Sie sind auf Seite 1von 69

Pump Alignment: Just The Facts

Proper alignment of the pump shaft with the driver can reduce vibration and significantly improve reliability. For appropriate applications, the time, expertise and instruments needed to achieve precision alignment (tolerances of less than 0.005 in) will prevent seal leakage and extend bearing life. Depending on such factors as operating speed and coupling type, not all pumps will require such precise alignment. When assessing a plant's alignment needs, it helps to understand basic shaft alignment concepts and procedures, as well as applicationspecific factors that dictate the required tolerances.
Effects of Misalignment

A common misconception about pump shaft/driver misalignment is that it increases bearing load, causing bearings to fail prematurely. In fact, except in cases of extreme misalignment, the resulting vibration is what damages bearings and seals. Since some vibration is normal for pumps, it is best to have an experienced vibration technician determine if the vibration is due to shaft misalignment, and whether it is severe enough to affect pump reliability.
Alignment Basics

The purpose of shaft alignment is to minimize the vibration resulting from forces transmitted across the coupling. The goal is to have both shafts rotating on a common axis, referred to as collinear. All misalignment of shaft centerlines (i.e., deviation from the collinear condition) can be described in terms of offset and angularity. Theoretically, two perfectly aligned shafts would rotate in the same axis, and if properly balanced and coupled, would not generate abnormal vibration during operation. If instead the two shafts are misaligned in the horizontal or vertical plane (or both), or are at an angle with respect to one another, they will rotate in different axes. The amplitude of the resulting vibration will vary, depending on such factors as the severity of the misalignment, operating speed and coupling type. In addition, the relative positions of a horizontal pump and driver can be viewed independently in the horizontal and vertical planes. Reducing alignment conditions to offset and angularity, independently in the horizontal and vertical planes, simplifies manual calculation of required "correction moves." Automated techniques for calculating corrections also use this convention. (Vertical pumps, solid couplings and hollow-shaft motors present unique concerns and require special procedures not discussed here.) Alignment (or misalignment) is measured at the coupling-the point of power transmission-not at the feet. The amount of shims to be added or removed beneath the feet does not directly indicate the alignment condition at the coupling.

Figure 1. Alignment tolerances in relation to operating speed.


Tolerances

Alignment tolerances specify how close the pump and driver shaft centerlines should be to collinear at running conditions. Offset tolerances are measured in thousandths of an inch (or mils), centerline-to-centerline at the coupling. Angularity tolerances are expressed as pitch or slope (mils/inch). Alignment tolerances for pumps range from the "rough alignment" that a conscientious technician can accomplish with visual indicators (accuracy of about 0.02 in) to precision alignment (accuracy of 0.0005 in or greater). The latter requires an experienced technician and accurate instruments (e.g., dial indicators or a laser alignment system). Accuracy of about 0.005 in can be accomplished with a simple straightedge and feeler gauge. The degree of precision required for a specific pump/driver will depend on the pump's rotating speed, the distance between the pump and driver shafts (spool-piece length) and the application's thermal characteristics. The required precision increases exponentially with operating speed; proportionally less precision is necessary with longer coupling spool pieces. For applications where temperature changes occur during operation, evaluation of thermal effects is also needed to determine target values. Another important factor is the coupling type. Industrial users generally agree that nonsegmented elastomer boot couplings produce less damaging vibration than jaw or gear couplings, given equal amounts of misalignment. Other kinds of couplings fall between these extremes.

Figure 2. Offset and angular tolerances in relation to operating speed.

Alignment Procedures
Rough Alignment: When installing the pump and driver, experienced technicians will perform a "rough alignment" based on visual indicators. They also will ensure that all machine feet are in good foot-plane and have 0.025 in to 0.050 in shims under them. Good foot-plane must be established and maintained throughout the alignment procedure to avoid stressing the machine cases. Target Values: Pumps and drivers are moving targets due to torque strains and thermal effects, so evaluation of these factors is an important step in pump shaft alignment. Just as a marksman anticipates the location of a moving target, proper alignment procedures must predict the relative running position of the machine cases (i.e., differences between cold and running alignment positions). These target values may be determined for the relative positions at the coupling or at the feet.

Figure 3. Sample alignment target values.

Measurement: Alignment tolerances and misalignment are measured at the coupling, where the power is transmitted. The simplest way to measure these parameters is with a

straightedge and feeler gauges. A taper gauge or caliper can also be used to measure angularity between the coupling faces. These methods can achieve accuracy of about 0.005 in, which is acceptable for many pumps that operate at 1,200 rpm or less. If precision alignment is required, careful use of dial indicator methods (e.g., rim-andface and reverse-dial), including compensation for bracket sag, can achieve accuracy of 0.0005 in. This will suffice for most pumps that run at 5,000 rpm or slower. Laser alignment systems accomplish the task quicker and with more accuracy, eliminating math errors and other common mistakes. Most of these systems also provide graphics that show the direction of the "correction move." Regardless of the alignment method, correction moves must be determined from measured misalignment data-whether the calculations are done manually, or automatically with a calculator, computer program or laser alignment system. Attempts to "guesstimate" correction moves often waste time and cause frustration. Foot-plane: To maintain good foot-plane during the alignment procedure, both front feet should be raised or lowered the same amount, and similarly, both rear feet should be moved in equal amounts. Shim adjustments that would tilt the motor side to side should be avoided. If a machine is bolt-bound or base-bound and cannot be adjusted sufficiently, it will be necessary to move the opposite (fixed) machine case. Most calculator and laser systems provide the means to recalculate for base- and bolt-bound conditions. Documentation: It is essential to record pre-alignment data, target values, tolerances and final aligned condition. This information can help maintenance personnel determine when to perform maintenance tasks and spot developing problems that may otherwise result in unexpected, costly failures and downtime. The documentation features included with many alignment calculators and laser systems may not be adequate for recording data about foundations, machine feet, shims, coupling components and observations. If so, a paper or an electronic work order system should be used to capture this information (see an example data form below and click here for an enlarged version and a downloadable pdf).

Assessing Alignment Needs

Pump and alignment tool manufacturers offer "suggested" alignment tolerances, most of which do not consider the coupling type. But application-specific variables make a single tolerance for pumps unrealistic. The best approach is to evaluate each installation based on operating speed, thermal movement, spool-piece length and coupling type. For a large 1,200 rpm pump with an elastomer-in-shear coupling, alignment with a straight edge and feeler gauge to precision of 0.005 in would suffice. A medium-size 3,600 rpm hot water circulation pump with a jaw coupling, however, would require precision alignment with dial indicators or a laser system. A high-temperature (400 deg F) refinery pump may have a spool-piece coupling to accommodate thermal movement, and would require an operational assessment of target values. While target value assessment and precision alignment techniques could be used for any of these pumps, the required time and expense would outweigh the benefits for the 1,200 rpm pump. Thermal growth analysis also would be unwarranted on the hot water circulation pump, because the driver and pump will have similar thermal growth. Common sense dictates that evaluation of the alignment needs of individual pumpsincluding coupling type, thermal changes, rotating speed and spool piece length-will ensure use of the most cost-effective shaft alignment procedure.

Conclusion

Successful pump alignment requires careful planning and execution, beginning with evaluation of tolerances and target values based on pump speed, thermal characteristics, coupling type and spacing. The technician must also be adequately trained and systematically document the entire procedure, including the original misalignment, the final alignment condition, and any observations related to machine reliability. The expense or sophistication of the alignment tools-whether laser, dial indicator or manualwill not produce the desired results if these essentials are ignored. Eugene Vogel is a pump and vibration specialist at the Electrical Apparatus Service Association (EASA), St. Louis, MO, 314-993-2220, Fax: 314-993-1269, www.easa.com. EASA is an international trade association of more than 2,100 firms in 58 countries that sell and service electrical, electronic and mechanical apparatus

Understanding NPSH for the Non-engineer


Tuesday, November 18th, 2008

A lesson learned sixty years ago. When I was a little boy in the late 1940s my father drove a 1939 Ford coupe, new cars were not available because of the war effort. The Ford coupes were good cars for the day however the windshield defrost system was almost nonexistent. Everyone added a small rubber bladed fan to the dash that blew air across the windshield acting as a defroster. Most fans were powered by 6 volt systems as 12 volt systems were not yet standard on todays cars. Seatbelts were not even used in race cars let alone in passenger cars and kids road in the front seat when there was room. Many families only had one car which dad drove to work leaving mother home to take care of the house and kids. Many women didnt even have driver licenses as there was no need. However, women whose husband was fighting in the war did drive and work outside the home to support the family and war effort. Enough of history let me get back to the lesson. One cold winter day the family was going to town on Saturday for the weeks shopping trip and I got to ride in the front seat. Standing on the seat between mother and dad I was told not to put my fingers in the fan. Being a typical little boy what do you think I did? Of course I put my finger in the fan, now remember it was rubber so I didnt get hurt only a sore finger. Years later I related that experience to NPSH. How does a sore finger relate to NPSH? The amount of energy I used to push my finger into the fan. NPSHr (required) is like the energy it takes to push liquid into the impeller eye of the impeller and just past the vane tips. NPSHa (available) is the energy in the piping system that is used to push the liquid into the eye of the impeller and just past the vane tips. That is how a sore finger years ago relates to NPSH.

So for non-engineers you now have an explanation of NPSH that you can use to explain NPSH to an engineer.

How long will a mechanical seal last? Part one of this two-part series analyzes the first two perfor
Written by Fluid Sealing Association Pumps & Systems, March 2007 "How long will I be able to run my equipment before it has to be taken down for servicing?" is a question often asked when a buyer or user decides to purchase a new pump, agitator, or other type of rotating equipment. In some cases, a user may consider upgrading his existing pump to improve its reliability, capacity, efficiency or to bring it into compliance with federal or local emission regulations. The question will usually be the same. In many industrial pump applications, the seal is targeted as the most critical component. The seal usually will force removal of the pump from service because of an unacceptable leak rate or, in some cases, a level or pressure alarm. Although the seal is often blamed for the equipment failure, the real reasons are typically found in the operation and mechanical condition of the pump or the seals' environmental control system. Companies that have successfully implemented programs to improve Mean Time Between Repair (MTBR) are achieving three years or more average seal life, and consequently, have reduced the Life Cycle Cost (LCC) of their pump population. LCC projections for pump applications may vary widely based on the seal MTBR. In cases where the risk of failure is high, it is prudent to look for expert advice since errors in life estimates may have costly consequences. The key to any seal MTBR projection is to understand the factors that are most likely to drive the performance of the seal. Key Performance Indicators of a mechanical seal are normally the leak rate and power consumption.
Looking Into the Crystal Ball?

Predicting the service life of a mechanical seal is not an easy task, since the performance of the seal on the macro level is affected by many phenomena on the micro level. The reliability of a seal depends to a very large extent on its ability to maintain a thin fluid film in the gap between the mating faces while simultaneously minimizing the duration and extent of mechanical contact between asperities on the rubbing areas of these faces. Too much contact may overheat the materials; not enough contact may cause high leakage rates.
Key Components

Mechanical seal faces work much like bearings. The main difference is that the lubricant is usually the pumped fluid itself and may be dirty, volatile, viscous, toxic, or explosive while under pressure and sometimes at elevated temperatures. It is evident that the seal faces are the most vulnerable parts of any mechanical seal, but

other parts such as the secondary sealing elements (O-rings, bellows, polymer wedges) or metal parts such as springs, drive pins, or set screws also may affect the life of the seal when subjected to excessive movements or high temperatures and pressures. This article focuses on the factors that will affect the life of the seal faces. In order to make a reasonable assessment of the expected life of a face seal, one needs to be aware of some basics about how a mechanical seal functions. Remembering five key performance factors can serve to guide determination of an optimum replacement cycle for a mechanical seal. This month we will focus on the first two of these considerations.
Seal Face Wear

First, it is imperative to know that seals seldom leak excessively because the faces have completely worn. Should this occur, the narrower of the two faces will have worn to the extent that contact would have been lost and a leak path created for the fluid to escape to the atmosphere. In fact, normal wear rates of seal faces are extremely low and thus not a typical reason for excessive leakage. Seals that do wear out may be found in high pressure applications and fluids that cause abrasive wear of the face materials. In the majority of cases where seal removal is required, seal faces do show signs of distress. In these circumstances, the wear nose of the narrow face still has sufficient material, but the rubbing area of the faces is damaged to the extent that it caused an unacceptable leak rate. These distress symptoms can vary widely, from barely visible to the naked eye to severe damage such as the fracturing of a seal face. Common distress symptoms are pitting, blistering, chipping, grooving, heat and thermal shock cracks. Wear-out seal removals are the exception, since in such cases the equipment will most likely have been taken out of service for some other reason not related to the seal or simply for preventative maintenance reasons. Excessive mechanical seal leakage is often the symptom, i.e. visible evidence of a deeper problem in the equipment, process or control system.
Types of Seal Failure

Seal failures may be categorized on the basis of MTBR as either infant, wear-out or random type. Statistically, random failures (also called chance or mid-life failures) are the predominate type. Infant failures are most often caused by incorrect seal selection, installation or startup and have been practically eliminated with the introduction of cartridge design mechanical seals. Random failures tend to occur rather unexpectedly after a considerable run period and are typically characterized by a rapid change of the seal's leakage behavior from normal to excessive. Under operating conditions, this may manifest itself as a change from a non-visible leak rate to a drip or from a steady drip to a small stream of leakage in high pressure or speed applications. Random failures are quite difficult to predict because they usually are the consequence of a process operation or equipment-induced transient, which could not be tolerated by the face materials. Transient operating conditions may be induced deliberately or be the result of a malfunction of a component in the pump or an unintended process deviation.

Verification whether the seal can handle expected transients is a crucial step in the service life assessment because it is most likely the transient condition that may have a destructive effect on the seal faces. Applications that are continuous, steady state in nature are relatively easy as compared with cyclic applications. In fact, it is safe to assume that if the lubrication film at the seal faces is stable at all times, the seal faces will last virtually forever, and thus the seal will not be the driver of the equipment's service life. Application limits, minimum leakage requirements, and the seal environment also must be considered when estimating the life of a mechanical seal. Next month we will address these other three key factors that affect seal life and consequent need for replacement. Next Month: How long will my mechanical seal last? - Part Two We invite your questions on sealing issues and will provide best efforts answers based on FSA publications.

What Determines Seal Leakage? Written by Fluid Sealing Association

How can I extend the performance range of standard cartridge mechanical seals?
Written by Fluid Sealing Association Pumps & Systems, January 2009 In today's global environment, the "one size fits all" approach is used for so many products that we expect it to apply to just about everything. In the mechanical seal industry, one problem with that philosophy is the definition of the word all. Mechanical seals must live in literally every environment under the sun-from extreme hot to extreme cold, wet to dry, solid to gas, acidic to caustic and so on. Until relatively recently, a vast variety of mechanical seals were needed to accommodate these pump applications.
Cartridge Seals

The advent of standard cartridge designs allowed for easy, trouble-free installation on equipment, but they were limited as to the variety of applications each design could handle. With the introduction of universal or modular cartridge seals, one seal family utilizing various combinations of common cartridge seal hardware and seal faces can be applied virtually throughout a plant, increasing seal life. It reduces complexity while

conserving valuable resources time and money. The implications of this technology are far-reaching. Plant-wide seal standardization streamlines operations. The number of variables in installation and maintenance is reduced, and the process is simplified. This increased operational efficiency reduces maintenance costs. Streamlining the process also ensures that the seals are installed correctly, and correctly installed seals increase plant reliability. Using a standard seal family can also reduce a plant's spare parts inventory and material costs.
Modularity

Universal/modular cartridges are based on a modular concept that allows applicationspecific seal arrangements that incorporate interchangeable seal face material, elastomer bellows, O-ring or metal bellows seal heads fitted to standard adaptive hardware (typically the gland plate, sleeve, collar and setting tools). Universal or modular seals typically fit both standard ANSI B73.1M and DIN 24960 pumps, so users who are already comfortable with a particular seal type and have both ANSI and DIN pumps benefit from the technology. The flexibility from modularity also enables the universal cartridges to meet a wide range of service requirements.

Modularity allows the building of seals appropriate for a given application. By standardizing the adaptive hardware into which the appropriate seal is applied, the seal technology can change to fit the individual pump's environment. Handling temperatures to 400-deg F, pressures to 400-psig and speeds to 5,000-rpm, these seals accommodate fluids such as aqueous solutions, caustics, hydrocarbons, acids and even dry running.
Configurations

Basic configurations can cover an extensive range of applications. The unique design of universal/modular cartridge seals increases component interchangeability across a

variety of configurations. Single and dual arrangements typically incorporate the same seal head for both the inboard and outboard seal, which applies to both O-ring and metal bellows designs. In fact, the O-ring seal heads can be directly interchangeable with the metal bellows seal heads. This flexibility offers some important advantages. On site, a technician can easily upgrade or change the capabilities of the seal in minutes without discarding the existing seal. Within a common set of adaptive hardware, an O-ring seal can be transformed into a metal bellows seal, and a metal bellows can become an O-ring seal. With other combination of parts, either can change into an elastomer bellows design. Upgrading or changing the seal with interchangeable components conserves resources since no components are wasted. The heads and mating rings are also interchangeable in some designs. This flexibility gives the user the option to select whether the seal head or seat should rotate according to specific needs or preference. Operating with the inboard seal head rotating helps remove any solids build-up that creates, in essence, a self-cleaning rotational action. On the other hand, a rotating mating ring and stationary seal head accepts more gland or shaft misalignment, which allows higher shaft speeds and provides greater cooling efficiency. Single seal versions can typically accept a quench gland with a restriction bushing or lip seal, and dual seals can use pumping rings to assist barrier fluid circulation. This flexibility also makes universal/modular seals both ANSI and DIN pump compliant. The seal's gland is designed with slots to accommodate multiple bolts and bolt circle diameters. Pipe taps are freely accessible to allow universal fitting to either ANSI or DIN standard pumps rotating clockwise or counterclockwise.
Seal Design Advantages

Seal manufacturers continually enhance designs to provide maximum reliability, emission containment and cost-efficiency. Each of the basic versions-the O-ring pusher, metal bellows and elastomer bellows non-pusher-has inherent benefits. The O-ring pusher seal, perhaps the most versatile of seal designs, accommodates a wide range of elastomer materials and handles the most difficult chemical service conditions. The elastomer bellows seal is an excellent general purpose application seal, providing flexibility and hang up resistance as well as self-alignment, which is an important maintenance-saving feature. Automatic adjustment compensates for potential problems commonly found on older equipment such as abnormal shaft end play and runout. Advantages of the metal bellows seal include self-cleaning and increased reverse pressure capability. The reverse pressure capability feature found on most modern modular seal designs allows the use of the same dual seal for pressurized barrier or unpressurized buffer fluid applications, and will handle system upsets
Efficiency and Flexibility

The advanced technologies incorporated in today's universal/modular seals increase efficient seal operation and reduce overall costs. This is even more evident in the advanced capabilities now available, such as wet and/or dry running seals. This

advanced technology is perfect for applications that may run dry on occasion, such as tank car loading or unloading, or even full-time dry running for single seals and secondary containment on dual seals. Narrow-faced seals with rugged drive mechanisms and interchangeable hard (tungsten or silicon carbide) seal face material allow seals to perform longer in higher viscosity fluids and/or applications with a high number of starts and stops. Non-contacting spiral groove upstream pumping technology eliminates or significantly reduces the costs of complex, high pressure buffer fluid systems needed in dual seal systems. This technology reduces seal face heat, seal support systems cost and has proven to reduce operating and maintenance costs.

Summary

Modular seal designs do not lessen the need for proper selection of the core seal whether it is a single or dual, unbalanced or dual balanced, ID or OD pressurized, pusher or non-pusher, rotating head or seat and so on. A comprehensive design and materials selection for a given application far outweighs the benefits of modularity. However, pairing best practices with seal design selection and the latest universal/modular technologies offers plant and maintenance engineers the opportunity to create a safer, more productive plant. Installation and maintenance are simplified. The possibility of an incorrectly installed seal is reduced. Upgrades and repair can be made on site. These efficiencies translate to maximum seal reliability and a reduction in the time spent on seal installation, maintenance and repair. Not only does this represent cost savings, it also means more uptime and higher productivity plant-wide. Next Month: What is the impact of packing friction on equipment performance? Sealing Sense is produced by the Fluid Sealing Association as part of our commitment to industry consensus technical education for pump users, contractors, distributors, OEMs and reps. This month's Sealing Sense was prepared by FSA Member Mike Kraus. As a source of technical information on sealing systems and devices, and in cooperation with the European Sealing Association, the FSA also supports development of harmonized standards in all areas of fluid sealing technology. The education is provided in the public interest to enable a balanced assessment of the most effective solutions to pump technology issues on rational Total Life Cycle Cost (LCC) principles.

Where Mechanical Seals Meet Pumps: What Is the Next Generation?


Written by Fluid Sealing Association Pumps & Systems, August 2008 Almost all centrifugal and rotary pumps require a sealing system to provide sealing integrity of the drive shafts carrying the impellers and protect against pumped fluid leakage and the environment outside of the wetted areas. These sealing systems range from traditional braided materials packed around the shaft to complex mechanical seal systems used in many modern pumps.
Throat Bushings

Pump Standard API-610 requires that the pump casing contain a seal chamber in which mechanical seals compliant with API-682 can be mounted. A common component in both of these specifications is the so-called throat bushing, which is defined as: "A device that forms a restrictively close clearance around the sleeve (or shaft) between the (inner) seal and the impeller." The throat bushing can form a primary interface between the pump and the mechanical seal. According to the arrangement of the mechanical seal and the type of flush plan applied, such bushings are not always required. When they are used, the details of their design and use can be lost in the API-610 and API-682 standards. Here is a look at throat bushings in more detail.
Function

As its definition implies, the primary role of a throat bushing is to provide a restriction between the fluid being pumped by the impeller and the mechanical seal area, so that flow is reduced or controlled. Such a flow restriction can be required in either direction; for instance, the restriction would stop the flow into the seal chamber to avoid particulate contamination of the chamber. When the mechanical seal flush plan uses "processed fluid"-either cooled, filtered or physically different from the pumped mediathe flow from the seal chamber into the pump may need to be controlled. Throat bushings must be designed to provide the optimum environment for both the mechanical seal and pump since the fluid flow amount through a concentric annular restriction is proportional to the cube of the clearance and their axial length.
Clearances Critical

For pumps working at high pressures, fluid pressure control in the seal chamber, which provides loads on the mechanical seal faces, is important and often requires a very small throat bushing clearance. It is often debated how small the clearance should be. All parties agree that controlling the clearance between the shaft and throat bushing leads to increases in efficiency and performance of both the pump and mechanical seal since the isolation provided can allow optimum operation of both systems. Reducing the clearances between the throat bushing and the rotating element can often lead to an increased risk of contact between the two components, especially during times of start-up and any upset situations within the pump or mechanical seal. The

interface materials should be able to tolerate such contact without problems, including:

Galling or seizing of the two components, which can occur between metallic materials Particulate generation, which could possibly upset the micron level gaps found in the mechanical seal interfaces Heat generation, which can increase the temperature of the seal chamber In essence, the throat bushing must behave as a wear part, so its material selection must be considered in the same way as it's considered for a wear component within the pump. In addition to the bearing requirement for a throat bushing, effects like fluid erosion and thermal expansion differences between the pump, shaft and throttle bushing must also be considered.
Nonmetallic Composites

With micron levels of clearance between mechanical seal faces, high levels of vibration in the pump system can often be problematic. The use of composite wear rings to reduce these vibrations and assist with "Lomakin Effect" support of the rotating components has proven an asset to mechanical seals when used with composite wear ring equipped pumps. Extending the use of composites into the mechanical seal area can also provide similar benefits to those found in pumps. In the wetted area of the pumps, losses across the impeller eye can be reduced by closing the clearances between the wear gaps of the casing and impeller. The ninth edition of API-610 first recognized the use of nonmetallic wear rings, particularly carbon fiber reinforced Polyetheretherketone (PEEK) composites. Such materials have reduced the wear ring gaps below those possible with more traditional metallic systems and have done so without compromising the risks of contact-induced problems.

Figure 1. Typical pump and mechanical seal arrangement

The typical single stage pump and mechanical seal arrangement in Figure 1 shows the pump wear parts, impeller (A) and hub (B). The flow arrows indicate the direction of the media flow with pumped and flush (green) and barrier (orange) fluids in the system. The "joint" in the system is the throttle bushing (C) between the impeller and the mechanical seal chamber. As maintenance, reliability and efficiency become ever more critical and the costs of pumps and mechanical seals continue to increase, materials such as polymer composites can significantly improve system design. Other areas where such light weight, wear-resistant materials can provide tangible benefits within mechanical seals are:

Internal circulating devices such as those required in Plan 23 flush plans, where inertial effects and closer clearances may provide benefits Mechanical seal sleeves where thermal and centrifugal expansion of the supporting component must be matched with ceramic faces such as Silicon Carbide Throttle bushings located on the outboard of the mechanical seal to reduce the low leakages from the mechanical seal even further (see D in Figure 1) Labyrinth flow restrictors within the mechanical seal controlling and directing flow within the seal itself
Conclusion

The next generation of materials where pumps meet mechanical seals may already be here in polymeric composites. The pump and mechanical seal industries just need to recognize, define and standardize these materials to realize their benefits.

Features of Welded Metal Bellows Seals


Written by Jesse Fordyce, John Crane Inc. Pumps & Systems, February 2008 Welded metal bellows seals continue to have success as a core sealing technology and have gained popularity recently in new innovative sealing technologies such as high-temperature non-contacting gas lubricated seals and high-temperature corrosion resistant seals. This is especially important in the oil, gas and chemical industries where pumping liquid from one area to another is complicated by great temperature extremes. Unfortunately, not all seals are created equal. Differing features, materials and a host of other factors can impact the overall effectiveness of the seal. Only with an understanding of these differences can plant operators select the most effective product for the application. A welded metal bellows seal is made through a process of stamping disc-like plates in specific contoured shapes and welding them in pairs at the inside diameters to form individual convolutions of the bellows. A series of convolutions is then stacked and welded at the outside diameters to form the bellows capsule. Suitable end-fittings complete the assembly. The welded metal bellows assembly acts as a spring to keep the primary sealing faces together, acts as a dynamic seal and transmits torque from the set screw collar to the seal's rotating face.

Welded bellows have specific advantages, including:


Higher strength with the ability to withstand greater pressures Wider operating temperature range Ability to be given precise design characteristics Lower spring rate (the amount of force required to compress it a given distance) Lower stress in critical areas Welded bellows allow for the use of optimal plate shapes such as the Nesting Ripple design Static secondary seals Only one moving part. . . the bellows

If you've discerned that an edge-welded metal bellows seal is optimal for your application, consider that not all seals are created equal. Differentiating bellows features-the plate shape and thickness, vibration attributes, the impact of double-ply, face angle and more-all impact the effectiveness of the welded metal bellows seal. Operators should understand these differences in order to pick the most effective product for their application as well as extend mean-time between repair, standardize inventory, increase reliability and improve fugitive emission control and water conservation. The following are some key distinctions and features of today's welded metal bellows:

Plate Shape

The plate shape influences flexing, stroke and operating length. In the nesting ripple configuration, all plates in the bellows are identical and contoured to permit nesting when compressed. Contouring also improves the bellows' ability to withstand high pressure. The nesting ripple plate shape is more effective in achieving maximum flexing, long (axial motion) stroke with short operating lengths and a low spring rate. The sweep radius is optimized at 20-25 percent of span and it prevents a phenomenon known as oil-canning-the inversion of the plate geometry that results in a bulging in and out of the plate, similar to that on the bottom of an oil can when it is pressed. Each convolution is made up of a male and female plate, which allows the seal to be designed with a short axial space.

Angle

For bellows with straight flat segments, the variability of the microstructure in the heataffected zones results in less reliable weld joints. By theoretical analysis using linear, thinshell theory, it has been shown that tilting the bellows axis drastically reduces stresses at the welds and heat-affected zones. The analysis indicates that the stresses at the welds were predominantly bending stresses. With increasing tilt angles, the bending stresses are lowered. This design principle also has been thoroughly documented in both theoretical and empirical studies conducted by an independent government-sponsored agency and verified experimentally. With a 45-deg tilt angle, bending stresses are directed away from the heataffected zone of the weld. This results in plate rigidity, which adds reliability and reduces fatigue.

Weld Integrity

State-of-the-art manufacturing processes ensure integrity of the weld by preventing excessive root gap with bead geometry, bead thickness and roll-over control. Bellows units should be checked for leak tight performance with helium mass spectrometry and vacuumtested to 10-6 TORR. With helium mass spectrometry, the seal is evacuated internally and blanketed in helium. Traces of the gas, which then penetrate through either a break in the weld or a material flaw, are immediately picked up by the sensing probe and the seal is rejected.
Plate Thickness (Thin Plates)

Thin plates provide lower spring rates, which result in lower face loads, less unit loading, less heat generation and longer life than thicker plates. Thicker bellows plates have higher spring rates and are more susceptible to metal fatigue. Repeated plastic deformation of the plates (beyond their plastic limit) during deflection can result in fatigue and greatly reduce seal cycle life. Increasing the thickness of the plate material, though easier to weld, increases its stiffness and increases the spring rate of the bellows significantly. A high spring rate is undesirable because of the significant changes in loading of the sealing faces with only slight changes in

the operating length of the seal. This causes excessive closing force, which in turn causes loss of the lubricating film between the sealing faces, excessive face heat and eventual seal failure. This is especially critical in high temperature and poor lubricating environments. Bellows with high spring rates are also less capable of compensating for installation problems, shaft movements, impeller adjustments, pump end play, shaft growth due to heat and gradual wearing of the sealing faces. Thinner plates are more difficult to manufacture, which is why many manufacturers are forced to use thicker plates.
Controlling Vibration

Not all bellows in the industry are fitted with vibration dampeners, which help prevent any potential damage from harmonic vibration caused by episodes of dry-running. However, a vibration dampener is ideal. In certain seal designs, the vibration damper pad is a built-in design feature which allows protection-if and when it's needed-against vibrations of a potentially damaging nature.
Double-PlyTM Bellows

Double-ply bellows are typically utilized in higher pressure applications and often used in services in which the fluid is thermo-sensitive or has a tendency to set-up and solidify on the seal faces where more start-up torque strength may be required. The double-ply design principle can be illustrated by a simple leaf spring of the type used in light trucks and haulit-yourself trailers. A spring, comprised of a single, thick, metal member with the strength necessary to support the load would result in too stiff a spring. But, when the strength is obtained by using a "stack" of separate, individually-flexing thin leaf elements, the spring rate is well within desired limits. Similarly, the spring rates of two-ply bellows proved to be significantly lower than those of single-ply bellows with twice the plate thickness.

Pressure and Seal Balance

Before you purchase your metal bellows seal, be sure to discern the proper pressure rating required by your application in relation to temperature, speed and sealed fluid lubricity. Determine if the seal is capable of handling reverse pressure, especially in dual pressurized seal arrangements. Metal bellows seals are hydraulically balanced by locating the effective diameter of the bellows with respect to the seal face. Most standard bellows seals are typically balanced to approximately a 70/30 ratio. This means that 70 percent of the face contact area is above the effective diameter. More recently, 50 percent balanced bellows seals have been designed to handle both inner diameter (I.D.) and outer diameter (O.D.) pressure. Fifty percent balanced seals are able to handle reverse pressure, which is an upset condition that may occur in dual seal operation. Pressure-balanced by design, the welded metal bellows seal does not need a step in the shaft or sleeve required to balance the seal.

Additional Welded Metal Bellows Design Features:

Plate Span: Narrower plate spans typically provide greater stability under pressure than wider spans. Face Width: Narrow face width typically results in less heat generation at the seal faces. Less heat improves face stability and narrow faces are less susceptible to coking than wider faces. Multitude of Metallurgies and Face Materials Available: Be sure your manufacturer offers an extensive array of materials such as AM-350, Hastelloy-C, Inconel 718, Alloy 20, Monel and Titanium. Design Characteristics: Consider not just the bellows seal head assemblies, but also the cartridge seal designs. Compare design features such as inside- or outside-mounted, reverse pressure capability, API 682 designs and high-temperature corrosion resistant designs.

Conclusion

Since metal bellows products vary greatly in what they are able to offer, it is important to understand the benefits and trade-offs as they apply to your particular operation. Be sure to review your applications and select the metal bellows seal best suited to help you extend meantime between repair, standardize inventory and increase reliability. Jesse Fordyce is the oil & gas market manager for John Crane Inc., 6400 West Oakton Street, Morton Grove, IL 60053, 847-967-3505, Fax: 847-967-3911, www.johncrane.com.

How can I reduce consumption of seal water going to my packing and mechanical seals?
Written by Fluid Sealing Association Two imperatives for many of today's industrial plants are to reduce the cost of operations through the enhancement of rotating equipment reliability and enhanced energy efficiency of pumping systems. One place to look for a significant, yet relatively easy "quick win," is the seal flush water going to packing, single, and...

Two imperatives for many of today's industrial plants are to reduce the cost of operations through the enhancement of rotating equipment reliability and enhanced energy efficiency of pumping systems. One place to look for a significant, yet relatively easy "quick win," is the seal flush water going to packing, single, and double seals. In many industrial plants water is being used to provide lubrication, cooling and/or as a means to exclude a harmful process fluid from the stuffing box or seal chamber.

The means for providing an external water flush or quench are generally described as API/ISO piping plans 32 (ANSI 7332), 54 (ANSI 7354) or 62 (ANSI 7362). These plans have certain potential issues that can be opportunities for improving operating costs, namely:

Higher flow rates than required for optimal packing or seal performance. Lower flow rates than required for optimal packing or seal performance. Orifices plugging with no readily accessible means to clean them. Cumbersome and space consuming piping configurations in order to include basic requirements of an appropriate system, such as pressure and flow control as well as monitoring, check valve and low flow alarm.

A potential solution to address these issues would be a single compact unit that includes all of the functionalities below:

Allows the end-user to readily optimize seal water pressure and flow to maximize packing and mechanical seal MTBR. Enables the end-user to readily monitor the seal water pressure and flow. Conveniently and economically incorporates a low flow alarm while still maintaining a minimal piping footprint. Saves space. Typically will reduce seal water consumption by about 1/3rd while optimizing packing and mechanical seal performance. Reduces process dilution. Can enhance rotating equipment reliability by alerting plant personal to a low flow condition that could lead to an unplanned maintenance event. Easy to maintain with readily accessible cleaning button that does not disturb flow or pressure.

Water quality is an important consideration. Plant water quality can vary significantly from one location to another and affect these units, so it should be monitored.

Inspection and maintenance procedures should take water quality into consideration. It also is always best to consult with your seal supplier to ensure optimum cooling flow rates for your system.

The table to the right is an example of a way to achieve significant operating savings through optimizing seal water flush. The basis is reduction in seal water consumption by an average of 1-gpm per stuffing box served. The bottom line on how to reduce consumption of seal water going to packing and mechanical seals is to consider replacing your existing seal water flush piping plan with a single unit seal water flush control and monitoring device. You will save operating costs as well as space.

The Most Effective Seal Design When Misalignment Is Present


Written by Fluid Sealing Association From the Fluid Sealing Association

Pumps & Systems, March 2008 Shaft misalignment is one of the most common causes of premature mechanical seal failure. Higher leak rates typify these failures. Rapid failures can occur in high pressure and speed applications. Misalignment may be static or dynamic in origin. Static misalignment is measurable when the equipment is not in operation, while dynamic misalignment is detectable only during operation. Static misalignment is most prevalent and will be our focus. Causes of static misalignment include parts that are out of tolerance, static shaft deflection and improper installation. Evidence that misalignment was the cause of a seal failure includes one or more of the following: broken springs, worn and/or extruded dynamic gaskets, fretting corrosion of metal surfaces adjacent to the dynamic gasket or worn drive mechanisms. No single mechanical seal design can fully compensate for poor alignment. Some designs are more forgiving of misalignment than others, but there can be performance compromises when selecting the seal design that best compensates for poor machine alignment. The first priority should always be to correct the misalignment so optimum seal performance can be achieved.
Causes of Misalignment

Seal housing misalignment may be caused by a permanent bend in the shaft, excessive shaft deflection or by misalignment between the rotor assembly and the seal housing. Determining shaft condition should be the first step when checking for misalignment at the seal housing. Bent shaft A bent shaft is easy to detect but often difficult to repair. To determine if a shaft is bent at the seal housing, mount a dial indicator on a stationary surface and measure the runout on the shaft sleeve while turning it 360-deg. Total Indicated Runout (TIR) of more than 0.002-in (0.05-mm) warrants a repair or replacement of the pump shaft. It is possible for the shaft to be bent at a location that will not be evident when checked at the shaft sleeve. A shaft also should be checked for straightness if vibration readings indicate an out-of-balance condition. The detection procedure is the same. Mount the indicator on a stationary surface and take a reading from the shaft while turning it. It is usually less expensive and more reliable to replace a bent shaft than to straighten it. Shaft Concentricity and Parallelism Misalignment from a lack of concentricity or parallelism can be measured by a dial

indicator on the shaft. While rotating the shaft and dial indicator, measure runout on the seal housing bore or register for concentricity at the seal housing face for parallelism (angularity). If the seal housing concentric runout is greater than 0.005-in (0.125-mm) or the face runout is greater than 0.0005-in/in (15m/3cm) of seal housing bore diameter, further investigation and corrective action is warranted. Potential causes of an out-oftolerance seal housing runout reading include:

Excessive static shaft deflection Out-of-tolerance assemblies Deformation of the pump assembly due to high structural loads

Tolerances Misalignment from the tolerances of the assembled machined components is harder to detect. There are usually several register fits between the bearing supports and the seal housing. These fits need to be machined concentric and square. Measurements should be made on a machine. Although an individual register may appear to be out of tolerance by only a small amount, it is important to remember that the total misalignment of the assembled parts will be the sum, or stack up, of the individual component misalignments. Register fits are subject to wear and corrosion over time and should be checked as part of any major maintenance. Mounting Many pumps have four or more mounting pads for mounting the pump to the baseplate. A pump often does not have 100 percent contact with all of the baseplate mounting surfaces. This is often called a soft foot condition. Bolting a pump to a base with a soft foot condition may result in misalignment of the pump shaft to the seal housing and place stress on the bearings and other components. To check for soft foot, mount a dial indicator on the baseplate and take a measurement from the pump adjacent to a hold-down fastener. A change of the indicator of more than 0.002-in. (0.05-mm) while tightening or loosening a hold-down fastener indicates a soft foot. Use stainless precision shims between the baseplate and pump to compensate. It is generally good practice to monitor shaft runout at the seal housing while installing a pump on its baseplate and connecting the pump to its piping. Flanges Mating flanges should be concentric and parallel. Bolting of misaligned piping to a pump will cause both the piping and the pump to distort. The amount of distortion is relative to the size and strength of the pump and piping, as well as the distance from the nearest fixed hold down. A good rule of thumb is that if mechanical leverage is needed to bring the flanges into position, then the nozzle loads should be checked. If the piping connections are suspected, mount a dial indicator to check for shaft alignment at the seal housing to check shaft angularity and parallelism. If connecting or disconnecting the piping changes the shaft alignment, corrective action should be

taken.
Seal Designs

Most seal designs are manufactured to operate with a maximum angular misalignment of 0.003-in. (0.08-mm). This includes the out-of-square tolerance and the shaft-to-seal chamber bore concentricity. There are two primary designs used to accommodate misalignment. One design utilizes a rotating compression unit where the springs rotate with the shaft (see below).

The other is a stationary design (see below) that uses a stationary compression unit where the springs do not rotate with the shaft. The stationary design compensates for misalignment with one adjustment, while the rotating design must compensate at every revolution of the shaft. This movement predisposes the spring mechanism to fatigue, which can cause premature failure. The stationary design is the better choice when misalignment occurs.

The same concept also applies to welded metal bellows technology. Utilizing a stationary bellows allows for a one-time reposition of the stationary face, eliminating constant flexing/movement of the thin cross section bellows leaflets. Most of the seal manufacturers use the flexible stationary design in their modular or cartridge seals because of space limitations. The excessive flexing and fatigue of the many thin cross-section springs is avoided, as is contact with the process fluid. Stationary designs are gaining acceptance for several other reasons. First, they keep the small multiple springs out of direct contact with the product being pumped. Many fluids contain solids which can clog the small multiple springs, hang up the seal, and cause premature failure. Second, high speed applications-those with shaft speeds in excess of 4,500 feet per minute (23-m/s)-usually require special design attention. Above this speed, dynamic forces begin to exceed the limitations of a conventional rotating design. Utilizing the stationary design reduces the secondary seal and drive mechanism movement, which could cause excessive wear and fretting damage. The stationary design also provides better seal face tracking capability and improves seal life. Further, the drive mechanisms can be increased to handle the higher torques associated with these speeds.
Summary

Mechanical seal failures are usually the symptom and not the cause of maintenance problems, many of which are related to static misalignment. In operation, other forces such as radial loads, impeller balance and cavitation also affect seal alignment. While some seal designs are more tolerant of misalignment, they ultimately are not the solution for a misaligned system. Optimum performance, which includes maximum seal life, can only be achieved by correcting the root cause of the misalignment. Always try to obtain more detailed information about suitable seal designs from the seal manufacturers.

Circulation Systems for Single and Multiple Seal Arrangements (PartTwo)


Written by Gordon Buck and Ralph Gabriel, John Crane Inc. Article Index Circulation Systems for Single and Multiple Seal Arrangements (Part Two) undefined All Pages Page 1 of 2 Plan 12

Plan 12 is similar to a Plan 11. The flush is taken off of the pump discharge, or and intermediate stage in the case of multiple stage pumps, through a strainer or filter to remove solids, and then through an orifice to control flow, before being introduced into the seal chamber. API 682/ISO 221049 does not recommend this plan, as problems can arise if not closely monitored where the strainer can become clogged, the flush flow is lost, and the seal is damaged due to overheating. This can be avoided by using a differential pressure indicator or flow indicator to alert the user of impending problems. Some strainers utilize magnets or magnetic strainers to attract metallic particles like magnitite that will be present in most water systems.
Figure 1. Seal Flush Plan 12

Advantages

No product contamination from an external source. Relatively simple piping plan. No reprocessing of product.

Solids are removed from flush stream keeping the seal chamber clean.

Disadvantages

If the product in the pump is not a good face lubricant, or has extremely low or high viscosity, the seal can become damaged depending upon the seal face material combination. Flush is recirculated. Strainer or filter will plug over time. Plan 13

Plan 13 is a off-shoot of a Plan 11, where the flow comes out of the seal chamber and goes back to the pump suction. This also helps the seal to vent gas out of the seal chamber. Typically, this plan is used on vertical pumps where the seal chamber is subject to pump discharge pressure at the top of the pump. However, this plan can be used on horizontal pumps, depending upon the type of impeller used and the differential pressures between seal chamber pressure and pump suction pressure. It is useful on high differential pressure applications where the use of a Plan 11 would require the use of multiple orifices. On the other hand, for small or low speed pumps that have a low differential pressure, no orifice is required. Due to the flow path, Plan 13 is not as effective as a Plan 11 in removing seal generated heat. The path is such that it comes from the back of the seal and rises up to the gland plate outlet port with no direct impingement on the seal faces. In some cases, flow diverters may be incorporated to improve the flow path or flush rates can be increased to make up for the decrease in efficiency.
Figure 2. Seal Flush Plan 13

Advantages

No product contamination from an external source. Relatively simple piping plan. No reprocessing of product. Continuous venting of the seal chamber.

Disadvantages

If the product in the pump is not a good face lubricant, or has extremely low or high viscosity, the seal can become damaged depending upon the seal face material combination. Flush is recirculated. Less efficient flow pattern. Plan 14

Plan 14 is a combination of a Plan 11 and a Plan 13. The flush is taken off of the pump discharge and sent to the seal chamber like a Plan 11. A second set of piping takes the flush from the seal chamber, sending it back to pump suction like a Plan 13. It is often used in vertical pumps to provide adequate flush flow and vapor pressure margin independent of the throat bushing below the seal chamber. It is used on viscous products to provide a flow path out of the box in addition to the throat bushing that can be restrictive. It is also an effective plan when the throat bushing that partially controls flow rates is inadequate for the seals' needs.
Figure 3. Seal Flush Plan 14

Advantages

No product contamination. No reprocessing of product. Optimized cooling. The flush flow can be controlled so that the cooling is directed at the faces and adequate flow is maintained. Allows complete automatic venting provided that the "FO" port in the gland is properly located. With a properly sized orifice and throat bushing, the seal chamber pressure remains high, resulting in an adequate vapor pressure margin.

Disadvantages

If the product in the pump is not a good face lubricant or is dirty, the seal can become damaged or clogged. Flush is recirculated.

Sizing For the above three plans, the flush rate should be calculated based on the pumping conditions to maximize efficiency and seal life. For applications above 3600-rpm or box pressures above 500-psig, the flush rate should be calculated to avoid excessive temperature at the seal faces. In lower pressure/speed applications, a "rule of thumb" of 1-gpm per inch size can be used. Controlling The flow rate is controlled by an orifice or series of orifices in the flush line. API 682/ISO 21049 states that the orifices should not be less than 1/8-in. The orifices do not all have to be the same size and can be larger if there is a possibility that the flush stream may clog smaller orifice sizes. Another method is to use a "choke tube." This is a piece of tubing generally -in heavy wall. The length of the tubing is calculated using a piping pressure drop calculation such that the pressure drop across the tubing is equal to the difference between the discharge or suction pressure (depending upon the specific plan) and the seal chamber pressure at the flow rate desired.

Circulation Systems for Single and Multiple Seal Arrangements (Part Two) -undefined Written by Gordon Buck and Ralph Gabriel, John Crane Inc. Article Index Circulation Systems for Single and Multiple Seal Arrangements (Part Two) Page 2 of 2 Plan 21/22

Plan 21 is again an off-shoot of a Plan 11, where the product is taken from pump discharge and directed through an orifice to a heat exchanger to reduce the flush temperature before being introduced into the seal chamber. Plan 22 is the same as a Plan 21, with the addition of a strainer located before the orifice.

A temperature indicator should be included on the process side of the exchanger, normally on the downstream side. Additional temperature indicators are used in some installations to monitor process and cooling water temperatures on both sides of the heat

exchanger. Plans 21 and 22 are not preferred plans by the API and by many users. This is due to the high heat load that is placed on the heat exchanger in these plans. For example, taking 3-gpm water from 350-deg F down to 160-deg F consumes 270,600-Btu/hr. These high heat loads result in wasted energy and high rates of fouling of the heat exchanger on both the process and cooling water side, which often results in shortened seal life. Fouling normally occurs on the water side of cooling tower based cooling systems. This can be averted to some degree by maintaining a velocity that is sufficient to resist the deposit of cooling water sediment on the coils. The use of a throat bushing will reduce the effects of heat soak into the seal chamber and can also increase the seal chamber pressure slightly assisting in increasing the vapor pressure margin. The standard pump fixed throat bushing may be utilized on lower temperature applications where the process fluid is not volatile. A floating throat bushing will reduce the clearance by more than half over a fixed bushing and is recommended on applications involving higher temperatures and process fluids with high vapor pressures. Plants typically have general rules on temperature ranges and vapor pressure margins that they want to maintain. Both water- and air-cooled heat exchangers may be used with these plans. The use of an air-cooled heat exchanger has the advantage of eliminating water side fouling, but is less efficient and may not be suitable for low pressure differential pressure applications.
Figure 4. Seal Flush Plan 21

Advantages

No dilution of process stream. Provides cooled process to the seal chamber, improving lubricity and increasing the vapor pressure margin. Can be applied to any pump where Plan 11 can be applied.

Disadvantages

High heat load on the heat exchanger, resulting in high operating costs with the potential for fouling. Some fluids may congeal or become highly viscous, during idle periods, if flow is not maintained through the heat exchanger. Flush is recirculated.

Sizing The flow rate should be determined in a similar manner to that done for Plan 11. Depending on the process fluid, a desired injection temperature should be determined. For hydrocarbon based fluids, API 682 recommends 36-deg F below the vapor point. For water and other aqueous solutions, most seals require that the temperature of the flush being introduced to the seal chamber be maintained below 180-deg F. The heat exchanger should be sized based upon the above flow rate calculation, the capability of the cooling circuit along with some safety factor for fouling. Controlling The flow rate should be controlled by an orifice or series of orifices in the flush line. Improved results for process fluids near their vapor pressure can be realized by locating the orifice(s) between the heat exchanger and the seal chamber. In this arrangement, the fluid is first cooled, resulting in a lower vapor pressure before the pressure drop across the orifice occurs.
Plan 23

Plan 23 is a closed loop circulation system used on hot applications for providing cooled flow to single seals. In Plan 23, a pumping ring in the seal chamber circulates product through a heat exchanger and back to the seal chamber. A throat bushing is used to isolate the cool seal chamber from the hot pump. The use of a close clearance throat bushing is recommended to minimize the mixing of the hot process fluid with the cooler seal chamber fluid. In the past, the primary use of Plan 23 systems has been hot water services, but recently it has been become more popular in refineries, where it has been used on hydrocarbon services. In comparison to a Plan 21 or Plan 22, the heat load on the heat exchanger is considerably less. The heat load consists of heat soak from the pump, heat generated by the seal faces, along with turbulence or churning of the pumping ring and seal head within the seal chamber. As a comparison, with a 350-deg F water application with a 3-in seal at 3600-rpm and 500-psi, the total heat load is 13,300-Btu/hr versus the 270,600-Btu/hr noted above for a Plan 21. The breakdown of the Plan 23 heat load is 6200-Btu/hr for seal generated heat, 6,840-Btu/hr for heat soak, and 260-Btu/hr for turbulence.

Figure 5. Seal Flush Plan 23

Advantages

The process fluid is used to cool and lubricate the seal. The reduced operating temperature improves lubricity and reduces the possibility of vaporization in the seal chamber. Plan 23 is very efficient versus the alternative Plan 21. The cooler is less likely to scale or foul. Less direct cooling of process fluid than Plan 21. Can provide a cooled seal chamber thru effective thermosyphon effects when the pump is idle.

Disadvantages

The initial cost is more than Plan 21 because of ancillary components. Plan 23 is not used for fluids with high freeze points or for viscous fluids because the pumping ring may not be able to force circulation. Venting is essential for Plan 23. Selection, design/location of the pumping ring, inlet and outlet ports, and piping are crucial to the successful operation of Plan 23. Plan 31

In Plan 31 the product is introduced to an abrasive separator from the discharge of the pump. The clean fluid is routed out the top of the separator and into the seal chamber, while the process with the heavier solids is routed back to pump suction. This plan should only be used for services containing solids that have a specific gravity at least twice that of the process fluid. Throat bushings are a requirement when using this flush plan. In this case, the throat bushing separates the clean fluid in the seal chamber from the dirty process fluid. The rule of thumb is that the velocity of the fluid passing through the throat bushing be on the order of 15-ft/sec to prevent the particles from entering the seal chamber. The flow rate of the clean flush will dictate the type of bushing required, where floating type bushings with close clearances should be used on low flow rate applications.

In an abrasive separator, the fluid containing the abrasives are fed into the inlet at the top of the cylindrical cone directed tangentially to the wall of the separator at a velocity sufficient to create a spiral or vortex action. The developed centrifugal force from the rotation of flow throws the heavy abrasive particles to the wall of the separator, where the abrasives collect and pass downward and out of the unit through the separator discharge port. The clean fluid moves to an inner spiral and is displaced upwards and out through a vortex finder located at the top center of the unit. The proper installation of the abrasive separator is a necessity for this flush plan to work. Typically, the separator needs to have a minimum of a 15-psi differential to operate properly. The maximum particle size should be less than one-quarter the size of the inlet orifice. It is important that the lines running from the clean outlet at top of the separator and the dirty outlet at the bottom be at similar pressures to obtain proper separation and flow rates. It may be necessary to vary the length of tubing or piping or even to add an orifice to one or both discharge lines to get the proper flow established. The flow rate can be controlled by an orifice in the line running from the pump discharge to the inlet of the abrasive separator. This orifice should not be less than .125in and should usually be larger, depending upon the size and the concentration of the particles being removed. Either multiple orifices or a "choke tube" can be used to control the flow. On vertical pumps where the seal chamber is at discharge pressure a separate line needs to be added, going from the bushing at the bottom of the seal chamber back to suction to reduce seal chamber pressure to allow the clean fluid to enter the seal cavity. On small or low differential pumps no orifices are required.
Figure 6. Seal Flush Plan 31

Advantages

Solids are removed from the flush stream keeping the seal chamber clean. Unlike a strainer or filter, the abrasive separator does not have to be cleaned.

Disadvantages

It is sometimes difficult to obtain the desired pressure differential required for the abrasive separator to operate efficiently. Exceeding the published differential pressure will cause the separator to not function properly. Improper piping will cause the separator to not operate efficiently. Unless the pressure differential from the two discharges to the final sources are almost the same, the separator can either starve the seal or allow abrasives to flow into the chamber. The abrasive separator and the piping in the dirty outlet leg can become worn over time from the abrasives spiraling down the coned shaped bore. Not advisable on low vapor margin applications, as vapor bubbles have a more natural inclination to be channeled into the seal flush connection. Less effective with viscous fluids

Sizing Generally the flush rate sizing will be the same as a Plan 11, keeping in mind that only the clean fluid flows to the seal. Various models or sizes of separators are available to provide different flow rates depending upon the pressure differential available.
Plan 32

In Plan 32, the flush stream is brought in from an external source. This plan is almost always used in conjunction with a close clearance throat bushing. The bushing can function as a throttling device to maintain an elevated pressure in the stuffing box or as a barrier to isolate the pumped product from the seal chamber. Plan 32 is used when a process stream is difficult to condition in a way that will provide adequate cooling and lubrication to the mechanical seal. In addition, it is often employed when a process stream includes components, which may either result in abrasive wear or will impede free movement of critical seal components. The design of a Plan 32 flush system involves application of hardware and logic that will provide the seal with an environment conducive to long term service, while not compromising the operation and profitability of the process stream.

Figure 7. Seal Flush Plan 32

Advantages The external flush fluid, when selected and applied properly, can result in vastly extended seal life, resulting in improved MTBPM for the pump system. Disadvantages

Product degradation or dilution will occur when using this plan. Depending on overall system design, introduction of an external fluid to the process stream can result in increased energy and reprocessing costs. Support system costs can be very high and adds additional equipment to the system, which must be in operation whenever the involved pump is on-line.

Sizing The flush rate is critical to any seal, but takes on another dimension when Plan 32 is involved. When an outside flush source is used, concerns regarding product dilution, and/or economics almost always surface. For these reasons, it is imperative that the seal supplier be adequately informed with regards to any limitations that will be placed on the flush rate. With respect to the flush rate, three common scenarios should be considered. In all cases, the flow rate required to cool the seal should take precedence:

Exclusion of process from the seal chamber is the primary objective. In this case, a close clearance, floating throat bushing should be placed in the back of the seal chamber. The flow rate which will normally achieve solids exclusion is 15-ft/sec. The process is at or near its boiling point and other flush plans are not practicable. Processes in this category are often hot, thus heat soak is also a consideration. When calculating the flow rate required, one should not allow more than one-quarter of the available flush fluidtemperature margin to be used. The process is not at its boiling point, though it has properties which adversely affect seal life. This scenario is common in process streams which have a tendency to polymerize, congeal, or set up at various stages in a batch process. In this situation, simply diluting the process is often

all that is required to maintain reliable seal operation. The required flush rate may be quite low, and often actual flush rate determination may be derived more from experience. In this case, knowledge of the process and its interaction with the flush stream are key to success.

Controlling Conditioning and controlling the rate of flush in a Plan 32 system can range from simple and inexpensive to elaborate and costly. Simpler is usually better whenever possible. Finally, the device selected to control the rate of flush is the most critical decision to be made, some methods are:

A drilled orifice or choke tube is the simplest device and will normally be the least costly. However, careful attention must be paid to protecting the orifice from plugging, particularly if the size is less than .125-in. Also, if the supply pressure and/or box pressure are not well determined, or are variable, accurate sizing may not be attainable. A manually adjustable needle or globe valve with flow indicator can be employed, thus allowing "on the fly" control of the flush rate. This approach will allow accurate tuning of the flush rate, though consistent monitoring is required. A control valve is the ultimate control of flow rate, though cost can be substantially higher to purchase and maintain such a device.

General Communication between the seal company and the end user is key to success of Plan 32. As with any application, the fluid properties and service conditions of the process stream must be established in order to provide a successful seal design. However, to ensure success on Plan 32 applications, fluid properties of the external flush stream should also be presented with the main process details. When Plan 32 is applied to process streams which are hot, the flush liquid may have a higher vapor pressure than the process fluid when at the same temperature. In short, the introduction of a lower boiling point liquid into the process stream will lower the NPSHA at the impeller to some degree and will have a negative effect on pump capacity as the liquid vaporizes in the pump. In the worst case, the pump may vapor lock or be damaged by the resulting disturbances. In this case, the flush rate is often "fine tuned" in a way that provides adequate seal cooling and minimizes vaporization within the pump. Plan 32 is not recommended for cooling only, as the energy costs can be very high. Further, material selections made for the seal should be made based on the extremes of the process fluid conditions and the external flush fluid, not the flush by itself.
Plan 41

Plan 41 is a combination of flush Plans 21 and Plan 31. The flush comes off from pump discharge, is first directed through an abrasive separator to eliminate solid particles, and then goes to a heat exchanger to reduce the temperature before being introduced into the seal chamber.

Optional accessories are an orifice to control the flow and a temperature indicator on the

product outlet side of the heat exchanger. In some installations, temperature indicators are also used to monitor cooling water temperatures. This Plan should only be used for services containing solids that have a specific gravity at least twice that of the process fluid. Throat bushings are a requirement when using this flush plan. The flow rate can be controlled by an orifice in the line running from the discharge to the inlet of the abrasive separator. This orifice should not be less than .125-in and should usually be larger, depending upon the size and the concentration of the particles being removed. Either multiple orifices or a "choke tube" can be used to control the flow. On vertical pumps where the seal chamber is at discharge pressure a separate line needs to be added, going from the bushing at the bottom of the seal chamber back to suction to reduce seal chamber pressure to allow the clean fluid to enter the seal cavity. However, in some cases, the pump manufacturer uses the top bushing as a balance piston, so the pump manufacturer should be consulted. On small or low differential pumps no orifices are required. It is important that the lines running from the clean outlet at top of the separator and the dirty outlet at the bottom be at similar pressures to obtain proper separation and flow rates. It will be necessary to vary the length of tubing or piping in the dirty discharge from the abrasive separator back to pump suction to obtain the same pressure drop produced by the heat exchanger. This flush plan can be very difficult to pipe properly to obtain the correct flow rate through the abrasive separator and the heat exchanger. Due to this constraint and the potential for high heat loads as noted with Plan 21, it is not a popular option. As with Plan 21, the abrasive separator needs to have a minimum of a 15-psi differential to operate properly. The maximum particle size should be less than one-quarter the size of the inlet orifice.

Figure 8. Seal Flush Plan 41

Advantages

Solids are removed and product temperature is reduced to enhance the seal's environment. Unlike a strainer or filter the abrasive separator does not have to be cleaned.

Disadvantages

This plan is not suitable for very low head services as the pressure drop thru both the abrasive separator and heat exchanger may be too great. Piping this arrangement to get the proper pressure drops in order to get efficient operation of the abrasive separator and the correct flow through the heat exchanger is difficult. Exceeding the published differential pressure will cause the separator to not function properly. Depending upon the temperature of the process the heat load on the heat exchanger can be high, resulting in high operating costs for the cooling water and/or fouling of the heat exchanger. Plan 62

Plan 62 is a common flush plan to improve the environment on the atmospheric side of single seals. Typically, this is either low pressure steam or nitrogen to prevent coke formation on hot hydrocarbon services, water to prevent the formation of crystalline substances on fluids with solids in solution, or nitrogen to prevent icing on cold or cryogenic applications. It is typically used with a floating or segmented bushing to limit the leakage of the quench fluid to atmosphere, but can be used with a fixed bushing if the application permits. Information on quenches was introduced in the first article of this series.

Figure 9. Seal Flush Plan 62

Advantages

Low cost alternative to tandem seals to improve condition on low pressure side of process seal.

Disadvantages

Leakage of process past primary seal is not contained except with the throttle bushing. Leakage can then leak to atmosphere or go to a drain. Improper control of steam can allow condensation to form that can boil and cause seal damage on hot processes. Poor steam control can lead to a reverse pressure on the seal and/or bearing oil contamination. Plan 65

Plan 65 is a leakage detection plan that is normally utilized with a single seal. This should be used on applications where the seal leakage would normally be in a liquid state. From the gland drain connection, leakage is directed past or through a reservoir containing a level switch, through an orifice and finally into either a sewer or liquid collection system. If the leakage from the seal is excessive, the orifice downstream from the reservoir restricts the flow allowing the level in the reservoir to rise, setting off an alarm. The downstream orifice should be located in a vertical leg to avoid accumulation of leakage in the pipe. This system should have a line running from the upper section of the reservoir connected to the piping downstream of the orifice to excessive leakage to drain. The selection of a proper throttle bushing is important. A fixed bushing, especially on larger size seals incorporating larger clearances, can allow leakage to leak past the bushing. This can contaminate the surrounding area or even spray the area in the case of

a severely damaged seal. If axial space is available, it is recommended that either a floating or segmented bushing be used with this plan.
Figure 10. Seal Flush Plan 65

Advantages

Provides an indication of excessive seal leakage without manual inspection. Can provide an automatic shutdown of equipment.

Disadvantages

Cost of system. Leakage levels have to be relatively high to set off the alarm.

What procedures should I use when storing my mechanical seals? Written by Fluid Sealing Association Pumps and Systems, January 2007 A section of the plant has been shut down because of a temporary change in product mix. New pumps have been installed in a plant addition that will not go on line for some time. In either case, mechanical seal assemblies are installed in the pumps and need to be put in the "sleep" mode for storage. What about the mechanical seal assemblies? Can they be left in the pumps for extended periods and then be ready for startup? How can I ensure that they will be? In another situation, mechanical seal assemblies have been placed in spare parts inventory well in advance of the need to install them as replacements. All of the above speaks to the issue of proper seal assembly storage. The mechanical seal assembly is a complex mix of precision metallic and nonmetallic components. Nonmetallics include ceramics such as silicon carbide and flexible graphite as well as elastomers such as ethylene propylene rubber (EPR). The direct contact with the process or flush fluids can be particularly problematic for the storage and reuse of some elastomers. Some of these materials are also dynamic sealing elements, while some are static, so they can be affected differently in service. Each has its own unique resistance to aging, and each is essential to effective seal performance. Storage must be based on preserving the integrity of all components. While general rules are noted here, it always is advisable to contact the manufacturers for their recommendations to ensure that the most reliable current practices are employed.
Storage in Pumps: 3 to 24 Months

Generally, mechanical seal assemblies can be stored satisfactorily for this period, provided manufacturers' installation and plant shut down procedures have been followed. Appropriate environmental and plant safety regulations also must be met. Given the above assumptions, storage procedures include:

Drain all product and flush fluid from the equipment and seal chamber. Dry equipment and seal chamber with compressed air. Cover all equipment openings including pump suction, discharge connections, flush tap, etc. Plug all openings in the seal chamber and gland. Mask or cover the clearance between the seal gland and the shaft to prevent dirt and debris from entering the seal cavity. Turn the shaft one to two revolutions by hand every three months.

Storage for Over 24 Months

Storage beyond 24 months will require removal of the seal assembly. When storage is longer than 24 months, environmental conditions are more likely to affect the flatness of the rotating and stationary seal faces as well as deteriorate some O-ring and gasket materials. Plant safety procedures and all environmental regulations again must be followed. Specific steps include:

Drain all product and flush fluid from the equipment and seal chamber. Remove, disassemble, and thoroughly clean and decontaminate the mechanical seal. Ensure that trapped fluid is removed from disassembled Cartridge Seals. Dry all parts, package, and store as individual components. Store in clean, cool environment.

Secondary Seals

The resistance to deterioration of elastomeric secondary seal materials - such as Orings, V-rings, and gaskets - under ambient storage conditions varies with the type of elastomer. Standard SAE ARP5316 lists the maximum recommended shelf life for Oring and molded secondary seal materials. However, these recommendations are based on storage in an unused, as received, condition, so they are not directly applicable to storage after some period of service in a pump. Nevertheless, they can serve as a guide to susceptibility of these materials to aging. This specification also contains recommended storage practices. Table 1 shows the maximum shelf life of a number of typical secondary seal materials as specified in this standard

The premise for Table 1, as listed in the SAE Standard, is that any of the elastomers are properly packaged and stored under optimum conditions. These include:

Ambient Temperature not exceeding 100 deg-F (38 deg-C). Exclusion of contamination. Exclusion of any radiation, particularly sunlight. Exclusion of ozone from all sources including electrical devices.

Generally, storage in ultraviolet (UV) resistant polyethylene or polyethylene lined Kraft paper bags ensures optimum storage life.

Exposure to process and/or flush fluids can influence the sealing properties of elastomers, and consequently shelf life, so before reinstalling them into any seal assembly it is particularly important that they be thoroughly inspected. All O-rings, Vrings, gaskets, etc. must be checked for deterioration, cracks, swell, hardness and compression set. Over time constrained elastomers, such as those in a groove, will take a set. The resultant decrease in sealing force can impact performance. Secondary seals more at risk are those completely encased in hardware, such as those that seal the two faces. Those that mate with the pump components, such as the sleeve, are less at risk. Manufacturers should be consulted to ensure that their condition is acceptable for reuse, should there be any question. Replacement can be the most prudent course, considering the potential serious impact of an in-service failure of any of these critical components.

Figure 1. Cross section of O-ring showing compression set.


Wear Faces

The wear faces of the seal rings and inserts should be tested for flatness with a helium light source and optical flat, prior to installation. Normally, all surfaces should be within two light bands. Faces may have to be re-lapped by an approved method should they be out of flatness. The manufacturer should be contacted for specifics if this requirement is not met. Static testing of the assembly also may be required.
Reassembly

Once the components have been checked to be acceptable for reuse, they can be reassembled and reinstalled in the pump in accordance with manufacturers' recommended procedures.
Conclusion

The storage of mechanical seal assemblies, whether in an as received or in service condition, requires practices that prevent deterioration of critical sealing properties.

Exposure to process and flush fluids can limit shelf life and reuse of some elastomeric secondary seal and gasket materials, depending on type and length of exposure. This article is intended to provide general guidelines and convey important considerations for storage. The FSA Mechanical Seal Handbook referenced below also is a valuable source for troubleshooting conditions noted in this column. The most prudent course, always, is to contact your mechanical seal manufacturer for specific instructions on proper storage procedures to ensure reliable start up and long term operation.

Diamond Advances Seal Face Performance Written by Charles F. West, Advanced Diamond Technologies, Inc. Pumps and Systems, May 2009 Advances in diamond are bringing the well-known properties of the world's hardest material to mechanical shaft seals. Diamond has long been sought after as a seal face material because of its unsurpassed hardness, high thermal conductivity, chemical inertness and low friction. Due to the processing limitations of synthetic diamond, however, early applications were relegated to abrasives and tooling that do not require the same high quality surface finishes and tolerances as those found in mechanical shaft seals.

In the last decade, profound advances have been made in the development of economical, large-scale diamond synthesis processes. These processes have led to an ability to produce smooth diamond that meets the surface requirements of seal faces. Along with the capability of manufacturing a large number of seal faces simultaneously, these improvements have made diamond-surfaced seals commercially viable. The use of carbon coatings on seal faces is not new. For years, a form of softer amorphous carbon known as diamond-like-carbons (DLCs) has been providing scuff resistance to the faces of gas compressor seals. DLCs adequately reduce scuffing during the brief contact that occurs during start-ups and shutdowns of compressors but have failed to demonstrate sufficient improvement over ceramic material options such as silicon carbide (SiC) and tungsten carbide (WC) in contact sealing applications. John Crane Inc., Huhnseal AB, and Burgmann Industries GmbH & Co. KG are marketing products incorporating significantly harder crystalline diamond face materials in universal ANSI pump cartridge seals, as well as in several other seal designs, where the extreme properties of diamond are meeting the demanding requirements of customers' applications. These new products are used when SiC and WC are not performing adequately, increase seal life by reducing face temperatures during intermittent dry operation and reduce wear that results from abrasives and poorly lubricating conditions.
Structure of Diamond Materials

Two forms of carbon are of interest to users of mechanical seals-graphite and diamond. Materials that integrate these allotropes of carbon are typically characterized by two key parameters: crystallinity and the percentage of diamond relative to graphite. DLCs, which is the material used on gas compressor faces, are not crystalline (i.e., the grains are disordered) and vary in composition with about 10 to 80 percent of the carbon bonded as it is in diamond. Accordingly, the range of hardness values in DLCs is also wide. The typical hardness is similar or below that of SiC, but the literature has reported some values exceeding -SiC. In the last couple decades, researchers have developed new diamond materials that can be grown directly onto seal faces such as SiC, and provide significant improvement to SiC's properties. These synthetic diamond technologies are 100 percent crystalline and

nearly all of the carbon in these materials is diamond. Whereas a diamond in an engagement ring consists of a single crystal, these materials are polycrystalline (i.e., they consist of many small crystals chemically bonded to each other) and consist of diamond grains that are about 10 to 20 m (1 m is about 40 micro inches) in size. These new materials are much harder than SiC and DLCs and are nearly as hard as natural diamond. They are also, due to the crystalline nature, more chemically resistant and have higher thermal conductivities than other seal materials. Unfortunately, these synthetic diamond materials suffer from a few drawbacks-the surface roughness is far too high, and they exhibit high friction and excessive counterface wear. Conventional diamond coatings are rough as deposited with a surface roughness of 1 m Ra or more. It is cost prohibitive to finish these rough diamond surfaces to meet the requirements of seal faces, and early attempts to use these materials for sealing applications failed.
How Diamond Is Applied to Seal Faces

The diamond seal products now available from John Crane Inc., EagleBurgmann, Huhnseal and those supplied to seal manufacturers by Advanced Diamond Technologies, Inc. are fabricated by growing a polycrystalline diamond film onto the face of a conventional finished SiC ring. It is then placed into a chamber where the pressure, gas composition and temperature are accurately controlled. A carbon bearing gas such as methane is introduced into the chamber and, under the right combination of processing conditions, diamond crystals grow on the SiC. The process occurs under vacuum at temperatures around 800-deg C (1,472-deg F). The diamond is not precipitating out from the vapor phase but grows up from the surface of the SiC. As these small diamond crystals grow, they coalesce together and form a continuous diamond surface. Specific processing conditions determine the diamond's properties. The relatively high temperature of the process results in a significant chemical interaction and subsequent bonding of the diamond onto the SiC. Excellent bonding is critical to ensure the diamond adheres well. Tests have shown that the bond between the SiC and the diamond can be stronger than the strength of the SiC itself.
Smooth Diamond Faces

The U.S. Department of Energy's Argonne National Laboratory (Argonne) is internationally known for synthesizing thin, smooth diamond materials with controllable properties which are now available commercially as ultrananocrystalline diamond (UNCD). Ultrananocrystalline diamond has grain-sizes measured in nanometers and not micrometers. As a result of its fine grain structure, ultrananocrystalline diamond can be deposited on SiC with a surface roughness of 2 to 5 nm Ra, well below the surface finish requirements of contacting mechanical seals. Researchers tend to think of synthetic diamond as a family of materials similar to the different types of carbides and carbon-based materials available today for seals. The technology developed at Argonne allows for different types of diamond materials with different surface roughness and other engineering specifications (see Figure 2). By accurately controlling the surface roughness, these diamond materials can be made smooth enough to run against counterfaces of much softer carbon or conventional uncoated SiC, enabling a wide choice of design options for seal manufacturers.

Controllable surface roughness also allows for two diamond faces to be paired against each other without the ringing or seizing that can occur when the same hard materials are paired together.

Ultrananocrystalline diamond-faced SiC mating ring

Figure 2. Diamond structure can be controlled to meet various seal requirements


Conformality and Flatness

Vapor-deposited diamond, including UNCD, is applied to SiC faces as a thin film with thicknesses ranging from 2 to 20 m. The diamond's thickness is routinely applied with uniformity within 0.3 m (11.6 micro inches). Due to these small variations in thickness, the diamond face retains the flatness of the original within one light band of helium. Ultrananocrystalline diamond is also suitable for non-contacting seal face designs in which the seal face is precisely machined with spiral grooves or other patterns that give the face three-dimensional features. UNCD is applied to an asfinished, non-contacting SiC ring with the diamond film conforming to the size and shape of the grooves.

Figure 3. Inference pattern showing diamond face flatness


Wear and Friction

Diamond overcomes two of the major challenges of seal face materials. First, a seal face must be fabricated to micrometer or sub-micrometer precision, enabling the seal design to properly maintain the face lubrication film. Second, the seal face must maintain the required surface quality and geometry even in poorly lubricating conditions (e.g., pumping near a fluid's vapor pressure or intermittent dry running), during exposure to abrasive solids in the media and to highly corrosive environments. Diamond-faced seals have been shown to have significantly longer useful life in poor lubricating environments such as hot water and during extremely abrasive pumping applications. Diamond is an ideal engineering material due to its extreme chemical resistance and unsurpassed hardness. The ideal face material reduces both the heat generation that can result in thermal distortions of the faces and secondary seal failures. The low coefficient of friction ( = 0.04) between diamond and SiC results in less heat generated when face lubrication is interrupted, minimizing seal failures during start-up and improving seal life during dry operation. It is because of the energy saving benefits of low friction materials that the U.S. Department of Energy originally funded the development of ultrananocrystalline diamond at Argonne.
Applications

It is important to remember that a seal face is one of many elements within a properly operating mechanical seal and all the major elements should be considered when diagnosing problem pumps. Diamond, however, enables pumps to function with higher reliability in demanding applications. These new diamond products enable hard face materials to be used while only softer carbon materials are used today due to the requirements of dry running. High performance single seals are now available that are

practical alternatives to the higher costs and maintenance of dual seals and associated systems. As a result of these and other benefits, diamond-faced seals are finding their way into a wide range of applications such as pumping light hydrocarbons, boiler feeds, water recirculation systems, deionized (DI) water systems, refinery and material processing applications, pulp and paper processing, and pharmaceutical production. The added cost of a conventional ANSI pump seal with diamond is equivalent to other face upgrade options. If an application could benefit from increased hardness and reduced friction, particularly if it is susceptible to dry running, consider installing diamond-faced seals during the next maintenance cycle to experience the benefits firsthand. Charles F. West is the vice president of engineering for Advanced Diamond Technologies, Inc., 429 B. Weber Road, #286, Romeoville, IL 60446, 815-293-0900.

Various Ways to Determine Pump Flow in the Field


Written by Dr. Lev Nelik, P.E., APICS

Pumps & Systems, October 2007 In my August column ("Tricks to Taking Flow Measurements in the Field"), I compared the pros and cons of one of several techniques that can be used to estimate pump flow when troubleshooting pump operations under less-than-desirable conditions: directly measuring it with a flowmeter. One of our readers, Don Casada of Diagnostic Solutions, LLC (Knoxville, TN), responded to that discussion with criticism of the direct measurement method from several perspectives, including the correct positioning of the external flowmeter on a pipe with appropriate requirements for a certain length of straight pipe; the absence of obstructions and bends; and similar HI guidelines. This month, we continue our discussion by examining two other indirect methods of determining pump flow in the field:

Pressure (head) measurement Power (amps) measurement

In both of these cases, you must obtain the pump performance curve for the respective application, which normally comes in combined or single-line formats, as shown in Figures 1 and 2.

Figure 1. Pump performance curve in a combined format. Source: 2004 Goulds Pumps Manual, ITT Industries

Figure 2. Pump performance curve in a single-line format. Source: 2004 Goulds Pumps Manual, ITT Industries

A combined format curve is typically available from a pump OEM generic catalog, while a single-line curve is usually supplied with specific pump quotation, or better yet, a factory-tested solution. In this example, a red triangle denotes the pump-rated point (70-gpm at 100-ft head) where a pump is expected to operate, but the actual flow is found suspect by operators.
Pressure (Head) Method

Let's assume that the discharge gauge reads 55-psig and the suction gauge reads 10psig, thus a 45-psi pressure differential exists. This would correspond to 45 x 2.31 = 104-ft head (assuming cold water, SG = 1.0). A horizontal 104-ft head line intersects the H-Q curve (at the proper impeller diameter, which is 5.12-in here in this case) at a little less than rated flow, approximately at 60-gpm.
Power (Amps) Method

The power curve indicates approximately 3.2-hp at the rated point. Power meters (kWmeter) are rarely available, with amps and volts being more commonly displayed at the control panel. Power can be calculated from these readings, although some assumptions of the power factor and motor efficiency would be required: BHP = (I x V x 1.73 x EFFmotor x PF) / 1000 In our example, a 5-hp 460-V motor is used and we actually read 450-V and 3.9-amps. A typical assumption of the product (EFFmotor x PF) is 0.85, although a somewhat better

value can be obtained if one is willing to spend some more time on research work. Thus, in our example: BHP = (3.9 x 450 x 1.73 x 0.85) / 1000 = 2.6 hp This is slightly less than the expected 3.2-hp, meaning a straight horizontal line at 2.6hp intersects the power curve at flow approximately 50-gpm, depending how accurately you eyeball the curve. Obviously, too many assumptions and approximations in reading curves bring bad news. However, the good news is that based on two methods, we can state that the flow appears to be somewhere between 50-gpm and 60-gpm. For many troubleshooting purposes, this answer is sufficient. That's not all. If we also add to this information the approximately 55-gpm data that was registered from the field flowmeter using the technique we discussed (in August) for the less than optimum pipe location, our confidence of the flow actually being somewhere between 50-gpm to 60-gpm will increase even further - a very good thing. As a note on the power method, some people feel more comfortable simply taking the ratio of actual amps to the motor nameplate (if one is still attached!) amps rating, then multiplying the result on motor rated power. In our example, if the motor rated amps were, say, 8.5-amps, and rated motor power 5-hp, we could then assume the actual power is 3.9 / 8.5 x 5 = 2.3-hp. This is close to the 2.6-hp value we derived earlier by using a power factor and motor efficiency assumption. The power method can be applied very successfully for field troubleshooting of many pump types, but it has significant drawbacks and cannot be applied for high specific speed (Ns) pumps, such as mixed flow and vertical turbine pumps. As HI illustrates nearby when comparing impeller profiles for various specific speed designs, pump power is not a nice continuously rising curve, as is the case for most end suction and split case pumps. Instead, the shape of the power curve can be entirely different. It can rise, drop, or stay constant with flow, even making its shape so flat that it becomes difficult to distinguish the difference for a rather wide variation of flows. The bottom line is that each method has its own place, strength and limitations:

The pressure (head) method is the simplest and quickest, but requires one to have a pump curve and gauges that are not broken or out of calibration. In the realities of the field, these curves are unfortunately long lost or misplaced for the old pumps, and even if they do exist, it is often impossible to know the most recent impeller diameter inside the pump after numerous prior pump repairs and modifications. The power (amps) method does not require one to "get dirty" around the pump replacing broken gauges, but inaccuracy of the power factor and motor efficiency is a drawback. (Reference power factor fundamentals presented by Joe Evans in "Power Factor: Electricity Behaving Badly (Part One)" (Pump Ed 101, Pumps & Systems June 2007) Direct flow reading is the most sure way, but most pumps do not have in-line flowmeters installed. Cutting into lines to install them is impractical and expensive. External (ultrasonic) meters are simple, but accuracy is limited due to difficulties in locating a good (HI approved)

spot along the pipe of the real field installation.

Often, applying all three methods reduces the error by allowing the user to learn to intelligently interpret the reasons for the differences, be able to explain the peculiarities and inconsistencies of each method, and correct such inconsistencies by solid reason, some understanding of flow mechanics, and reasons for deviations of practice from the theory. As always, our habit for leaving a parting thought with you: What simplifying assumptions were made in describing the pressure (head) method? What additional errors can these assumptions introduce, and to what magnitude? The first three correct answers get you a winning ticket to our next Pump School session(s). Keep on pumping! Dr. Nelik (aka "Dr. Pump") is president of Pumping Machinery, LLC, an Atlanta-based firm specializing in pump consulting, training, equipment troubleshooting, and pump repairs. Dr. Nelik has 30 years experience in pumps and pumping equipment. He has published over 50 documents on pump operations, the engineering aspects of centrifugal and positive displacement pumps, and maintenance methods to improve reliability, increase energy savings, and optimize pump-to-system operations. With questions, comments, or to attend his Pump School, he can be contacted at www.PumpingMachinery.com.

10 Ways to Select A Happy Pump


Written by Robert X. Perez Pumps and Systems, February 2007 1) Only select pumps with suction specific speeds less than 11,000-less than 9,000 is even better. 2) Select your pump so it never operates below 70 percent to 80 percent of its best efficiency point. 3) Remember that 1800-rpm pumps are usually more reliable than 3600-rpm pumps 4) Hydraulic efficiency peaks at specific speeds between 2000 and 3000 and drops dramatically below a specific speed of 500. Higher efficiency means less vibration and noise and lower energy bills. 5) Use double suction impeller sparingly. They are less stable at off-design conditions than single suction impellers. 6) For single stage pumps never select a pump with a maximum diameter impeller. You may need to increase the impeller diameter in the future. 7) Select a driver that allows you to operate to the end of the pump curve. 8) Use hydraulic stability, not temperature rise, as criteria for setting the minimum acceptable pump flow. 9) Incorporate a healthy NPSH margin or ratio, i.e. NPSHr/NPSHa into your selection. This ratio should be anywhere from 1.1 to 2.0 depending on the fluid, criticality and suction energy level. A higher value is always better. 10) Consider fluid volatility when making your pump selection. Be more conservative when the fluid has a single boiling point, as opposed to a fluid with a wide boiling point range.

What is a Safe NPSH Margin for a Centrifugal Pump? Can You Provide Too Much NPSH? Written by Terry Henshaw, P.E. What is a Safe NPSH Margin for a Centrifugal Pump? Can You Provide Too Much NPSH? Page 1 of 2 Pumps and Systems, June 2009 Editor's Note: This is the fifth article in the year-long series, Understanding NPSH. To read the previous article, click here. To read the next article, click here.
What is a Safe NPSH Margin for a Centrifugal Pump?

The Hydraulic Institute NPSHA Margin Work Group (1) recommended NPSH margins ranging from zero to four times the NPSHR (NPSHANPSHR = 1 to 5). The higher values apply to pumps with high to very high values of suction energy, a term defined as the product of the diameter of the pump suction nozzle, pump RPM and the suction specific speed.

McGuire (2) has provided a similar set of recommendations, seen in Figure 1, based on the service, pump size and peripheral velocity of the impeller eye.

Figure 1. Recommended centrifugal pump NPSH margins (J.T. McGuire [2]) Sloteman, et.al., (7) reported field experience with a four-stage, barrel type, boiler-feed pump, rated at 33,500 hp, 11,000 gpm and 5,800 rpm. The 3 percent NPSHR was 250 ft. The power plant, with a booster pump, provided 500 ft, twice that "required" by the pump. Impellers were made of CA6NM stainless steel. Because of cavitation damage, the first-stage impeller experienced a life of about 12 months. Although a 100 percent NPSH margin was provided, the cavitation damage rate was unacceptable.

Can You Provide Too Much NPSH?

Well, it does seem so. As we have just proven, conventional wisdom is to provide some margin of NPSHA over NPSHR. Unfortunately, the facts indicate that if the pump is operating near the 3 percent head-drop NPSHA, an increase in NPSHA will result in a noisier pump with a higher rate of damage. Grist (3) reported that maximum cavitation-erosion rate corresponded with maximum cavitation noise. Based on the information available, he postulated that this occurs when the NPSHA is about two times the "3 percent" NPSHA. Figure 2 is a reproduction of Figure 2 from Grist's paper (3). Note that this ratio of "two" occurs only at (or near) the BEP. At both lower and higher capacities, the ratio increases. The ratio of two should be taken as typical, for the literature indicates lower ratios for pumps with low suction specific speeds and higher ratios for pumps with high suction specific speeds.

Figure 2. From Grist's "Nett Positive Suction Head Requirements for Avoidance of Unacceptable Cavitation Erosion in Centrifugal Pumps" (3)

Performance Curves and NPSH Tests


Written by Terry Henshaw, P.E. Pumps and Systems, May 2009 Editor's Note: This is the fourth article the year-long series, Understanding NPSH. To read the previous article, click here. To read the next article, click here. NPSHR Shown as a Single Line as a Function of Capacity Only. Figure 1 is an example of singleline performance characteristics of a centrifugal pump. The independent variable capacity is plotted on the horizontal scale. Power, head, efficiency and NPSHR are plotted on vertical scales-each as a function of capacity.

Figure 1. The centrifugal pump family of characteristic curves Note that the NPSH curve is roughly a lazy "U," reaching a minimum value at a capacity Q about 40 percent of the bestefficiency capacity. Although most published curves do not show the increased NPSHR at low flows, all centrifugal pumps exhibit such a characteristic. The NPSHR always rises as the pump capacity approaches shutoff (zero flow rate). Figure 2 is a typical published performance curve. In this example, the NPSH curve is shown as a single line. (Note that it incorrectly shows the NPSHR to be a minimum at shutoff.)

Figure 2. Typical published performance curve. Single-line NPSH curve.


The Effect of Impeller Diameter on NPSHR

Figure 3 shows all parameters in the same fashion except for NPSH. Instead of a single curve that applies to all impeller diameters, lines of "isoNPSH" are shown. Note that for the same capacity, smaller diameter impellers require more NPSH.

Figure 3. Typical published performance curve. NPSH values increase as impeller diameter decreases. It is normal to experience a rise in NPSHR as the impeller diameter is reduced since a 3 percent drop of a smaller head (resulting from a smaller diameter impeller) is a

smaller head drop. By our definition of NPSHR, we are therefore permitting less cavitation with smaller diameter impellers. Currently there is no known way to predict this increase in NPSHR. Testing must therefore be relied on to establish NPSHR at each impeller diameter. Is it fair to penalize the smaller diameter impellers? Some pump vendors reason that it is not-the singleline NPSHR curve, obtained with the maximumdiameter impeller, should apply to all diameters supplied for that pump. What if one of those smallerdiameter impellers is supplied in a smallerdiameter casing (with a duplicate inlet passage)? We would be required to base our NPSHR on a 3 percent drop of a lower pump head, resulting in a higher NPSHR for the same impeller. Such contradictions account for some of the confusion surrounding this confounding subject of NPSH. A Method for Eliminating the NPSH Variation Caused by Different Impeller Diameters Because pump head is a function of the impeller diameter, D2, the traditional 3 percentheaddrop NPSH varies as the impeller diameter varies. With the same degree of cavitation in the impeller eye, a pump with a maximumdiameter impeller will exhibit a smaller percentage of head drop than the same pump with a reduceddiameter impeller. Gongwer (3) recognized this shortcoming in our definition of NPSHR. To normalize the NPSH requirements in his tests, he created a "constant" pseudohead on which to base his head loss. Rather than use the actual head of the pump, he calculated the head that would be developed with an impeller diameter that was twice the diameter of the impeller eye (D2=2D1). A pump develops a head that is approximately U22/2g. This pseudohead is therefore (2U1)2/2g, or 2 U12/g. The 3 percent head loss therefore converts to 0.03x2UI2/g = 0.06 UI2/g. In U.S. units, this would convert to a head loss, for the 3 percent loss, of H = (DlxN/5300)2. A 5in diameter eye, rotating 3,600-rpm, would have an allowable head loss of 11.5ft (completely independent of impeller diameter and pump head). Adoption by the pump industry of this revised definition of head loss would be a significant step in normalizing NPSH characteristics. It would be of value to pump users and manufacturers. Such an adjustment would result in NPSH3 curves being lifted for all pumps with (discharge) specific speeds less than about 1,500, and a lowering of NPSH3 curves for pumps with (discharge) specific speeds above that value. Prior to the concept of suction specific speed, S, the pump industry attempted to use the ThomaMoody concept, sigma, which stated that NPSHR was proportional to pump head. It is not. Numerous authors have so confirmed. S enabled us to begin decoupling NPSHR from H. Adoption of the above suggested criterion would be a further step in decoupling these two independent characteristics. Prior to the development of the NPSH concept, a performance curve would sometimes

show the inlet characteristics of a centrifugal pump as lines of constant "maximum suction lift," as can be seen in Figure 4.

Figure 4. An older performance curve that shows maximum suction lift instead of NPSHR (Acknowledgement is made to the Center for Professional Advancement, Applied Pump Technology Course)
Actual NPSH Test of a Centrifugal Pump

NPSHR is a difficult characteristic to measure, and to further complicate matters, pump manufacturers often disagree on the appropriate testing method and acceptance criterion. The most widelyaccepted criterion is the 3 percent head drop, as specified by API 610 (1). (There is also disagreement on the definition of "noncavitating head.") Figure 5 shows the test results obtained during an NPSH test on a centrifugal pump.

Figure 5. Actual NPSH test points for a process centrifugal pump

To get an NPSH point, we held the capacity constant while reducing the NPSHA. For each point, we plotted head as a function of NPSHA. We then determined how much NPSH was available when the head dropped 3 percent, which was the NPSHR for that capacity. To obtain a single point for the NPSHR curve, it is necessary to run a pump at a fixed capacity for an extended period of time. Because of the problems of running a centrifugal pump near shutoff for any extended period (heating, radial thrust, recirculation, pulsations and vibrations), we avoid trying to get NPSHR values at low capacities. When we plotted our points and smoothed a curve through them, we got a curve as shown in Figure 1. We heard a loud crackling sound coming from the pump as we reduced the NPSHA. That destructive sound helped us to understand how cavitation could damage pump components. We also noted that the sound level reached a peak, then dropped as we approached the 3 percent point. So here is the question: If you bought this pump, installed it in your plant, and provided the "required" 33-ft of NPSH, would the pump be cavitating? The answer is yes! If pumping cool water, it would cavitate enough to cause the head to be down by 3 percent, which is a lot of cavitation. How much NPSH is required to eliminate all cavitation? If you provided 75-ft, would that stop the cavitation? That would get the head back up to "noncavitating head," but, unfortunately, even at that point, there is still a significant amount of cavitation in the impeller eye. But how much more NPSH is required to eliminate all cavitation? To address this question, the Hydraulic Institute formed the NPSHA Margin Work Group. Initial conclusions of this group were published in the August 1997 issue of Plant Services Magazine (2). Key points in the article include the following:

Cavitation exists in all centrifugal pumps. NPSHA at incipient cavitation can be from 2 to 20 times the "3 percent" NPSHR. Achieving full (100 percent) head requires from 1.05 to 2.5 times the "3 percent" NPSHR. (In Figure 5, we can see that this ratio is about 2.3.)

A study by this author revealed that the NPSH required to achieve incipient cavitation for the typical pump at a capacity correspondent with no prerotation at the impeller inlet is a function of the impeller inlet vane angle, and varies from 2 to 10 times the NPSH3. We can conclude that it is normally not practical to try to eliminate cavitation, but what is a reasonable margin that we should strive for in our system designs? (We will try to answer that in future articles.)

References 1. Centrifugal Pumps for General Refinery Service, API Standard 610, American Petroleum Institute, 1220 L Street NW, Washington, DC 20005 2. "Cavitation Problems?", Hydraulic Institute NPSHA Margin Work Group, Plant Services, August 1997 3. Gongwer, Calvin A., NA "Theory of Cavitating Flow in CentrifugalPump Impellers", ASME Hydraulic Div. SemiAnnual Meeting, Milwaukee, WI, June 1940

Understanding NPSH: NPSH Definitions


Written by Terry Henshaw, P.E. Pumps and Systems, March 2009 Editor's Note: This is the second article the year-long series, Understanding NPSH. To read the previous article, click here. To read the next article in the series, click here.
Definition of NPSH

The margin of pressure over vapor pressure, at the pump suction nozzle, is Net Positive Suction Head (NPSH). NPSH is the difference between suction pressure (stagnation) and vapor pressure. In equation form: NPSH = Ps - Pvap Where: NPSH = NPSH available from the system, at the pump inlet, with the pump running Ps = Stagnation suction pressure, at the pump inlet, with the pump running Pvap = Vapor pressure of the pumpage at inlet temperature Since vapor pressure is always expressed on the absolute scale, suction pressure must also be in absolute terms. In U.S. customary units, both pressures must be in psia. Gauge pressure is converted to absolute pressure by adding atmospheric pressure. In equation form: absolute pressure = gauge pressure + atmospheric pressure The above equation provides an answer in units of pressure (psi). This can be converted to units of head (feet) by the following equation: h = 2.31p/SG Where: h = Head, feet p = Pressure, psi SG = Specific gravity of the liquid

Problem No. 1: NPSH Stagnation suction pressure is determined to be 1-psig at a sea level installation. The vapor pressure of the liquid is 8-psia. Calculate NPSH in PSI and feet for a 0.9 SG liquid NPSH = Ps - Pvap = 1 + 14.7 - 8 = 7.7 PSI NPSH = 2.31p/SG = (2.31) (7.7)/0.9 = 19.8 FEET
Velocity Head is Included

Note that suction pressure is stagnation pressure (total pressure); it includes velocity head. Adding velocity head puts all pumps on the same basis, otherwise a pump would require different amounts of NPSH when tested with different sizes of suction lines (assuming that suction pressure is measured in the suction line, which is normally the case). It should also be noted that the velocity head is normally quite small, relative to the NPSH, so it can usually be ignored.
Units of NPSH

For centrifugal pumps, NPSH values are expressed in units of specific energy (equivalent column height) such as feet or meters. For displacement pumps (rotary and reciprocating), NPSH values are normally expressed in pressure units such as pounds per square inch (psi), kilopascals, or bars. NPSH values are neither gauge pressures nor absolute pressures. The g in psig means that the pressure is measured above atmospheric pressure. The a in psia means that the pressure is measured above absolute zero, a perfect vacuum. NPSH is a measurement of pressure above vapor pressure, so the units of NPSH (in the U.S.) are just psi or feet.
What Symbol Should We Use?

Because units of pressure are typically used to express the value of NPSH for a displacement pump, some authors, and companies, use symbols such as NPIP (for net positive inlet pressure) and NIP (for net inlet pressure), because the units are pressure units, not head units. For simplicity, I'll stick with NPSH, regardless of the units used.
Suction Pressure: The First Half of the NPSH Equation

Suction pressure must be determined at the pump suction nozzle when the pump is running. If suction pressure is measured with a gauge, the atmospheric pressure (at the pump location) must be added to the gauge reading to convert the reading to absolute pressure. The elevation of the gauge must also be added (if the gauge is above datum) or subtracted (if the gauge is below datum). Although often negligible, the velocity head in the pipe at the gauge connection should be added to obtain total (stagnation) pressure. For a reciprocating pump (and some rotaries), the acceleration head must be subtracted. (More on acceleration head later.)

Vapor Pressure: The Second Half of the NPSH Equation

Vapor pressure is more difficult to determine than suction pressure. It is a measure of the "desire" of a liquid to boil to a gas. Some liquids, such as butane and ammonia, have high vapor pressures. They must be kept under pressure, or they will boil (flash). An open container of pure ammonia would quickly boil away, filling the area with noxious ammonia gas. Cool water has a low vapor pressure. An open container of cool water, on the earth's surface, would not boil, but would evaporate slowly over a period of days. The desire of cool water to boil is therefore low. If we were to place that same open container of cool water on the surface of the moon, it would boil away, similar to the ammonia. Why? The atmospheric pressure on the moon is zero, a perfect vacuum. The vapor pressure of pure, air-free water at 80-deg F is about 1/2-psia. This means that if the pressure on the water is reduced below - psia, the water will boil.
A Function of Temperature

Vapor pressure is a function only of temperature. As the temperature of the liquid increases, its vapor pressure increases until the critical temperature is reached. At the critical temperature, vapor pressure vanishes. Above the critical temperature, there is no distinction between a liquid and a gas. It is all fluid.
Boiling Reestablishes Equilibrium Conditions

Any liquid at its vapor pressure is on the verge of boiling (flashing). In such a condition it is said to be in equilibrium, at its bubble point or saturated. If the pressure is reduced slightly, it will start to boil. If the temperature is held constant (which requires heat input) and the pressure held constant (below the vapor pressure), the liquid will continue to boil until it has all flashed to vapor. If heat is not provided to the liquid, the portion flashing to vapor will cool, and will also absorb heat from the remaining liquid, causing the liquid temperature to drop. The lower temperature will result in a lower vapor pressure. The boiling will continue only until the vapor pressure drops to the pressure which is imposed on the liquid. When that vapor pressure is reached, and the boiling stops, the liquid-vapor mixture is again in equilibrium. This is what happens in the suction passage of a pump. Cavitation will cool the liquid and stop the cavitation. Otherwise, all the liquid would flash to vapor.
NPSH Available: A System Characteristic

NPSHA stands for NPSH Available from the system. It can be calculated by measuring suction pressure at the pump suction nozzle, correcting to datum, adding atmospheric pressure, adding velocity head and subtracting vapor pressure. In equation form: NPSHA = Psg + Pz + Patm + Pvel - Pvap Where:

NPSHA = NPSH available to the pump, psi Psg = Gauge pressure measured at suction nozzle, psig Pz = Elevation of gauge above pump centerline, converted to pressure units, psi Patm = Atmospheric pressure, psia Pvel = Velocity head, convened to pressure units, psi Pvap = Vapor pressure of the pumpage, at the pump suction nozzle, psia If desired, all units can be convened to head (feet) prior to plugging into the equation. If the system has not been built, it is necessary to calculate the NPSHA by starting with the pressure in the suction tank. Add atmospheric pressure, add (or subtract) the liquid level above (below) datum, subtract all losses from the tank to the pump and subtract vapor pressure. With reciprocating pumps it is also necessary to subtract acceleration head, a term which will be explained later. In equation form: NPSHA = Pt + Patm + Pzt - Pf - Pvap Where: Pt = Tank pressure, psig Pzt = Elevation of liquid in suction tank, converted to pressure units, psi Pf = Friction losses at tank exit and in suction line, converted to pressure units, psi Problem No. 2: NPSHA A suction gauge with its centerline 2-ft below the centerline of a centrifugal pump reads 152-psig. Atmospheric pressure is 14.0-psia. The pipe is 3-in standard weight steel. Capacity is 100-gpm. Vapor pressure is 163-psia. SG is 0.5. Calculate the NPSHA in feet. Because the desired answer is in feet, rather than PSI, we will convert all pressure units to feet. Flow Area of Pipe = 3.14 x 1.52 = 7.07 square inches VEL = 0.321 x Q/A = 0.321 x 100/7.07 = 4.54 feet/sec Hvel = V2/2G = 4.542/64.4 = 0.3 feet NPSHA = Hsg + Hz + Hatm + Hvel - Hvap = (psg + patm - pvap)(2.31/SG) + Hz + Hvel

= (152 + 14.0 - 163)(2.31/0.5) + (-2) + 0.3 = 13.9 - 2 + 0.3 = 12 feet [Note: A new 300-psi gauge, used to measure the suction pressure, normally has an accuracy of 1 percent of full scale, or 3-psi (14-ft). Therefore, the error in the gauge could be more than the calculated NPSHA.] [Also note that the velocity head was negligible, and could have been ignored.]
NPSH Reguired: A Pump Characteristic

The letters NPSHR stand for the NPSH required by the pump. This characteristic must be determined by test. Test methods and acceptance criteria for different types of pumps will be discussed later.
System Requirement

For proper operation of the pump, it is necessary that NPSHA > NPSHR. The system must provide more NPSH than the pump requires. Terry Henshaw is a retired consulting engineer who designs centrifugal pumps, reciprocating pumps and related high-pressure equipment and conducts pump seminars. For 30 years, he was employed by Ingersoll-Rand and Union Pump. Henshaw served in various positions in the Hydraulic Institute, including chairman of the Reciprocating Pump Section and chairman of the Metrication Subcommittee. He also served as a member of ANSI Subcommittee B73.2, chairman of the API 674 manufacturers' subcommittee and a member of the ASME Performance Test Code Committee PTC 7.2. He authored a book on reciprocating pumps, several magazine articles and the two sections on pumps in Marks' Handbook (11th Edition). He has been awarded six patents. Henshaw is a registered professional engineer in Texas and Michigan, is a life follow of the ASME, and holds engineering degrees from Rice University and the University of Houston.

Das könnte Ihnen auch gefallen