used to indirectly measure, at a reasonable cost, the suction side pressure loss inside a pump before mechanical action increases the liquid pressure. The measurement is taken while pumping water with a minimum of inlet stream turbulence (i.e. no close connected, double elbows), no entrained gas, and frequently, with water that has most dissolved gas removed. These factors will increase the NPSH-Required values obtained during the test. While this may appear to be an idealistic system, and is not realistic for actual operating conditions, it is the only method that can provide reproducible test results. The user must apply a margin between the liquid in let total pressure and its vapor pressure greater than this pressure loss to prevent vapor formation at the impeller inlet. Even though Figure 1 illustrates the typical vendors published noncavitating Total Dynamic Head curves (head vs. Flow) and the NPSH-

Required curve (NPSH-Required vs. Flow) on the same graph, it must be understood the NPSH-R values applied to the pump will reduced the Total Dynamic Head developed by the pump by 3% at any flow rate. While the hydraulic loss due to reduced TDH is typically not significant, the resulting shock on the equipment can reduce the mechanical life of the seal, bearings, and impellers.

Figure 1 Published generalized performance curve compared to TDH with NPSH-A/NPSH-R = 1.0 NPSH-Available Manufacturers expect the user to supply an NPSH-Available which exceeds the NPSH-Required value that is published. The margin between the NPSH-Required becomes a commercial decision and should be properly evaluated by the person selecting the pump and developing the piping system. Systems which are pumping liquids at their equilibrium condition, i.e. ready to flash with an increase in temperature or a decrease in pressure, require special consideration when pump is being selected and the system is being designed. The NPSH-Available in such systems normally can be significantly increas3d only by: increasing the supply tank elevation above the pump; lowering the pump relative to the supply vessel; providing a booster pump; cooling the liquid to reduce the vapor pressure characteristic.

PUMP APPLICATION CONSIDERATIONS As a guideline, the NPSH-Available should exceed the NPSH-Required by a minimum of 5 feet, or be equal to 1.35 times the NPSH-Required, whichever is the greater value. As an example, for an NPSH-Required of 20 feet, the NPSH-Available should be a minimum of 27 feet. This is considered the minimum acceptable margin, and even then, some degree of mechanical and erosion damage can be experienced. Increasing the margin will improve the reliability and acceptable operating range of the pump. In order to totally eliminate cavitation, which affects the Total Dynamic Head, the Net Positive Suction HeadAvailable must be from two to five tomes the NPSH-Required, depending on the operating flow rate relative to the Best Efficiency Point flow rate of the pump design, as Figure 2 illustrates.

Figure 2 Mechanical damage requires little elapse time (hours to weeks). Erosion damage requires longer elapse time (days to months) Pumps that operate with NPSH-A/NPSH-R ratio above 1.35, but below the 2X to 5X parameter, can have what is considered acceptable (but not optimal) seal and bearing life; however, they may be vulnerable to erosion damage to the impeller, which will require more frequent

impeller replacement than would otherwise be experienced had the cavitation been totally eliminated. Anyone developing pumping systems designed with these minimum recommended margins should consider testing the equipment to be delivered to confirm that it meets the published data. The following information should be considered when ordering the pumps to decide if tests should be conducted: Most manufacturers do not hydraulically test pumps unless required to do sop by the purchaser The first test5 of a specific pump may result in higher NPSHRequired and Total Dynamic Head values than indicated by the published data due to casting variations of the casing and impeller and the manufacturers documentation to test procedures. To correct this condition, the manufacturer may have to grind the controlling surfaces of the impeller and case to reduce the NPSH-Required and reduce the impeller diameter to reduce the TDH values within acceptable tolerances. After such modifications are made, the pump is retested to confirm the result of the rework. The testing procedures followed by most manufacturers (per hydraulic Institute Standards) typically yield the minimum information, which is not considered adequate for most critical services, especially as it relates to NPSH-Required testing. Recommended test procedures, the critique of proposed test procedures and/or witnessing of critical equipment tests can be supplied on requests. In todays business environments, plant investment must reap the maximum return. It is advisable that those making decisions regarding new pumping systems or troubleshooting existing systems take heed of this information. Assistance on thee topics can be supplied on request. Note: Bob Hart has spent 27 years as a Principal Consult in the Rotating Machinery Group with the Engineering Department of theDuPont Company, Wilmington, DE DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page

ARTICLES LIST

ARTICLE 16: IS YOUR PUMP BURNING MONEY AWAY?


Obviously, everyone wants their equipment to last long time and not break down or wear out. Unfortunately, sooner or later, everything does. The issue is how to extend the life of a pump, and to evaluate if the existing machinery operates at the conditions conducive to long life. Our April Editorial addressed this issue. As you know, when a pump operates at flow away from BEP (best efficiency point), its efficiency is low: the lower the flow the lower the efficiency. A centrifugal pump, operating at 80% efficiency at BEP, will operate at perhaps 30% efficiency at half flow. The larger the pump the more this translates to lost dollars and, surprisingly this can be a very big lost dollars indeed. Consider, as an example, a 1,000 hp motor, which runs continuously, 24 hours a day, 360 days a year. At an estimated energy cost of say $0.07, these 1000 hp (746 kW), will consume 746 x 24 x 360 x 0.07 = $451,181 per year. If this happens at 30% pump efficiency, then, if the pump efficiency is improved or restored to the original peak of 80%, then the energy cost would be only 451,181 x (30/80) = $169,193 The difference is $281,988 per year! You can substitute your actual numbers, but the point is it is not a small matter, but a big dollar value, and an unhappy pump. Why do pumps operate off-peak? Several reasons. We find that often the flow usage decreased through the years due to plant downsizing less energy demand, workforce reduction, surrounding areas industry moved out, etc. Or, sometimes the pumps were oversized originally just in case, but, with running experience, a proper flow is lower then at best efficiency flow point. The extra capacity is thus not needed so the pump is throttled, - and that is at big efficiency loss. Of course, the simplest thing to stop wasting these dollars is to get a smaller, better sized-for-proper-flow pump. Pump manufacturers can do it, why not? But it is an expensive solution. And not only because of the pump purchase cost, but the associated changes in piping can be, as you know, a nightmare. Or prohibitive. Would it not be much better to keep the same casing, same piping, and just change over the impeller? Of course it is. A new impeller design can be fitted - into same casing, no piping changes, no hustle - the pump BEP point is essentially shifted to the left smack on where the operating point is - and the efficiency is again restored to 80% (or close to), instead of the 30%, as assumed in our example. The cost of the impeller change? Not much comparing to the whole new pump. Price? Often a 3-4 months payback.

Why does not everyone do this if so obvious? First of all it is not so obvious, because the inertia and a habit of running the pumps to the ground is the fact of life. The pump does not cry out loud that it is being abused and runs inefficiently no one measure the dollars being burned away by the wasted-money-meter at the pump the lost dollars are hidden in the overall electric bill. Accountants see them and want to make the bills smaller but accountants are not hydraulic designers, and the pumps are left until next year and the next. Some of these wasted dollars are passed on to the consumers, via energy surcharge or product price but, when the surcharges reach certain point, it is not so easy to keep rising the rates indefinitely to recover the wasted energy. Besides, at a given surcharge to the public why not save on energy anyway? changing out the impeller is not a very difficult thing to do, especially if there is no piping changes required. Another reason for inefficiently running pumps is that big pump companies are not in business retrofitted impeller designs and sizing them to right conditions. To do that means spending time and having their engineers involved. Designing impellers is easy, but you have to have people that not only know how, but have time to do that. Goulds, Flowserve, Sulzer, etc. - are not in business of keeping a huge staff of hydraulics engineers looking out for customers to fix problems with a, say, $10K impeller, when they can sell a complete pump for $50K or more. And, - piping changes are not pump manufacturers problem! these expenses come out from the users pockets. This is why some companies has been successful doing such retrofits. Impellers can be machined from solid blocks of structural engineered composites: strong as steel, good to 300 deg. F (special grades to 550 deg. F), and superior from cavitation standpoint better then bronze or stainless steel. Handling to 15% of solids particulate is not an issue, and resistance to most chemicals is excellent. Sea water, brine and brackish water do not attack the simsite composites cooling water pumps, river raw water intake, etc. are examples of excellent success of applications. For chemical plants excellent opportunities for most acids, caustics and other nasty chemicals. And, - with 80% lighter then metal these composites improve pump rotordynamics, and reduce shaft deflections solving seal leak problems, improving bearings life, and saving couplings from failing. We can help you identify and evaluate such potential money-burning pumps, and recommend a solution. What we need is a performance curve of your pump, if you have it, and its desired operating point. We will estimate the radial thrust load, onset of the suction recirculation, and determine the best impeller design right to the bulls eye of the best efficiency point! You have nothing to loose, but save a lot of money. If, after having our evaluation, you still decide not to implement the upgrade not a problem and the decision is yours and we do not charge for the technical evaluation. Give us a call. We will provide you references where other users, like your company, were faced with inefficient operation of pumps and decided to do something about that. They can provide testimony from their own experience. Your operators will get a better pump and easier to operate. Your mechanics will thank you for 80% parts weight reduction. And you management will love the money saved.

DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article 17: What Happens When a Pump Operates Off-Peak?


Last week I got a call from a friend asking for help with a pump at a Paper Mill. The problem: piping vibrates violently, pump is noisy, and failures are constant. Besides, pulsations started to affect the quality of the paper, obviously an unacceptable thing to the mills customers. The question was - what can be done? As we looked closely at the details of operation, it became clear that this double suction pump, initially specified and sized to operate at the 3000 gpm (which is where its BEP is), actually runs at about 500 gpm. Several options were presented to the mill in the past, such as by-passing the excess flow, or buying a smaller pump although would solve the problem with vibrations are an expensive way to solve problems. Getting a new pump is easy, - as long as you do not pay for it from your own pocket. Piping modifications alone could exceed the cost of a new pump, and sometimes even that is not possible, and prohibitively disruptive. The question was could something be done to pump internals, while keeping the same casing, and not having to change the piping? Fortunately, the answer was yes. A new impeller, with modified hydraulics was designed to fit the existing casing, and shift the pump best efficiency point to 500 gpm. When a centrifugal pump operates way to the left of its BEP, many nasty things happen. First, a radial thrust grows exponentially, resulting in significant shaft deflections thus quick seal failures, reduced bearings life, worn out bushings andrings. Also, a hydraulic phenomenon called rotating stall sets-in, which is essentially a backflow, leaving the impeller eye, and progressing backwards, resulting in violent piping vibrations, pressure pulsations, and wear-out of the components. The problem becomes worth especially when a hydraulic parameter called suction specific speed (NSS) is high. Suction specific

speed is an indirect indication of the impeller eye being too large, but also depends on several other factors, related to design, installation and application. There are certain engineering rules and principles related to minimum allowable flow - as function of pump energy, specific speed (NS), suction specific speed (NSS), and other factors, which when violated can cause trouble.

A properly redesigned impeller can thus be an effective way to solve such hydraulic problems, and is only a fraction of a cost of a complete pump replacement. And, these days, when funds are tight, and maintenance and purchasing departments are struggling to find better ways to extend the equipment life, - equipment upgrades through hydraulics optimization come to therescue. Paper mills, chemical plants, refineries, municipalities, - all have the same goal: maintain production, while reducing expenditures by extending equipment life. Working with the end users, such improvements in equipment reliability are possible, and require sound engineering approach, where attention to details, and appreciation of hydraulics and mechanical interactions, can be realized effectively, and economically. . Send you comments to: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article #18: Pump Reliability What Does this Term Mean to You?
This proposed discussion is for the ultimate purpose of making the appropriate decisions in developing, installing and operating a pumping system. I personally narrow the responsibility for 'reliability' with the following statement: "Fundamental reliability is an engineering responsibility, not a maintenance function. Maintenance cannot be expected to correct problems that require major changes in system design requiring capital investment". Management should not hold the maintenance department responsible for equipment failures over which they (maintenance) have no control. Example: mechanical seal failures frequently can be traced to improper operating conditions, piping configurations and/or inadequate NPSH margin. None of these are typically in the control of the maintenance department. Maintenance personnel need to understand the issues in these areas that are creating an impossible task for them to correct and take the issue to management to provide the necessary corrective action and funding. Per Webster's Seventh New Collegiate Dictionary - "Reliability: the quality or state of being reliable. Reliable: suitable or fit to be relied on". If Webster's cannot offer a more precise definition, we have a challenging task ahead of us. Developing a reliable pumping system requires the interaction of individuals with the expertise in at least the following areas:

Process Engineering Civil (concrete, space layout) Piping Design Procurement Power (electric motors) Instrumentation (controls and instruments)

Pump Applications Maintenance Operations One of the more difficult challenges in a task involving a number of individuals that must contribute to an interacting functional system (such as a pumping system) is to understand the definitions used by various groups for the common terms that are used in communication. There have been major errors that have cost significant money, time and even injury because of this issue. It is nearly impossible to achieve agreement to use a common definition of even the most moderately complex terms between the various technical and trade groups without a major reeducation effort. This is frequently complicated by the fact these groups frequently communicate only by information contained in data sheets. The expedient task is to identify the various definitions being used for a given term and attempt to design the system to compensate for these variations. This requires that the one selecting and specifying the requirements of the supporting functions (the pump Application Engineer) that make up the pumping system be knowledgeable in the definitions and practices that tend to be used by these various groups and then - attempt to compensate for the inadequacy of the definitions being used. In some cases it may be necessary to request a specific group to alter their design approach to provide for the 'tolerances' their specific definition has and the negative effect it can have on the final product (the pumping system).

My definition of Reliability is:


A system that will function for a specified period of time, that is of reasonable length and compatible with the process requirements, without unpredictable shutdown or failure.

Robert J. Hart Robert J. Hart Enterprises, LLC Note: Bob Hart has spent 27 years as a Principal Consultant in the Rotating Machinery Group with the Engineering Department of the DuPont Company, Wilmington, DE What is YOUR definition of Pump Reliability? Comments: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Pumping Machinery, LLC


Article 25: Submersible Pumps (dry or wet?) for Water and Wastewater Applications
A Maintenance Manager of a large municipal waste treatment plant asked me to help with the noisy and problematic pumps for his aeration basins recirculation. Presently, he has an old end-suction centrifugal pumps, with a long segmental shaft, snaking up vertically and connecting to a regular dry motor at the surface floor:

Pumps have been problematic for awhile. Repair is always an option, and, although Bob got the quotes from the repair shops, he was more inclined to replace these old pumps (and hard-to-get parts) with a newer design. He was also interested to look at the "dry submersible" pump options. A dry-submersible pump is essentially a regular submersible pump, used for wet wells, but then modified and fitted with options to run dry. Such submersible pump can then be installed exactly at the same place, in a dry pit area, where the existing regular pump (shown on photo above) was, - an easy, simple and straightforward replacement, with little piping changes, and no need for a long shaft, or the motor itself sitting far away, at the top floor surface. A regular submersible pump' biggest problem is cooling of the motor. It depends on the liquid in the wet-well to cool the motor, and thus the wet well can not be allowed to run completely dry - some water must be present, to cover the pump to almost the top of the hook, to which the pull-out chain is attached. If such pump is installed at the dry pit, instead of a wet pit, there would be no water to cool it, and the motor would burn up quickly. Some companies, however, make submersible pumps that "do not need to be submersed", i.e. the can run dry, and the cooling of their integral motor is done by the pumped liquid, via diverting a portion of it to a cooling jacket, designed around the outside part of the motor, as shown on the illustration below.

The illustration above comes from the Flygt Pump product brochure, which brings another point to a discussion. Bob wanted to know who makes such pumps, and what are my recommendations. This is not an easy question to answer. There are many companies that make submersible pumps, so which one to recommend. I decided to do a little study, based on three criteria: 1) Good engineering design with quality manufacturing 2) Customer service and support 3) Price I then looked up several pump companies that manufacture submersible pumps: Flygt, ABS, Barnes, Ebara, Polaris, and Hydroflo. I knew about the Flygt and ABS, as I came across their names and installations many times. I had no problems locating their web sites, getting sales and technical brochures, and finding local representatives quickly. What impressed me about Flygt was not just the name recognition, but quality of publications, the extend of the product coverage, as well as familiarity and fluency of their technical sales staff with the pumping stations auxiliaries, such as: guide rail system, control panels, auxiliaries, float switches and other electronic control options, and more. Flygt people seemed to "speak the application language", and not just selling a pump from a limited catalog. In fact, I got their brochure about submersible pump applications, in a simple, layman language, easy to follow, and easy to learn the terminology. When I got done my study of Flygt product, I was impressed with the company technical know-how, the quality of publication and educational material, and the turn-key approach to the pump application request. They acted more like a solution-oriented company, and not just a sales office, represented a remotely manufactured product with obscure name, and just as little commitment to answer questions. Barnes Pumps literature was decent, although with a flavor a "good-pumpcompany" outfit, and not necessarily a flow solution organization. I suppose someone who knows exactly what he or she wants in a pump, and knows how to install it, hook up the floats, fine-tune the switches, etc. will do well with them, but if you have questions, I am not so sure the answers would be just as complete. ABS pump company is also a known player in the industry, and they manufacturer a variety of pump types, and not only for the water and waste treatment submersible applications:

Ebara had a nice web site as well, but it was not clear if they are focused on the municipal submersible market, or on large vertical and other types of pumps, made in Japan. It looks like their US submersible division is in South Carolina, and large vertical pumps division in Houston, with little integration and independent

operation. Their web site did not tell me as much as I wanted to know about the pump specifics, such as performance curves, sectional drawings, and other related technical data, although hopefully it will eventually develop some of these capabilities. Their focus seem to be on selling pumps only, and while having certain knowledge about the supporting systems and controls, their main forte appears to be strictly pumps, with the auxiliaries left to the user or the users contractor. Another company was Polaris, although I never heard the name before, but listed for reference, since it appeared to be distributed close to my home stage, - in the nearby Alabama. Their 2-page brochure did not tell me much, other then an invitation to quote a pump.

The last company I looked was Hydroflo, which appears to be also a representation of the off-shore low-cost manufacturer, and while their quality is probably fine, there is little information available on any technical issues, such as seal options, motor details, sectional drawings, wet or dry pit options, etc. There was no web site available either, but I was able to obtain a quotation on a 30 horsepower pump via my friend in Chicago, who runs a pump distributor.

With the quote, I went to Bob, and showed it to him. He liked the number, and said it is cheaper then some other companies, that he is more familiar. He then asked a few questions: "Will the spare parts be just as competitive? Where are pumps manufactured and who is in charge? Will they offer an extra warranty? Who makes their motor, and what is the detail of the windings? What are seal options? Where is a bill-of-materials, what are the material options, and where is a Recommended Spare Parts List? Where are the curves and is there a catalog?"
Unfortunately, none of Bob's questions could be answered, and so Bob put the quote just where it belonged - to the waste basket!

The question was, I wondered, is any one of these companies, or perhaps any additional others as well, is "better" then others? No. I could not say that, and the products may well be equally good. But the only way for the user to know for sure - is to be convinced by action. The lesson here is that, even at best price, if the product quality, support, customer service, spares, documentation, and auxiliaries are not in good shape a pump company would doubtfully get very far in a today competitive world, where customer service, and problem solving, are the name of the game, and not just a sales, on a phone, or a fax machine. Pump users are looking for the solution of their water treatment issue, - a complete system, and just one of its elements:

Article 19: WHAT HAPPENS WHEN A PUMP NO LONGER OPERATES AT OPTIMUM CONDITIONS? (Part 1) Last month, we discussed general implications of pumps operating to the left of the best efficiency point (BEP). Low efficiency, high radial loads, noise, vibration - become a real problem when that happens. Damage to the seal, shaft, couplings and poor reliability are a real and direct result of such operation. This month, we will explore the effect of such operation on pump efficiency and will estimate wasted energy. In the following Editorials, we will also examine the effect on radial load, cavitation damage, and other aspects. The larger the pump, the more energy is wasted when a pump operates off-peak. A full range of ANSI pumps, for example, as offered by a pump manufacturer, may consist of many sizes, to span a

wide range of flows with larger sizes reaching over 4000 gallons per minute:

The graph above is called an Overall Hydraulic Coverage Chart. For each size, a head-capacity curve at the maximum and minimum impeller diameter is plotted (sometimes minimum diameter is not shown in order to make the chart less cluttered). This allows to make an approximate selection of a pump size, and then to look up the individual detailed hydraulic performance curve for that size, to finalize the details. Lets consider a case of a relatively small pump first. For example, if a pump user is looking for a pump to pump 40 gpm at 140 feet head, a 1x1.5-6 pump size (with approximately 6 impeller diameter) would be picked. The pump will work, but unfortunately will not be operating at its optimum design point.

As evident from the hydraulic curve for this pump size, this pump will have 40% efficiency (yellow circle). Its design point (red angle), however, is at 58% efficiency, i.e. the pump operates to the left of its BEP point for the impeller diameter required to achieve the desired head. As we said, the issues of radial thrust, vibrations, etc. will be addressed at later Editorials (you may take a sneak preview based on Articles #16 and 17), but for now lets see what this means strictly from the energy point of view. In our example, note that a horsepower line that passes near the operating point is approximately at 4 hp, which is roughly 3 kW. How much does it cost to operate this pump if running continuously, 365 days per year, at, say $0.07 per kilowatt-hour? 3 (kW) x 24 (hr/day) x 360 (days/year) x 0.07 ($/kW-HR) = $1814 Now, what would it cost if the efficiency was somehow improved to the same 58% that this pump enjoys when operating at the design point? Obviously, if a pump runs more efficiently, it will take less power. In fact, the power (and thus cost) would be inversely proportional to efficiency:

1814 x (40/58) = $1251 The net savings would thus be 1814 1251 = $563, which is 31% less Next, lets consider a somewhat larger pump. Say we have a 4x610H size operated at 600 gpm and producing 100 feet of head:

Again, the pump is off the efficiency peak. It operates at approximately 65%, whereas its peak efficiency at that diameter (10.25) could be 82% ! Now, the energy dollars become more pronounced. Its power consumption is approximately 25 hp (19 KW), according to horsepower lines in the proximity to operating point: 19 x 24 x 360 x 0.07 = $11,491 But it would be less if efficiency was restored to the designed 82%:

11491 x (65/82) = $9,108 The savings would be: 11,491 9,108 = $2,382, i.e. about 21% in this case Lets next take even larger size, 8x10-17:

Lets assume this pump operates at 2000 gpm (280 feet head), instead of a peak point of 4000 gpm. The efficiency at the actual operating point is only 70% instead of the potentially achievable 83% by this pump. The horsepower at the operating point is roughly 225 hp (168 KW), and the yearly energy bill is: 168 x 24 x 360 x 0.07 = $101,516 At restored efficiency, this would be: 101516 x (70/83) = $85,616

The nest savings would be: 101516 85616 = $15,900 For larger sizes, the energy savings could be even greater. As you can see, the net savings depend on how far back away from the Best Efficiency Point the pump operates. Unfortunately, this problem exists in all too many actual installations in the field. Many pumps, procured and installed years ago, often no longer operate at the originally intended hydraulic conditions. As operating conditions change, the pump is simply throttled further and further away from the BEP. The result dollars literally burned, - not to mention other problems (high loads, shaft breakage, etc.). Obtaining a smaller pump is not a good answer. First of all, a smaller pump may still not (and usually does not) have the hydraulics sized to hit the operating point dead on. It may help somewhat, but is expensive and not as efficient. The user choice is limited only to the pump sizes available, as standard, from the pump manufacturers catalog, and even with a large number of sizes in the catalog, it is virtually impossible to cover each and every variation of the operating conditions. So, the user is forced to settle for the second best, but not the optimum. More importantly, however, is the issue of economics and feasibility of piping change, to accommodate a proposed pump downsizing. Piping changes alone can often cost more then a pump. A better solution is to have a new impeller, custom-designed and sized for your operating conditions. By doing that, a pump performance will essentially shift or slide to exactly where the Best Efficiency Point is, - and the net losses become zero. Such approach is effective, and the investment is minimal, with a payback of less then a year, and often just a few months. Not only ANSI single stage overhung-impeller pump designs can benefit from this approach. Cooling water between-bearingpumps, are known to have benefited greatly with improved impeller hydraulics. When a metal impeller is replaced with structural engineered composite material (80% lighter then metal), the combined effect of hydraulic fine-tuning with reduced weight (and thus load), can be dramatic. Rotordynamics benefits of such approach are obvious, and savings immediate. Other pump types, such as vertical multistage river

intake pumps, condenser, circulating, etc. can have similar issues, and could be likewise retrofitted with improved hydraulics designs, - quickly, efficiency, and economically. More on this in the next edition of our Editorial. In the meantime, if you suspect that your pump is not operating at the optimum conditions, - send us your hydraulic curve and indicate the desired operating condition. We will evaluate the potential energy savings, as a function of your operating conditions in relation to the actual pump BEP point. We will then evaluate the costs, and impact of rotordynamics, and will provide our engineering recommendations. You may be surprised how much money you may potentially save.

. Dr. Lev Nelik, P.E., Apics Pumping Machinery June, 2003 Send your comments with request to evaluate efficiency improvements and potential savings via retrofit program to: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST Article 20: WHAT HAPPENS WHEN A PUMP NO LONGER OPERATES AT OPTIMUM CONDITIONS (Part 2) In June of 2003, we started to discuss hydraulic implications of pumps operating to the left of the best efficiency point (BEP). Low efficiency, high radial loads, noise, vibration - become a real problem when that

happens. Damage to the seal, shaft, couplings and poor reliability are a real and direct result of such operation. However, is it possible to quantify Reliability? What is the impact on the equipment Life Cycle Cost, when a pump operates, say, 40% off-peak? How much does this cost the plant? not just in energy alone (that aspect we covered in June), but in terms of seal replacement bearing life coupling repairs cavitation and recirculation damage cost and so on. Is it possible to derive at some factors which would relate the maintenance dollars spent to the inefficiency of the pump operation?! In Part 1 (Article 19), we talked about the effects of pumps operating at off-peak flow on efficiency and did an estimate of the wasted energy. That discussion now has been published as so please feel free to examine it there. This month, we are now examining the effect of such off-peak operation on radial load, cavitation damage, and other aspects and linking these to the estimate of the actual plant costs. To do that, some assumptions had to me made, and we would like to hear from you if you agree, or disagree, with these assumption. As times goes on, this fledgling theory of operation-to-reliability costing method may develop into a more comprehensive method, and the input from the users will help make the next step.
In the recent years, the importance of improving the overall reliability and plants uptime resulted in renewed attention to the reliability assessment of the individual system components. Pumps constitute one of the major classes of the plant components, and directly contribute to the overall economics of the life cycle evaluation. Lets take a look, for example, at the advantages of the use new materials, such as structural composites, with regard to parts upgrade program (pump impellers, wear rings and bushings), as well as complete pumps. The reliability and savings can be achieved with focus on the following: Significantly reduced weight (80% lighter then metal) with excellent tensile strength, approaching metals Superior chemical resistance for most demanding tough chemicals Abrasion resistance up to 15% solids Dry-running continuous 3-D interwoven woven graphite fibers providing good self-lubricating properties Cavitation resistance exceeding bronze and stainless steel Improved rotordynamics, longer seal life and bearing life extension Quality: these thermoset parts 5-axis machined from solid block (no unbalance), as compared to injection-molded or cast thermoplastics (voids and crevices result in unbalance)

Higher efficiencies due to tighter clearances allowed and superior finish Lets also compare open versus closed impellers. Open impellers have traditionally been accepted as a standard design configuration of the ANSI pumps for chemicals. Initiated as far back as 1961 (originally called an AVS Standard), ANSI pumps have been installed in numerous plants, and established a convenient and accepted standard to which both manufacturers, and the end users, could readily comply. ANSI pumps, made by different pump manufacturers would adapt to the same piping and baseplate dimensions, thus making them dimensionally interchangeable, although the internal parts geometry differs from one manufacturer to another. The main advantage of the open impeller design is cost. They are easier to cast and clean up at the foundry, especially in case of sand casting process, which is often applied to iron construction. Stainless steel designs are typically made using precision patterns, and cleanup of passages from the mold residue is less of an issue, but still a less expensive operation. With the advent of the mag-drive designs, however, closed impellers became a necessity, mainly due to a need to reduce the axial thrust. Carbon or silicon carbide thrust washers, lubricated by the products, do not have the same load capability as antifriction ball bearings. Thus, a need to reduce the loads necessitated the change toward the closed impeller designs. However, substantial reliability benefits can be achieved by replacing open impellers by closed impellers, - even for the standard designs that utilize single or double seals, or packings. Although the benefits of this approach have always been understood, the quantifiable justification has become possible only recently, as new reliability methods and approaches became available, using a Life Cycle Cost Method (LCM). Benefits The benefits of the closed impellers are materialized in (4) following areas: Bearings Life Seals Life Efficiency improvement NPSH 1) Bearings As a typical example, consider a typical MTX frame ANSI end suction overhung open impeller pump. Typical radial load FR = 400 lbs and axial load is FA = 900 lbs. The equivalent dynamic load is calculated as:

P = XFR + YFA = 0.63 x 400 + 1.24 x 900 = 1368 lbs This load is carried by the 5306A double row ball bearing, which has a dynamic load coefficient C = 16,400 lbs, which results in L10life calculation as: L10 = (C/P)3 x 106/(60xRPM), and adjusted by the a23 coefficient, reflecting oil optimization (typically a23 = 2.5) Thus, Lna = 2.5 x (16,400/1368)3 x 106/(60x3600) = 20,052 hrs, i.e. 2.3 years. ANSI spec requires 17,500 hrs bearings life, which is approximately in agreement with calculations. However, the axial load (F A) used in these calculations, can change dramatically, depending on the position of the impeller within the volute, the height of the pump-out vanes (POV), and the gap between the pump-out vanes and the casing wall. The design value, used at above calculations, is at the assumed design value of 0.060 gap between the POV and the wall. The main reason to use pumpout vanes (POV) is to change the pump axial hydraulic thrust:

The rotation of the impeller results in dragging into rotation of the fluid in the gap between the impeller and casing walls. This is similar to a motion of a teaspoon in a cup, or a disk spinning inside containment. The resulting motion is referred to as forced vortex. Such vortex sets-in in the front and back gaps between the casing walls and the impeller front (shown on the right) and a back hub (shown on the left). The pressure distribution in the gap is parabolic higher at the impeller OD, and gradually reducing towards the shaft centerline.

Pressure time area is force which is exerted on the impeller from both sides (FR and FL). The difference between these forces is hydraulic axial thrust, which is ultimately transmitted to the bearings, and thus is desirable to be small. This pressure at a given position in a gap depends on the radius, rotational speed of the fluid (divided by the impeller rotational speed), and the gap. Curve (1) shows the static pressure distribution behind the impeller hub without the pump-out vanes. As we know from basic hydraulics, - the faster the fluid moves, the lower the static pressure is. So, if we could make the fluid in the gap to rotate faster, the static pressure would be reduced, and the force FL to become smaller closer, and hopeful equal to the FR to reduce, or eliminate the net thrust. Without the POV, the fluid in the gap is rotated only by the friction (drag) of the impeller hub wall. It rotates with the same speed as the impeller right at the impeller wall surface, but (so called no-slip-condition) is not rotating at the casing wall, since that wall is stationary. Thus, on the average, the bulk of the fluid in the gap is spinning at the angular velocity equal to half of the angular velocity of the impeller. But, if we add the POV, the fluid becomes trapped within the POV space, and thus rotates with the same speed as the impeller i.e. double of what the fluid does in the absence of the POV. This, of course, assuming the gap between the POV and the casing wall is (theoretically) zero (x=0). The number of the POVs actually does not have to be equal to the number of the impeller main blades, but often is, due to casting production process, for simplicity. Obviously, the gap x can not be zero, so the actual reduction of the pressure profile (curve marked as 2, is less, depending on the gap x. And, if this gap becomes too large, the effect of the POV diminishes, and eventually disappears. It turns out that, the POV are most effective at x=0, and become completely noneffective when x=t, i.e. when the liquid gap (x) becomes equal to theheight of the pump-out vanes (t). (Papers are written on this subject, such as a well known Zankers paper), explaining whys and whats, and those of you who have a couple of sleepless nights to study them let us know, and we will get you the material). The balancing holes are also used to reduce pressure distribution, i.e. a similar idea as the pump-out vanes. In fact, this is why they are called balancing. To be effective, the impeller must have a tight clearance between it and a casing (not shown on a picture above), to separate the higher pressure zone, from a lower pressure zone. The balancing holes thus connect the back of the impeller with the inlet area, where pressure is low (close to suction). Some leakage will occur, reducing the efficiency. If there is no clearance, such as shown above, the leakage will be greater, reducing the efficiency even more.

Another reason for the POV is to reduce the pressure at the mechanical seal area. The effect of these vanes can be very strong, and, sometimes even result in creating vacuum, and boiling out of the liquid. This could be trouble, as mechanical seals do not like to operate in a vapor environment. Regarding the performance there is price to pay for the thrust balance, as the additional power required to spin the liquid faster takes away the efficiency. This is why, the higher energy pumps, such as API, or boiler feed, rarely have the pump-out vanes, while the ANSI pumps, which are relatively lower horsepower units, have these. Note the picture above shows an open impeller. The front gap, between the impeller and the casing, must be tight (typically 0.005 0.015, depending on a pump size). The challenge is thus to maintain both front and back gaps small from the efficiency and thrust standpoint. Closed impellers solve this problem, but at the expense of reduced ability to handle stringy, such as fibrous, solids. Regarding the reverse vane impeller. The impeller shown above is a standard design, such as, for example, manufactured by Goulds, model 3196. To adjust the front clearance, the rotor must be pushed against the casing, and then backed-off by the amount of design clearance. It is sometimes desirable to keep the casing piped-up, as it is somewhat a chore to re-pipe it. This requires re-setting of the front gap, during maintenance, on site and, if it rains or snows you freeze and catch a cold! If the impeller is turned around, such that the clearance gap is between the impeller and the stuffing box (or a sealing chamber), then this clearance can be set at the shop, and the rotor can then be simply brought to a casing and bolted on quickly. Example of such design is former Durco Mark III design. Clearly, the perceived advantage, widely publicized in the past by the manufacturers, of the ability to adjust the impeller front clearance to compensate for wear is quickly, can do more harm then good: as impeller is adjusted forward to close the worn out gap, the back clearance increases by the same amount. The increased gap between the POV and the casing wall can effect pressure distribution dramatically. The resultant axial thrust can actually triple, i.e. F A = 3 x 900 = 2700 lbs Additionally, as impeller wears out, the removed metal causes radial unbalance, which adds to a hydraulic thrust. Estimates vary, but 50% increase in radial load is quite possible, so that FR = 1.5 x 400 = 600 lbs The equivalent dynamic load then becomes: P = 0.63 x 600 + 1.24 x 2700 = 3736 lbs, which impacts bearings life dramatically: Lna = 2.5 x (16,400 / 3736)3 x 106/(60x3600) = 979 hrs = 1.3 months !

2) Seals The impeller wear, and the resulted unbalance, results in increased radial load and shaft deflection:

Seal manufacturers have done research which shows exponential deterioration of seal life, including leakage and failures, when the angular misalignment exceeds 0.002. At 50% increased radial load, a normal projected seal life of 2 years can be reduced to less then 6 months. The problem can in fact become worse if the impeller is adjusted (as was discussed above) axially due to wear, because this results in reduced seal spring tension, lowering closing force and initiating premature leak. 3) Efficiency Theoretically, an open impeller with negligible front clearance is somewhat more efficient then a closed impeller. However, efficiency drops very quickly as actual front clearance is increased due to wear. When initial 0.005 front clearance is opened up to 0.010, a pump can loose 10% efficiency due to increased leakage, and at 0.015 there would be approximately another 10% efficiency reduction. Obviously, none of this is an issue for closed impellers. 4) NPSH The impact on NPSHR is similar to the efficiency. The increased front leakage effects the NPSHR and could add another 2-4 feet when clearance doubles. The R-ratio (NPSHA/NPSHR) varies form one installation to another, but a typical rule is to have at least 5 feet NPSHA above the NPSHA required. When almost half of this margin is taken away to front leakage, the R-ratio can become too close to the NPSHR and the cavitation bubbles that begin to form even before the NPSHA reaches the value of performance loss, become active enough to cause caviation damage:

It is difficult to quantify the decrease in impeller useful life due to worsen cavitation condition, since there is no known statistics published on the average life of the impeller as function of NPSHA, and some reasonable assumptions may be required. If we assume that a non-cavitation impeller could last 10 years until replacement, the reduction of the NPSHA margin from 5 feet to 2.5 feet would affect the life in direct proportion, i.e. reducing it to 5 years. More studies, however, would be required to quantify this particular effect on reliability. LIFE CYCLE APPROACH In order to compare different designs using Reliability methods, a Cycle Basis needs to be established, which is also applied to other types of components. A reasonable basis for comparison is a 5 year interval. Another assumption is made that, during this time cycle, a pump operates in several descretised steps of operation, for example 20% in regime 1, another 20% in regime 2, and so on. This allows a more uniform distribution of various regimes of operation throughout the cycle and a better averaged comparison. We will apply this method to each of the selected elements that are affected by the forcing function, such as: a) Bearings life decreasing from 2.3 years to 1.3 (0.21 years) months within five zones for the purpose of averaging:

The above distribution can be a safe assumption when a continuous statistical data is limited. Otherwise, if the load is known at each interval, and the bearing life can be calculated at each of the intervals, the averaging can be done by summing the life at the intervals and dividing by the total time span (5 years in case if Cycle Basis is selected as 5 years). For linear distribution, the averaged weighted life would be: (2.3 + 1.75 + 1.2 + 0.65 + 0.1) / 5 = 1.2 years b) Seals using similar linear function (2 years life at interval 1, reducing linearly to a calculated 0.6 years life at interval 5), we get: (2 + 1.62 + 1.25 + 0.88 + 0.5) / 5 = 1.1 years c) Efficiency - assuming pump efficiency drops from 70% to 58% (70, 65, 60, 55, 50), we get the weighted average: (70+65+60+55+50) / 5 = 60% To relate this to the life cycle factor, we need to evaluate the energy savings due to efficiency difference. Assuming a pump with a 10 hp motor, running 100%, at $0.06 per kWxhr, this results in: 10 x 0.746 x 24 x 365 x 0.06 x 0.70 = $2755 10 x 0.746 x 24 x 365 x 0.06 x 0.60 = $2352 Annual energy savings $402 per pump, which is roughly a 1 year payback period basis a typical cost of a replacement impeller. d) NPSH margin assume reducing from 5 feet margin to 2.5 feet (5, 4.4, 3.8, 3.1, 2.5), we get the averaged cycle value of: (5+4.4+3.8+3.1+2.5)/5 = 3.8 feet If a 10 year life is assumed with a full specified margin (5 feet), then the reduced life is 3.8/5 x 10 = 7.6 years Note: the actual life of impellers also varies depending on applications and pump types. Double suction cooling water pumps, for example, are notorious for NPSHrelated problems (Ref. [10]). COMPARISION Bearings life factor Seals life factor Efficiency factor NPSH life factor Open 1.2 yr 1.1 yr 0 yr 7.6 yr Closed 2.3 yr 2 yr 1 yr 10 yr

Total: MTBF

9.9 yr

15.3 yr (54% improvement)

Having established life factor comparison, the next step is relate these to the Mean Time between Failures (MTBF), which is typically known relatively accurately for a given plant. Assume for example, a plant averages 3 years MTBF. The 54% improvement, calculated above, would increase the MTBF to 3 x 1.54 = 4.6 years. What does an improvement in 1.6 years, per pump, in MTBF mean to a typical plant? 1.6 added years is equivalent to an improvement of 1/1.6 = 0.62 failures pre year. The cost of equipment failures is a total of individual components: parts, labor and lost production. While replacing of impellers, seals, bearings, and other parts is costly, a cost of lost production can even more significant. For a typical chemical plant with thousands of pumps, this could be a very significant number. The cost of a lost production also varies from one plant to another, and literature has values anywhere from $5,000 per hour to $200,000 per hour. At a $10,000 per hour value of lost production, each pump would result in 10,000 x 0.62 = $6,200 difference, and, with a population of 1000 pumps at a plant, this amounts to 1000x6200 = $6.2M for a plant! Clearly, these vary from one plant to another, but the importance of the issue is clearly there, and it can not be ignored.

Example above: metal impeller, reduced life due to corrosion attack. Composite (continuous graphite fibers in epoxy matrix), not affected by chemical attack. Also, an ANSI pump, installed at the chemical plant, resulted in extended life due to lowered weight, as well as improved chemical resistance. These cases are currently being prepared, and will be included with the final manuscript. A retrofit program of converting open impellers to closed impeller design present substantial, real and immediate benefits to the end user. For a typical chemical plant, or similar operating facility, this could be thousands, or even millions of dollars, saved in maintenance, repair and production budgets. The conversion process is straightforward and technically sound. The best approach is to

establish a planned program of replacing, starting from the worst operating units and continuing to the next level, driving the plant reliability record continuously upward. New technologies, such as engineered structural composites, present a timely and effective opportunity to achieve maximum benefits and quickly. The combined benefits of significantly reduced weight, substantial improvements in chemical resistance, excellent abrasion characteristics, and superior cavitation resistance, makes Simsite a material of choice for such program. We welcome your input on this topic, and will be glad to post your comments and opinions at the Pump Magazine On-Line commentary board.
Dr. Lev Nelik, P.E., Apics Pumping Machinery, LLC August, 2003

DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article #26: Proper Pump Piping Procedure 10 steps

By Dr. Lev Nelik, P.E., APICS Pumping Machinery, LLC www.PumpingMachinery.com

It should be realized that piping issues directly affect the pumps life and its performance. Bringing the pump to the pipe in one operation and expecting a good pump flange or vessel fit is a very difficult, if not impossible, task. When bringing the pipe to the pump the last spool (suction side and discharge side, each) should always be left until the pump has been leveled in placed and rough aligned. The final alignment will be a free bolt condition, and, as may sound like a surprise to some, no come-alongs would be needed. As an ultimate investement in common sense and proper attention to details, - your pumps will last longer, with fewer failures of seals, shafts, bearings and couplings. More equipment uptime, and less lost production, will result in significant savings in dollars, and fewer headaches.

Step 1 (only for cases where there is NO thermal growth otherwise see Step 2 discussion)

At this point the pipe should be securely anchored just before the last spool, to prevent future growth towards the pumps flanges.

Figure 1 Occasional usage of anchors (only if there is NO thermal growth (which is practice very rare) for the pump piping example would be a very short run of suction pipe connecting it to a cold water tank, which keeps the pipe at essentially constant (e.g. ambient) temperature, same as water in the tank. In most cases, however, anchors should not be used (see Step 2 discussion)

The final piping lay out should not be finalized until certified elevation drawings are received from the engineering group or from the pump vendor. Once the final certified prints are received the final isometrics can be completed and the piping takeoff can be done.

The delivery of the equipment can either be early or it can be late in arriving at the site. When the equipment is late it is critical to have certified elevation prints of the equipment. The certified prints that the isometrics required for the piping takeoffs can be made without impacting the construction schedule. If the equipment is early, it will arrive at the site prior to the construction team needing it for installation. In such cases, early preparations must be made for long term storage. It is customary to use oil mist lubrication to keep the equipment in as-shipped conditions during the storage. The pressurization of the bearing housing and the casing with just 10 to 20 H2O pressure prevents moisture and contaminants from entering the sealed areas and damaging the components. The early delivery of equipment to the site has the advantage of allowing for verification of the actual measurements.

Step 2 Once the location of the equipment is set, the baseplate can be put in place, leveled and rough-aligned, with the equipment mounted. Rough alignment of the equipment should be done prior to building the grout forms. To avoid stresses caused by thermal expansion of pipes, expansion loops should be

installed in suction and discharge lines. The sliding pipe supports near pump suction and discharge are required to eliminate weight loads of piping on them pipe, which can cause excessive loads and misalignment, leading to seal failures, bearings, couplings, etc. However, anchors (three dimensional restraints) should not be used, as these could cause significant stresses and casing distortions due to thermal expansion. Consider, an example (Fig. 2C) of incorrectly placed anchor (restraining growth in ALL directions, i.e. not simply a vertical sliding support), even 2 feet away from the pump suction, and the case where the pipe expands by only 30 degrees F (morning to afternoon):

For the pipe you use, the area of contact between the pump and pipe flanges depends on the size of the 2 pipe. Assume, for example, a 20 in contact area (or use your pipe/flange number). The resultant force on the pump will be: F = 6000 x 20 = 120,000 lbs very high. It will distort the pump casing, feet, shafts, etc., causing problems. If, in addition to that, you are pumping hot product, the piping expansion problem could be so much worse. But even the daily fluctuations of ambient temperature alone could cause problems, as shown in a sample calculation above.

A. Correct: sliding support does not retrain the piping to slide up/away B. Piping restrained (can not slide up/away), high thermal expansion loads C. Anchor will (problem!) allow pipe to expand towards/into the pump,

Figure 2 Rough alignment phase (note that the motor and the pump are not coupled yet and the baseplate is still sitting free, not grouted

Step 3 Once you are satisfied with the rough alignment, remove all the equipment (pump, motor gearbox, etc) from the baseplate. Level the baseplate to maximum out of level of 0.025" (0.06 mm) from end to end in two planes. Use machined pads as the base for the leveling instruments. Inspect the foundation for cleanliness, and if not clean, use solvent to remove grease and oil.

Figure 3 Baseplate leveling pads and grout location

Step 4 Allow time for the cleaning substances to evaporate. Form the base using the appropriate techniques to allow for the weight, temperature rise and fluidity of the grout material. Grout the base using epoxy grout. Allow the grout to cure, following the grout manufacturers recommendations. This normally requires 24 hrs at 80 F (27C). Remove the forms and clean all sharp residue and edges from the foundation.

Figure 4 Typical anchor bolt and leveling wedges

Step 5 The rough alignment step, which we mentioned above, is critical to minimize the changes that will be required to appropriately fit the piping to the pump. At the last stage, when the final spools are installed, the final alignment will be achieved with small adjustments. This will minimize the adjustments required on the motor feet/bolts. Unfortunately (motor manufacturers take heed!), motor hold-down bolts are often too tight and allow only for small adjustments to the motor before becoming bolt bound. Motor manufacturers could improve this situation significantly if motor feet were slotted, by design, rather than drilled for bolts. Figure 5 shows the tightness of space available to insert the foot hold-down bolt.

Figure 5 Potential bolt-bound situation due to tight clearances between bolt, feet and base

This illustrates once again why good alignment at step 3 can save time and the cost of having to alter motor feet (a nightmare) by slotting or reaming.

Step 6

Reinstall the pump and the motor on the baseplate. Rough align the equipment again, using reverse indicator or laser alignment or similar accurate techniques.

Figure 6 Rough alignment after grouting

It should be now easy to fine-tune the motor movement within the allowable alignment target without becoming bolt bound. This is possible because of the rough alignment during the prior step (Step 4) was completed. Note: Never install shims under the pump feet. If the shims are lost or misplaced then alteration to the piping may be required to get the pump within the required alignment specification. The normal procedure is to place 0.125" (3.2 mm) thick shims under the motor feet. This allows for adjustments that will be required during final alignment.

Step 7 Make up the final spool pieces for the suction and discharge spaces. Bring the piping to the pump now.

Figure 7 Illustration of the final connection of the suction piping.

Step 8

Warning! anchor is placed erroneously - it will restrain the pipe thermally moving away from the pump free: the pipe will expand from the anchor into the pump! (see discussion in Step 2)

Figure 8 Final piping

As a final alignment step, bring the piping to the equipment; take final measurements, tack weld the spools in place. At this time the spools can be removed and taken back to the hot work permit area to finalize the weld. Leave a square and parallel gap between the flange faces. The gap should be wide enough to accommodate the size of the gasket required, plus 1/16 - 1/8, depending on piping sizing. (This is the only distance over which the piping will be pulled. However, because it is properly anchoredbefore the spool pieces, this length is short, and stresses are minimized). Final align the equipment, taking into account hot and cold operating conditions, using two indicators on the pump shaft coupling area.

Step 9

Figure 9 Overhead view of the motor and pump

As the piping is tightened into place, the shaft shall not be moved more than 0.002" (0.005 mm), otherwise modify the spool pieces until the piping misalignment is fixed.

Several clues are common to piping misalignment. These clues come via the way of mechanical seal and or bearings running hot, and failures. A quick analysis of the failed parts can clearly show the evidence of piping misalignment. To make a final confirmation of the symptoms, unbolt the piping while measuring the movement in the vertical and horizontal plan. Again, the piping that moves more than 0.002" (0.005 mm) must be modified to correct the situation.

Step 10 Place and indicator in horizontal and vertical planes, using the motor and pump coupling. Uncouple the pump and motor, while watching the indicator movement. Start unbolting the flanges, and continue watching for movement in the indicators. If the needle jumps over 0.002" (0.005 mm) the piping has to be modified to improve the pumps performance.

Figure 10 Piping alignment verification

References

1. 1) Pump Standards, Hydraulic Institute publication, ANSI/HI 1.1-1.5-1994, Parsippany, NJ, 1994

2. 2) API 610 Standard for centrifugal pumps, 8th Edition, American Petroleum Institute, Washington, DC, August, 1995

3. 3) API 676 Standard for rotary pumps, 2nd Edition, American Petroleum Institute, Washington, DC, December, 1994

4. 4) Equipment Testing Procedure for Centrifugal Pumps (Newtonian liquids), 2nd Edition, AlChE, New York, 1984

5. 5) AlChE Equipment Testing Procedure for Rotary positive displacement pumps (Newtonian liquids), Second printing, New York, 1968

6. 6) Nelik L., "Centrifugal and Rotary Pumps: Fundamentals with Applications", CRC Press, Boca Raton, FL, March, 1999

7. 7) AlChE Equipment Testing Procedure, 1999, New York, NY

8.

8) L. Rizo, L. Nelik, Piping-to-Pump Alignment, Pumps & Systems, April 1999

E-Mail your questions and suggestions to:

DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

Article #30: API 8th versus 9th versus 10th versus ISO

Below is a document that was put together by members of the API 610 / ISO 13709 JWG. It shows the changes that were made between API 610, 8th edition and ISO 13709 (which is essentially API 610, 9th edition). API 610, 10th edition is ISO 13709 adopted back with minor editorial corrections. All these standards are technically equivalent.

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

General

Format Document in ISO format.

Scope Lower limit application guidelines (chemical pump coverage) deleted. Not applicable to sealless pumps.

4.1.10

Fig 1.2

Pump types Classification BB4 now a ring section pump.

5.1

2.1

General

5.1.12

2.1.10

Viscosity correction factors shall be submitted with proposal and test curves. If specified, the vendor shall provide both maximum sound pressure and sound power level data per octave band for the equipment. Higher energy level pumps defined for the purpose of reducing vane pass frequency vibration and low frequency vibration at reduced flowrates and above 3600 r/min and absorbing more than 300 kW (400 hp) per stage.
Requirement for fully machined mating faces of pump casing and bearing housing assembly deleted.

5.1.16

2.1.14

5.1.18

2.1.16

5.1.26

2.1.24

None

2.1.27

This clause in 8th edition regarding vendor involvement in field

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

support, when specified, deleted.

5.1.30 5.3

2.1.29

Equipment shall be designed for outdoor installation (default).


Pressure casings

5.3.1

2.2.2

Maximum discharge pressure = maximum suction pressure + maximum pressure rise with furnished impeller at rated speed and normal SG.

5.3.2

None

Option for maximum SG, maximum impeller or number of stages, trip speed.

5.3.4

2.2.1

Design tensile stresses defined in this standard, removing reliance on the traditional ASME Code Section VIII or alternative pressure vessel codes. Allowable design tensile stress is 0.25 Su with specific casting factors.

5.3.5

2.2.2

MAWP maximum discharge pressure + 10% maximum pressure rise, but not less than: a) rating of PN20 (class 125 or 150) flange for type BB1 and VS15 pumps, b) 40 bar (600 psig) at 38C (100F) or rating of PN50 (class 300) flange for all other types.

5.3.7

2.2.3

Vendor is encouraged to propose alternative corrosion allowance if materials with superior corrosion allowance are used and if they result in lower cost without affecting safety and reliability.

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

5.3.9

2.2.6

Tutorial added for guidance in application of axially split casings above limitations listed in this clause. Specific requirement for casing component alignment dowels or rabbeted fits deleted.
Nozzles and pressure casing connections

None

2.2.11

5.4

2.3

5.4.2.2

2.3.2.2

Cast iron flanges shall meet the surface finish requirements of ASME B16.1 or B16.42. Class 125 flanges shall have a minimum thickness equal to that of Class 250 flanges for sizes NPS 8 and smaller.

5.4.2.3

2.3.2.6

Flanges other than cast iron shall meet the flange finish requirements ofASME B16.5 or B16.47. (Note: 8th ed. Clause 2.3.2.6 has been deleted.) This standard no longer has special flange finish requirements.

5.4.3.5

2.3.3.5

All connection welding shall be completed before the casing is hydrostatically tested.

5.4.3.7

2.3.3.7 and 3.5.1.14

Combined two 8th ed. Paragraphs. Added: A lubricant/sealant that is suitable for high temperature duty shall be used to ensure that the threads are vapour-tight.

5.4.3.8

2.3.3.8

At interfaces with the purchasers equipment, the use of machined and studded connections requires the approval of the purchaser.

5.4.3.11

None

All purchasers connections shall be accessible for disassembly without requiring the pump, or any major part of the pump, to be

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

moved.

5.6

2.5

Rotors

5.6.1

2.5.1

Tutorial explains use of enclosed, semi-open and open impellers relative to pump type.

5.6.2

2.5.1

Impellers may be one-piece castings, [machined] forgings, or fabrications, without specific purchaser approval.

5.6.6

From 2.7.3.8

The shaft-to-seal sleeve fit(s) shall be h6/G7 (this should be changed to F7/h6 0.001-0.003 in. loose for 2 inch shaft diameter) to be consistent with API 682 / ISO 21049). REVIEW Target rings may be furnished for shafts that exhibit inconsistent electrical properties in shaft sensing areas.

5.6.11

None

5.6.14

None

All shaft keyways shall have fillet radii conforming to ASME B17.1.

5.6.15

5.1.12

Rotor of one- and two-stage pumps shall have 1st dry bending critical speed 120% pumps maximum continuous speed. [Mistakenly applied to only OH2 pumps in 8th Edition.]

5.7

2.6

Wear rings and running clearances

5.7.1

2.6.1

Close axial clearances shall not be used to balance axial thrust.

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

Impellers may have integral wear surfaces or removable wear rings.

5.7.3

2.6.3

Wear rings may now be held in place by locking pins, screws or by tack-welded.
5.7.4 2.6.4.1

Non-metallic wear rings may be proposed guidelines (materials in Annex H, Table H-4)

5.8

2.7

Mechanical shaft seals


5.8.1 2.7.2/2.7.3 All mechanical seals and related auxiliary systems now defer to ISO 21049 (API-682, 2nd Edition).

5.8.3

2.7.3.6

Seal chamber dimensions are now contained only in ISO 13709; no longer contained in API 682, 2nd ed. (ISO 21049); diagrams (Figure 25) and dimensions (Table 6) same as API 610, 8th ed. Needs review for next edition.

5.9.3

2.8.3

Vibration
None 2.8.3.3 / 2.8.3.4 The need for taking true peak vibration readings has been deleted. The references to instrumentation have also been deleted. Allowable housing vibration for pumps > 3,600 rpm or > 300 kW (400 hp) per stage rises to 4.5 mm/s (0.20 in/s) by chart in Figure

Table 7

Table 2-5

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

29.

Tables 7, 8

Tables 2-5, 2-6

Error corrected; Au taken directly from measurement, not FFT. Also formulas for Au limit corrected.

5.9.4

2.8.4

Balancing
5.9.4.1 2.8.4.1

Major rotating components dynamically balanced to ISO 1940-1 grade 2.5 instead of 4W/n - as standard.
5.9.4.4 None

IF SPECIFIED, components shall be balanced to grade G1 (equivalent to 4W/n)


5.10.1 2.9.1

Bearings
5.10.1.5 2.9.1.5

New note: There are applications where alternative bearing arrangements may be preferable, particularly where bearings operate continuously with minimal axial loads.
5.10.1.6 None

If loads exceed the capability of paired angular-contact bearings as described in 5.10.1.5, alternative rolling-element arrangements may be proposed.

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

5.10.2 5.10.2.4 b)

2.9.2.1 2.9.2.3. b)

Bearing Housings

For ring-oil or splash systems, if bearing temperature sensors are supplied, the outer ring temperatures shall not exceed 93 C (200 F), same as for pressurized systems.
5.11 2.10

Lubrication
5.11.4 None

If specified, rolling-element bearings shall be grease-lubricated in accordance with a) through d) (which covers grease life requirements).
5.12 2.11

Materials
5.12.1.3 2.11.1.9 The material specification of all gaskets and O-rings exposed to the pumped fluid shall be identified in the proposal. O-rings shall be selected and their application limited as specified in ISO 21049.

5.12.1.10

None

If specified, coatings of a type agreed shall be applied to impellers and other wetted parts to minimize erosion. Special balancing procedures and cautions.

5.12.1.12

2.11.1.11

For sour services, requiring NACE compliance for maximum hardness and yield strength: a) No longer specifically applies to impellers and balancing drums but does apply to bowls (Note: in some applications, it may be

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

desirable to apply the NACE requirements to impellers). b) Through-hardened impeller wear rings, HRC > 22, not allowed. c) Double-casing pump inner casing parts that are in compression, such as diffusers, are not considered pressure casing parts.

5.12.2

2.11.2

Castings
5.12.2.5 2.11.2.5 If casting weld repair procedures are specified, only those for repairs in pump manufacturers shop are required; repairs at the foundry level are deemed covered by producing specification (the casting material specification.

5.12.2.6

None

Pressure-containing castings of carbon steel shall be furnished in the normalized and tempered condition.

5.12.3

2.11.3

Welding
5.12.3.1 2.11.2.1/ 2.11.3.3

Default welding requirements are listed in new Table 10, showing requirement and applicable code or standard. Alternative standards may be proposed by the vendor and included on the welding and material datasheet inAnnex N.

5.12.4

2.11.4

Low temperature service

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

5.12.4.3

2.11.4.2

Purchaser to specify whether EN 13445 or ASME VIII, Division 1 to apply for impact testing.

5.13

2.12

Nameplates and rotation arrows

None

2.12.5

Deleted: Nameplates and rotation arrows (if attached) shall be of austenitic stainless steel or of nickel-copper alloy. Attachment pins shall be of the same material. Welding is not permitted.

6.1

3.1

Drivers
None

3.1.12

Deleted: The equipment feet shall be drilled with pilot holes that are accessible for use in final doweling.

6.2

3.2

Couplings
6.2.1 3.2.1

Couplings and guards between drivers and driven equipment shall be supplied and mounted by themanufacturer of the pump.
6.2.2 e) 3.2.8 Couplings operating at speeds in excess of 3800 r/min shall meet the requirements of ISO 10441 OR API 671 for component balancing and assembly balance check.

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

6.2.3

None

If specified, coupling balance to ISO 1940-1, grade G6.3. [Now have 3 grades of coupling balance.]

6.2.8

3.2.4

Coupling hubs with cylindrical bores may be supplied with slip fits to the shaft and set screws that bear on the key.

6.2.9

3.2.4

Coupling hubs designed for interference fits to the shaft shall be furnished with tapped puller holes at least 10 mm (0,38 in) diameter to aid in removal.

6.2.11

None

If specified, couplings shall be fitted with a proprietary clamping device.

6.2.12

None

Provision shall be made for the attachment of shaft alignment equipment without the need to remove the spacer or dismantle the coupling in any way.

6.2.14

3.2.12

a) Allowable access dimensions shall comply with the specified standards, such as ISO 14120, EN 953 or ASME B15.1. b) Sufficiently rigid to withstand a 900 N (200 lbf) static point load in any direction without the guard contacting moving parts c) Fabricated from solid or perforated sheets with openings not exceeding 10 mm (0,375 in) and constructed of steel, brass or non-metallic (polymer) materials.

6.3

3.3

Baseplates

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

6.3.3

3.3.3

Added: Installed baseplate flatness can be affected by transportation, handling and installation procedures beyond the vendors scope. Installation practices in API RP 686 should be followed.

6.3.4

3.3.4

Shims shall not be used under the pump. Shim packs shall not be thicker than 13 mm (0,5 in) nor contain more than 5 shims. All shim packs shall straddle the hold-down bolts and vertical jackscrews, and extend at least 5 mm (1/4 in) beyond the outer edges of the equipment feet.

6.3.8 None

3.3.7 3.3.8

All joints, including deck plate to structural members, shall be continuously seal-weldedto prevent crevice corrosion.
J hooks on underside of drip pans deleted.

6.3.14

3.3.14

Transverse and axial alignment positioning jackscrews required for drive train components weighing more than 250 kg (500 lb) was 200 kg (450 lb) for transverse horizontal adjustment and 400 kg (900 lb) for longitudinal adjustments. Screws shall be at least M12 (1/2-13).

6.3.17

3.3.17

Underside of baseplates sand blasted to ISO 8501 Grade Sa2 or SSPC SP 6 and coated with a primer compatible with epoxy grout.

6.3.20

None

Lifting lugs attached to the equipment shall be designed using a maximum allowable stress of one-third of the specified minimum yield strength of the material.

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

6.4

3.4

Instrumentation
6.4.1 3.4.1/3.4.2

Gauges Temperature indicators and pressure gauges, if furnished, shall be in accordance with ISO 10438 (API 614).

6.5

3.5

Piping
6.5.1.1 None

Piping shall be in accordance with ISO 10438 (API 614). Most piping details are removed from this standard.
6.5.1.6 3.5.1.6 If specified, each piping system shall be manifolded to a single purchasers inlet or outlet connection near the edge and within the confines of the baseplate. (this was required in the 8th edition)

6.5.1.7

None

New clause covers bolting requirements for piping, same as 5.1.31.

None

3.5.2.10.3

Block and bleed valves are no longer specified for pressure gauges when service is flammable or hazardous. [An omission? API 614 shows as an if specified item.]

Inspection, testing, and preparation for shipment

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

7.1.4

4.1.4

a)

Expected dates of testing shall be communicated at least 30 days in advance and actual dates confirmed as agreed. At least five working days advanced notification of witnessed or observed inspection or test. (moved from 4.3.1.3)

Smaller pumps may need to be removed from the test stand between preliminary and witness tests. b) If specified, witnessed mechanical and performance tests shall require a written notification of a successful preliminary test. The vendor and purchaser shall agree if the machine test set-up is to be maintained or if the machine can be removed from the test stand between the preliminary and witnessed tests.

7.2.1.1

4.2.1.1

List of data the vendor shall keep for at least 20 years has been modified.

7.2.1.4

None

All preliminary running tests and mechanical checks shall be completed by the vendor before the purchasers inspection.

7.2.2

4.2.2

Materials inspection section revised and NDE standards tabulated (Table 13) by type of inspection, methods, and acceptance criteria (defaulting to updated ASME Code standards). Casting defects maximum severity level now deferred to ASME Code. New: Hydrotesting is permitted without the seal gland plate or seal chamber installed. If a cast material gland plate or seal chamber is used, it shall be separately hydrotested to the same pressure requirements as the pressure casing.

7.3.2.1 f)

4.3.2.1

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

7.3.3.2 c)

4.3.3.1.3

Seal leakage rate during any phase of testing is to be in accordance with ISO 21049. Separate statement that liquid seals, suitable for testing on water, shall exhibit no visible signs of leakage.

8.1

5.1

Single stage overhung pumps


8.1.1 None Rear support under bearing housing not allowed.

8.1.3

5.1.3

Integral gear-driven (Type OH6) pumps

8.1.3.1

None

Impeller shall be keyed or splined to the gearbox output shaft.

8.1.3.6

None

Temperature and pressure gauges mounted directly on the gearbox shall be in accordance with ISO 10438 except that the diameter of the gauges shall be 50 mm (2,0 in). If specified, separable threaded solid-bar thermowells shall be supplied for temperature gauges.

8.1.3.7

None

Inducers, impellers and similar major rotating components shall be dynamically balanced to ISO 1940-1 grade G2.5, or to a residual unbalance of 7 gmm (0,01 oz-in), whichever is greater. Vibration measured during performance test shall not exceed Table 7 levels.

8.2

5.2

Between bearings pumps (types BB1 through BB5) Should be BB1-3 and 5; BB4 does not comply with specification.

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

8.2.2.2

None

Rotors with clearance-fit impellers shall have mechanical means to limit impeller movement in the direction opposite to normal hydraulic thrust to 0,75 mm (0.030 in) or less.

8.2.2.3

5.2.2.2

If specified, rotors with shrink-fit impellers shall have mechanical means to limit movement in the direction opposite to normal hydraulic thrust to 0,75 mm (0,030 in) or less. [Minor change in wording.]

8.2.5.1.4

None

If the shaft contains more than 1,0 % chromium and the journal surface speed is above 20 m/s (65 ft/s), the shafts journal shall be hard-chromium-plated, hard-coated, or sleeved with carbon steel (to avoid damage from wire wooling).

8.2.5.2.4

5.2.5.2.4

Thrust bearing shall be sized for the maximum continuous applied load. Size determined by: a) minimum oil film thickness of 8 m (0.0003 in), b) maximum unit pressure (load divided by area) of 35 bar (500 psi), and c) maximum calculated babbit surface temperature of 130C (265F). Based on design factor of 2 on ultimate capacity. Bearing metal temperature limits on shop test and in the field: Shop test 93C (200F) Field alarm or trip 115C (240F)

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

8.2.6.2

5.2.6.2

External pressure-lubrication systems shall comply with ISO 104438-3 andAnnex B, Fig B-10 and Table B.1. [ISO 104438-3 is API-614, 4th Edition, chapter 3. Table B.1 of ISO 13709 provides a system comparable to that of paragraph 5.2.6.2 of API 610, 8th Edition.]

8.2.7.1

5.2.8.1

For pressure-lubricated bearings, test stand oil and oil system components downstream of the filters shall meet the cleanliness requirements of ISO 10438-3.

8.2.8.3

None

Option of shipping and storage container for vertical storage.

8.2.8.4

None

Option of nitrogen purge of container in clause 8.2.8.3.

8.3

5.3

Vertically suspended pumps (types VS1 through VS7)

8.3.1.1

None

Specified discharge pressure shall be at the purchaser discharge connection. Hydraulic performance shall be corrected for column static and friction head losses. Bowl or pump casing performance curves shall be furnished with the correction indicated.

8.3.1.2

None

Bearing housings for vertically suspended pumps need not be arranged so that bearings can be replaced without disturbing pump drives or mountings.

8.3.3.1

5.3.2.1

The requirement for fully enclosed impellers (5.6.1) does not apply to vertically suspended pumps.

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

None

5.3.4.2

Deleted: Pump rotors shall be designed such that their first dry critical speed is the following percentage above their maximum allowable continuous speed: for rotors designed for wet running only 20%; for rotors designed to be able to run dry 30%.

8.3.8.2.2

None

Vertical pumps without integral thrust bearings require rigid adjustable-type couplings. [Restored from 7th Edition, paragraph 3.2.10.]

8.3.8.3.1

5.3.7.3.1

If specified, the mounting plate for double casing pumps shall be separate from main body flange and located sufficiently below it to permit the use of through-bolting on the body flange (see Figure 33). Typical mounting for VS6 and VS7 pumps with soleplate.

Fig 33

None

None

5.3.7.3.3

Deleted: The pump-to-motor mounting surface shall contain a rabbeted fit.

8.3.10.6

5.3.9.8

Default is integral bushing spiders and rabbet fits for all column sizes.

8.3.11.2

5.3.10.3

Default is integral bushing spiders and rabbet fits for all column sizes.

8.3.12.2, c)

None

Rotor of VS5 pumps shall have 1st dry bending critical speed 130% pumps maximum continuous speed.

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

8.3.13.2

None

Option of hydrotesting bowls and column pipe at 1.5 times maximum differential pressure developed by the bowl assembly.

Vendors data
6.1.3 Co-ordination meeting within 4-6 weeks after order commitment (was 4 weeks). Also, additional items added to the coordination meeting agenda.

9.1.3

9.2.3

6.2.3

NEW: p) a list of any components that can be construed as being of alternative design, hence requiring purchasers acceptance (4.2).

9.3.1.3

6.3.1.2

The purchaser and vendor shall agree to the timing and extent of drawing and data review. Deleted: after data review, the vendor shall furnish certified copies in the quantity specified.

9.3.2

6.3.2 / 6.3.3

Drawings and technical data


9.3.2.1 6.3.2 Reference to ISO 31, 128, 129 and 3098 and ASME Y14.2M for drawings has been eliminated.

9.3.2.1

None

Dimensional outline drawings shall indicate the tolerance for pump suction and discharge nozzle face and centerline locations referenced from the centerline of the nearest baseplate anchor bolt hole. The centerline of baseplate anchor bolt hole locations shall indicate the tolerance from a common reference point on the baseplate.

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

Annexes
A None Specific speed and suction specific speed. Unit of flow is m3/s, resulting in values 1/51.64 or 0.019 times that for customary US units. Cooling water and lubrication system schematics. Seal piping now per ISO 21049 (API-682, 2nd Edition).

[Fig B.10, Lube-oil system schematic, needs revision.] C E Hydraulic power recovery turbines. Only change is identification.

Standard baseplates. Dimension F, foundation bolt extension above top of foundation, deleted.

Inspectors checklist. Revised for ISO13709 clause numbers. Following specific changes made to items listed: Level 1 Motor and electrical components area classification. Casing jackscrews. Baseplate requirements. Restrained rotor. Storage preservation instructions. Level 3

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

Material identification deleted.

Criteria for piping design. No change in text; typographical errors from API 610, 8th Edition corrected.

Materials.

Table G.1

Table G.1

Material class selection guide. Class D-2 added for produced water, formation water and brine.

Table H.1

Table H.1

Material classes for pump parts. Class D-2, super duplex, added.

Table H.2

Table H.2

Material specifications for pump parts - general update, plus addition of super duplex stainless steel.

Table H.3

Miscellaneous material specifications. All proprietary materials deleted [ISO rule].

Table H.4

None

Non-metallic wear-part materials. [seal materials in ISO 21049 / API-682, 2ndEdition].

Table H.5

Table 3.4

Piping materials. Covers only process, steam and cooling water. Lubricating oil now in ISO 10438 (API-614, 4th Edition).

Lateral analysis

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

I.1.2

I.1.2

Report to include rotors 1st, 2nd, and 3rd dry critical speeds and details of items considered in the analysis.

I.2.4.b)

[Measured vibration amplitudes within 135% of calculated values; was 35%] J Procedure for determination of residual unbalance. No technical change.

Seal chamber runout illustrations. No technical change.

Vendor drawing and data requirements. No technical change.

Test data summary. No technical change.

Pump data sheets. Rearranged in graded form to match pump type: Pages 1, 2: process data and notes. Pages 3, 4: single stage overhung pumps. Pages 3, 4: between bearings pumps. Pages 3, 4: vertically suspended pumps. Page 5: pressure design codes, welding requirements, purchaser defined material inspection.

None

Purchasers checklist deleted.

ISO 13709 Clause

8th Edition Paragraph

Subject & Change

None

Standardized electronic data exchange file specification deleted.

None

Conversion factors deleted.

None

Calibration and performance verification of true peak and RMS measurement instruments used for test stand acceptance deleted.

Bibliography

Bibliography. Revised to include ISO and EN standards; now 86 references total.

E-Mail your questions and suggestions to:

DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page ARTICLES LIST

To Grease or Not to Grease That is the Question!

Lev Nelik, Ph.D., P.E., APICS Pumping Machinery, LLC Atlanta, GA

When examining actual field installations (not just theory) of grease-lubricated pumps, I find three most common scenarios. Deadheaded arrangement (a) is the most common, with no bearing shields. New grease is injected via top grease fitting (if it is not broken!) with nothing restricting it from flowing through the bearing spaces and on to the bearing housing, to be burned off or to eventually accumulate there as a pile of hardened junk. (This can be particularly problematic for motors, with grease reaching the windings, degrading varnish, shorting the leads. The motor manufacturers, however, seem to pay attention to this issue, by not using arrangement (a), and with purge plugs provided and better marked). Lip seal is not the only (and not the best) available option in the market, but I find it installed (if still functioning) most typically.

At companies that pay some attention to the importance of lubrication, a thru-flow arrangement (b) is usual. Bearing is in the cavity, and the cavity itself (not the bearing) is sealed from both sides, and new grease is allowed to exit through the purge port at the bottom of the housing during re-greasing. The best practice is to make sure fresh grease shows at the purge port, and run a pump for about an hour with the purge port open. This would allow bearing to expel the excess grease. Then the purge port should be plugged.

In practice, however, I found very few folks actually do that, and in some cases they do not bother to even open the bottom purge plug when re-greasing (in some cases a regreaser does not even know that there is a purge connection at the bottom). The mixture of new and old grease, forced by the grease gun, then bursts out through the housing/cap lip seals, often taking the lip seals themselves along the way. It is little wonder that, soon after that, water and dirt make an easy way inside the bearing housing through the damaged lip seal, quickly bringing the bearing life to an end.

There are instance where a cross-lubrication pattern (c) is applied. The idea here is to ensure the path of grease through the bearing, while, in comparison, in a thru-flow case (b) a concern is that the old hardened grease remains within the spaces between the balls, and essentially directing away the new grease from the intended delivery.

A shielded bearing arrangement (d) is rare in pump practice, although I have heard of cases where companies specifically wanted it, and claim to have excellent results. However, I found that there is a confusion of whether the shield should be on the inboard or outboard side, or both. (The shield is supposed to act as a metering orifice, and is supposed to be on the inner side where grease is supplied. The idea behind a double shielded arrangement is not clear). I found that those that thought they knew the right shield position were not sure why, and could not guarantee that the bearing shield does not erroneously end up at the wrong side after each repair. Interestingly enough, several pump manufacturers and even bearing manufacturers that I inquired about the differences in shield(s) locations, could not clearly explain these differences. Technical (or perhaps application-related) reasons behind these differences appear to be even more obscure.

What do you actually use at your plant? For example, if, say, you used to have an arrangement (b) and switch to (c) or (d) why?

To grease or not to grease? that is the question!

Dr. Nelik has 25 years experience with pumps and pumping equipment. He is a Registered Professional Engineer, and has published over fifty documents on pumps and related equipment worldwide. He is a President of Pumping Machinery, LLC company, specializing in pump consulting, training, and equipment troubleshooting. His experience in engineering, manufacturing, sales, field and management includes: Ingersoll-Rand, Goulds Pumps, Roper Pump and Liquiflo Equipment. He teaches pump training courses in the US and worldwide, and consults on pumps operations, engineering aspects of centrifugal and positive displacement pumps, maintenance methods to improve reliability, improve energy savings, and optimize pump-tosystem operation.

Send you comments to: DrPump@Pump-Magazine.com Back to PUMP MAGAZINE Page

ARTICLES LIST

We do live in a small world these days, and manufacturers crossing country boundaries is no longer a surprise, as companies search for best prices, and good

profits. However, with a reduction in cost, we can not afford to sacrifice quality, and - very importantly - need to maintain support and assistance of customers, before, during, and after the sale. This assistance does not end with a sale of a product: in fact - that is where it just starts. Follow-up, parts availability, literature, drawings, documentations, instruction manuals, and other help, as needed, with potential system troubles -should that happen - is the stuff that will differentiate the successful companies from the rest. Send you comments to: DrPump@Pump-Magazine.com

Das könnte Ihnen auch gefallen