(PDF) Pump-Handbook.pdf | Prabir Datta - Academia.edu
Pump Handbook EDITED BY Igor J. Karassik Joseph P. Messina Paul Cooper Charles C. Heald THIRD EDITION McGRAW-HILL New York San Francisco Washington, D.C. Auckland Bogotá Caracas Lisbon London Madrid Mexico City Milan Montreal New Delhi San Juan Singapore Sydney Tokyo Toronto In memory of our good friends and colleagues William C. Krutzsch Warren H. Fraser Igor J. Karassik Copyright © 2001, 1986, 1976 by The McGraw-Hill Companies, Inc. All rights reserved. Printed in the United States of America. Except as permitted under the United States Copyright Act of 1976, no part of this publication may be reproduced or distributed in any form or by any means, or stored in a data base or retrieval system, without the prior written permission of the publisher. 1 2 3 4 5 6 7 8 9 0 DOC / DOC 0 6 5 4 3 2 1 0 ISBN 0-07-034032-3 The sponsoring editor for this book was Linda Ludewig, the editor supervisor was David Fogarty, and the production supervisor was Pamela A. Pelton. It was set in Century Schoolbook by D&G Limited, LLC. Printed and bound by R. R. Donnelley & Sons Company. McGraw-Hill books are available at special quantity discounts to use as premiums and sales promotions, or for use in corporate training programs. For more information, please write to the Director of Special Sales, Professional Publishing, McGraw-Hill, Two Penn Plaza, New York, NY 10121-2298. Or contact your local bookstore. Information contained in this work has been obtained by The McGraw-Hill Companies, Inc. (“McGraw-Hill”) from sources believed to be reliable. However, neither McGraw-Hill nor its authors guarantee the accuracy or completeness of any information published herein and neither McGraw-Hill nor its authors shall be responsible for any errors, omissions, or damages arising out of use of this information. This work is published with the understanding that McGraw-Hill and its authors are supplying information but are not attempting to render engineering or other professional services. If such services are required, the assistance of an appropriate professional should be sought. This book is printed on recycled, acid-free paper containing a minimum of 50 percent recycled de-inked fiber. OTHER MCGRAW-HILL REFERENCE BOOKS OF INTEREST Handbooks Baumeister • Marks’ Standard Handbook for Mechanical Engineers Bovay • Handbook of Mechanical and Electrical Systems for Buildings Brady and Clauser • Materials Handbook Brater and King • Handbook of Hydraulics Chopey and Hicks • Handbook of Chemical Engineering Calculations Croft, Carr, and Watt • American Electricians’ Handbook Dudley • Gear Handbook Fink and Beaty • Standard Handbook for Electrical Engineers Harris • Shock and Vibration Handbook Hicks • Standard Handbook of Engineering Calculations Hicks and Mueller • Standard Handbook of Professional Consulting Engineering Juran • Quality Control Handbook Kurtz • Handbook of Engineering Economics Maynard • Industrial Engineering Handbook Optical Society of America • Handbook of Optics Pachner • Handbook of Numerical Analysis Applications Parmley • Mechanical Components Handbook Parmley • Standard Handbook of Fastening and Joining Peckner and Bernstein • Handbook of Stainless Steels Perry and Green • Perry’s Chemical Engineers’ Handbook Raznjevic • Handbook of Thermodynamic Tables and Charts Rohsenow, Hartnett, and Ganic • Handbook of Heat Transfer Applications Rohsenow, Hartnett, and Ganic • Handbook of Heat Transfer Fundamentals Rothbart • Mechanical Design and Systems Handbook Schwartz • Metals Joining Manual Seidman and Mahrous • Handbook of Electric Power Calculations Shand and McLellan • Glass Engineering Handbook Smeaton • Motor Application and Control Handbook Smeaton • Switchgear and Control Handbook Transamerica DeLaval, Inc. • Transamerica DeLaval Engineering Handbook Tuma • Engineering Mathematics Handbook Tuma • Technology Mathematics Handbook Tuma • Handbook of Physical Calculations Encyclopedias Concise Encyclopedia of Science and Technology Encyclopedia of Electronics and Computers Encyclopedia of Engineering Dictionaries Dictionary of Mechanical and Design Engineering Dictionary of Scientific and Technical Terms LIST OF * CONTRIBUTORS Able, Stephen D., B.S. (M.E.), MBA, M.S. (Eng) SECTION 3.6 DIAPHRAGM PUMPS Senior Engineering Consultant, Ingersoll-Rand Fluid Products, Bryan, OH Addie, Graeme, B.S. (M.E.) SUBSECTION 9.16.2 APPLICATION AND CONSTRUCTION OF CENTRIFUGAL SOLIDS HANDLING PUMPS Vice President, Engineering and R&D, GIW Industries, Inc., Grovetown, GA Arnold, Conrad L., B.S., (E.E.) SUBSECTION 6.2.3 FLUID COUPLINGS Director of Engineering, American Standard Industrial Division, Detroit, MI Ashton, Robert D., B.S. (E.T.M.E.) SUBSECTION 2.2.4 CENTRIFUGAL PUMP INJECTION-TYPE SHAFT SEALS Manager, Proposal Applications, Byron Jackson Pump Division, Borg-Warner Industrial Products, Inc., Long Beach, CA Bean, Robert, B.A.(Physics), M.S. (M.E.) SECTION 3.6 DIAPHRAGM PUMPS Engineering Manager, Milton Roy Company, Flow Control Division, Ivyland, PA Beck, Wesley W., B.S. (C.E.), P.E. CHAPTER 13 PUMP TESTING Hydraulic Consulting Engineer, Denver, CO. Formerly with the Chief Engineers Office of the U.S. Bureau of Reclamation Benjes, H. H., Sr., B.S. (C.E.), P.E. SECTION 9.2 SEWAGE TREATMENT Retired Partner, Black & Veatch, Engineers-Architects, Kansas City, MO *Note: Positions and affiliations of the contributors generally are those held at the time the respective contributions were made. ix x LIST OF CONTRIBUTORS Bergeron, Wallace L., B.S. (E.E.) SUBSECTION 6.1.2 STEAM TURBINES Senior Market Engineer, Elliott Company, Jeannette, PA Birgel, W. J., B.S. (E.E.) SUBSECTION 6.2.1 EDDY-CURRENT COUPLING President, VS Systems, Inc., St. Paul, MN Birk, John R., B.S. (M.E.), P.E. SECTION 9.6 CHEMICAL INDUSTRY Consultant, Senior Vice President (retired), The Duriron Company, Inc., Dayton, OH Brennan, James R., B.S. (M.I.E.) SECTION 3.7 SCREW PUMPS; SECTION 9.17 OIL WELLS Manager of Engineering, Imo Pump, a member of the Colfax Pump Group, Monroe, NC Buse, Frederic W., B.S. (Marine Engrg.) SECTION 2.2.7.1 SEALLESS PUMPS: MAGNETIC DRIVE PUMPS; SECTION 3.1 POWER PUMP THEORY; SECTION 3.2 POWER PUMP DESIGN AND CONSTRUCTION; SECTION 5.2 MATERIALS OF CONSTRUCTION FOR NONMETALLIC (COMPOSITE) PUMPS; SECTION 9.6 CHEMICAL INDUSTRY Retired Senior Engineering Consultant, Flowserve Corporation, Phillipsburg, NJ Cappellino, C. A., B.S. (M.E./I.E.), M.S. (Product Dev’t.), P.E. SECTION 9.8 PULP AND PAPER MILLS Engineering Project Manager, ITT Fluid Technology Corp., Industrial Pump Group Chaplis, William K., B.S. (M.E.), MBA SECTION 3.1 POWER PUMP THEORY; SECTION 3.2 POWER PUMP DESIGN AND CONSTRUCTION Product Engineering Manager, Flowserve Corporation, Phillipsburg, NJ Clopton, D. E., B.S. (C.E.), P.E. SECTION 9.1 WATER SUPPLY Assistant Project Manager, Water Quality Division, URS/Forrest and Cotton, Inc., Consulting Engineers, Dallas, TX Cooper, Paul, B.S. (M.E.). M.S. (M.E.), Ph.D. (Engrg.), P.E. CHAPTER 1 INTRODUCTION: CLASSIFICATION AND SELECTION OF PUMPS; SECTION 2.1 CENTRIFUGAL PUMP THEORY; SECTION 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS; SECTION 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS; SECTION 9.19.2 LIQUID ROCKET PROPELLANT PUMPS Retired Director, Advanced Technology, Ingersoll-Dresser Pumps, now Flowserve Corporation, Phillipsburg, NJ Costigan, James L., B.S. (Chem.) SECTION 9.9 FOOD AND BEVERAGE PUMPING Sales Manager, Tri-Clover Division, Ladish Company, Kenosha, WI Cunningham, Richard G., B.S. (M.E.), M.S. (M.E.), Ph.D. (M.E.) SECTION 4.1 JET PUMP THEORY Vice President Emeritus for Research and Graduate Studies and Professor Emertitus of Mechanical Engineering, Pennsylvania State University, University Park, PA Cutler, Donald B., B.S. (M.E.) SUBSECTION 6.3.1 PUMP COUPLINGS AND INTERMEDIATE SHAFTING Techncial Services Manager, Rexnord Corporation, Warren, PA Cygnor, John E., B.S. (M.E.) SUBSECTION 9.19.1 AIRCRAFT FUEL PUMPS Retired Manager, Advanced Fluid Systems, Hamilton Sundstrand, Rockford, IL Czarnecki, G. J., B.Sc., M.Sc. (Tech.) SECTION 3.7 SCREW PUMPS Chief Engineer (Retired), Imo Pump, a member of the Colfax Pump Group, Monroe, NC Dahl, Trygve, B.S. (M.E.), M.S. (M.E. Systems), Ph.D. (M.E.), P.E. CHAPTER 11 SELECTING AND PURCHASING PUMPS Chief Technology Officer, IntellEquip, Inc., Bethlehem, PA. Formerly with Ingersoll-Dresser Pump Co., now part of Flowserve Corporation. LIST OF CONTRIBUTORS DiMasi, Mario, B.S. (M.E.), M.B.A. SECTION 9.4 FIRE PUMPS District Manager, Peerless Pump, Union, NJ Divona, A. A., B.S. (M.E.) SUBSECTION 6.1.1 ELECTRIC MOTORS AND MOTOR CONTROLS Account Executive, Industrial Sales, Westinghouse Electric Corporation, Hillside, NJ Dolan, A. J., B.S. (E.E.), M.S. (E.E.), P.E. SECTION 6.1.1 ELECTRIC MOTORS AND MOTOR CONTROLS Fellow District Engineer, Westinghouse Electric Corporation, Hillside, NJ Dornaus, Wilson L., B.S. (C.E.), P.E. SECTION 10.1 INTAKES, SUCTION PIPING, AND STRAINERS Pump Consultant, Lafayette, CA Drane, John, C.Eng., M.I. Chem.E. SECTION 9.9 FOOD AND BEVERAGE PUMPING Technical Support Engineer, Mono Pumps Limited, Manchester, England, UK Eller, David, B.S. (A.E.), P.E. SUBSECTION 6.3.2 HYDRAULIC PUMP AND MOTOR POWER-TRANSMISSION SYSTEMS President and Chief Engineer, M&W, Pump Corporation, Deerfield Beach, FL °Elvitsky, A. W., B.S. (M.E.), M.S. (M.E.), P.E. SECTION 9.7 PETROLEUM INDUSTRY Vice-President and Chief Engineer, United Centrifugal Pumps, San Jose, CA *Foster, W. E., B.S. (C.E.), P.E. SECTION 9.2 SEWAGE TREATMENT Partner, Black & Veatch, Engineers-Architects, Kansas City, MO °Fraser, Warren H., B.M.E. SECTION 2.3.2 CENTRIFUGAL PUMP HYDRAULIC PERFORMANCE AND DIAGNOSTICS Chief Design Engineer, Worthington Pump Group, McGraw-Edison Company, Harrison, NJ Freeborough, Robert M., B.S. (Min. E.) SECTION 3.3 STEAM PUMPS Manager, Parts Marketing, Worthington Corporation, Timonium, MD Furst, Raymond B. SUBSECTION 9.19.2 LIQUID ROCKET PROPELLANT PUMPS Retired Manager of Hydrodynamics, Rocketdyne, now The Boeing Company, Canoga Park, CA Giddings, J. F., Diploma, Mechanical, Electrical, and Civil Engineering SECTION 9.8 PULP AND PAPER MILLS Development Manager, Parsons & Whittemore, Lyddon, Ltd., Croydon, England Glanville, Robert H., M.E. SECTION 9.15 METERING Vice President Engineering, BIF, A Unit of General Signal, Providence, RI Guinzburg, Adiel, B.Sc. (Aero. E.), M.S. (Aero.), Ph.D. (M.E.) SUBSECTION 9.19.2 LIQUID ROCKET PROPELLANT PUMPS Engineering Specialist, Rocketdyne Propulsion & Power, The Boeing Company, Canoga Park, CA °Gunther, F. J., B.S. (M.E.), M.S. (M.E.) Subsection 6.1.3 ENGINES Late Sales Engineer, Waukesha Motor Company, Waukesha, WI Haentjens, W. D., B.M.E., M.S. (M.E.), P.E. SECTION 9.10 MINING Manager, Special Pumps and Engineering Services, Hazleton Pumps, Inc., Hazleton, PA °Deceased. xi xii LIST OF CONTRIBUTORS Hawkins, Larry, B.S. (M.E.), M.S. (M.E.) SUBSECTION 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS Principal, Calnetix, Torrance, CA Heald, Charles C., B.S. (M.E.) SUBSECTION 2.2.1 CENTRIFUGAL PUMPS: MAJOR COMPONENTS; SUBSECTION 9.7 PETROLEUM INDUSTRY; SECTION 10.1 INTAKES, SUCTION PIPING, AND STRAINERS; CHAPTER 12 INSTALLATION, OPERATION, AND MAINTENANCE Retired Manager of Engineering, Ingersoll-Dresser Pumps, now Flowserve Corporation, Phillipsburg, NJ Hendershot, J. R., B.S. (Physics) SUBSECTION 6.1.1 ELECTRIC MOTORS AND MOTOR CONTROLS; SUBSECTION 6.2.2 SINGLE-UNIT ADJUSTABLE-SPEED ELECTRIC DRIVES President, Motorsoft, Inc., Lebanon, OH Honeycutt, F. G, Jr., B.S. (C.E.) P.E. SECTION 9.1 WATER SUPPLY Assistant Vice President and Head Water Quality Division, URS/Forrest and Cotton, Inc., Consulting Engineers, Dallas, TX House, D. A., B.S. (M.E.) SECTION 9.2 SEWAGE TREATMENT Product Engineering Manager, Flowserve Corporation, Taneytown, MD Ingram, James H., B.S. (M.E.), M.S. (M.E.) SUBSECTION 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE, INSTRUMENTATION, AND DIAGNOSTICS Senior Engineering Specialist, Monsanto Fibers and Intermediates Company, Texas City, TX Jackson, Charles, B.S. (M.E.), A.A.S. (Electronics), P.E. SUBSECTION 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE, INSTRUMENTATION, AND DIAGNOSTICS Distinguished Fellow, Monsanto Corporate Engineering, Texas City, TX Jaskiewicz, Stephen A., B.A. (Chemistry) SUBSECTION 2.2.7.2 SEALLESS PUMPS: CANNED MOTOR PUMPS Product Manager, Chempump (Division of Crane Pumps & Systems, Inc.), Warrington, PA Jones, Graham, B.S. (M.E.), M.B.A. SUBSECTION 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS Former Project Manager for Magnetic Bearings, Technology Insights, San Diego, CA Jones, R. L., B.S. (M.E.), M.S. (M.E.), P.E. SUBSECTION 9.7 PETROLEUM INDUSTRY Senior Staff Engineer, Shell Chemical Company, Shell Oil Products Company, Houston, TX Jumpeter, Alex M., B.S. (Ch.E.) SUBSECTION 4.2 JET PUMP APPLICATIONS Engineering Manager, Process Equipment, Schutte and Koerting Company, Cornwells Heights, PA Kalix, David A., B.S. (C.E.), P.E. (M.E.) SUBSECTION 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS Senior Product Development Engineer, Yarway Corporation, Blue Bell, PA °Karassik, Igor J., B.S. (M.E.), M.S. (M.E.), P.E. SUBSECTION 2.2.1 CENTRIFUGAL PUMPS: MAJOR COMPONENTS; SECTION 2.4 CENTRIFUGAL PUMP PRIMING; SECTION 9.5 STEAM POWER PLANTS; CHAPTER 12 INSTALLATION, OPERATION, AND MAINTENANCE Chief Consulting Engineer, Worthington Group, McGraw-Edison Company, Basking Ridge, NJ °Deceased. LIST OF CONTRIBUTORS xiii Kawohl, Rudolph, Dipl. Ing. SECTION 9.4 FIRE PUMPS Retired Engineering Manager, Ingersoll-Dresser Pumps, now Flowserve Corporation, Arnage, France Kittredge, C. P., B.S. (C.E.), Doctor of Technical Science (M.E.) SUBSECTION 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS Consulting Engineer, Princeton, NJ Koch, Richard P., B.S. (M.E.) SECTION 9.5 STEAM POWER PLANTS Manager of Engineering, Pump Services Group, Flowserve Corporation, Phillipsburg, NJ Kron, H. O., B.S. (M.E.), P.E. SUBSECTION 6.2.4 GEARS Executive Vice President, Philadelphia Gear Corporation, King of Prussia, PA °Krutzsch, W. C., B.S. (M.E.), P.E. SI UNITS—A COMMENTARY; CHAPTER 1 INTRODUCTION: CLASSIFICATION AND SELECTION OF PUMPS Late Director, Research and Development, Engineered Products, Worthington Pump Group, McGraw-Edison Company, Harrison, NJ Landon, Fred K., B.S. (Aero. E.), P.E. SUBSECTION 6.3.1 PUMP COUPLINGS AND INTERMEDIATE SHAFTING Manager, Engineering, Rexnord, Inc., Houston, TX Larsen, Johannes, B.S. (C.E.), M.S. (M.E.) SECTION 10.2 INTAKE MODELING Retired Vice President, Alden Research Laboratory, Inc., Holden, MA Lippincott, J. K., B.S. (M.E.) SECTION 3.7 SCREW PUMPS Vice President, General Manager (Retired), Imo Pump, a member of the Colfax Pump Group, Monroe, NC Little, C. W., Jr., B.E. (E.E.), D. Eng. SECTION 3.8 VANE, GEAR, AND LOBE PUMPS Former Vice President, General Manager, Manufactured Products Division, Waukesha Foundry Company, Waukesha, WI Maxwell, Horace J., B.S. (M.E.) SUBSECTION 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS Director of Engineering, Yarway Corporation, Blue Bell, PA Mayo, Howard A., Jr., B.S. (M.E.), P.E. SUBSECTION 6.1.4 HYDRAULIC TURBINES; SECTION 9.13 PUMPED STORAGE Consulting Engineer, Hydrodynamics Ltd., York, PA McCaul, Colon O., B.S., M.S. (Metallurgical Engrg.), P.E. SECTION 5.1 METALLIC MATERIALS OF PUMP CONSTRUCTION Senior Engineering Consultant, Flowserve Corporation, Phillipsburg, NJ Messina, Joseph P., B.S. (M.E.), M.S. (C.E.), P.E. SECTION 8.1 GENERAL CHARACTERISTICS OF PUMPING SYSTEMS AND SYSTEM-HEAD CURVES; SECTION 8.2 BRANCH-LINE PUMPING SYSTEMS Consultant Miller, Ronald S., B.Sc. (M.E.), B.Sc (Metallurgical Engrg.) SECTION 5.1 METALLIC MATERIALS OF PUMP CONSTRUCTION Manager, Advanced Materials Engineering, Ingersoll-Rand Company Moll, Steven A., B.S. (E.E.) SECTION 6.2.1 EDDY-CURRENT COUPLINGS Senior Marketing Representative, Electric Machinery, Minneapolis, MN °Deceased. xiv LIST OF CONTRIBUTORS Moyes, Thomas L., B.S. (M.E.) SUBSECTION 9.18 CRYOGENIC LIQUEFIED GAS SERVICE Chief Engineer - R&D, Flowserve Corporation, Tulsa, OK Netzel, James P., B.S. (M.E.) SUBSECTION 2.2.2 CENTRIFUGAL PUMP PACKING; SUBSECTION 2.2.3 CENTRIFUGAL PUMPS: MECHANICAL SEALS Chief Engineer, John Crane, Inc., Morton Grove, IL Nolte, P. A., B.S. (M.E.) SECTION 9.20 PORTABLE TRANSFER OF HAZARDOUS LIQUIDS Director of Agricultural Business, Flowserve Corporation, Memphis, TN Nuta, D., B.S. (C.E.), M.S. (Applied Mathematics and Computer Science), P.E. SUBSECTION 9.14.2 NUCLEAR PUMP SEISMIC QUALIFICATIONS Associate Consulting Engineer, Ebasco Services, Inc., New York, NY O’Keefe, W., A.B., P.E. SUBSECTION 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS; CHAPTER 7 PUMP CONTROLS AND VALVES Editor, Power Magazine, McGraw-Hill Publications Company, New York, NY °Olson, Richard G., M.E. M.S., P.E. SECTION 6.1.5 GAS TURBINES Late Marketing Supervisor, International Turbine Systems, Turbodyne Corporation, Minneapolis, MN Padmanabhan, Mahadevan, B.S. (C.E.), M.S. (C.E.), Ph.D., P.E. SECTION 10.2 INTAKE MODELING Vice President, Alden Research Laboratory, Inc. Holden, MA Parmakian, John, B.S. (M.E.), M.S. (C.E.), P.E. SECTION 8.3 WATERHAMMER Consulting Engineer, Boulder, CO Parry, W. E., Jr., B.S. (M.E.), P.E. SUBSECTION 9.14.1 NUCLEAR ELECTRICAL GENERATION; SUBSECTION 9.14.2 NUCLEAR PUMP SEISMIC QUALIFICATIONS Project Engineering Manager, Nuclear Equipment/Vertical Pumps, Flowserve Corporation, Phillipsburg, NJ Patel, Vinod P., B.S. (M.E.), M.S. (Metallurgical Engrg.), P.E. CHAPTER 11 SELECTING AND PURCHASING PUMPS Senior Principal Engineer, Machinery Technology, Kellogg Brown & Root, Inc., Houston, TX Peacock, James H., B.S. (Met.E.) SECTION 9.6 CHEMICAL INDUSTRY Manager, Materials Division, The Duriron Company, Inc., Dayton, OH Platt, Robert A., B.E., M.E., P.E. SECTION 3.8 VANE, GEAR AND LOBE PUMPS General Manager, Sales and Marketing, Carver Pump Company, Muscatine, IA Potthoff, E. O., B.S. (E.E.), P.E. SUBSECTION 6.2.2 SINGLE-UNIT ADJUSTABLE-SPEED ELECTRIC DRIVES Industrial Engineer (retired), Industrial Sales Division, General Electric Company, Schenectady, NY Prang, A. J. SECTION 3.7 SCREW PUMPS Manager, Engineering and Quality Assurance, Flowserve Corporation, Brantford, Ontario, Canada Ramsey, Melvin A., M.E., P.E. SECTION 9.12 REFRIGERATION, HEATING, AND AIR CONDITIONING Consulting Engineer, Schenectady, NY °Deceased. LIST OF CONTRIBUTORS xv °Rich, George R., B.S. (C.E.), C.E., D.Eng., P.E. SECTION 9.13 PUMPED STORAGE Late Director, Senior Vice President, Chief Engineer, Chas. T. Main, Inc., Boston, MA Robertson, John S., B.S. (C.E.), P.E. SECTION 9.3 DRAINAGE AND IRRIGATION Chief, Electrical and Mechanical Branch, Engineering and Construction, Headquarters, U.S. Army Corps of Engineers Roll, Daniel R., B.S. (M.E.), P.E. SECTION 9.8 PUMP AND PAPER MILLS Vice President, Engineering & Development, Finish Thompson Inc, Erie, PA Rupp, Warren E. SECTION 3.6 DIAPHRAGM PUMPS President, The Warren Rupp Company, Mansfield, OH Sellgren, Anders, M.S. (C.E.), Ph.D. (Hydraulics) SUBSECTION 9.16.1 HYDRAULIC TRANSPORT OF SOLIDS; SUBSECTION 9.16.2 APPLICATION AND CONSTRUCTION OF CENTRIFUGAL SOLIDS HANDLING PUMPS Professor, Division of Water Resources Engineering, Lulea University of Technology, Lulea, Sweden Sembler, William J., B.S. (Marine Eng.), M.S. (M.E.) SECTION 9.11 MARINE PUMPS Tenured Associate Professor, United States Merchant Marine Academy, Kings Point, NY Shapiro, Wilbur, B.S., M.S. SUBSECTION 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS Consultant, Machinery Components, Niskayuna, NY Shikasho, Satoru, B.S. (M.E.), P.E. SECTION 9.21 WATER PRESSURE BOOSTER SYSTEMS Chief Product Engineer, Packaged Products, ITT Bell & Gossett, Morton Grove, IL Smith, L. R. SECTION 9.18 CRYOGENIC LIQUIFIED GAS SERVICE Retired, formerly of J. C. Carter Company, Costa Mesa, CA Smith, Will, B.S. (M.E.), M.S. (M.E.), P.E. SECTION 3.5 DISPLACEMENT PUMP FLOW CONTROL; SUBSECTION 9.16.3 CONSTRUCTION OF SOLIDS-HANDLING DISPLACEMENT PUMPS Engineering Product Manager, Custom Pump Operations, Worthington Division, McGraw-Edison Company, Harrison, NJ Snyder, Milton B., B.S. (B.A.) SUBSECTION 6.2.5 ADJUSTABLE-SPEED BELT DRIVES Sales Engineer, Master-Reeves Division, Reliance Eectric Company, Columbus, IN Sparks, Cecil R., B.S. (M.E.), M.S. (M.E.), P.E. SECTION 8.4 PUMP NOISE Director of Engineering Physics, Southwest Research Institute, San Antonio, TX Szenasi, Fred R., B.S. (M.E.), M.S. (M.E.), P.E. SECTION 3.4 DISPLACEMENT PUMP PERFORMANCE, INSTRUMENTATION, AND DIAGNOSTICS; SECTION 8.4 PUMP NOISE Senior Project Engineer, Engineering Dynamics Inc., San Antonio, TX Taylor, Ken W., MIProdE. CEng. SECTION 9.15 METERING Vice President, Global Business Development, Wayne Division, Dresser Equipment Group, a Halliburton Company Tullo, C. J., P.E. SECTION 2.4 CENTRIFUGAL PUMP PRIMING Chief Engineer (retired), Centrifugal Pump Engineering, Worthington Pump, Inc., Harrison, NJ Vance, William M., M.B.A. SECTION 9.17 OIL WELLS Senior Project Sales Engineer, Weir Pumps Limited, Glasgow, Scotland, UK °Deceased. xvi LIST OF CONTRIBUTORS VanLanningham, F. L., SECTION 6.2.4 GEARS Consultant, Rotating and Turbomachinery Consultants Wachel, J. C., B.S. (M.E.), M.S. (M.E.) SECTION 3.4 DISPLACEMENT PUMP PERFORMANCE, INSTRUMENTATION, AND DIAGNOSTICS; SECTION 8.4 PUMP NOISE Manager of Engineering, Engineering Dynamics, Inc., San Antonio, TX Webb, Donald R., B.S. (M.E.), M.S. (Engrg Administration) SUBSECTION 6.1.4 HYDRAULIC TURBINES Plant Assessment Manager, Voith Siemens Hydro, York, PA Wepfer, W. M., B.S. (M.E.), P.E. SUBSECTION 9.14.1 NUCLEAR ELECTRICAL GENERATION Consulting Engineer, formerly Manager, Pump Design, Westinghouse Electric Corporation, Pittsburgh, PA Whippen, Warren G., B.S. (M.E.), P.E. SUBSECTION 6.1.4 HYDRAULIC TURBINES Retired Manager of Hydraulic Engineering, Voith Siemens Hydro, York, PA Wilson, Kenneth C., B.A.Sc. (C.E.), M.Sc.(Hydraulics), Ph.D. SUBSECTION 9.16.1 HYDRAULIC TRANSPORT OF SOLIDS Professor Emeritus, Dept. of Civil Engineering, Queen’s University, Kingston, Ontario, Canada Wotring, Timothy L., B.S. (M.E.), P.E. SUBSECTION 2.2.4 CENTRIFUGAL PUMP INJECTION-TYPE SHAFT SEALS Engineering Manager, Flowserve Corporation, Phillipsburg, NJ Zeitlin, A. B., M.S. (M.E.), Dr.-Eng. (E.E.), P.E. SECTION 9.22 HYDRAULIC PRESSES President, Press Technology Corporation, Mamaroneck, NY SI UNITS— A COMMENTARY W.C. KRUTZSCH P. COOPER Since the publication of the first edition of this handbook in 1976, the involvement of the world in general, and of the United States in particular, with the SI system of units has become quite common. Accordingly, throughout this book, SI units have been provided as a supplement to the United States customary system of units (USCS). This should make it easier, particularly for readers in metric countries, who will no longer find it necessary to make either approximate mental transpositions or exact mathematical conversions. The designation SI is the official abbreviation, in any language, of the French title “Le Système International d’Unités,” given by the 11th General Conference on Weights and Measures (sponsored by the International Bureau of Weights and Measures) in 1960 to a coherent system of units selected from metric systems. This system of units has since been adopted by the International Organization for Standardization (ISO) as an international standard. The SI system consists of seven basic units, two supplementary units, a series of derived units, and a series of approved prefixes for multiples and submultiples of the foregoing. The names and definitions of the basic and supplementary units are contained in Tables la and lb of the Appendix. Table 2 lists the units and Table 3 the prefixes. Table 10 provides conversions of USCS to SI units. As with the second edition, the decision has been made to accept variations in the expression of SI units that are widely encountered in practice. The quantities mainly affected are pressure and flow rate, the situation with each being explained as follows. The standard SI unit of pressure, the pascal, equal to one newton* per square meter†, is a minuscule value relative to the pound per square inch (1 lb/in2  6,894.757 Pa) or to the old, established metric unit of pressure the kilogram per square centimeter (1 kgf/cm2  * The newton (symbol N) is the SI unit of force, equal to that which, when applied to a body having a mass of 1 kg, gives it an acceleration of 1 m/s2. † In countries using the SI system exclusively, the correct spelling is metre. This book uses the spelling meter in deference to prevailing U.S. practice. xxiii xxiv SI UNITS—A COMMENTARY 98,066.50 Pa). In order to eliminate the necessity for dealing with significant multiples of these already large numbers when describing the pressure ratings of modern pumps, different sponsoring groups have settled on two competing proposals. One group supports selection of the kilopascal, a unit which does provide a numerically reasonable value (1 lb/in2  6.894757 kPa) and is a rational multiple of a true SI unit. The other group, equally vocal, supports the bar (1 bar  105 Pa). This support is based heavily on the fact that the value of this special derived unit is close to one atmosphere. It is important, however, to be aware that it is not exactly equal to a standard atmosphere (101, 325.0 Pa) or to the so-called metric atmosphere (1 kgf/cm2  98,066.50 Pa), but is close enough to be confused with both. As yet, there is no consensus about which of these units should be used as the standard. Accordingly, both are used, often in the same metric country. Because the world cannot agree and because we must all live with the world as it is, the editors concluded that restricting usage to one or the other would be arbitrary, grossly artificial, and not in the best interests of the reader. We therefore have permitted individual authors to use what they are most accustomed to, and both units will be encountered in the text. Units of pressure are utilized to define both the performance and the mechanical integrity of displacement pumps. For kinetic pumps, however, which are by far the most significant industrial pumps, pressure is used only to describe rated and hydrostatic values, or mechanical integrity. Performance is generally measured in terms of total head, expressed as feet in USCS units and as meters in SI units. This sounds straightforward enough until a definition of head, including consistent units, is attempted. Then we encounter the dilemma of mass versus force, or weight. The total head developed by a kinetic pump, or the head contained in a vertical column of liquid, is actually a measure of the internal energy added to or contained in the liquid. The units used to define it could be energy per unit volume, or energy per unit mass, or energy per unit weight. If we select the last, we arrive conveniently, in USCS units, as footpounds per pound, or simply feet. In SI units, the terms would be newton-meters per newton, or simply meters. In fact, however, metric countries weigh objects in kilograms, not newtons, and so the SI term for head may be defined at places in this volume in terms of kilogram-meters per kilogram, even though this does not conform strictly to SI rules. Similar ambiguity is observed with the units of flow rate, except here there may be even more variations. The standard SI unit of flow rate is the cubic meter per second, which is indeed a very large value (1 m3/s  15,850.32 U.S. gal/min) and is therefore really only suitable for very large pumps. Recently, some industry groups have suggested that a suitable alternative might be the liter per second (11/s  1/0—3 m3/s  15.85032 U.S. gal/min), while others have maintained strong support for the traditional metric unit of flow rate, the cubic meter per hour (1 m3/h  4.402867 U.S. gal/mm). All of these units will be encountered in the text. These variations have led to several forms of the specific speed, which is the fundamentally dimensionless combination of head, flow rate, and rotative speed that characterizes the geometry of kinetic pumps. These forms are all related to a truly unitless formulation called “universal specific speed,” which gives the same numerical value for any consistent system of units. Although not yet widely used, this concept has been appearing in basic texts and other literature, because it applies consistently to all forms of turbomachinery. Equivalencies of the universal specific speed to the common forms of specific speed in use worldwide are therefore provided in this book. This is done with a view to eventual standardization of the currently disparate usage in a world that is experiencing globalization of pump activity. The value for the unit of horsepower (hp) used throughout this book and in the United States is the equivalent of 550 foot pounds (force) per second, or 0.74569987 kilowatts (kW). The horsepower used herein is approximately 1.014 times greater than the metric horsepower, which is equivalent to 0.735499 kilowatts. Whenever the rating of an electric motor is given in this book in horsepower, it is the output rating. The equivalent output power in kilowatts is shown in parentheses. Variations in SI units have arisen because of differing requirements in various user industry groups. Practices in the usage of units will continue to change, and the reader will have to remain alert to further variations of national and international practices in this area. ABOUT THE EDITORS Igor Karassik, now deceased, was an original editor of this book. His extensive contributions to the earlier editions remain a signal feature of this edition. A major figure in the pump industry for the greater part of the past century, he also authored or co-authored six books in this field. Beginning in 1936, he wrote more than 600 articles on centrifugal pumps and related subjects, which appeared in over 1500 publications worldwide. For the greater part of his career, he held senior engineering and marketing positions within the Worthington Pump & Machinery Company, which after a number of permutations became part of the Flowserve Corporation. Igor Karassik received his B.S. and M.S. degrees in Mechanical Engineering from Carnegie Mellon University. He was a Life Fellow of the American Society of Mechanical Engineers and recipient of the first ASME Henry R. Worthington Medal (1980). Joseph P. Messina, also one of the original editors, has spent his entire career in the pump industry; and his past contributions on pump and systems engineering continue to be presented in their entirety in this edition. He served as Manager of Applications Engineering at the Worthington Pump Company. He became a Pump Specialist at the Public Service Electric and Gas Company in New Jersey, serving as a committee member of the Electric Power Research Institute to improve the performance of boiler feed pumps. He assisted in updating the Hydraulic Institute Standards and taught centrifugal pump courses. He also taught Fluid and Solid Mechanics at the New Jersey Institute of Technology and holds a B.S. in Mechanical Engineering and an M.S. in Civil Engineering from the same institution. Now a pump technology consultant, he has been a contributor to the technical journals and holds pump-related patents. Paul Cooper has been involved in the pump industry for over forty years. He began by specializing in the hydraulic design of centrifugal pumps and inducers for aerospace applications at TRW Inc. This was followed by a career in research and development on pump hydraulics and cavitation at the Ingersoll-Dresser Pump Company, now part of the Flowserve Corporation, where he conducted investigations at the Ingersoll-Rand Research Center and later served as the director of R&D for the company. A Life Fellow of the ASME, he received that society’s Fluid Machinery Design Award (1991) and the Henry R. Worthington Medal (1993). He received his B.S. (Drexel University) and M.S. (Massachusetts Institute of Technology) degrees in Mechanical Engineering and a Ph.D. in Engineering from Case Western Reserve University. Now a consultant, he is the author of many technical papers and holds several patents on pumps. Charles C. Heald has spent his entire career in the pump industry. He conducted the hydraulic and mechanical design of several complete lines of single and multistage pumps for the Cameron Pump Division of Ingersoll-Rand, which became part of the IngersollDresser Pump Company. He served as Chief Engineer and Manager of Engineering. Currently a consultant, he continues to function as the editor of the company’s Cameron Hydraulic Data Book. The petroleum industry has always been the focus of his efforts, and he has served for over 35 years as a member of the API 610 specification task force, receiving a resolution of appreciation from API in 1995. A Life Member of the ASME, he obtained the B.S. degree in Mechanical Engineering from the University of Maine, and he is the author of several technical articles and the holder of patents pertaining to pumps. PREFACE TO THE THIRD EDITION It is difficult to follow in the footsteps of Igor J. Karassik, whose vision and leadership played a major role in the concept of a handbook on pumps that is broad enough to encompass all aspects of the subject—from the theory of operation through design and application to the multitude of tasks for which pumps of all types and sizes are employed. That vision was realized in the first edition of the Pump Handbook, which appeared a quartercentury ago, with the capable and dedicated co-authorship of William C. Krutzsch, Warren H. Fraser, and Joseph P. Messina. Acceptance of this work globally soon led these distinguished pump engineers to assemble a second edition that not only contained updated material but also presented all numerical quantities in terms of the SI system of units in addition to the commonly used United States customary system of units. Worldwide developments in pump theory, design and applications have continued to emerge, and these have begun to affect the outlook of pump engineers and users to such an extent that a third edition has become overdue. Pumps have continued to grow in size, speed, and energy level, revealing new problems that are being addressed by innovative materials and mechanical and hydraulic design approaches. Environmental pressures have increased, and these can and are being responded to by the creative attention of pump engineers and users. After all, the engineer is trained to solve problems, employing techniques that reflect knowledge of physical phenomena in the world around us. All of this has led the current authors to respond by adding new sections and by revising most of the others as would be appropriate in addressing these developments. Specifically the following changes should be noted. Centrifugal pump theory, in the rewritten Section 2.1, proceeds from the basic governing fluid mechanics to the rationale that underlies the fundamental geometry and performance of these machines—while maintaining the concrete illustrations of design examples. A new subsection on high-energy pumps is included. An update has been made to Section 2.2.1 on major components of centrifugal pumps. Section 2.3.1 on centrifugal pump general performance characteristics has been updated. xvii xviii PREFACE TO THE THIRD EDITION The emerging technology of magnetic bearings is presented in the new Section 2.2.6. Section 2.2.7, is a new treatment of sealless centrifugal pumps that includes both the canned-motor and magnetically-coupled types. Chapter 3 on displacement pumps has been reorganized and includes updates of the sections on both reciprocating and rotary positive displacement pumps. A new Section 4.1 on jet pump theory begins the chapter on jet pumps and deals with liquids and gases for the motive and secondary flows as well as the basics of design optimization. Chapter 5 on materials of construction, including the Sections 5.1 and 5.2 on metallic and nonmetallic materials respectively, has been completely rewritten and updated. Chapter 6 on pump drivers has been updated, Section 6.1.1 on electric motors and Section 6.2.2 on adjustable-speed electric drives having been substantially rewritten. In Chapter 9 on pump services, most of the applications sections have been updated, including those for fire pumps (Section 9.4) and pumps for steam power plants (9.5), pulp and paper (9.8), mining (9.10), metering (9.15), pumped storage (9.13), and nuclear services (9.14). Section 9.11 on marine applications has been rewritten. Sections 9.16.1 on hydraulic transport of solids and 9.16.2 on centrifugal slurry pumps are completely new and include several examples. A new section on aerospace pumps has been added, which includes Sections 9.19.1 on aircraft fuel pumps and 9.19.2 on liquid rocket propellant pumps. Section 9.20 on handling hazardous liquids is new. Chapters 10 on intakes and suction piping, 11 on selecting and purchasing pumps and 12 on installation, operation, and maintenance have been updated. We recognize that further developments are going on apace and that more could have been done. Computational fluid dynamics (CFD) and finite-element structural and rotordynamic analysis techniques, as well as the revolution in information management and utilization, already promise to profoundly transform pump design, application, and operational practice—and indeed all other areas of engineering endeavor. Nevertheless, we offer this third edition of the Pump Handbook as a practical tool for the present day. In this sense, we hope that it will fulfill the vision of the authors of prior editions while at the same time serving as a stepping stone to the future world of pumping. PAUL COOPER PREFACE TO THE SECOND EDITION Once more, the dubious honor of writing a preface has been bestowed upon me by my three co-editors. And while they are perfectly willing to share the pluses and minuses of collective editorship, they refused to engage in collective “prefaceship,” if I may be allowed to coin a word. At best, they reserved for themselves the right of looking over my shoulder and criticizing the spirit of levity with which I chose to approach the task for which they had unanimously volunteered me. I should add parenthetically that the preface of the first edition (which you can read on the following pages) is actually my fourth draft; the first three were judged too irreverent by my co-editors. (I have preserved these first three drafts for whoever inherits my collection of unpublished material.) Assuming that my co-editors are more charitable this time, or alternately that our publisher is pressed for time, what follows (if not what precedes) will appear more or less as written. First of all, we would like to assure the readers of this second edition of the Pump Handbook that it is not merely a slightly warmed-over version of the first edition, with such errata as we have spotted corrected and with a few insignificant changes and additions. Actually, the task of rewriting and editing the material in a form that would correspond to what was planned for this second edition proved to be a monumental, not to say awesome, undertaking. To begin with, in concert with the publishers, it was decided that all data given here would appear in both USCS and SI units. This was not as simple a task as it may appear, for the reason that “absolute” pure SI units do not lend themselves well to the scale of numbers generally encountered in industrial processes. To give but one example, the pascal, which is the SI unit of pressure, corresponds to 0.000145 lb/in2, and even the kilopascal is only 0.145 lb/in2. Although this might be a reasonably satisfactory unit for scientific work, the case is hardly such for centrifugal pumps used in everyday life. This led us to choose what might be called a modified set of SI units, all as explained in “SI Units—A Commentary,” on page xxi. Even conveying this desirable concept of a practical set of SI units to the authors of various sections proved to be somewhat difficult. xix xx PREFACE TO THE SECOND EDITION As a result, we have permitted these authors some leeway in their specific choice, understanding full well that what is desirable in one industry may differ from the preferred choice in another. We decided that a number of sections and subsections in the first edition could benefit by being significantly expanded. This, for instance, is the case with the following: 2.2.1 2.3.1 2.4 8.1 8.4 9.4 9.15.1 9.17.1 10.1 Appendix “Centrifugal Pumps: Major Components” “Centrifugal Pumps: General Performance Characteristics” “Centrifugal Pump Priming” “General Characteristics of Pumping Systems and System-Head Curves” “Pump Noise” “Fire Pumps” “Nuclear Electric Generation” “Hydraulic Transport of Solids” “Intakes, Suction Piping, and Strainers” “Technical Data” At the same time, we felt that some material originally included in the subsection “Centrifugal Pumps: Major Components” should be excised from there and treated in greater depth separately. This expanded coverage includes the following: 2.2.2 2.2.3 2.2.4 2.2.5 “Centrifugal Pump Packing” “Centrifugal Pump Mechanical Seals” “Centrifugal Pump Injection-Type Shaft Seals” “Centrifugal Pump Oil Film Journal Bearings” Finally, a large amount of subject matter has been added to the second edition: 2.3.2 2.3.3 2.3.4 3.3 3.6 3.7 5.2 6.3.2 6.3.3 9.15.2 9.17.3 9.18 9.19 9.20 10.2 “Centrifugal Pump Hydraulic Performance and Diagnostics” “Centrifugal Pump Mechanical Performance, Instrumentation, and Diagnostics” “Centrifugal Pump Minimum Flow Control Systems” “Diaphragm Pumps” “Displacement Pump Performance, Instrumentation, and Diagnostics” “Displacement Pump Flow Control” “Materials of Construction of Nonmetallic Pumps” “Magnetic Drives” “Hydraulic Pump and Motor Power Transmission Systems” “Nuclear Pump Seismic Qualifications” “Construction of Solids-Handling Displacement Pumps” “Oil Wells” “Cryogenic Liquefied Gas Service” “Water Pressure Booster Systems” “Intake Modeling” In brief, the editors have attempted to increase the usefulness of this handbook. The extent to which we have achieved this objective, we will leave to the judgment of our readers. IGOR J. KARASSIK PREFACE TO THE FIRST EDITION Considering that I had written the prefaces of the three books published so far under my name, my colleagues thought it both polite and expedient to suggest that I prepare the preface to this handbook, coedited by the four of us. Except for the writing of the opening paragraph of an article, a preface is the most difficult assignment that I know. Certainly the preface to a handbook should do more than describe minutely and in proper order the material that is contained therein. Yet I submit that the saying “a book should not be judged by its cover” should be expanded by adding “and not by its preface.” If the reader will accept this disclaimer, I can proceed. As will be stated in Section 1, “Introduction and Classification of Pumps,” it can rightly be claimed that no machine and very few tools have had as long a history in the service of man as the pump, or have filled as broad a need in his life. Every process which underlies our modern civilization involves the transfer of liquids from one level of pressure or static energy to another. Pumps have played an essential role in our life ever since the dawn of civilization. Thus it is that a constantly growing number of technical personnel is in need of information that will help them in either designing, selecting, operating, or maintaining pumping equipment. There has never been a dearth of excellent books and articles on the subject of pumps. But the editors and the publisher felt that a need existed for a handbook on pumps that would present this information in a compact and authoritative form. The format of a handbook permits a selection of the most versatile group of contributors, each an expert on his particular subject, each with a background of experience that makes him particularly knowledgeable in the area assigned to him. This handbook deals first with the theory, construction details, and performance characteristics of all the major types of pumps—centrifugal pumps, power pumps, steam pumps, screw and rotary pumps, jet pumps, and many of their variants. It deals with prime movers, couplings, controls, valves, and the instruments used in pumping systems. xxi xxii PREFACE TO THE FIRST EDITION It treats in detail the systems in which pumps operate and the characteristics of these systems. And because of the many services in which pumps have to be applied, a total of 21 different services—ranging from water supply to steam power plants, construction, marine applications, and refrigeration to metering and solids pumping—are examined and described in detail, again by a specialist in each case. Finally, the handbook provides information on the selection, purchasing, installation, operation, testing, and maintenance of pumps. An appendix provides a variety of technical data useful to anyone dealing with pumping equipment. We are greatly indebted to the men who supplied the individual sections that make up this handbook. We hope that our common task will have produced a handbook that will help its user to make a better and more economical pump installation than he or she would have done without it, to install equipment that will perform more satisfactorily and for longer uninterrupted periods, and when trouble occurs, to diagnose it quickly and accurately. If this handbook does all this, the contributors, its editors, and its publisher will be pleased and satisfied. No doubt a few readers will look for subject matter that they will not find in this handbook. Into the making of decisions on what to include and what to leave out must always enter an element of personal opinion; therefore we will feel some responsibility for their disappointment. But we submit that it was quite impossible to include even everything we had wanted to cover. As to our possible sins of commission, they are obviously unknown to us at this writing. We can only promise that we shall correct them if an opportunity is afforded to us. IGOR J. KARASSIK CONTENTS List of Contributors / ix Preface to the Third Edition / xvii Preface to the Second Edition / xix Preface to the First Edition / xxi SI Units—A Commentary / xxiii Chapter 1 Introduction: Classification and Selection of Pumps 1.1 Chapter 2 Centrifugal Pumps 2.1 2.1 2.2 Centrifugal Pump Theory / 2.3 Centrifugal Pump Construction / 2.97 2.2.1 Centrifugal Pumps: Major Components / 2.97 2.2.2 Centrifugal Pump Packing / 2.183 2.2.3 Centrifugal Pump Mechanical Seals / 2.197 2.2.4 Centrifugal Pump Injection-Type Shaft Seals / 2.239 2.2.5 Centrifugal Pump Oil Film Journal Bearings / 2.247 vi CONTENTS 2.2.6 2.2.7 2.3 2.4 Centrifugal Pump Magnetic Bearings / 2.277 Sealless Pumps / 2.295 2.2.7.1 Magnetic Drive Pumps / 2.297 2.2.7.2 Canned Motor Pumps / 2.315 Centrifugal Pump Performance / 2.327 2.3.1 Centrifugal Pumps: General Performance Characteristics / 2.327 2.3.2 Centrifugal Pump Hydraulic Performance and Diagnostics / 2.397 2.3.3 Centrifugal Pump Mechanical Performance, Instrumentation, and Diagnostics / 2.405 2.3.4 Centrifugal Pump Minimum Flow Control Systems / 2.437 Centrifugal Pump Priming / 2.453 Chapter 3 Displacement Pumps 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 Power Pump Theory / 3.3 Power Pump Design and Construction / 3.21 Steam Pumps / 3.37 Displacement Pump Performance Instrumentation and Diagnostics / 3.63 Displacement Pump Flow Control / 3.75 Diaphragm Pumps / 3.85 Screw Pumps / 3.99 Vane, Gear, and Lobe Pumps / 3.123 Chapter 4 Jet Pumps 4.1 4.2 5.2 6.2 5.1 Metallic Materials of Pump Construction (and Their Damage Mechanisms) / 5.3 Materials of Construction for Nonmetallic (Composite) Pumps / 5.49 Chapter 6 Pump Drivers 6.1 4.1 Jet Pump Theory / 4.3 Jet Pump Applications / 4.23 Chapter 5 Materials of Construction 5.1 3.1 Prime Movers / 6.3 6.1.1 Electric Motors and Motor Controls / 6.3 6.1.2 Steam Turbines / 6.37 6.1.3 Engines / 6.57 6.1.4 Hydraulic Turbines / 6.77 6.1.5 Gas Turbines / 6.89 Speed-Varying Devices / 6.99 6.2.1 Eddy-Current Couplings / 6.99 6.1 vii CONTENTS 6.3 6.2.2 Single-Unit Adjustable-Speed Electric Drives / 6.109 6.2.3 Fluid Couplings / 6.127 6.2.4 Gears / 6.143 6.2.5 Adjustable-Speed Belt Drives / 6.167 Power Transmission Devices / 6.175 6.3.1 Pump Couplings and Intermediate Shafting / 6.175 6.3.2 Hydraulic Pump and Motor Power Transmission Systems / 6.191 Chapter 7 Pump Controls and Valves 7.1 Chapter 8 Pump Systems 8.1 8.1 8.2 8.3 8.4 General Characteristics of Pumping Systems and System-Head Curves / 8.3 Branch-Line Pumping Systems / 8.83 Waterhammer / 8.91 Pump Noise / 8.109 Chapter 9 Pump Services 9.1 9.2 9.3 9.4 9.5 9.6 9.7 9.8 9.9 9.10 9.11 9.12 9.13 9.14 9.15 9.16 9.17 9.18 Water Supply / 9.3 Sewage Treatment / 9.25 Drainage and Irrigation / 9.45 Fire Pumps / 9.57 Steam Power Plants / 9.73 Chemical Industry / 9.113 Petroleum Industry / 9.133 Pulp and Paper Mills / 9.157 Food and Beverage Pumping / 9.187 Mining / 9.197 Marine Pumps / 9.215 Refrigeration, Heating, and Air Conditioning / 9.253 Pumped Storage / 9.261 Nuclear / 9.279 9.14.1 Nuclear Electric Generation / 9.279 9.14.2 Nuclear Pump Seismic Qualifications / 9.301 Metering / 9.313 Solids Pumping / 9.321 9.16.1 Hydraulic Transport of Solids / 9.321 9.16.2 Application and Construction of Centrifugal Solids Handling Pumps / 9.351 9.16.3 Construction of Solids-Handling Displacement Pumps / 9.369 Oil Wells / 9.377 Cryogenic Liquefied Gas Service / 9.399 9.1 viii 9.19 9.20 9.21 9.22 CONTENTS Aerospace / 9.409 9.19.1 Aircraft Fuel Pumps / 9.409 9.19.2 Liquid Rocket Propellant Pumps / 9.431 Portable Transfer of Hazardous Liquids / 9.441 Water Pressure Booster Systems / 9.447 Hydraulic Presses / 9.463 Chapter 10 Intakes and Suction Piping 10.1 10.2 10.1 Intakes, Suction Piping, and Strainers / 10.3 Intake Modeling / 10.39 Chapter 11 Selecting and Purchasing Pumps 11.1 Chapter 12 Installation, Operation, and Maintenance 12.1 Chapter 13 Pump Testing 13.1 Appendix Technical Data A.1 Index I.1 C • H • A • P • T • E • R • 1 Introduction: Classification Selection Pumps AND OF W. C. Krutzch Paul Cooper 1.2 CHAPTER ONE INTRODUCTION ______________________________________________________ Only the sail can contend with the pump for the title of the earliest invention for the conversion of natural energy to useful work, and it is doubtful that the sail takes precedence. Because the sail cannot, in any event, be classified as a machine, the pump stands essentially unchallenged as the earliest form of machine for substituting natural energy for human physical effort. The earliest pumps we know of are variously known, depending on which culture recorded their description, as Persian wheels, waterwheels, or norias. These devices were all undershot waterwheels containing buckets that filled with water when they were submerged in a stream and that automatically emptied into a collecting trough as they were carried to their highest point by the rotating wheel. Similar waterwheels have continued in existence in parts of the Orient even into the twentieth century. The best-known of the early pumps, the Archimedean screw, also persists into modern times. It is still being manufactured for low-head applications where the liquid is frequently laden with trash or other solids. Perhaps most interesting, however, is the fact that with all the technological development that has occurred since ancient times, including the transformation from water power through other forms of energy all the way to nuclear fission, the pump remains probably the second most common machine in use, exceeded in numbers only by the electric motor. Because pumps have existed for so long and are so widely used, it is hardly surprising that they are produced in a seemingly endless variety of sizes and types and are applied to an apparently equally endless variety of services. Although this variety has contributed to an extensive body of periodical literature, it has also tended to preclude the publication of comprehensive works. With the preparation of this handbook, an effort has been made to create just such a comprehensive source. Even here, however, it has been necessary to impose a limitation on subject matter. It has been necessary to exclude material uniquely pertinent to certain types of auxiliary pumps that lose their identity to the basic machine they serve and where the user controls neither the specification, purchase, nor operation of the pump. Examples of such pumps would be those incorporated into automobiles or domestic appliances. Nevertheless, these pumps do fall within classifications and types covered in the handbook, and basic information on them may therefore be obtained herein after the type of pump has been identified. Only specific details of these highly proprietary applications are omitted. Such extensive coverage has required the establishment of a systematic method of classifying pumps. Although some rare types may have been overlooked in spite of all precautions, and obsolete types that are no longer of practical importance have been deliberately omitted, principal classifications and subordinate types are covered in the following section. CLASSIFICATION OF PUMPS___________________________________________ Pumps may be classified on the basis of the applications they serve, the materials from which they are constructed, the liquids they handle, and even their orientation in space. All such classifications, however, are limited in scope and tend to substantially overlap each other. A more basic system of classification, the one used in this handbook, first defines the principle by which energy is added to the fluid, goes on to identify the means by which this principle is implemented, and finally delineates specific geometries commonly employed. This system is therefore related to the pump itself and is unrelated to any consideration external to the pump or even to the materials from which it may be constructed. Under this system, all pumps may be divided into two major categories: (1) dynamic, in which energy is continuously added to increase the fluid velocities within the machine INTRODUCTION: CLASSIFICATION AND SELECTION OF PUMPS 1.3 FIGURE 1 Classification of dynamic pumps to values greater than those occurring at the discharge so subsequent velocity reduction within or beyond the pump produces a pressure increase, and (2) displacement, in which energy is periodically added by application of force to one or more movable boundaries of any desired number of enclosed, fluid-containing volumes, resulting in a direct increase in pressure up to the value required to move the fluid through valves or ports into the discharge line. Dynamic pumps may be further subdivided into several varieties of centrifugal and other special-effect pumps. Figure 1 presents in outline form a summary of the significant classifications and subclassifications within this category. Displacement pumps are essentially divided into reciprocating and rotary types, depending on the nature of movement of the pressure-producing members. Each of these major classifications may be further subdivided into several specific types of commercial importance, as indicated in Figure 2. Definitions of the terms employed in Figures 1 and 2, where they are not self-evident, and illustrations and further information on classifications shown are contained in the appropriate sections of this book. 1.4 CHAPTER ONE FIGURE 2 Classification of displacement pumps OPTIMUM GEOMETRY VERSUS SPECIFIC SPEED _________________________ Optimum geometry of pump rotors is primarily influenced by the specific speed NS or S, defined as shown in Figure 3. This parameter is one of the dimensionless groups that emerges from an analysis of the complete physical equation for pump performance. In this equation, performance quantities such as efficiency h and head H (or just H) are stated to be functions of the volume flow rate Q, rotative speed N or angular speed , rotor diameter D or radius r, viscosity, NPSHA, and a few quantities that have lesser influence. For low viscosity (high Reynolds number) and NPSHA that exceeds what the pump requires (namely NPSHR), the performance in terms of the head coefficient c  gH/(2r2) is influenced only by the flow coefficient or “specific flow” Qs  Q/(r3). Now, if one divides Qs1/2 by c3/4, the rotor INTRODUCTION: CLASSIFICATION AND SELECTION OF PUMPS 1.5 FIGURE 3 Optimum geometry as a function of BEP specific speed (for single stage rotors). radius r ( D/2) drops out (which is convenient because we don’t usually know it ahead of time), and we get the universal specific speed S as the major dependent variable—in terms of which the hydraulic design is optimized for maximum efficiency, as shown in Figure 3. This optimum geometry carries with it an associated unique value of the head coefficient c, thereby effectively sizing the rotor. For “rotodynamic” or impeller pumps, imagining speed N and head H to be constant over the NS-range shown yields increasing optimum impeller diameter as shown. This size progression shows that the optimum head coefficient c decreases with increasing specific speed. Outside the NS range shown in Figure 3 for each type of rotor, the efficiency becomes unsatisfactory in comparison to that achievable with the configuration shown for this NS. Rotary positive displacement machines such as vane pumps, gear pumps, and a variety of screw pump configurations are more appropriate for the lower values of NS, the lowest NSvalues requiring reciprocating (piston or plunger) positive displacement pumps. Regarding units for these relationships, the rotative speed N is in revolutions per second (rps) unless stated to be in rpm because the quantity of gH usually has the units of length squared per second squared. The diameter D has the same length unit as the head; for example, in the rotor size equation, head in feet would imply diameter in feet. The universal specific speed S has the same value for any combination of consistent units, and similarly shaped turbine and compressor wheels have similar values of S —making it truly “universal.” Note that for the unit of time of seconds,  is given as radians per second [ N(rpm)  p/30], where radians are unitless. SELECTION OF PUMPS _______________________________________________ Given the variety of pumps that is evident from the foregoing system of classification, it is conceivable that an inexperienced person might well become somewhat bewildered in trying to determine the particular type to use in meeting most effectively the requirements for a given installation. Recognizing this, the editors have incorporated in Chapter 11, “Selecting and Purchasing Pumps,” a guide that provides the reader with reasonable familiarity regarding the details that must be established by or on behalf of the user in order to assure an adequate match between system and pump. 1.6 CHAPTER ONE FIGURE 4 Approximate upper limit of pressure and capacity by pump class Supplementing the information contained in Chapter 11, the sections on centrifugal, rotary, and reciprocating pumps also provide valuable insights into the capabilities and limitations of each of these classes. None of these, however, provide a concise comparison between the various types, and Figure 4 has been included here to do just that, at least for the basic criteria of pressure and capacity. The lines plotted in Figure 4 for each of the three pump classes represent the upper limits of pressure and capacity currently available commercially throughout the world. At or close to the limits shown, only a few sources may be available, and pumps may well be specially engineered to meet performance requirements. At lower values of pressure and capacity, well within the envelopes of coverage, pumps may be available from dozens of sources as pre-engineered, or standard, products. Note also that reciprocating pumps run off the pressure scale, whereas centrifugals run off the capacity scale. For the former, some highly specialized units are obtainable at least up to 150,000 lb/in2 (10,350 bar)1 and perhaps slightly higher. For the latter, custom-engineered pumps would probably be available up to about 3,000,000 U.S. gal/min (680,000 m3/h), at least for pressures below 10 lb/in2 (0.69 bar). Given that the liquid can be handled by any of the three basic types and given conditions within the coverage areas of all three, the most economic order of consideration for a given set of conditions would generally be centrifugal, rotary, and reciprocating, in that order. In many cases, however, either the liquid may not be suitable for all three or other considerations—such as self-priming or air-handling capabilities, abrasion resistance, control requirements, or variations in flow—may preclude the use of certain pumps and limit freedom of choice. Nevertheless, it is hoped that the information in Figure 4 will be a useful adjunct to that contained elsewhere in this volume. 1 1 bar  105 Pa. For a discussion of bar, see “SI Units—A Commentary” in the front matter. C • H • A • P • T • E • R • 2 Centrifugal pumpS SECTION 2.1 CENTRIFUGAL PUMP THEORY PAUL COOPER INTRODUCTION ______________________________________________________ A centrifugal pump is a rotating machine in which flow and pressure are generated dynamically. The inlet is not walled off from the outlet as is the case with positive displacement pumps, whether they are reciprocating or rotary in configuration. Rather, a centrifugal pump delivers useful energy to the fluid or “pumpage” largely through velocity changes that occur as this fluid flows through the impeller and the associated fixed passageways of the pump; that is, it is a “rotodynamic” pump. All impeller pumps are rotodynamic, including those with radial-flow, mixed-flow, and axial-flow impellers: the term “centrifugal pump” tends to encompass all rotodynamic pumps. Although the actual flow patterns within a centrifugal pump are three-dimensional and unsteady in varying degrees, it is fairly easy, on a one-dimensional, steady-flow basis, to make the connection between the basic energy transfer and performance relationships and the geometry or what is commonly termed the “hydraulic design” (more properly the “fluid dynamical design”) of impellers and stators or stationary passageways of these machines. In fact, disciplined one-dimensional thinking and analysis enables one to deduce pump operational characteristics (for example, power and head versus flow rate) at both the optimum or design conditions and off-design conditions. This enables the designer and the user to judge whether a pump and the fluid system in which it is installed will operate smoothly or with instabilities. The user should then be able to understand the offerings of a pump manufacturer, and the designer should be able to provide a machine that optimally fits the user’s requirements. 2.3 2.4 CHAPTER 2 The complexities of the flow in a centrifugal pump command attention when the energy level or power input for a given size becomes relatively large. Fluid phenomena such as recirculation, cavitation, and pressure pulsations become important; “hydraulic” and mechanical interactions—involving stress, vibration, rotor dynamics, and the associated design approaches, as well as the materials used—become critical; and operational limits must be understood and respected. NOMENCLATURE ____________________________________________________ The units for each quantity defined are as stated in this nomenclature, unless otherwise specifically stated in the text, equations, figures, or tables. NOTE: A  area, in2 (mm). a  constant of the diffuser or volute configuration in Pfleiderer’s slip relation. a  radius of impeller disk, ft (m),  rt, 2. Ap  area of flow passage normal to the flow direction, ft2 (m2). b  width of an impeller or other bladed passage in the meridional plane, ft (m). NOTE: When dealing with radial thrust, b2 includes also the thickness of the shrouds. Cp  specific heat of liquid being pumped, Btu/(lbm-degF); [kcal/(kg-degC)]. c or V  absolute velocity, ft/sec (m/s). D  diameter; unless otherwise subscripted  impeller exit diameter, ft (m). d  diameter, ft (m). Dh  hydraulic diameter of flow passage ( 4Ap/ ), ft (m). F  thrust force, lbf (N). g  acceleration due to gravity,  32.174 ft/sec2 (9.80665 m/s2) at earth sea level. go  constant in Newton’s Second Law,  32.174 (lbm-ft)/(lbf-sec2). (There is no SI equivalent; use the dimensionless constant 1 in place of go in SI computations.) {gp}  set of fluid properties associated with gas-handling phenomena H  head of liquid column, ft (m) (Eq. 3); can also have the same meaning as the change in head H (that is, the same meaning as “pump head”). H  change in head across pump or pump stage, also called the “pump head” or “total dynamic head” ft (m). H  the small reduction in pump head (usually 3%) in testing for NPSHR, ft (m). He  Hi,   the ideal head for an infinite number of blades that produce no blockage. Hi  ideal head [ H  g(HL)], ft (m) (Eq. 15b); sometimes called the “input head.” HL  head loss, ft (m). gHL  all losses in the main flow passages from pump inlet to pump outlet, ft (m). h  static enthalpy in Btu/lbm times goJ, ft2/sec2; (or in kcal/kg times J, m2/s2  J/kg). 2.1 CENTRIFUGAL PUMP THEORY 2.5 hsv or NPSH  net positive suction head, ft (m). ID  inner diameter. J  the mechanical equivalent of heat, 778 ft-lbf/Btu (4184 N-m/kcal). /  blade, vane, or passage arc length, ft (m). M or T  torque, lbf-ft (N-m). m  distance in streamwise direction in meridional plane (Figure 14), in or ft (m). # m  mass flow rate, lbf-sec/ft (kg/s),  rQ. MCSF or Qmin  minimum continuous stable flow, ft3/sec (m3/s). N or n  rotative speed of the impeller, rpm. NPSH or hsv  net positive suction head, ft (m). NPSHA or NPSHA  available NPSH. NPSHR or NPSHR  required NPSH to prevent significant loss ( 3%) of pump p or to protect the pump against cavitation damage, whichever is greater. NPSH3% or NPSH3%  required NPSH to prevent significant loss ( 3%) of pump p; this is the “performance NPSH” defined in Section 2.3.1. nb or Zi  number of impeller blades. nq  specific speed in rpm, m3/s, m units (Eq. 38b)  Ns/51.64 (Eq. 39c). nv or Zd  number of vanes in diffuser or stator. Ns or Ns,(US) or ns  specific speed in rpm, gpm, ft units (Eq. 38a). Nss or S  suction specific speed in rpm, gpm, ft units (Eq. 42). OD  outer diameter. P  total pressure, lbf/ft2 (Pa). p  pressure, lbf/ft2 [Pa (N/m2)] ( “static pressure”). p  pressure rise, lbf/ft2 (Pa). pL  pressure loss, lbf/ft2 (Pa). pL, i  impeller pressure loss from its inlet to the point of interest, lbf/ft2 (Pa). pL, i  I/L  pressure loss pL, i in impeller plus pressure loss in inlet passage, lbf/ft2 (Pa). gpL  all losses in the main flow passages from pump inlet to pump outlet, lbf/ft2 (Pa). pv or pvp  vapor pressure of liquid being pumped, lbf/ft2 (Pa). PI  power delivered to all fluid flowing through the impeller, ft-lbf/sec (kW). PS  shaft power, ft-lbf/sec (kW).  perimeter of flow passage cross section normal to the flow direction, ft (m). Q  volume flow rate or, more conveniently, “flow rate” or “capacity,” ft3/sec (m3/s). QDR  flow rate below which discharge recirculation exists, ft3/sec (m3/s). QL  leakage from impeller exit to inlet, ft3/sec (m3/s). Qmin or MCSF  minimum continuous stable flow rate, ft3/sec (m3/s). QR  flow rate below which recirculation exists, ft3/sec (m3/s). 2.6 CHAPTER 2 QSR  flow rate below which suction recirculation exists, ft3/sec (m3/s). Q3D (quasi-3D)  quasi-three dimensional. R  radius of curvature of meridional streamline, ft (m) (Figures 13, 14, and 25). r  radial distance from axis of rotation, ft (m). rb  radial distance from axis of rotation to center of circle defining impeller passage width, ft (m) (Figures 13 and 25). re  maximum value of r within the “eye plane.” (Figures 13 and 25). s  width of gap between impeller disk and adjacent casing wall, ft (m). S  Nss, suction specific speed in rpm, gpm, ft units (Eq. 42). sp. gr.  specific gravity, namely, the ratio of liquid density to that of water at 60°F (15.6°C). {S}  set of flow properties associated with solids in the pumpage T  axial thrust, lbf (N). T or T or M  torque, lbf-ft (N-m). T  temperature, °F or °R (°C or °K). Tc  temperature rise due to compression, °F (°C). t  time, sec (s). t  blade or vane thickness, ft (m). u  internal energy in Btu/lbm multiplied by goJ, ft2/sec2; (or in kcal/kg times J, m2/s2). U  tangential speed r of the point on the impeller at radius r, ft/sec (m/s). Ue  the value of U at the maximum radial location re within the “eye plane.” V  volume, ft3 (m3). V or c  absolute velocity of fluid, ft/sec (m/s). Ve  the average value of the meridional velocity component Vm in the eye ( Q/Ae), ft/sec (m/s). Vs  slip velocity (Figure 15), ft/sec (m/s). W or w  velocity of fluid relative to rotating impeller, ft/sec (m/s). Wg  the one-dimensional value of W that would exist if there were no slip. w1  throat width (Figure 21), ft (m). y  transformed distance along blade from trailing edge (Figure 19), in or ft (m). z  axial distance in polar coordinate system, ft (m). Z or Ze  elevation coordinate, ft (m). Zd or nv  number of vanes in diffuser or stator. Zi or nb  number of impeller blades. a  angle of the absolute velocity vector from the circumferential direction. b  angle of the relative velocity vector or impeller blade in the plane of the velocity diagram (as seen, for example, in Figure 3) from the circumferential (tangential) direction. 2.1 CENTRIFUGAL PUMP THEORY g  * 0* e e2 e2,b h u 2.7  fluid weight density, lbf/ft3 (N/m3)  rg. (1N  1 kg-m/s2). = clearance, ft (m)  displacement thickness of the boundary layer, ft (m).  displacement thickness of the zero-pressure-gradient boundary layer, ft (m), ( 0.002 / for turbulent boundary layers at n  1 cs in typical centrifugal pumps).  absolute roughness height, ft (m)  fraction of impeller discharge meridional area (that is, the area normal to the velocity component Vm, 2) that is not blocked by the thickness of the blades and the boundary layer displacement thickness on blades and on hub and shroud surfaces.  fraction of the circumference at the exit of the impeller that is not blocked by the thickness of the blades and boundary layer displacement thickness on blades. (See computation in Table 4.)  hp  pump efficiency; or a component efficiency (different subscript, Eqs. 8–11).  rotational polar coordinate or central angle about the impeller axis, radians. In a polar-coordinate description of impeller blades or stationary vanes, u becomes the construction angle and is usually regarded as positive in the direction of the blade development from inlet to exit of the impeller or other blade row. NOTE: m  slip factor  Vs/U2 (1 h0, where h0 is the slip factor as defined by Busemann18.) m  absolute viscosity, lbf-sec/ft2 (Pa-s or N-s/m2); often quoted in centipoises, abbreviated to “cp” (1 cp  0.001 Pa-s). [(m in cp)  sp. gr. (n in cs).] n  kinematic viscosity ( m/r), ft2/sec (m2/s); often quoted in centistokes, abbreviated to “cs” (1 cs  1 mm2/s). [(n in cs)  (m in cp)/sp. gr. r  fluid mass density, lbf-sec2/ft4 (kg/m3), g/g. s  solidity (Eq. 53). s  Thoma’s cavitation parameter  hsv/H. T or T or M  torque, lbf-ft (N-m). f  flow coefficient. fe  Ve/Ue  impeller inlet or eye flow coefficient. fi (or fi, 2)  impeller exit flow coefficient  Vm, 2/U2 (Figure 12). c  head coefficient (Figure 12); stream function (Figure 14). ci  ideal head coefficient [ ci, 2  Vu, 2/U2 for zero inlet swirl (Vu, 1  0)]. ci, 2  Vu, 2/U2 [ ci for zero inlet swirl (Vu, 1  0)].   angular speed of the impeller in radians per second (1/s)  Np/30. s  universal specific speed (unitless) (Eq. 37)  Ns/2733 (Eq. 38a)  nq/52.92 (Eq. 38b). ss  universal suction specific speed (unitless) (Eq. 41)  Nss/2733 (Eq. 42). {2-ph}  set of fluid properties associated with vaporization 2.8 CHAPTER 2 Subscripts b D DF d e ex f h i i (or ideal) i (or imp) in (or s) I/L I L m m mean n o out p R S s (or in) SE s/o r rms s s s sh stg T  impeller blade.  drag due to disk friction, bearings, and seals.  disk friction.  discharge flange or exit (ex) of the pump.  at the “eye” of the impeller. The “eye” is the throat (minimumdiameter point) at the entrance into the impeller and is the area defined by the “eye plane,” which is normal to the axis of rotation. “e” can refer more specifically to the shroud or maximum-diameter point within the eye, as with re (Figure 13) or Ue. The inlet tips of the impeller blades are generally at or near this location.  exit of diffuser or the discharge flange or port of the pump (d).  the direction of the flow.  hub.  inner limit of region or gap (Tables 4 and 5)  ideal.  impeller.  pump inlet flange or port.  inlet passage; that is, the passage from the pump inlet flange or port to the impeller.  input to fluid.  losses.  “mechanical” (pertaining to efficiency, Eq. 9).  component of velocity in the meridional plane (that is, the axialradial plane containing the axis of rotation).  the 50% or rms meridional streamline.  normal or BEP value.  outer limit of region or gap (Tables 4 and 5).  pump outlet flange or port.  pressure side of blade or passage.  value of r at the impeller ring clearance.  shaft.  suction flange or inlet of the pump.  shockless entry (that is, inlet velocity vector aligned with blade camber line).  shut-off or zero flow rate Q.  in the radial direction.  the 50% or mean meridional streamline.  shaft.  suction side of blade or passage.  same meaning as sh and t.  shroud (also at the eye plane at inlet—and in general “t” at outlet).  stage.  entry throat of volute or diffuser. 2.1 CENTRIFUGAL PUMP THEORY 2.9 t  the tip or maximum radial position of the impeller blades at inlet or outlet (same meaning as s and sh). t  tongue or cutwater. u (see u, below). v  volumetric (pertaining to efficiency, Eq. 11). v  volute. z  in the axial direction. u or u  component of velocity in the circumferential direction (that is, the tangential direction in the polar view that is perpendicular to the axis of rotation). 1  impeller inlet at the blade leading edge—at the mean unless further defined. 2  impeller outlet at the blade trailing edge—at the mean unless further defined. 3  volute base circle or entrance to diffuser.   for an infinite number of blades that also produce zero blockage of the flow. Superscripts ¯¯  average value ENERGY TRANSFER__________________________________________________ Hydraulics or fluid dynamics has the primary influence on the geometry of a rotodynamic pump stage—of all the engineering disciplines involved in the design of the machine. It is basic to the energy transfer or pumping process. Staging is also influenced by the other disciplines, especially in high-energy pumps. The basic energy transfer relationships need to be thoroughly understood to achieve a credible design and to understand the operation of these machines. Action of the mechanical input shaft power to effect an increase in the of energy of the pumpage is governed by the first law of thermodynamics. Realization of that energy in terms of pump pressure rise or head involves losses and the second law of thermodynamics. The First Law of Thermodynamics Fluid flow, whether liquid or gas, through a centrifugal pump is essentially adiabatic, heat transfer being negligible in comparison to the other forms of energy involved in the energy transfer process. (Yet, even if the process were not adiabatic, the density of a liquid is only weakly dependent on temperature.) Further, while the delivery of energy to fluid by rotating blades is inherently unsteady (varying pressure from blade to blade as viewed in an absolute reference frame), the flow across the boundaries of a control volume surrounding the pump is essentially steady, and the first law of thermodynamics for the pump can be expressed in the form of the adiabatic steady-flow energy equation (Eq. 1) as follows: PS  m c a h  # where hu p r V2  gZe b 2 out ah  V2  gZe b d 2 in (1) 2.10 CHAPTER 2 FIGURE 1 Energy transfer in a centrifugal pump # Here, shaft power Ps is transformed into fluid power, which is the mass flow rate m times the change in the total enthalpy (which includes static enthalpy, velocity energy per unit mass, and potential energy due to elevation in a gravitational field that produces acceleration at rate g) from inlet to outlet of the control volume (Figure 1). When dealing with essentially incompressible liquids, the shaft power is commonly expressed in terms of “head” and mass flow rate, as in Eq. 2: PS #  g¢H  ¢u m where H p V2   Ze rg 2g (2) (3) The change in H is called the “head” H of the pump; and, because H (Eq. 3) includes the velocity head V2/2g and the elevation head Ze at the point of interest, H is often called the “total dynamic head.” H is often abbreviated to simply “H” and is the increase in height of a column of liquid that the pump would create if the static pressure head p/rg and the velocity head V2/2g were converted without loss into elevation head Ze at their respective locations at the inlet to and outlet from the control volume; that is, both upstream and downstream of the pump. The Second Law of Thermodynamics: Losses and Efficiency As can be seen from Eq. 2, not all of the mechanical input energy per unit mass (that is, the shaft power per unit of mass flow rate) ends up as useful pump output energy per unit mass gH. Rather, losses produce an internal energy increase u (accompanied by a temperature increase) in addition to that due to any heat transfer into the control volume. This fact is due to the second law of thermodynamics and is expressed for pumps in Eq 4: g¢H 6 PS # m or h 6 1 (4) 2.1 CENTRIFUGAL PUMP THEORY 2.11 FIGURE 2 Determining component efficiencies. (This is a meridional view.) # where m  rQ The losses in the pump are quantified by the overall efficiency h, which must be less than unity and is expressed in Eq. 5: # h g¢Hm  Overall Pump Efficiency PS (5) It should be pointed out here that real liquids undergo some compression—which is accompanied by a reversible increase in the temperature Tc of the liquid—called the “heat of compression.” This portion of the actual total temperature rise T is in addition to that arising from losses and must therefore be taken into account when determining efficiency from measurements of the temperature rise of the pumpage.1 See the discussion on this subject in Section 2.3.1. To pinpoint the losses, it is convenient to deal with them in terms of “component efficiencies.” For the typical shrouded- or closed-impeller pump shown in Figure 2, Eq. 5 can be rewritten as follows: h PI PS Noting that g¢H1m  mL 2 # # PI Pi  g¢Hi 1m  mL 2 # and # m  rQ Hi  Ideal Head one may rewrite Eq. 6 as follows: # #m m m# L # (6) (7) 2.12 CHAPTER 2 h PI PS ¢H ¢Hi Q  hm Q  QL hHY hv (8) where PS PD PI “Mechanical” f hm   Efficiency PS PS ¢Hi gHL ¢H Hydraulic  fh  Efficiency HY ¢Hi ¢Hi Q Volumetric f hv  Efficiency Q  QL (9) (10) (11) Approximate formulas for the three component efficiencies of Eq. 8 will be given further on. Their product yields the overall pump efficiency as defined in Eq. 5, and reflects the following division of the pump losses: a. External drags on the rotating element due to i) bearings, ii) seals, and iii) fluid friction on the outside surfaces of the impeller shrouds—called “disk friction”; the total being PD  PS PI. Generally, the major component of PD is the disk friction, and the “mechanical efficiency” is that portion of the shaft power that is delivered to the fluid flowing through the impeller passages. b. Hydraulic losses in the main flow passages of the pump; namely, inlet branch, impeller, diffuser or volute, return passages in multistage pumps, and outlet branch. The energy loss per unit mass is ggHL  g(Hi H), the ratio of output head H to the input head Hi being the hydraulic efficiency. This is the major focus of the designer for typical centrifugal pump geometries (which are associated with normal “specific speeds”—to be defined later). The other two component efficiencies are then quite high and of relatively little consequence. c. External leakages totaling QL leaking past the impeller and back into the inlet eye. This leakage has received its share of the full amount of power PI  rg Hi (Q  QL) delivered to all the fluid (Q  QL) passing through the impeller. This leakage power is PL  rg Hi QL, which is lost as this fluid leaks back to the impeller inlet. The remaining fluid input power is thus (PI PL)  rg Hi Q, the ratio of this power to the total (PI) being the volumetric efficiency. There are exceptions to this convenient model for dividing up pump losses. The main exception is that if the pump has an open impeller, that is, one without either or both shrouds, that portion of the total leakage QL disappears. The leakage now occurs across the blade tips and affects the main flow passage hydraulic losses. The volumetric efficiency is now higher, but the hydraulic efficiency is lower. In that case disk friction is still present, as the impeller still has to drag fluid along the adjacent stationary wall(s). Another exception—for closed impellers—is that disk friction is fundamentally an inefficient pumping action, the fluid being flung radially outward2; and this can result in a slight increase in pump head if the fluid on the outside of an impeller shroud or disk is pumped into the main flow downstream of the impeller. VELOCITY DIAGRAMS AND HEAD GENERATION __________________________ The mechanism of the transfer of shaft torque (or power) to the fluid flowing within the impeller is fundamentally dynamic; that is, it is connected with changes in fluid velocity. This requires the introduction of Newton’s second law, which when combined with the first 2.1 CENTRIFUGAL PUMP THEORY 2.13 FIGURE 3 Impeller velocity diagrams (1 = inlet; 2 = outlet) law of thermodynamics, yields Euler’s Pump Equation. Fluid velocities at inlet and exit of the impeller are fundamental to this development. Fluid flowing along the blades of an impeller rotating at angular velocity  and viewed in the rotating reference frame of that impeller has relative velocity W. Vectorially adding W to impeller blade speed U  r yields the absolute velocity V, as shown in the velocity diagrams of Figure 3. Newton’s Second Law for Moments of Forces and Euler’s Pump Equation Relating impeller torque T to fluid angular momentum per unit mass rVu is the convenient way of applying Newton’s second law to centrifugal pumps. This is stated as follows for the control volume V that contains the pump impeller (Eq. 12): T  µµµV 1 0 1rrVu 2> 0t2>dV  µµrrVudQ (12) where T  TS TD is the summation of torques acting on the impeller; namely, the net torque TI acting on the fluid flowing through it. The volume integral (first term on the right side) of Eq. 12 is the unsteady term, which is zero for steady operation. It comes into play during changing or transient conditions, such as start up and shutdown; that is, when the angular momentum per unit volume rrVu is changing with time within the impeller volume V. The surface integral (second term on the right hand side) of Eq. 12 is the one that pump designers and users are mainly concerned with. Its integration over the exterior surface of the control volume V is effectively accomplished for most impellers by combining one-dimensional results from inlet to outlet on each of several stream surfaces—imagined to be nested surfaces of revolution bounded by the hub and shroud stream surfaces (indicated in Figure 2). Insight into the power of this term can be gained by taking the mean value of the integrand in terms of the velocities on a representative stream surface; that is, essentially the surface of revolution lying at an appropriate mean location between hub and shroud. Each of the two velocity diagrams of Figure 3 lies in a plane tangent to this mean stream surface. For flow through an impeller, the torque delivered to the fluid is therefore given by the following relationship involving these average quantities: TI  1m  mL 2 # or : # 1r2Vu, 2 r1Vu, 1 (13) 2.14 CHAPTER 2 PI  1m  mL 2 # # 1U2Vu, 2 U1Vu, 1 2 (14) Eq. 13 says that the torque is equal to the mass flow rate times the change of angular momentum per unit mass (rVu). This becomes the “power” statement of Eq. 14 when both sides are multiplied by . Following the statement of the second law of thermodynamics in Eq. 4, we now can similarly say that gH must be less than the power input to the fluid per unit of mass flow rate, namely (UVu) from Eq. 14. So, we now arrive at Euler’s Pump Equation—expressed three different ways as follows: or g¢H 6 ¢1UVu 2 (15a) g¢H  hHY ¢1UVu 2 (15c) Q  2prbVm (16) g¢Hi  ¢1UVu 2 (15b) The inequality (Eq. 15a) is quantified by Eq. 15b, which follows in view of Eq. 7. Eq. 15c then follows from the definition of hydraulic efficiency (Eq. 10). Euler’s Pump Equation makes one of the most profound statements in the field of engineering, because it determines the major geometrical features of the design of a rotodynamic machine. By reversing the inequality in Eq. 15a, the same principle applies to turbines; hence, the more encompassing title, “Euler’s Pump and Turbine Equation.” So, to design or analyze a pump, one needs to a) obtain the velocity diagrams that will produce the ideal head at the design flow rate and b) determine how the shape of these diagrams affects the hydraulic efficiency hHY, so as to obtain the desired pump stage head. Step (a) for a given pump is a simple one-dimensional exercise that utilizes the principles of continuity and kinematics (Eqs. 16 and 17) to construct the velocity diagrams for a given total impeller volume flow rate Q and pump rotative speed ( or N): Continuity: where W  Vm/sin bf Kinematics: Vu  U (17) W cos bf Step (b) is in essence the evaluation of the hydraulic losses gHL in Eq. 10, which depend mainly on the relative and absolute velocities, the associated flow passage dimensions, and incidence angles. Eq. 15c then gives the head that the pump stage will generate. Performing steps (a) and (b) at several other flow rates at the same speed enables one to develop the pump performance characteristics. STATIC PRESSURE GENERATION ______________________________________ The Extended Bernoulli Equation To estimate the losses, it is convenient first to investigate the static pressure and velocity head portions of the total head. Eq. 15c can be written in terms of the total pressure P, which equals rgH. Similarly, we may speak of hydraulic losses as losses of static pressure gpL, which equals rggHL; so P  Pin  r1UVu U1Vu, 1 2 pL (18) where, from Eq. 3, the static, dynamic and potential energy components of the total pressure are brought into evidence: 2.15 2.1 CENTRIFUGAL PUMP THEORY FIGURE 4 Velocity triangle and flow within impeller passageway Pp 1 rV2  rgZe  rgH 2 (19) Eq. 18 is the turbomachine form of the “extended Bernoulli equation,” which states that along a streamline the total pressure P—also known as the Bernoulli constant and defined in Eq. 19—is a) decreased by losses and b) increased by energy addition that occurs along the streamline. Centrifugal and Diffusion Effects in Impellers Changes in potential energy across a pump stage are small; so the static pressure rise is found essentially from subtracting the dynamic (velocity) pressure change from the total pressure change. Within the impeller, the static pressure in turn arises from a) centrifugal and b) passage diffusion effects. Fluid in the impeller passage of Figure 4b flows from low to high radius r or blade speed U, often also experiencing a decrease in passage relative velocity W. The geometry of the velocity diagram (Figure 4a) leads to the following combination of Eqs. 18 and 19 applied across the impeller: 1 1 ¢ a p  rV2m  rV2u b  ¢ 3rU 2 2 1U Wcosb2 4 pL, i ¢1rgZe 2 (20) and, because Vm  Wm, this simplifies to the following form of the extended Bernoulli equation, which applies along a streamline from the inlet of the impeller: 1 r1U2 2 U21  W21 W2 2 Static Pressure Change Centrifugal Effect Passage Diffusion pL, i { p1  { { { p rg1Ze Ze,1 2 Losses: Incidence Friction ∞ Secondary Flow ∂ 1r¢u2 Tip Leakage Mixing (21) Here, the U-increase corresponds to the centrifugal contribution to the static pressure rise, and the W-decrease to the diffusion contribution. There is no U-change along an axial streamline in an axial-flow impeller or propeller; so, static pressure rise is due only to diffusion. Radial-flow impellers, on the other hand, often have little or no net W-change, the centrifugal effect being paramount. [A study of the velocity diagrams of Figure 3 suggests that such impellers possess a high “degree of reaction.” The degree of reaction is defined as that fraction of the total energy addition within the impeller (Eq. 15b) that does not include the change in absolute velocity energy, (V 2/2). This fraction is, therefore, the sum 2.16 CHAPTER 2 of the static pressure energy change, that due to elevation change, and the energy losses in the impeller, as can be seen from Eqs. 18 and 19.] Collector Static Pressure Rise, Inlet Nozzle Drop A similar form of Eq. 21 applies in the stationary flow elements, where W is also the absolute velocity V and blade speed U is zero. (This would be a more recognizable form of the extended Bernoulli equation, wherein only losses modify the Bernoulli constant.3) A further static pressure increase occurs in the stationary collection system downstream of the impeller (with the attendant losses); namely, in the stationary volute or diffuser. This static rise is generally about a third of that in the impeller, and it is due only to diffusion, that is, the decrease in velocity in that passage. To complete the picture, there is often an increase in the comparatively small velocity in the approach passageway or nozzle or suction branch from the pump inlet port or flange to the impeller eye or blade leading edge. This is accompanied by an attendant small pressure drop. Internal Static Pressure Distribution If the fluid enters the pump from a stagnant pool, the total pressure at the impeller eye P1 will be very nearly the static pressure of the upstream pool (plus the pressure equivalent of the elevation of the pool above the eye). This is one reason why the local static pressure p within the pump is often referenced to P1, as indicated in the following form of Eq. 21: p where P1  1 r1U2 2 W2 2 P1  p 1  r pL, i rgZe rU1Vu, 1 V21  rgZe, 1 2 (22) (23) Figure 5 is an illustration of the internal static pressure development. The difference between P1 and p1 is due to the impeller inlet absolute velocity head or dynamic pressure rV12/2, a much larger difference existing at the impeller exit, namely rV22/2. Losses in the collector result in the pump or stage outlet total pressure Pout being less than P2, the rise in total pressure Ppump from port to port being Pout Pin. In the figure, Pin is very nearly the same as P1, the inlet passage loss being comparatively small for most pumps. NET POSITIVE SUCTION HEAD _________________________________________ Local reduction of the static pressure p to the vapor pressure pv of the liquid causes vaporization of the liquid and cavitation. Internal pressure drops are due to a) impeller inlet velocity head and inlet passage loss and b) blade loading and loss within the impeller. In order to prevent a substantial decrease of impeller pressure rise, the sum of these pressure drops should not exceed the difference between Pin and pv, the head equivalent of which is called “net positive suction head” or NPSH: Pin pv  NPSH 1 Pin  rgNPSH  pv rg (24) Insufficient NPSH leads to cavitation and loss of pump pressure rise. That is because the impeller can become filled with vapor, in which case the density r of the fluid within the impeller is then reduced by orders of magnitude. This in turn, as can be seen in Eqs. 18–22, results in essentially zero pump pressure rise; that is, total loss of pump performance. Eq. 24 substituted into Eq. 22 yields the local static pressure above vapor pressure in terms of the NPSH: 2.1 CENTRIFUGAL PUMP THEORY 2.17 FIGURE 5 Pump stage internal pressure development. Total pressure rise P = rgH. p pv  rgNPSH  pL, iI>L 1 r1U2 2 rgZe W2 2 (25) rU1Vu, 1 This, together with the foregoing pressure drops, which occur in the inlet region of the pump, is illustrated in Figure 5. The figure contains three plots of p along the representative streamline from 1 to 2, m being distance along this line in the meridional plane. These plots are for the suction side or trailing face of an impeller blade, the pressure side or driving face, and the average or mid-passage position. The middle or average pressure plot is readily described by Eq. 25 in terms of the local average W-distribution. The local bladeto-blade static pressure difference pp ps arises from the torque exerted on a strip of fluid between the blades and approximated here via blade-to-blade average velocity components in Newton’s Second Law for Moments of Forces: p dm r1pp ps 2bdm  rVmbr # ¢u # d1rVu 2 pp ∆ r Ω ps  2p rVm d1UVu 2 nb Æ dm (26) 2.18 CHAPTER 2 where, for ease of illustration, the blade-to-blade polar angle difference u is taken equal to 2p/nb, the actual value of u being slightly less than this due to the thickness of the blades. Thus, for example, too small a number of blades nb results in a larger value of pp ps and a lower minimum static pressure in the inlet region of the impeller. The density reduction in a cavitating impeller is difficult to predict analytically; therefore, empirical relationships for acceptable levels of NPSH have been developed and will be presented further on, as guidelines for design and performance prediction are developed. PERFORMANCE CHARACTERISTIC CURVES _____________________________ Velocity diagrams and ideal head-rise vary with flow rate Q as illustrated in Figure 6 for the typical case of constant rotative speed N or angular speed . Flow patterns in Figure 6b correspond to points on the characteristic curves of Figure 6a. The inlet velocity diagrams (just upstream of the impeller) are shown there for high and low flow rate—with zero swirl being delivered by the inlet passageway to the impeller; that is, Vu,1  0. The outlet velocity diagrams on Figure 6a are found one-dimensionally, the magnitude of the exit relative velocity vector W2 varying directly with Q and its direction being nearly tangent to the impeller blade. From these diagrams are found the absolute velocity vector V2 and its circumferential component Vu, 2. Because blade speed U2 is constant, the resulting plot of the ideal head Hi  U2Vu,2/g (from Eq. 15b) is a straight line, rising to the point U22/g at zero-Q or “shutoff head.” This is twice the impeller OD tip speed head U22/2g. The right-most velocity diagram in Figure 6a has zero Vu,2; however, the maximum or “runout” flow rate happens at lower Q than this. That is because the actual head H is less than Hi due to losses (as seen in Eq. 10), and H  0 at runout—where overall pump efficiency (Eq. 8) is also zero. This one-dimensional analysis works well in the vicinity of the best efficiency point (b.e.p. or BEP) and at higher Q because the fluid flows smoothly through the impeller pas- FIGURE 6A and B Characteristic performance curves of a pump stage, related to velocity diagrams 2.1 CENTRIFUGAL PUMP THEORY 2.19 sages as illustrated in Figure 6b for “high Q.” However, it fails at “low Q,” where recirculating flow develops—indicated by a substantial one-dimensional deceleration or reduction in the fluid velocity relative to those passages—that is, W2 V W1. This is analogous to a diffuser with side walls that diverge too much: the main fluid stream separates from one or both walls and flows along in a narrow portion of the passage in a jet—the rest of the passage being occupied with eddying fluid that can recirculate out of the impeller inlet and exit. Consequently, the real outlet velocity diagram at low Q is the one with the dashed lines and the smaller value of Vu, 2, rather than the solid-lined, one-dimensional diagram superimposed on it. This in turn reduces the ideal head at the low-Q point of the curves. To complicate matters further at low Q, one-dimensional application of this “corrected” outlet velocity diagram via Eq. 14 would produce a pump power consumption curve that passes through the origin of Figure 6a. Such a result (assuming negligible external drag power PD), is known not to occur in a real pump. Rather, superimposed on the jet flow pattern just described is recirculating fluid that leaves the impeller, gives up its angular momentum to its surroundings, and re-enters the impeller to be re-energized. In other words, the one-dimensional simplifications mentioned after Eq. 12 do not hold at low Q; rather, there is an added “recirculation power,” which is the UVu-change experienced by the recirculating fluid integrated over each element of re-entering mass flow rate4. The complexity of this recirculation destroys one’s ability to interpret pump performance under such conditions by means of velocity diagrams. Instead, a transition is made from empirical correlations for head and power at “shutoff” or zero net flow rate to the high-Q, one-dimensional analysis, enabling one to arrive at the complete set of characteristic curves for efficiency, power, and head illustrated in Figure 6a. In fact, impeller pressure-rise at shutoff is very nearly what would be expected due to the centrifugal effect of the fluid rotating as a solid body, namely rU22/2. The recirculating flow patterns seem to be merely superimposed with little effect on impeller pressure-rise. This recirculation, on the other hand, does produce some additional shutoff pressure rise in the collecting and diffusing passages downstream of the impeller. SCALING AND SIMILITUDE ____________________________________________ When a set of characteristic curves for a given pump stage is known, that machine can be used as a model to satisfy similar conditions of service at higher speed and a different size. Scaling a given geometry to a new size means multiplying every linear dimension of the model by the scale factor, including all clearances and surface roughness elements. The performance of the model is then scaled to correspond to the scaled-up model by requiring similar velocity diagrams (often called “velocity triangles”) and assuming that the influences of fluid viscosity and vaporization are negligible. The proportions associated with Eqs. 27, 29, and 32 illustrate this. The blade velocity U (Eq. 30) varies directly with rotative speed N or angular speed —and directly with size, as expressed by the radius r. For the velocity V (or W) to be in proportion to U, the flow rate Q must therefore vary as r3; hence, the “specific flow” Qs must be constant (Eq. 28). Further, as the head is the product of two velocities, it must vary as 2r2; hence, the head coefficient c must be constant (Eq. 31). Finally, as power is the product of pressure-rise and flow rate, shaft power Ps must vary as r3r5; hence, the power coefficient must be constant. Q  AVe 1Q and Q r32 A V r22 r2 ND3or  Constant  Qs D2 ND r32 (27) (28) 2.20 CHAPTER 2 g¢H  hHY ¢1UVu 2 At above Q, h  Constant, and N2D2 1 ¢H (29) (30) U  r and c PS  g¢H U22  constant rQg¢H h rN3D5 1 PS and P̂S  (31) PS r3r52 (32)  constant (33) Figure 7a is the result of following these similarity rules for a given pump that undergoes a change in speed from full speed to half speed without a change in size. The similar Q at half speed for a given Q at full speed is half that at full speed. At each such halfspeed Q-value, the head H is accordingly one-fourth of its full-speed value and the efficiency h is unchanged. One can avoid replotting the characteristic curves in this manner for every change in speed (and size) by expressing them nondimensionally in terms of Qs, c, h, and P̂s. They then all collapse on one another as illustrated in Figure 7b. Note that a change in pump geometry or shape of the hydraulic passageways destroys this similitude and necessarily produces a new set of curves—shaped differently but similar to each other. Similitude enables the engineer to work from a single dimensionless set of performance curves for a given pump model. This is a practical but special case of the more general statement that pump performance as represented by efficiency, head, and power, is more generally expressed in terms of the complete physical equation as follows: h, ¢H, PS  fct’s. 1Q, r2, , r, y, NPSH, 52-ph6, 5gp 6, 5S6,5/i 6 2 (34) h, c, P̂S  fct's. 1Qs, Re, t2, 52-£6, 5p 6, 5 g 6, 5Gi 6 2 (35) where {/i} is the infinite set of lengths that defines the pump stage geometry. A common group of these lengths is illustrated in Figure 8. Dimensionlessly, Eq. 34 becomes where the dimensionless quantities containing flow rate, viscosity and NPSH are respectively defined as follows: where Qs  Q r32 Re  r22 y t2  2gNPSH 2r22 Machine Cavitation Specific ReynoldsCoefficient Flow Number (36) and • {Gi}  {/i/r} defines the dimensionless geometry or shape. • {2- }  the dimensionless quantities arising from the set of fluid, thermal, vaporization, and heat transfer properties {2-ph} that influence the flow of two-phase vapor and liquid. These quantities come into play when the NPSH is low enough for such flow to be extensive enough to influence pump performance. 2.1 CENTRIFUGAL PUMP THEORY 2.21 FIGURE 7 Similar performance curves: a) dimensional; b) dimensionless • {g}  the dimensionless quantities arising from the set of properties associated with entrained solids and emulsifying fluids that affect the performance of slurry pumps and emulsion pumps. SPECIFIC SPEED AND OPTIMUM GEOMETRY ____________________________ The hydraulic geometry or shape of a pump stage can in principle be chosen for given values of the other independent variables in Eqs. 34 or 35 to optimize the resulting performance; for example, to maximize the best efficiency hBEP under certain conditions on the head and power. Two such conditions that are common are a) no positive slope is allowed anywhere along the H-vs.-Q curve of Figure 7 (called the “no drooping nor dip” condition) and b) the maximum power consumption must occur at the BEP (often called the “non-overloading” condition). A fundamental and generally typical pumping situation involves a) negligible influence of viscosity, (that is, high Reynolds number) b) the absence of two-phase fluid effects, (that is, the existence of sufficient NPSH or t) and c) the absence of solid particles and emulsion-related substances in the fluid. In this situation, Eq. 35 has one remaining significant independent variable; namely, the specific flow Qs, which in the definitions of Eq. 36 contains the volume flow rate Q, the pump speed , and the characteristic radius r2. Most users don’t know the 2.22 CHAPTER 2 FIGURE 8 Defining the geometry of a pump stage size of the pump stage a priori; so, r2 is eliminated by replacing Qs in Eq. 35 with a new quantity that is the result of dividing the square root of Qs by the 43-power of the head coefficient c. Thus, from the definitions just given of Qs and c, one arrives at the specific speed ⍀s as the independent variable in terms of which the geometry is optimized3: 1g¢H2 2Q s  3>4  2Qs c3>4 (37) For convenience, specific speed is usually expressed in terms of the conventional quantities N, Q, and H that correspond to the factors in Eq. 37, which quantities are expressed in the units commonly used commercially. For example, forms found in the United States and in Europe and the relationship of these to the truly unitless “universal specific speed” s defined in Eq. 37 are as follows: s  s  Thus, Ns, 1U.S.2 N1rpm2 2Q1USgpm2> 3 ¢H1ft2 4 3>4  2733.016 2733.016 nq N1rpm2 2Q1m3>s2> 3 ¢H1m2 4 3>4  52.919 52.919 nq 1m3>s, m2  Ns 1USgpm, ft2>51.64 (38a) (38b) (38c) 2.1 CENTRIFUGAL PUMP THEORY FIGURE 9 2.23 Optimum geometry as a function of BEP specific speed (for single-stage rotors) Rotor Shape as a Function of Specific Speed Optimization of pump hydraulic geometry in terms of the BEP specific speed has taken place empirically and analytically throughout the history of pump development. An approximate illustration of the results of this process for pump rotors or impellers is shown in Figure 9. Not only does the geometry emerge from the optimization process but also the head, flow, and power coefficients for each shape as well. Approximate values for the optimum BEP head coefficient c are shown on the figure. The actual rotor diameter can then be deduced as noted—from the c-definition of Eq. 31. The relationship among the various rotors is illustrated in the figure—assuming that they all have the same speed and head; that is, as one moves along the abscissa or specific speed axis of Figure 9, only the flow rate is changing as far as the illustrations of the rotors are concerned. As would be expected, therefore, high-specificspeed impellers need to have large passages relative to their overall diameter. This is powerfully illustrated if one contrasts the propeller (high-s) with the low-s centrifugal impeller. At the lower end of the specific speed range shown in Figure 9, rotodynamic pumps (that is, centrifugal pumps, in which category mixed- and axial-flow geometries are generally included) would be too low in efficiency to be practical. Rather, rotary positive displacement pumps take over because there is a transition through the drag pump domain. Sometimes called a regenerative or periphery pump, the drag pump is actually a rotodynamic machine, developing head peripherally around the impeller through successive passes radially through the blades on both sides until a barrier is reached at some point on the periphery, where the fluid is then discharged.5 The screw pump, on the other hand, is a truly positive displacement (rotary) machine. It can have two, three, or more meshing screws and can move large quantities of fluid— both single- and multiphase—against a large pressure difference p, giving it a specific speed range that extends well into centrifugal pump territory. Not shown is the progressive cavity pump, which has a single screw surrounded by an elastomer sealing member. Lower flow rates are readily accommodated by the vane pump, whereas gear pumps handle a higher range of pressure differences at such flow rates. Finally, extremely high pressures are produced by reciprocating pumps, the specific speed range of which extends off the figure on the left. 2.24 CHAPTER 2 Positive displacement pumps appear in Figure 9 in order to provide perspective. The concept of specific speed is not generally applied to these machines, because a given positive displacement pump can have such a wide range of pressure-rise capability at a chosen flow rate and speed as to make it difficult to associate a given rotor geometry with a particular value of specific speed. On the other hand, a unique rotodynamic pump geometry is readily associated with the specific speed of the BEP of such a machine. Performance of Optimum Geometries Figure 9 enables one to easily identify the pump stage types associated with required pumping tasks in terms of head, flow rate, and rotative speed. Beyond this general picture is the related performance of a real pump geometry in a real fluid. Although, for centrifugal pumps, the specific speed has the major effect on performance, the available NPSH and the viscosity of the pumpage also have an influence. These are evident in the following formal statement of the efficiency of an optimized pump (cf Eq. 35) 1hmax 2 5GI6opt  f 1s, ss, 52-£6, 5p 6, 5 g 6, Re, H, Q 2 (39) where the radius r2, representing the size, has been eliminated from the other variables in Eq. 35 by introducing the suction specific speed ss and the head-flow Reynolds number Re,H,Q, which are defined in Eqs. 40 and 41: Re, H, Q  2Q 1g¢H2 1>4  Re 2Qs c1>4 n 1g NPSH2 3>4 2Q ss   (40) 1t2>22 3>4 2Qs (41) where, the common form of the suction specific speed, called Nss, is given in commercial U.S. units by (Eq. 42) Nss, 1U.S.2  3NPSH 1ft2 4 3>4 N1rpm2 2Q1gpm2  ss (42) 2733.016 Size Effect For sufficiently high NPSH (or sufficiently low suction specific speed) and low viscosity (or high Reynolds number), real pumps also possess a strong size effect on efficiency. This is because, in normal manufacturing processes, the clearances d preventing internal leakage QL (for example, past the impeller sealing rings in Figure 2) do not scale up as rapidly as the size (represented by r2), nor do the surface roughness heights e. Thus, a larger pump tends to be more efficient. Strictly speaking, however, the geometry of the larger pump is not the same as that of the smaller pump, and this forces one to modify Eq. 39 by reintroducing two of the length ratios Gi that were part of the set {Gi} in Eq. 35 which characterize the hydraulic shape of the machine. Thus, Eq. 39, revised to reflect these realities, becomes e d , b r2 r2 (43) c log10 a (44) 1hmax 2 5Gi6opt.  f1 a s, ss, Re, H, Q, 52-£6, 5p 6, 5 g 6, A study of a large number of commercial centrifugal pumps by H. H. Anderson6 has quantified Eq. 43 for such machines. These pumps were all operating in water and had sufficient NPSH for performance not to be influenced by ss. The results are given by Eq. 44, which is plotted in Figure 10: h  0.94 0.08955 c Q1gpm2 N1rpm2 Xd 0.21333 0.29 2286 2 bd Ns 2.1 CENTRIFUGAL PUMP THEORY 2.25 FIGURE 10 Efficiency of centrifugal pumps versus specific speed, size, and shape–adapted from Anderson6. Note: Actual experience for Ns  2286 shows higher efficiency, as indicated by the dashed line. where X c 2 140 d ␧1m in.2 Eq. 44 is a combination of separate relationships described by Anderson for efficiency and speed as functions of flow rate6. Included is a correction for specific speed that is too conservative for Ns, (U.S.)  2286 or s greater than about unity. With this qualification, Figure 10 is a useful representation for centrifugal pumps and is often as far as many users go in determining the performance of these machines. Viscosity Effects Centrifugal pump geometries have not generally been optimized versus Reynolds number—often because the effect on hydraulic shape is not very great except for the highest viscosities of the pumpage, and a given application can sometimes experience a substantial range of viscosity. Studies of conventional centrifugal pumps over a range of Reynolds number have been combined in nomographic charts in the Hydraulic Institute Standards, which yield correction factors to the head, efficiency, and flow rate of the BEP of a low-viscosity pump in order to obtain the BEP of that pump when operating at higher viscosity7. Figure 11 is a presentation of these correction factors in terms of the head-flow Reynolds number. Strictly speaking, in view of Eq. 43, each pump geometry has a unique set of such correction factors, yet the data presented in Figure 11 have been widely utilized as reasonably representative of conventional centrifugal pumps. NPSH Effects In many cases, the available NPSH is low enough, or the suction-specific speed ss at which the pump stage must operate is high enough for significant two-phase activity to exist within the impeller. This is to be expected in centrifugal impellers of water pumps if the available ss is greater than about 3 to 4 (or Nss,(U.S.)  8,000 to 11,000). In such a case, ss and the vaporization quantities {2- } in Eq. 35 dictate a profound change in the impeller geometry into that of an inducer. The inducer has an entering or “eye” diameter that is significantly enlarged—together with tightly wrapped helical blading. Often the inducer is a separate stage that pressurizes the two-phase fluid as needed to provide a sufficiently low value of ss at the entrance of the more typical impeller blading that is immediately downstream of the inducer. If the two-phase fluid is near its thermodynamic critical point, the {2- } operate to greatly reduce the amount of two-phase 2.26 CHAPTER 2 FIGURE 11 Viscous fluid effects on centrifugal pumps—adapted from Hydraulic Institute ANSI/HI 2000 Edition Pump Standards, Reference 7. activity within the pump. (At the critical point, the liquid and gas phases are identical, and therefore both have the same specific volume.) An example is the pumping of liquid hydrogen, for which an inducer is unnecessary until much higher values of ss are reached. Moreover, inducers—typically limited to ss-values of about 10 (Nss,(U.S.)  27,000) in water—can, at sufficiently low tip speed, operate at zero NPSH, which corresponds to an infinite value of ss8. Pumping Entrained Gas In addition to the liquid’s own vapor (which is the gas involved in the NPSH-effects discussion), many pumping applications deal with a different gas; that is, a different substance from the liquid being pumped. The effects of this gas on performance arise from a) the volume flow rate of the gas at the inlet, b) the pressure ratio of the pump, which determines how far into the impeller this gas volume persists; that is, how much it gets compressed, and c) how much of the gas dissolves in the liquid as the pressure increases within the pump, which depends on both the solubility and the degree of agitation of the fluid produced by the pump. The set of fluid properties associated with these gas-handling phenomena are represented by {gp} in Eq. 34, the dimensionless form (Eqs. 35 and 43) of this set being {p}. Generally, for typical commercial centrifugal pumps, the performance under such conditions usually manifests itself as a loss of pressure rise, which is reasonably stable up to an inlet volume flow rate fraction of gas to liquid of 0.04 to 0.079. Inducers can handle larger inlet volume fractions of gas, and, under Dalton’s law for partial pressures, the liquid’s own vapor also occupies the volume of the gas bubbles. Single and multistage centrifugal pumps have been built that handle far greater gas volume than these single-stage values10,11; moreover, multiphase rotary positive displacement screw pumps can handle gas volume fractions up to 1 (100 percent gas)11. Effects of Slurries and Emulsions Finally there is the influence of the dimensionless quantities {g} in Eq. 43. Impeller and casing design are altered so as to reduce wearproducing velocities if the pumpage is a slurry of solids contained in a carrier liquid. (Slurry pumps are usually single-stage machines with a collector or volute casing sur- 2.1 CENTRIFUGAL PUMP THEORY 2.27 rounding the impeller.) This usually means a smaller impeller eye diameter (which, as can be seen in Figure 6, reduces the inlet relative velocity W1,) and a larger radial distance from the impeller to the surrounding volute because the circumferential velocity component Vu of the fluid emerging from the impeller (also seen in Figure 6) slows down with increasing radial position and is then lower in the volute passageway12. Performance also is altered, depending on the composition and concentration of the slurry. These are complicated non-Newtonian flows and are covered in detail elsewhere in this book in conjunction with a thorough treatment on solids-handling pumps. Emulsions are another example of such flows, many of which are destroyed by excessive local shear in the fluid. For this reason, screw pumps are sometimes utilized for emulsions rather than oversized, slow-running centrifugal pumps. Except for thin layers of the fluid at the clearances, most of the flow in a screw pump experiences very little shear in comparison to the flow through a centrifugal pump. Electromagnetic Effects Not appearing in Eq. 43 are quantities associated with electromagnetic phenomena. For example, electric current flowing radially outward through fluid contained in an axially directed magnetic field is capable of producing a rotating flow. Called a hydromagnetic pump, this device is therefore “centrifugal,” yet it has no moving parts. Such pumps have been used for liquid metals and could be made reasonably efficient for any pumpage with high conductivity. DESIGN PROCEDURES _______________________________________________ Establishing the Pump Configuration The first step in designing a pump is to determine the type and number of stages that are needed to meet the given set of operating conditions, usually Q, p1, p2, available NPSH ( NPSHA) and specific gravity of the fluid. If the pump must meet several such sets of operating conditions, one set has to be chosen for the BEP or design point so all the others are satisfied, if possible. Making the proper choice of this BEP may require some iteration: first making a trial choice, doing a preliminary design, and determining the corresponding pump performance characteristic curves, and then repeating these steps if necessary. Pump rotative speed N (rpm) must be chosen in order to proceed. Selection of the highest practical rpm is desirable because it yields the smallest size and therefore usually the lowest cost and easiest containment of system pressure. If, for the chosen number of stages, the stage specific speed is too low, Figure 10 indicates that efficiency is generally improved with greater speed. The maximum possible rpm is that which yields a value for the suction specific speed ss (Eq. 41) or Nss (Eq. 42) that the first stage of the pump can accommodate, where NPSH is found from Eq. 41 as follows: NPSH  1  4>3 2>3 a b Q g ss (45) Typically, the ss-capability of an impeller does not exceed a value that is somewhere in the range 3 to 4.5 (Nss  8,200 to 12,300), depending on minimum (off-design) flow rate requirements, and it is typically 10 (27,000) or less for inducers. Generally, the supplier furnishes a pump that has a value of NPSH-capability (or “NPSHR”) smaller than the stated NPSHA. This difference or “NPSH-margin” is desirable if there is any uncertainty as to the true value of NPSHA. It is essential for high values of pstg; (see discussion of “high-energy pumps” further on in this section.) Often a higher speed can be employed if a double-suction impeller (entry of fluid from both sides) can be used, as then only half of the given Q enters each side of this impeller, and only that half can be used in the ssequation applying to that fluid and impeller type. The double-suction configuration is popular for large, single-stage pumps because the axial thrust is nominally zero. It is also found as the first stage in some multistage pumps, in which case the arrangement of the remaining impellers can be “back-to-back” (half of them facing in one direction and half opposite) to achieve axial thrust balance. On the 2.28 CHAPTER 2 other hand, the interstage flow passages are simpler if these impellers are all facing in the same direction—in which case the thrust is opposed by a balancing drum or disk as described in Section 2.2.1. Forming the specific speed from N, Q, and H (from Eq. 3) and referring to Figure 9 for the kind of impeller (which then gives one an idea of the other hydraulic components (inlet passage, diffuser and collector, and so on) and Figure 10 for the expected efficiency, one can decide whether the pump would perform better and still meet any installation and size restrictions with more than the one stage implied by this first Ns-calculation. If the user requires the performance characteristics of a positive displacement pump— such as a wide range of pressure-rise at a nearly constant flow rate Q—selecting such a pump may be possible if the specific speed is not too high. (See Figure 9.) The choice of speed for positive displacement pumps is sometimes determined by mechanical considerations rather than suction capability. Sizing the Pump The next step is to determine the approximate size of a pump or pump installation. Beginning with the impeller, the agent of the energy transfer to the fluid in a pump, one utilizes the results of the pump hydraulic geometry optimization process. This also yields the proportions of the velocity diagrams that correspond to the geometry associated with the desired value of specific speed. These in turn lead to two major sizing factors for an impeller, which are plotted for typical commercial pumps in Figure 12; namely the head coefficient c (originally defined in Eq. 31) and the outlet flow coefficient fi defined on this figure4. Equations in the figure show how these factors are used to determine the exit radius r2 (or diameter D2) and the width b2 (Figure 8) respectively. The flow coefficient fi is obviously a velocity component ratio. For the head coefficient of a pump with zero inlet prewhirl to the impeller (that is, Vu, 1  0), the relevant FIGURE 12 Sizing factors for the diameter and width of typical impellers—adapted from Stepanoff 4. 2.1 CENTRIFUGAL PUMP THEORY 2.29 velocity ratio (see Figure 3) is Vu, 2/U2, which in view of Eqs. 15b and 31 is equal to the ideal head coefficient ci, as this corresponds to the ideal head Hi. Referring to Eq. 15c, it is obvious that the actual head coefficient c  hHY x ci, or, in the (common) case of zero prewhirl, ci  Vu, 2/U2. Overall pump dimensions are conveniently viewed in terms of factors times the impeller diameter. The overall diameter exceeds that of the impeller due to the surrounding casing. For single stage pumps, this casing will include a volute and perhaps a set of diffuser vanes immediately surrounding the impeller; so, the casing diameter can be 50% greater than D2 or more. On multistage pumps (many with radial-outflow diffusers and return vanes), this excess is often less than 50%. Single-stage pump axial length includes provision for inlet passageways and bearing housings and is therefore approximately the same as the overall pump diameter. On the other hand, the axial length of a stage in a multistage pump is often less than half of the impeller diameter. Minimizing this “stage length” (and diameter) is a goal of competitive pump designers in the quest to create a machine of light weight and low cost. But too small a stage length is accompanied by inferior hydraulic performance because not enough room is provided for the passages around and within the impeller to turn the fluid with minimum loss. Further, the bearing housings and the suction and discharge “heads” or end pieces must be considered in arriving at the overall length of a multistage pump. This sizing discussion so far has focused on pumps with a radial-outflow geometry. For the axial-flow geometries of higher specific speeds, the situation is simpler. The diameters of the propeller (Figure 9) and any stationary vanes downstream (or upstream) are essentially the same. The approach and discharge piping is of about the same size. A simple guideline for the maximum radius of the propeller is found from the following approximate extrapolation of the c-curve of Figure 12: ct  0.53 s (46) where the outer radius of the propeller rt,2 should be substituted for the mean exit radius r2 in the c-definition. Thus the c-curve is in reality a ct-curve for non-radial-flow pumps. For small departures from radial flow, as illustrated in Figure 8, the c-curve of Figure 12 can be used either way. These guidelines are often all that pump users or fluid system designers need to plan their installations. To get beyond this overview, one must pursue the hydraulic design in detail—as afforded by the following development and examples. Designing the Impeller Determination of the geometrical features of the impeller is generally accomplished in the following order: a) the “eye” radius re, b) the exit radius r2 or rt,2, and c) the exit width b2 or, in the case of mixed- and axial-flow impellers, the hub exit radius rh,2 —all of which form the starting point for d) shaping the hub and shroud profiles (Figure 13); and, finally, e) construction of the blades. a) The eye. The inlet radius of the impeller eye re (Figure 13) is nearly the same as rt,1, which is the diameter of the tips of the impeller blades at the inlet. This emerges after the eye flow coefficient fe  Ve/Ue [the ratio of the one-dimensional axial velocity entering the eye (Figure 8) to the tangential speed of the impeller eye Ue  re] is known: fe  Q>Ae re Ae  p1re 2  Ve Ue rs 2 2 Thus, re can be found from the following combination of Eqs. 47 and 48: (47) (48) 2.30 CHAPTER 2 FIGURE 13 Hub and shroud profiles of centrifugal pump impeller re  £ Q pfe a 1 rs b re2 2 § 1>3 (49) where the shaft-to-eye ratio rs/re can be estimated at first. Typical values for fe vary from 0.2 to 0.3 for impellers and down to 0.1 or less for inducers, depending on suction conditions, as can be seen from its relationship to suction-specific speed: 1g NPSH2 2Q ss  3>4  B pfe a 1 r2s r2e b n 1t>22 3>4 (50) where the cavitation coefficient t (cf. t2 in Eq. 36) is defined in terms of the eye speed Ue  re (Table 1). t is related to fe through empirical correlations, such as those given in Table 1. (Gongwer’s13 values for the correlation factors k1 and k2 apply to large pumps. The larger “typical” values shown for 3% breakdown apply to the more common smaller sizes. The inducer correlation is a curve fit to the data of Stripling and Acosta14 for the breakdown value of t.) Thus one can solve for fe from a given suction-specific speed and, through Eq. 49, obtain the eye size. However, the value of fe at the BEP or design point rarely exceeds 0.3, regardless of how much NPSH is available. This fe-limit therefore applies to impellers that follow and are in series with the first stage in a multistage pump. These are variously referred to as “series” or “intermediate” stages. (Slurry pump impellers are an exception to this guideline, for then the relative velocity is minimized to avoid excessive wear. In this case fe can be as high as 0.4 and NPSHA is generally more than adequate for these slowrunning machines.) b) The exit radius r2 (or diameter D2). This is found from head-coefficient c by means of the equation for r2 in Figure 12. The upper curve for c can be used unless detailed performance analysis or a desired non-typical performance characteristic curve indicates otherwise. Eq. 46 can be used for specific speeds s greater than 1.6 (Ns  4373), where the maximum radius rt,2 is computed per the previous discussion accompanying that equation. c) The exit width b2. The equation for exit passage width b2 in Figure 12 can be used for radial-outflow and mixed-flow impellers, r2 being located halfway across the passage. This 2.1 CENTRIFUGAL PUMP THEORY 2.31 TABLE 1 NPSH correlations involves the exit flow coefficient or meridional velocity ratio fi  Vm,2/U2, the lower curve of Figure 12 being for typical values of this quantity. The “openness” factor e allows for blockage due to blade thickness and to the buildup of boundary layers on the surfaces of the passageways (blades and hub and shroud). The value of e is generally between 0.8 and 0.9, the higher figure applying to larger machines. For axial-flow impellers or propellers and inducers, a choice of the hub-to-tip radius ratio at the exit defines the passage width instead. This ratio decreases with specific speed from about 23 at the right end of Figure 12 to 13 or less at the highest specific speeds. d) Hub and shroud profiles. With the eye and the outlet sizing established, the two are connected by specifying the hub and shroud profiles. Some texts illustrate the variation of hub and shroud profiles with specific speed15. Although these are excellent guidelines coming from experience, what follows is the approach one would take to synthesize these 2.32 CHAPTER 2 shapes on the basis of fundamental fluid dynamical considerations, at the same time taking experience into account. Referring to Figure 13, an acceptable geometry can be achieved by following these guidelines: i. Maintaining the meridional flow area 2prb,1b1 at the blade leading edge at about the same as it is at the eye, namely p(re2 rs2), but then gradually increasing it versus meridional distance to the generally larger value already established at the exit, namely 2pr2b2. ii. Choosing the minimum radius of curvature Rsh of the shroud to be about half the radial opening at the eye. This avoids excessive local velocity V1,sh at the blade leading edge. This has two consequences. Shaping the impeller blades to match a widely varying meridional approach velocity can complicate the construction of these blades. Also, on first-stage impellers, if V1,sh is too great, the local pressure at that location will be closer to the vapor pressure, increasing the required NPSH (or NPSH3%). This is due to a larger resulting value of the empirical factor k1, presented in Table 1 as the nth power of the velocity ratio V1,sh/Ve. Single-phase theory would require that the exponent n  2, but two-phase activity in the pump reduces the local pressure reduction that a single-phase application of Bernoulli’s equation would indicate. Figure 14 is a plot of the meridional streamlines in the space between the hub and shroud surfaces of the first-stage impeller of a high-energy boiler feed pump in the absence of blades. This was obtained from a computer solution via Katsanis’ program16 of the inviscid axisymmetric flow field, which is governed by the following equation: FIGURE 14 Axisymmetric flow analysis for the distribution of meridional velocity Vm along the blade leading edge (Eq. 51a) 2.1 CENTRIFUGAL PUMP THEORY 2.33 dVm Vm  R dn (51a) As the average value of Vm at the blade leading edge is about the same as its average Ve in the eye, V1,sh can be estimated from the finite-difference form of this equation, which expresses the change in Vm from shroud to hub in terms of an average radius of curvature R of the meridional streamlines across the passage of width n from shroud to hub in the n-direction (normal to the streamlines in Figure 14): ¢Vm Vm  ¢n R (51b) The estimated average R in Figure 14 is about twice the passage width n; so, by the estimate of Eq. 51b, Vm/Ve  12. If half this difference is between Ve and V1,sh, then V1,sh/Ve ⬇ 1.25. The computer solution of the inviscid axisymmetric-flow Eq. 51a in this bladeless passage yields Vm/Ve  0.73 and V1,sh/Ve  1.45 at this location16. Now, referring to Table 1, raising V1,sh/Ve to the power 1.4 would produce the typical value of 1.69 given for k1, which implies that the exponent n  1.4. However, two real effects operate to bias V1,sh/Ve toward lesser values; namely, a) greater loss of total pressure of flow entering the impeller along the shroud—due to wall friction and higher velocity, and b) shifting of the flow away from the shroud due to the presence of the impeller blades, which in conventional designs present the incoming flow with the greatest incidence at the hub. This serendipitous state of affairs tends to bring the value of n back toward 2. Moreover, the above estimate of V1,sh/Ve  1.25 from Eq. 51b is more typical of the flow for which designers tend to set the impeller blades at the inlet. These results are strongly influenced by the radius of curvature at the shroud Rsh, which in Figure 14 is about half of the passage width n. This accords with the guideline for Rsh. iii. Shaping the hub profile compatibly with the guidelines as stated earlier. This is best done after making an initial estimate of the shroud profile as outlined previously. The distribution of meridional flow area from the eye to the exit should then be specified. From this, a hub profile will emerge. As seen in Figures 13 and 14, the hub of a radialoutflow impeller becomes essentially radial over the outer portion of its extent. If this does not result from this procedure, appropriate adjustments can be made to the shroud profile and the process repeated. iv. For a high-specific-speed, axial-flow impeller, or inducer, the hub profile is often a cone or a reverse curve between a smaller radial location at inlet to a larger one at outlet, the latter radius decreasing with increasing specific speed as mentioned earlier. A cone or cylinder for the shroud profile is often found in such machines. e) Construction of the blades. The blades are designed by i) selecting the locus of the leading and trailing edges in the meridional plane, ii) establishing the surfaces of revolution (streamwise lines in the meridional plane) from inlet to outlet along which the construction proceeds, iii) selecting the inlet angles, iv) selecting the outlet angles, v) establishing the number of blades, and vi) obtaining the blade coordinates from inlet to outlet: i. Leading and trailing edge loci. If every point along the leading and trailing edges is revolved about the axis of rotation so as to lie in one meridional plane, the loci of these edges appear as shown in Figure 13 or 14. The outer or shroud end of the blade leading edge is positioned at or near the minimum radial location; that is, at or near the eye plane, whereas the inner or hub end is typically well back and largely around the corner along the hub profile. These locations are desirable; first, at the shroud, because the absolute velocity V (typically  Vm) approaching the blade begins to decelerate beyond the eye plane, so starting the blade ahead of this decelerating region tends to 2.34 CHAPTER 2 prevent separation of the fluid from the shroud surface due to the pumping action in the blade channels17; and secondly, being far enough along the hub in the streamwise direction to avoid impractical blade shapes (excessive twist, rake, and so on) that would make both the construction and the flow inefficient. The locus of the blade trailing edges is normally straight in the meridional plane and is axial in orientation for most centrifugal pumps. At the higher specific speeds, this locus becomes more and more slanted until it takes on the nearly radial orientation it has for a propeller (Figure 9). ii. Surfaces of revolution for blade construction. Developing the coordinates of the blades along three streamwise surfaces of revolution—the hub, mean, and shroud, whose intersections with the meridional plane appear as streamwise lines in that plane—usually provides a sufficient framework for shaping the blades of an impeller. However, for high specific-speed impellers, where the passage width in the meridional plane n (Figure 14) is large (about equal to or greater that the meridional distance from leading to trailing edge), definition along two intermediate surfaces of revolution is also needed to achieve a satisfactory design. The “mean” line is one that is representative of the flow from a one-dimensional standpoint as well as for the construction of the blades. Precisely, this is the massaveraged or “50%” streamline (that is, the streamline for c  0.5 in Figure 14)—which evenly divides the mass flow17. This line is reasonably and conveniently approximated by the “rms streamline;” that is, the line that would result in a uniform meridional velocity distribution from hub to shroud and therefore equal areas 2prn normal to the meridional velocity component Vm. In this case, n ( b) is the spacing between the rms streamline and the hub or shroud line. This would put each point on the mean line at the root mean square radial position along a true normal to the meridional streamlines; hence, the “rms” terminology. iii. Inlet blade angles. The blade angles are set to match the inlet flow field. This is done where each of the previously chosen surfaces of revolution (that intersect the meridional plane in the streamwise lines just described) crosses the chosen locus of the blade leading edges in the meridional plane. At each such crossing point, an inlet velocity diagram of the type shown in Figure 3 is plotted in a plane tangent to the surface of revolution at that point. (Figure 3, representing a purely radial-flow configuration, is a view of such a plane, as the surfaces of revolution are then simply disks.) Each such velocity diagram or triangle contains a specific value of the angle bf,1 between the relative velocity vector W1 and the local blade speed vector U1  r1. The corresponding blade angle bb,1 between the mean camber line of the blade and the circumferential direction is set equal to bf,1 or slightly higher than this to allow for the higher Vm,1 caused by non-zero blade thickness at the leading edge and to allow for higher flow rates that may be called for at off-design conditions. To construct the triangle, one first plots U1 and then Vm,1, which is taken from an analysis such as that of Figure 14 (altered as noted previously for the effect of the blades) or is chosen as the mean value Q/2prb,1b1 (Figure 13) at the rms streamline. It is adjusted from experience at the shroud and hub. Likewise, if any prewhirl Vu,1 is delivered to the impeller, it must be taken into account as illustrated in Figure 3. iv. Outlet blade angles. Whereas the inlet velocity diagrams enable the designer to correctly set the blades to receive the incoming fluid with minimum loss, the outlet velocity diagram displays the evidence—through the magnitude of the circumferential velocity component Vu,2 that the intended head will be delivered by the pump in accordance with Eq. 15c. As shown in Figure 3, Vu,2 is determined—for the given impeller tip speed U2 —by the exit relative flow angle bf,2 in conjunction with the exit meridional velocity component Vm,2. This value of Vm is somewhat larger than that given by Eq. 16 because of a) blockage due to blade thickness and boundary layer displacement thickness and b) the presence of any leakage flow QL (Figure 2 and Eq. 11) that may also be flowing through the impeller exit plane or Station 2. Well inward of the exit plane, the direction of the one-dimensional relative velocity vector W can be assumed to be parallel to the blade surface; however, in the last third 2.1 CENTRIFUGAL PUMP THEORY FIGURE 15 2.35 Impeller outlet velocity diagram of the passage, the blade-to-blade distribution of the local relative velocity changes due to the unloading of the blades at the exit. This produces a deviation of the direction of W2 from that of the blade. This deviation, called “slip” in pumps, results in less energy being delivered to the fluid by the impeller than would be the case if there were “perfect guidance” such as would occur with an infinite number of blades. Accordingly, in the outlet velocity diagram of Figure 15, the relative flow angle bf is less than the blade angle bb. This deviation is quantified by the “slip velocity” Vs. The magnitude of Vs depends on the distribution of loading along the blades from inlet to exit and therefore on the geometry of the flow passages and the number of blades. (Without slip, W2 is the same as the “geometric” relative velocity Wg,2 shown in the figure.) The slip factor m  Vs/U2 —typically between 0.1 and 0.2—was determined theoretically by Busemann for frictionless flow through impellers with logarithmic-spiral blades (constant-b from inlet to exit) and a two-dimensional, radial-flow geometry with parallel hub and shroud18. Applicability of this theory to typical impellers, despite the differences in geometry and the real fluid effects, was found to be good by Wiesner, who represented Busemann’s results by the following convenient approximation19: m VS 2sin b2  U2 n0.7 b (52) A broader, empirical slip correlation for pumps was developed by Pfleiderer, taking into account impeller geometry and blade loading, as well as the influence of the downstream collecting system (volute or diffuser)20. Pfleiderer computes the slip velocity as the product of a slip factor p and the impeller exit tangential velocity Vu,2, where p is computed as shown in Table 2. This table also contains a simple example; namely, a radial flow impeller of a volute pump, for which the resulting value of m is 0.1826— versus 0.1498 via Eq. 52; however, in this case the latter result is low by about 15 percent. A study of the Busemann plots in Wiesner’s paper yields m  0.18. Yet, if this had been a vaned-diffuser pump, Pfleiderer would have predicted m  0.1468 for the same impeller, as it would have delivered more Vu,2 for the same bb,2, Wg,2, and, therefore, Vu,2, (Figure 15). This stems from the factor “a” in Table 2 having the value 0.6 (for a vaned diffuser) instead of 0.8 (for a volute). So by this combination of circumstances—and in this example—Eq. 52 describes the slip of a diffuser pump impeller. But, despite the simplicity of Eq. 52, Pfleiderer’s method (Table 2) would appear to be a more rational, comprehensive, and satisfying method for estimating slip in real pumps. So, to find the outlet blade angle, the designer begins by deciding upon the required value of Vu,2; finds the exit flow angle and other elements of the diagram assuming the existence of slip. Next, the designer computes the slip and then obtains the value of the outlet blade angle bb,2. The process is iterative because the forementioned blockage depends on the blade angle as well as the thickness. 2.36 TABLE 2 CHAPTER 2 Pfleiderer’s slip formula v. Number of blades. The choice of the number of impeller blades is influenced by a) interaction of the flow and pressure fields of the impeller and adjacent vaned structures such as the volute tongues or diffuser vanes and b) the need to maintain smooth, attached—and therefore efficient—fluid flow within the impeller passages. The effect of the number of blades on the interaction phenomenon is addressed in the latter part of this section under the topic of high-energy pumps, where this issue becomes critical. Smooth, attached flow is assured if the product of the number of blades and their total arc length / along a given meridional streamline, as illustrated in Figure 16, is of sufficient magnitude. Divided by a representative circumference on that streamline, usually that of the impeller outer diameter (OD), this product is called the solidity s: s nb / 2pr2 (53) In practice, solidity varies from about 1.8 at low specific speed (s  0.4 or Ns  1093) to slightly less than unity at s  3 (Ns  8199). For example, Dicmas’ curve21 is 2.37 2.1 CENTRIFUGAL PUMP THEORY FIGURE 16 Intersection of impeller blades with mean surface of revolution FIGURE 17 Relative velocity distributions. useful for s  1 (Ns  2733). This limits the relative velocity reduction that occurs on the blade surfaces. Illustrated in Figure 17, this reduction or diffusion arises from the loading on the blades expressed in terms of the blade-to-blade relative velocity difference Wb: Ws Wp 1¢Wb 2  2p nb Vm, o d1UVu 2 W dm (54) 2.38 CHAPTER 2 Vm,o is the local meridional velocity component neglecting blockage. (One-dimensionally, Vm,o is the value of Vm found from Eq. 16, where the radius r is that from the axis of rotation to the center of the circle of diameter b in Figure 8, which in turn lies on an imaginary line in Figure 8 that is normal to the hub, shroud, and intermediate stream surfaces.) Here, Wb emerges by applying Bernoulli’s equation [Eq. 21 with no change in radius (that is, no change in U) or loss as one traverses from pressure side to suction side of the passage] to the static pressure difference pp ps arising from the delivery of angular momentum to the fluid (Eq. 26). This in turn results from the application of the shaft torque to the blades. It is also assumed in the derivation of Eq. 54 that the blade-to-blade average relative velocity W lies halfway between the surface velocities Ws and Wp, (which would exist just outside the boundary layers on the blades,) as illustrated in Figure 17. This is a good assumption for efficient flow well within a bladed channel22. Wb is inversely proportional to the solidity because, on the average, from inlet to outlet, Eq. 54 becomes Ws Wp  ¢Wb  a 2pr2 sin b r Vb b¢ a r2 u nb ¢m (55) where it can be seen from Figure 16 and Eq. 53 that the fraction involving the number of blades nb is the reciprocal of the solidity s because /  ¢m>sin b (56) For unconventional impeller geometries, the foregoing solidity guidelines may be inadequate to assure efficient flow. For any geometry, though, the concept of a diffusion factor D, utilized by NACA researchers23 to assess stationary cascades of airfoils can be employed. In view of Eqs. 53—56, their equation for D takes the following form for both axial- and non-axial-flow geometries, rotating or not: D1 W2  W1 ¢a r Vb r2 u (57) 2sW1 This can be deduced from Figure 17 as follows: D W1 W2 W1  ¢Wb ¢Wdiffusion a b 2W1 W1 (58) Then, Eq. 57 is obtained through the definitions of the average value of Wb (Eq. 55 with Eq. 56) and s (Eq. 53). NACA researchers found that losses increase rapidly if D  0.6. However, many centrifugal pump impellers have virtually the same value of relative velocity W at in and at outlet—along the rms streamline (Figure 17), so D from Eq. 57 is less than 0.6 on the rms streamline and even negative along the hub streamline. This situation was encountered in accelerating (turbine) cascades and led to the use of local diffusion factors, one for each side of the blade, namely Dp and Ds. Here, inspection of Figure 17 and Eq. 58 leads to Dp  1 Wp, min W1 ; Ds  1 W2 Ws, max (59a and b) where the 0.6 limit applies individually to Dp and Ds —or to the sum of the two, in which case the limit is 1.2. Eqs. 59a and 59b, therefore, constitute a more useful form 2.1 CENTRIFUGAL PUMP THEORY 2.39 of the diffusion factor concept for assessing the blade loading and the choice of the number of blades in centrifugal pump impellers24. Finally, the total blade length or number of blades, should not exceed that necessary to limit the diffusion as just described, as this adds unnecessary skin friction drag, which causes a reduction in efficiency. Thus the solidity values given in conjunction with Eq. 53 should not be appreciably exceeded, unless blade load needs to be reduced to lower levels, as with inducers to limit cavitation8 or impellers for pumps that must produce lower levels of pressure pulsations. vi. Development of the blade shape. Blades are developed by defining the intersection of the mean blade surface (really an imaginary surface) or camber line on one or more nested surfaces of revolution. Two such surfaces are formed by the hub and shroud profiles. If the blade shape is two-dimensional (that is, the same shape at all axial positions z), the mean blade surface is completely defined by constructing it on only one such surface of revolution. Generally, however, the shape is three-dimensional and is a fit to the shapes constructed on two or more of these surfaces of revolution; namely, the hub and shroud and usually at least one surface between them. After this final shape is known, half of the blade thickness is added to each side. (Sometimes the full blade thickness is added to one side only, meaning that the constructed surface just mentioned ends up—usually—as the pressure side of the finished blade rather than the mean or “camber” surface. The effective blade angles are then slightly different from those of the pressure side used in the construction process.) The construction along a mean surface of revolution is illustrated in Figure 18. The distribution of the local blade angle b (or more precisely, bb) is found first by either the “point-by-point” method or the conformal transformation method—both of which yield the polar coordinates of the blade, r, u, and z. These coordinates also depend on the chosen shapes of the intersections of the surfaces of revolution with the meridional plane; that is, the hub, shroud, and mean meridional “streamline” or rms line, as in Figure 18c, and the fact that, on the surface of revolution Figure 18a, tan b  arc bc/arc ac  dm/dy. The elemental tangential length dy ( arc ac) is the same on both the surface of revolution (Figure 18a) and in the polar view (Figure 18b). From Figure 18b, it is seen that dy  rdu, so the “wrap” angle u is found from FIGURE 18 Blade construction: a) view of construction surface of revolution; b) polar view; c) meridional view 2.40 CHAPTER 2 tan bb  dm rdu (60) and r and z are found from the fact that the coordinate m along each of the construction surfaces is a function of r and z (Figure 18c). If the blade is two-dimensional, its mean surface consists of a series of straight-line axial elements, each having a unique r and u at all z. Such a blade is typical of low-specific-speed, radial-flow impellers, and can be easily constructed by the “point-by-point” method. Here, one specifies the distribution of Wg —often linear as in Figure 17—after determining the hub and shroud profiles and the corresponding distribution of Vm15. In effect, one obtains the distribution of the blade angle bb by constructing a velocity diagram like the one in Figure 15 at every m-location from inlet (1) to outlet (2) in Figure 18c, dealing only with the “geometric” or non-deviated velocities, in order to get a smooth variation of the blade angle bb vs m. Allowance is made for blockage due to the thickness of the blades and the displacement thickness of the boundary layers in the passage. The resulting wrap angle u for each m-point—as well as the corresponding r and z—is then found from Eq. 60. (For convenience in designing the blades, the construction angle u is often taken as positive as one advances from impeller inlet to exit. For most impellers, this turns out to be opposite to the direction of rotation; and u is taken in the direction of rotation for most other purposes of pump design and analysis.) As discussed previously in Paragraph iv and illustrated in Figure 17, the actual flow will deviate from the resulting blade via the “slip” phenomenon. The point-by-point method allows the designer to exercise control over the relative velocity distributions on the blade surfaces (Eq. 54 and Figure 17) via specification of the distribution of Wg or other velocity component in Figure 15; for example, Vu,#. This becomes more important if an unconventional impeller geometry is being developed17. The point-by-point method can also be used for three-dimensional blades. A simple approach in this respect would be to use this method to determine the blade shape along the rms- or 50%-streamline (that is, on the mean surface of revolution depicted in Figure 18). The shapes on the other streamlines, generally the hub and the shroud, can also be found by this method. The resulting overall blade shape, however, is subject to the condition that the resulting wrap u2 u1 cannot greatly differ on all streamlines without the blade taking on a shape that is difficult to manufacture and which may turn out to be structurally unsound or create additional flow losses. This is because the final blade shape is the result of stacking the shapes that have been established on the nested stream surfaces defined by these meridional streamlines. Blade forces due to twists arising from this stacking could modify the expected flow and cause unexpected diffusion losses. One way to generate blade shapes along the hub and shroud that have the same (or nearly the same) wrap as that obtained from point-by-point construction of the blade on the mean surface of revolution is to establish the desired inlet and outlet blade angles bb on each such surface and then mathematically fit a smooth shape y(m) to these end and wrap conditions, where y is the tangential coordinate seen in Figure 18 and defined in Figure 19. A conformal representation of the shapes of the blades resulting from such a procedure on each of the three surfaces is seen in Figure 19. These shapes are sometimes called “grid-lines” or simply “grids”—from the description of the graphical procedure that relates these shapes in the conformal representation to those on the actual, physical surfaces4. In such a representation, the blade angles are the same as they are on the physical surface of revolution because tan b  dm/dy and dy  rdu, also yielding Eq. 60. If the associated distributions of Wg and Vm are smooth, one can expect to have a satisfactory result if these conformal representations are also smooth. Thus, many skilled designers bypass the computations just described for the point-by-point method and use the conformal transformation method of blade design. Here, one simply establishes the grid-line shapes by eye in the conformal plane of Figure 19, specifying the blade angles b at inlet and outlet by the previous procedures as the starting point for 2.1 CENTRIFUGAL PUMP THEORY FIGURE 19 2.41 Conformal transformation of blade shape: “grid-lines” drawing each grid-line. This conformal blade shape is then transformed onto the physical surface, the differential tangential distance dy becoming rdu on the physical surface (Figure 18) and the differential meridional distance dm being identical in both the conformal and physical representations. If the resulting blade shape appears to be unsatisfactory, the designer repeats this process, possibly first altering the hub and shroud profiles or the blade leading and trailing edge locations on these profiles and recomputing the b’s. Designing the Collector The fluid emerging from the impeller is conducted to the pump discharge port or entry to another stage by the collecting configuration, which can employ one or more of the following elements in combination: a) volutes, which can be used for designs of all specific speeds, b) diffuser or stator vanes, which are often more economical of space in high-specific-speed single-stage pumps and in multistage pumps, and for the latter, c) return or crossover passages, which bring the fluid from the volute or diffuser to the eye of the next-stage impeller. Generally, the most efficient impeller has a steady internal relative flow field as it rotates in proximity to these configurations. This is assured by all of these elements because they are designed to maintain uniform static pressure around the impeller periphery—at least at the design point or BEP. An exception to this rule is the concentric, “doughnut”-type, “circular-volute” collector, which is used on small pumps or in special instances where the uniform pressure condition is desired at zero flow rate. The proximity of stationary vanes in these collecting configurations to the impeller must be considered in their design. Called “Gap B,” the meridional clearance between the exit of the impeller blades and adjacent vanes ranges from 4 to 15 percent of the impeller radius, volutes having higher values in this range than diffusers, and pumps of higher energy level requiring the larger values. If these gaps are too small, the interactions of the pressure fields of the adjacent blade and vane rows passing each other can cause vibration and structural failure of impeller blades, diffuser vanes, and volute tongues. a) Volutes. A volute is built by distributing its cross-sectional area on a “base circle” that touches the tongue or “cutwater” and is meridionally removed from the impeller exit by 2.42 CHAPTER 2 FIGURE 20 Volute casing: a) polar view; b) meridional view including Section A-A of throat T Gap B. (For radial-discharge impellers, as in Figure 20, this is a radial gap, and the base circle has radius r3.) Beginning at the tongue, the cross-sectional area Av of the volute passage is zero, but it increases with angle u in the direction of rotation, ending up at area AT in the “throat” T, as depicted in Figure 20. Worster demonstrated that the desired peripheral uniformity of static pressure can be achieved if the product rVu is constant 2.1 CENTRIFUGAL PUMP THEORY 2.43 everywhere in the volute25. One-dimensionally, this means that rTVu,T  r2Vu,2; and, if the velocity VT is essentially tangential (in the u-direction), rTVT  r2Vu,2. The diffusion or reduction of the velocity V from the impeller periphery at r2 to the larger rT of the throat produces a static pressure increase above that at the impeller exit; however, friction losses in the volute would cause a reduction in static pressure around the impeller at r2 from tongue to throat unless the throat area AT  Q/VT is slightly enlarged, creating a little more diffusion to compensate for this loss. Thus, in practice, at the BEP, rTVT  10.9 to 0.952 r2Vu, 2 (61) At off-BEP conditions, the volute will be either too large or too small and Eq. 61 will not be satisfied. When the flow coefficient (or Q/N) drops below the BEP value, there will be excessive diffusion and an increase of static pressure around the volute from the zero area point around to the maximum area point at the throat. Proceeding around further, past the throat, a sudden drop in pressure occurs across the tongue to bring the pressure back to what it was at the starting point25. The opposite situation occurs above BEP. Each of these off-BEP circumferential static pressure distributions is properly viewed as the consequence of a mismatch between the head-versus-flow characteristics of the impeller and volute26. For the impeller, there is the falling, straight, Hi-versus-Q line or “impeller line” of Figure 6, whereas the volute characteristic or “casing line” would be a straight line starting at the origin of Figure 6 and crossing the impeller line at the match point, which is generally at or close to the BEP flow rate. This casing line is straight because the throat velocity VT varies directly with flow rate Q and, through Eq. 61, directly with the ideal head Hi —because  r2Vu,2  Hi (Eq. 15b for Vu,1  0). In other words, the same volute could be optimum at a different value of Q if it were paired with another impeller whose Hi-versus-Q line crossed this same casing line at that different Q. To essentially eliminate the consequent radial thrust on the impellers of large pumps at off-BEP conditions, a double volute is used; that is, there are two throats, 180 degrees apart, there being either two discharge ports or a connecting “back channel” to carry the fluid from one of the throats around to join the flow emerging from the other—to form a single discharge port. The value of the volute cross-sectional area Av at a given polar position uv can be found for the portion of the total pump flow rate Q being carried in the volute at that uv-position together with the condition rVu  constant versus radius. This will produce a distribution Av(u) that is slightly below a straight-line variation versus uv from zero to AT. Often, the practice is to use the latter straight-line design because this produces larger values of Av where the hydraulic radius of the volute is small, thus compensating for the greater friction loss in that region through lower velocity—particularly for the smaller pump sizes. The cross-sectional shape of the volute is dictated by the need to make a minimum-loss transition from a small area at the beginning of the volute where the height (as can be deduced from Figure 20b) is much smaller than the width b3 to the throat, for which the height (to the outer casing wall at rw) and the width bmax are more nearly equal. Too small an aspect ratio (height/width) decreases the hydraulic diameter too much and increases the loss. There is another transition from the throat through an essentially conical diffuser (which may negotiate a turn) to a larger, circular exit port. This diffuser can be designed with the help of charts of flow elements and will normally have a 7-deg. angle of divergence and a discharge area up to twice that of the throat AT27. Thus, there is a substantial diffusion from the impeller periphery to the pump or stage exit port. This generally produces a static pressure rise in the collection system that is 20 to 25 percent of that of the whole stage. b) Vaned diffusers. A vaned diffuser is rotationally symmetric and, if properly applied, produces minimal radial thrust over the whole flow rate range of a pump. Although diffusion can be accomplished in a radial outflow configuration without vanes due to the essential constancy of the angular momentum per unit mass rVu, one rarely finds a pump with a vaneless diffuser, partly because so much radial distance is needed to effect the reduction of tangential velocity required, as well as the still larger volute needed on single-stage 2.44 CHAPTER 2 FIGURE 21 Vaned diffuser pumps to collect the fluid at the exit of such a diffuser. Also, the absolute flow angle a2 (Figure 15) of the fluid leaving the impeller is usually too small to satisfy the conditions for stall-free flow in a vaneless diffuser28. A vaned diffuser, on the other hand, can accomplish the reduction of velocity in a shorter radial distance. Also it can diffuse axially and, to a degree, even with radially inward flow. Vaned diffusers are similar to multiple volutes in concept, except they are subject to offdesign flow instabilities if not shaped correctly. Width b3 is usually slightly greater than b2 in order to accommodate discrepancies in the axial positions of impellers that feed them. With reference to Figure 21, “Gap B” ( r3 r2) is in effect a short vaneless diffuser, and by the time the fluid has reached the throat (the dashed line at Station 4), it has gained a substantial portion of the static pressure recovery that takes place via diffusion from Station 2 to Station “ex.” This “pre-diffusion” is enhanced by the fact that the throat area at Station 4 ( b3w1 per passage for parallel-walled radial-flow diffusers) is larger than it is for volutes, the following relation applying to diffusers29: rTVT  0.8r2Vu, 2 (62) Therefore, more diffusion than would result from applying Eq. 61 occurs in a vaned diffuser, the skin friction loss due to an otherwise higher velocity at the throat being offset by an efficient reduction of the velocity up to that point and a lower velocity onward. The fully vaned portion from throat to exit (Figure 21), which performs most of the rest of the diffusion and associated static pressure recovery of the stage, is designed to perform efficiently and maintain stable flow. For typical radial-flow geometries with parallel walls, the vanes can be of constant thickness and comparatively thin or can thicken up to form “islands.” The latter approach usually produces a channel that is two-dimensional with straight sides diverging at an included angle, length-to-entrance width //w1, and area ratio Aex/AT in a combination that avoids appreciable stall30. A typical combination is an included angle of 1121 deg., //w1  4, and Aex/AT  1.8, which also applies for vanes of constant thickness, as illustrated in Figure 21. Constant thickness vanes have curvature. This modifies the performance somewhat31,32, but it allows a smaller overall radius ratio of the diffuser, rex/r3. Also, this ratio rex/r3 will be smaller as the number of vanes nv increases. The best experience seems to be with diffusers that have only a few more vanes than the number of impeller blades nb (nb rarely exceeds 7 in traditional commercial pumps). For pumps of higher energy levels (or high head per stage, as discussed further on in connection with high-energy pumps), it is important that nv be chosen so as to avoid a difference of 0 or 1 between nb and nv or their multiples—up to at least the third multiple or “order” of each. A difference of 2 should also be avoided for at least the lower orders33. At off-design flow-coefficients (or off-design flow rate at a constant speed), the angle a of the absolute velocity vector V (Figure 15) approaching the diffuser will vary; yet, for typ- 2.1 CENTRIFUGAL PUMP THEORY 2.45 ical stages, a wide range of flow coefficient is possible without damaging instabilities, even at high energy levels. This is likely the case because a is rather small at the design point or BEP (except for designs having high specific speed), so variations of the angle that occur with flow changes are within the unstalled performance range of the diffuser vane system. c) Return passages. Conducting the relatively low-velocity fluid from the diffuser to the eye of the next impeller in a multistage pump is accomplished with return vanes or passages that also deswirl the fluid wholly or partially. Except for development of stall in the diffuser, these passages will not see a changing angle of the approaching velocity vector because the diffuser feeding them is a stationary element. In radial-flow pumps, there is a sharp turn in the meridional plane in order to redirect the fluid inward. The fluid, still possessing a circumferential component of velocity that is greater than the meridional component, actually sees a much gentler turn. However, downstream of this point, a sharp turn of the blades is invariably a feature of a return passage; and this, together with the need to ensure undistorted flow into the following impeller, often dictates that the vane system accelerate the fluid as it approaches the eye. Although losses in the return passages—being related to the low velocity within them—have a minor effect on the overall stage efficiency, the design of such passages must ensure unstalled flow into the impeller in order to avoid the negative impact of a distorted inlet flow on the efficiency and to promote pulsation-free operation of the impeller. A variety of return-passage geometries exist, some of which are presented in the literature 29,34. The continuous-vane type is integral with the upstream diffuser, thereby eliminating the entry losses into yet another vane system after the diffuser4,34. Improvements in manufacturing technology have made this potentially more efficient approach more viable for radial machinery. The continuous-vane concept is standard practice in the design of mixed-flow “bowl”-type pumps21. The diffusing stator vane row that receives the fluid from the impeller of an axial-flow pump—being an axial-flow element itself—possesses the return feature already. Diffusion in axial-flow stators is typically accomplished by a reduction in velocity of about 30 to 40 percent. The actual value is governed by an acceptable level of the diffusion factor, Eq. 57. (A similar reduction in relative velocity is needed for an axial-flow impeller to generate static pressure, as can be seen from Eq. 21. By comparison, centrifugal impellers, on the rms streamline, usually have W2 about equal to W1 —as seen in Figure 17.) Axial-Flow Pumps The preceding development, though general, is applicable mainly to centrifugal and mixed-flow pumps. In that procedure, the impellers have appreciable solidity, and original blade shapes are constructed from the viewpoint of one- or two-dimensional channel flow. The collectors are often volutes or non-axial-flow vane systems. Performance is not known a priori and so must be estimated, as outlined further on. On the other hand, the extensive two-dimensional, experimental, axial-flow cascade data amassed by NACA researchers23 and others enables the designer to adopt existing airfoil blade shapes and so predict the performance with greater confidence. The procedure for utilizing these shapes and the corresponding experimental results has long been the basis for designing axialflow compressors for gas-turbine engines and is clearly described by Hill and Peterson35. This approach is widely used, especially for high-specific-speed, low-solidity axial-flow propeller pumps—in designing both rotating and stationary blade rows. Insights for propeller pump design and performance characteristics can be found in Stepanoff4. An exception to this axial-flow pump design approach is the case of inducers. Although they are axial flow pumps, they have high solidity and are usually designed as channelflow machines. The design philosophy outlined in the preceding paragraphs is applicable, except that the blades usually approximate constant- or variable-pitch helices. Performance prediction is generally accomplished via one-dimensional calculations and the correlations described in the following paragraphs. PREDICTING THE PERFORMANCE CURVES ______________________________ The choices made in the foregoing design procedures can and should be verified analytically, the objectives being first to generate the performance characteristic curves for head 2.46 CHAPTER 2 and power at constant speed and second to ensure stable behavior of the various systems in which the pump is to be applied. For the first objective, the solution involves analytical or empirical approaches: a) at non-recirculating flow conditions; that is, from flow rates Q somewhat below QBEP out to the maximum “runout” flow rate, b) at shut-off (Q  0) and low flow, or c) the complete set of curves for a given pump predicted by means of computational fluid dynamics (CFD). Generating Performance Curves The fluid dynamical limitation on the deceleration of the relative velocity W determines the shape of the head-versus-flow curves. This is inherent in the choice made for the head coefficient c in Figure 12, which sizes the impeller and is illustrated in Figure 22. The typical situation of zero (or nearly so) inlet whirl Vu,1  0 means that the ideal head coefficient ci equals the most significant ratio of the outlet velocity diagram because from Eqs. 15 and 31 (with for Vu,1  0): cideal  U22 1Vu, 2>U2 2 (63) U22 c  hHY cideal  hHY Vu, 2 U2 (64) Figure 22 illustrates how specific speed s affects the BEP value of ci and therefore c. Overall, only a small reduction of W occurs in most impellers. So, at low s, the low value of W1 associated with the small eye relative to the maximum diameter (Figure 9) enables the outlet velocity diagram (Figure 22a) to have a high value of Vu,2/U2. On the other hand, this ratio drops as s increases and the eye grows to be as large as the maximum diameter of the wheel. Figure 22b is the result because the value of c at shut-off (about 12) is not based on the one-dimensional concept of velocity diagrams but primarily on the pressure generated by solid body rotation of stalled (though recirculating) fluid contained within the impeller. The BEP values of c in Figure 22b are consistent with Figure 12 and illustrate why a high-specific-speed impeller has such a substantial “rise to shut-off” of the head curve. This is dramatically illustrated in Figures 8–10 of Section 2.3.1 in which the head curves are normalized to that of the BEP36. a) Non-recirculating flows. The BEP efficiency and head can be determined from correlations for typical pumps or from computation of the losses. Fluid dynamic procedures described in this section can be used to determine the shapes of the head and power curves at all flow rates to runout, using the BEP as an anchor point for such computations. For pumps designed conventionally, beginning with Figure 12, Anderson’s overall (BEP) efficiency correlation (Eq. 44) as modified in Figure 10 is useful. Other similar charts, especially Figure 6 in Section 2.3.1, are in widespread use. The breakdown of the losses involved, as expressed by Eqs. 8–11, is quantified through the development of the three component efficiencies hHY, hm, and hv in Table 3. All three decrease with decreasing specific speed—as might be expected from the charts just mentioned. This can be seen in the hHY-expression (a) of Table 3 because ci is greater at low s as discussed relative to Figure 22. Jekat’s hHY-expression (b) of the table works surprisingly well, largely because of the flow effect in Figure 9 (explained there as the “size effect” of larger relative roughness and clearances in smaller pumps) and because low s tends to go hand-in-hand with low flow rate Q. To compute hHY at Q  QBEP (and, if required, at Q  QBEP as well), it is necessary to go deeper into the prediction of hHY by developing expressions for the losses noted in Eq. 21, which are basically expansions of the expression for the collector loss coefficient zc26 and for the impeller loss expression (c) of Table 337. In this expression, the incidence loss coefficient k can be obtained from cascade data or developed as a combination of a turning and a sudden expansion loss4,8,27. The “pipe-type friction factor” f can be increased to include secondary flow and diffusion losses due to blade loading (or turning38 of the absolute velocity vector V). The resulting f-value can thus be twice the usual pipe value associated with the skin friction losses in the passage. (The pipe value of f is found from the well-known 2.1 CENTRIFUGAL PUMP THEORY FIGURE 22 2.47 Performance versus specific speed: a) velocity diagrams at BEP; b) head-versus flow curves. pipe friction chart—Figure 31 in Section 8.1—by substituting a representative average passage hydraulic diameter Dh  4Ap/ for the pipe diameter d.) A further increase in this f-value occurs if the impeller is missing one or both rotating shrouds; that is, it is a semi- or fully-open impeller with blade tip leakage losses appearing in the main flow stream39. Multiphase flows in pumps often are accompanied by greater than normal hydraulic losses; for example, increasing the concentration of solids in the carrier liquid flowing through a slurry pump increases the f-value still further40 (see Section 9.16.2). 2.48 CHAPTER 2 Quasi three-dimensional (Q3D) analysis41 affords an assessment of the secondary flow and diffusion losses and gives results similar to inviscid three-dimensional (3D) flow analysis. Q3D analysis starts by solving the 2D meridional (hub-to-shroud) flow field—as in Figure 14, but with blades present. This is followed by a series of 2D blade-to-blade TABLE 3 Component efficiency expressions developed 2.1 CENTRIFUGAL PUMP THEORY 2.49 TABLE 3 Continued. inviscid solutions42, each on a surface of revolution generated by one of the meridional streamlines of the hub-to-shroud 2D solution and producing results like that of Figure 17. From this, one computes the diffusion factors (Eqs. 57—59) and decides whether the diffusion losses are significant—in which case a redesign is in order, followed by a further Q3D evaluation. This type of iterative design approach for impeller blading has led some designers to combine Q3D analysis with an “inverse” design approach and a performance prediction scheme as discussed in this subsection. Here, in distinction to the more common “direct” choice of the conformal blade shape (Figure 19) between inlet and outlet as 2.50 CHAPTER 2 described in paragraph (e) (vi) under “Designing the Impeller,” one specifies the distribution of fluid dynamical quantities from inlet to outlet—such as UVu or W—and finally produces the corresponding blading17,43. In this sense, specifying Wg as described in the same paragraph (e) (vi) is an inverse design procedure. Mechanical efficiency hm, as stated earlier, is largely the result of impeller disk friction. If the drag of bearings and seals is added, as in Eq. (d) of Table 3, the moment coefficient Cm in the disk friction formula (e) can be increased over known disk friction values44,45 to include these effects. (On the other hand, the drag power loss of shaft seals, though usually quite small, is generally directly proportional to speed. Such losses can therefore be significant in small pumps running at lower-than-normal speeds.) The Cm-expression given in Formula (f) reflects this adjustment and includes the drag on both sides of a smooth impeller for a typical clearance ratio s/a  0.05, where a is the disc radius. This works well for most impellers: The drag at the ring fits roughly compensates for the fact that the impeller eye has been cut out of the disk, and so on. (There is very little influence on Cm of the gap width s between impeller shroud and casing wall, Cm being proportional to (s/a)0.1 in general44. For very small s/a, Cm instead grows as s/a decreases; see Refs. 44 and 45 for formulas.) The value of Cm can be even larger for semi- or fully-open impellers, if the neighboring fluid is rotating faster relative to the wall—as is the case with radial-bladed open impellers. The fluid between a shrouded impeller and adjacent wall, on the other hand, rotates at half speed44. (In cases where the impeller surface and adjacent wall are both rough, Cm is larger than just discussed45.) Finally, notice in Eq. (h) that very low specific speed s produces a dramatically low value of hm. This drives c to the larger values of Figure 12 at low s —also dictated by the W-deceleration considerations per Figure 22. Overall there is a benefit, despite possibly lower hHY [Eq. (a)] due to the consequently greater ci and collector loss. Volumetric efficiency hv applies to leakage across impeller shroud rings or “neck rings” and balancing drums. Eq. (j) in Table 3 is an approximation for the leakage across a typical ring of a closed-impeller pump, assuming orifice-type flow at a discharge coefficient of 1 4 2 , as reported by Stepanoff . Referring to Figure 2, leakage QL occurs at r  rR, (rR being approximately 1.2 times re) under a pressure difference across the ring of about 32 that of the pump stage. If the shroud is removed and the open blades are fitted closely to the adjacent wall, as with open impellers, the consequent leakage from one impeller passage to the next across the blade tips does not affect hv, and Eq. (j) should be modified accordingly. Rather, the tip leakage causes a hydraulic efficiency loss as previously discussed. Finally, as with hm, Eq. (j) indicates that low-s pumps have low hv. At flow rates Q other than QBEP, the analytical methods described previously for computing the hydraulic efficiency are utilized, together with computation of the inlet and outlet velocity diagrams, which yield the ideal head and power curves as illustrated in Figure 6. In this procedure, the slip velocity Vs (Figure 15) applies to the BEP and, at other flow rates, the exit relative flow angle bf,2 can be assumed constant. This accords with the fact that Vs for the narrower active jet at low flow rates must be smaller. A blockage model for the thickening wakes and narrower active jets that develop as Q is decreased can be introduced to compute the one-dimensional velocity diagrams, but ignoring this at nonrecirculating flow rates appears not to be serious in determining the shapes of the head and power curves. b) Shut-off and low flow. The foregoing analyses apply over that portion of the flow rate range that does not involve recirculation, as illustrated in Figure 6. The complexity of recirculation has not been readily handled analytically, and this has forced pump designers to estimate the low-flow end of the H-Q curve with the help of empirical correlations. Nevertheless, insightful fluid dynamical reasoning about the physics of the flow have led to useful expressions for the head developed and the power consumed at shut-off. Shutoff, then, in addition to the BEP, becomes the other anchor point of the head and power curves; and this—together with the shapes established for these curves at the higher flow rates—gives the analyst an idea of the intervening shapes. Shut-off head Hs/o can be viewed as the sum of two effects occurring at Q = 0, each being represented by a term in this equation: 2.51 2.1 CENTRIFUGAL PUMP THEORY Hs>o  1U2t,2 kimp U2h,1 2 2g  U22 kex 2g (65) or kimp cs>o  a r2t,2 r2h,1 r22 2 r22 b kex 2 (66) where the first term is the centrifugal effect of essentially solid body rotation of the fluid confined within the impeller; and the second term is the pitot effect of the recirculating fluid from the impeller that impinges against the volute or diffuser throats which in turn are connected through stagnant fluid to the exit port of the pump. While the factors kimp and kex associated with these effects vary with the hydraulic configuration, the values involved can be estimated as follows: kimp ≈ 1, as the radial equation of motion3 would indicate for fluid rotating at r everywhere within the blades, i.e., for rh,1 6 r 6 rt,2 (Fig. 8). Thus, as indicated by Eqs. 65 and 66, increasing the minimum radius of the blades at inlet rh,1 tends to reduce the shut-off head. However, the presence or absence of fluid swirl in the region upstream of the impeller blades at shut-off has been found experimentally to affect the value of kimp in surprising ways—sometimes increasing it above unity in such a way as to minimize the effect on shut-off head of any non-zero value of rh,1. The value of kex depends on (r3 r2)/r2, or “Gap B” and other features of the impeller exit and collector geometry. It is usually in the range 0.2  0.1, any change in the geometry that increases the shut-off power coefficient (see below) raising kex by driving more recirculating flow from the impeller against the volute or diffuser throats. Thus the shut-off head coefficient cs/o (Eq. 66) for typical radial-flow pumps generally exceeds 21, the value of 0.585 being advanced by Stepanoff 4. Estimates for cs/o are also indicated in Fig. 22b. Shut-off power consumption Ps/o includes disk, bearing, and seal drag power PD and that which drives the recirculation Precirc. The latter is generally dominant by far. From similarity arguments (Eq. 33), the shut-off power coefficient P̂s>o  Ps>o r3r52 (67) is a constant for a given pump geometry. Mockridge, in a discussion attached to an ASME paper by Stepanoff, reasoned that a wider impeller (larger b2 at the same diameter D2) would recirculate more fluid at shut-off and therefore have a higher value of this coefficient. His correlation is shown in Figure 23 and is probably the most significant quantitative result available for predicting the performance of centrifugal pumps at shut-off conditions36. c) Complete prediction via CFD. The uncertainties that have characterized the prediction of pump performance are now being overcome through advances in computational fluid dynamics. CFD entails three-dimensional solution of the flow fields within pumps via the Reynolds-averaged Navier-Stokes equations. Graf demonstrated the ability of a CFD computer code to calculate recirculation, the consequent prediction of the head curve for the impeller comparing favorably with experimental data46. The resulting distorted flows entering and leaving adjacent systems of impeller blades and stator vanes produce time-varying boundary conditions on each, the associated computational grids also moving relative to each other. This involves extensive, time-dependent computation. To provide solutions quickly on conventional, storage- and speed-limited workstations, some steady-flow codes treat these interfaces by circumferentially averaging the conditions at each point of the blade and vane leading and trailing edges as they appear in the meridional plane. Even with this simplification, pump analysts can now predict the entire performance curve of head within about two percent and the power curve with slightly less accuracy47. 2.52 CHAPTER 2 FIGURE 23 Shut-off power coefficient The design task therefore resolves itself into an iteration between an efficient geometrygenerating scheme and a rapid CFD flow and performance analysis of the geometry resulting from each iteration48. This is especially useful if a non-traditional geometry is involved, or if an efficient design is sought that will produce a desired performance curve shape. Nevertheless, many turbomachinery designers can make more rapid and valid judgments about their respective classes of machines through the time-honored iteration between a proprietary direct or inverse design and performance-prediction scheme and inviscid quasi-3D analysis 41,43. They have developed reliable diffusion criteria (computed, for example, from Eqs. 59a and 59b) for interpreting the acceptability of the free-stream relative velocity distributions Ws and Wp on the blade surfaces (Figure 17) produced by the Q3D blade-to-blade solutions43. Because CFD codes solve the actual viscous flow field, the boundary condition on the blade surface is zero relative velocity. This can be at least partly overcome by displaying the CFD-distributions of pressures on the blade surfaces, the interpretation of which would require knowledge of the corresponding criteria for these pressures46. Also, the velocities at the edge of the boundary layer could be extracted from the CFD solution and displayed in familiar terms. A useful design approach for the present may therefore be to a) produce the final design by the more traditional methods and b) predict the performance curves via CFD49. Predicting Axial Thrust The prediction of pump performance is not truly complete without the corresponding prediction of the hydrodynamic axial and radial thrust that the impeller(s) can be expected to encounter. A comprehensive treatment of radial thrust appears in Section 2.3.1, and a review of axial thrust and thrust balancing devices is cov- 2.1 CENTRIFUGAL PUMP THEORY 2.53 ered in Section 2.2.1. However, obscure flow phenomena can profoundly affect the radial distributions of pressure on the outside surfaces of a shrouded impeller that give rise to the net axial thrust. These phenomena become even more complex when discharge recirculation occurs and can cause adverse mechanical response in high-energy pumps, as will be explained further on. As a basis for tackling such problems, the fundamentals of axial thrust are presented in Table 4 for shrouded centrifugal impellers that have leaking fluid flowing in the gaps between the impeller shrouds and the adjacent casing walls. The positive direction of the thrust T is taken toward the suction or eye of the single-suction impeller shown. The incoming axial momentum rQVz,1 is generally quite small for radial impellers and has been omitted from the Table. It serves, however, to reduce T. The centrifugal effect of the fluid spinning in the sidewall gaps causes a reduction in static pressure from the outer periphery (OD) of the impeller to the sealing ring, and this TABLE 4 Leakage effects on axial thrusts 2.54 CHAPTER 2 TABLE 4 Continued. is quantified in the expressions given for the swirl velocity component Vu. These expressions are curve fits to experimental data for the leakage flowing radially inward on either or both sides of the impeller50 and for outflowing leakage as occurs on the back side (away from the “front” or suction side) of multistage pump impellers due to the higher pressure arising from the diffusing system downstream of each impeller51. In the absence of leakage, the fluid in the sidewall gap rotates at about half the local impeller wheel speed; that is, Vu  r/2, and this half speed is typical of the gap flow near the impeller OD, even in the presence of leakage. The greater the inflow leakage, the lower the pressure becomes at the entrance to the ring clearance. The major effect is that of swirl or the tangential component of velocity Vu, which varies inversely with radius unless casing wall drag interferes. More leakage flow is less influenced by this drag and so experiences a greater centrifugal effect. This in turn means more pressure drop from OD to ring. (The leakage rate, of course, is affected, the solutions for both leakage and pressure distribution being linked and usually requiring iteration.) The opposite effect happens for outflow on the back side of multistage pump impellers. Here the fluid enters the sidewall gap at a small radius (see Section 2.2.1) and so with negligible swirl. It flows outward without picking up much swirl, especially if there is substantial radial outflow leakage, which means the centrifugal effect is small, yielding a nearly constant pressure versus radius. The overall result is more net thrust than might be expected from a cursory look at the pressure-loaded surfaces. If wear ring clearances increase during the life of the pump, the net thrust of multistage pump impellers increases. Likewise, unequal ring wear leads to uncertain changes in the thrust of a “balanced” single- or double-suction impeller with inflow to wear rings on both sides. Similarly, these theories can be applied to balancing drums and other such devices described in Section 2.2.1. Integration of the pressure equation in Table 4 becomes a chore unless the whole theory is computerized. A quick estimate of the thrust is possible, however, if the distributions of pressure in the separable domains of the surfaces are assumed to be linear; in that case, 2.1 CENTRIFUGAL PUMP THEORY 2.55 TABLE 5 Approximate axial thrust calculation the integration is simple and yields the closed-form results of Table 5. This table also indicates how to account for each element of the thrust, including the axial momentum terms, which become significant for higher-specific-speed mixed-flow impellers. In all cases, in order to proceed with the calculation, the static pressure at the impeller OD must be known, as it is from the boundary condition imposed at the OD that the rest of the pressure distribution emerges. Even for substantial leakage, the pressure drop of the fluid entering the sidewall gaps from the impeller exit is small or negligible; therefore, the impeller pressure essentially applies in the gap (at the OD) as well. The foregoing methods of predicting pump head also yield the impeller OD pressure, which is usually between 75 and 80% of the stage pressure rise above inlet. Thrust computations can therefore be coupled with the head-curve prediction scheme being employed for the pump, thereby yielding predicted thrust curves together with the predictions of hydraulic performance. 2.56 CHAPTER 2 Ensuring Stable Performance The ability for a pump to run smoothly with minimal pressure-rise and flow-rate excursions is dependent on the shape of the pump head-flow performance curve and the characteristics of the system in which it operates. There are two types of pump-system instability; namely, a) static instability, which can be ascertained by studying the pump and system head curves, and b) dynamic instability, which requires more detailed knowledge of the system52. (In addition to these system-related instabilities, there is the unsteady behavior of the separated and recirculating flows that occur when a pump operates a flow rate substantially below the BEP. Called hydraulic instability, this becomes important in higher-energy applications and is therefore discussed later.) a) Static stability and instability. Figure 24 illustrates two pumping systems; namely, a) a piping system in which the flow is turbulent and largely independent of Reynolds number; so, the head drop H through it is proportional to Q2, and b) two reservoirs with a constant difference H between the two liquid surfaces and comparatively negligible head loss in the pipes connecting them. In each case, the pump is designed to produce head H, as required to deliver the desired flow rate Q. The influence of the pump head curve shape is immediately appreciated in Case (b): the curve “droops” as Q is reduced to shut-off, thereby producing two vastly different flow rates at the same head. In fact, however, the pump will not operate at the lower-Q intersection point of the two curves. The pump shutoff head is less than H, so it will produce no positive flow rate. Instead, as discussed in Section 2.3.1, fluid will flow backwards through the pump. Further, if circumstances could allow operation at this lower-Q point, even a vanishingly small increase of Q would cause a further, divergent increase because the head of the pump exceeds the H of the system. Likewise, a small decrease leads to even lesser Q because the system H exceeds that of the pump. This is called “static instability.” Conversely, the higher-Q point of Figure 24b is “statically stable,” small departures in Q being suppressed by algebraically opposite signs of the difference between the system and pump heads. Both intersection points of Figure 24a are seen by this type of analysis to be statically stable. If the operator increases the frictional resistance by closing up a valve in the piping system, the operating point simply moves to the left on the curve and remains stable. Thus, it is concluded that if the slope of the pump H-versus-Q curve is less than that of the system, operation will be statically stable—and vice versa. Most pumping systems are combinations of the “pure friction” type of Figure 24a and the “purely elevation” type of Figure 24b. In this case of static stability, the drooping pump head curve presents no problem. Theoretically, it is possible to have a pump head curve with a kink that could have a more positive slope than that of a system curve, which might intersect it at such a kink. The high-s curve of Figure 22b depicts a kink or dip, which is due to stalled flow within a mixed- or axial-flow impeller that is not sufficiently confined by the impeller to be maintained in solid body rotation. (It is the centrifugal effect of such rotation that maintains the pressure rise in radial-flow impellers despite the stalling.) That particular pump, if applied to a system that never intersects the pump head curve at the kink, would never experience static instability. On the other hand, the designer may want to take on the challenge of designing a machine without such a kink. The previously mentioned design procedure utilizing CFD in both the impeller and the diffuser to check whether the kink is gone would be a way of tackling this problem53. In-depth discussion of the variety of systems that can be encountered, including multiple systems and parallel operation of multiple pumps, can be found in Sections 2.3.1, 8.1, and 8.2. Purely from a static stability standpoint, most of these situations demand a substantially negative slope of the head curve throughout the range of flow rate Q—or what is commonly specified as a pump with head “continuously rising to shut-off.” (This “rise in head” versus a drop in Q should not be confused with the pump developed head H, which is also properly termed the “head rise” H, produced by the pump at a given value of Q.) b) Dynamic instability. If the system has appreciable capacitance, operation may not be stable if the slope of the pump head curve is positive or even zero54. This is true even though the slope of the pump head curve is less than that of the system head curve as required above for static stability—as with the lower-Q intersection point of Figure 24a. Dynamic instability can be manifested as pump surge, a phenomenon wherein the flow 2.1 CENTRIFUGAL PUMP THEORY 2.57 FIGURE 24A and B Pump-system stability rate oscillates and can even be alternately positive and negative through the pump54. This is characteristic of “soft” systems that contain vessels with free surfaces and, therefore, appreciable capacitance. Two-phase flow increases the capacitance of a system and can cause dynamic instability. For example, fluctuating vapor volume within the propellant pump inducers can contribute to the dynamic instability of a rocket propulsion system55. On the other hand, a “hard” piping system with no capacitance is theoretically capable of accommodating a pump that has a flat or drooping head curve and that operates on that flat or drooping section of the curve. Low-specific-speed pumps can have drooping curves (Figure 22b), especially if designed with a high head coefficient c at the BEP. Figures 5 and 7 of Section 2.3.1 depict flat and drooping curves of low-s radial-bladed pumps, a type that is widely used in low-cost, small sizes. If properly applied, such machines will operate with stability. On the other hand, a conservative approach that guarantees both static 2.58 CHAPTER 2 and dynamic stability for the widest range of applications is to design all pumps without flat spots or droops in the head curves. DESIGN EXAMPLE ___________________________________________________ To illustrate the application of the preceding design information, the basic hydraulic design requirements are presented in Table 6 for a single-stage centrifugal water pump with a volute collector. The chosen conditions compute to a universal specific speed s of 1 (Ns  2733, and nq  52.9 from Eqs. 38a and 38b). The pump is to be designed for a suction specific speed ss of 4.5 (Nss K S  12,300 by Eqs. 41 and 42), therefore, a “performance-NPSH” (see NPSH discussion in Section 2.3.1) or NPSH3% of 14 ft (4.27 m). As such, this pump is readily applicable to taking water from an atmospheric reservoir with some suction lift. Although acceptable for 1780 rpm specified for this pump, this suction-specific speed ss of 4.5 is regarded as high for pumps with more head (energy level) than this one (see the energy-level discussion later). Impeller Design Beginning upstream, an ideal, axial-flow approach passage is assumed —this is known as an “end-suction” configuration. The suction-approach passage or suction branch (not shown) is simply a conical nozzle that increases the velocity of the fluid from the suction port to the impeller eye by about 50 percent in an axial distance of about half the diameter of the eye. This helps to ensure the existence of uniform flow at the eye. Too short a nozzle would mean excessive local meridional (axial in this geometry) velocity at the impeller shroud and possible separation. SUCTION NOZZLE IMPELLER INLET Beginning with ss of 4.5, computations for this example are carried out in Table 7 for the impeller eye geometry. The choice of the NPSH3% correlation of Table 1 is used, with k1  1.69 and k2  0.102. The local maximum velocity at the eye V1,sh (Vm,1,sh) is assumed to be 25% greater than the one-dimensional average Ve  Q/Ae. This is smaller than the bladeless result of Figure 14, but is typical of this end-suction configuration, TABLE 6 pump) Given conditions for design example (single stage, end-suction volume 2.1 CENTRIFUGAL PUMP THEORY 2.59 TABLE 7 Impeller inlet (design example) including the effect of the blading. Moreover, the value of k1 should be more than adequate for this value of V1,sh. The ss and t-fe relationships yield the eye flow coefficient fe, which in turn sizes the eye. fe  0.29 implies a t-value of 0.253, which is typical. However, lower fe- and t-values are common, especially for the case of a shaft through the eye, because this tends to maintain the level of ss in the face of the rs-effect in Eq. 50. The nominal velocity diagram at the eye—substantially the shroud-end or tip of the blade leading edge—shows Ve rather than V1,sh for the meridional component of velocity and so is not the actual velocity diagram at that location. Rather, this triangle serves to identify the geometry through the basic ratio fe  Ve/Ue —without having to deal with the uncertain choice of V1,sh/Ve. Moreover, fe is the tangent of the tip relative flow angle bf,1 as it would be for a uniform axial velocity profile in the eye. With the eye radius re established, the local shroud radius of curvature Rsh follows from the guidelines associated with Figure 13. The geometry established so far is illustrated in Figure 25. Before the full picture shown there can be established, the outlet must be sized. The computations in Table 8 for the impeller exit begin with the choice of the typical value of 2212 degrees for the outlet blade angle bb,2. This enables the head coefficient c to be chosen under the guidance of the upper curve in Figure 12. The value 0.385 is selected, and this yields the impeller diameter of 12 in. (304.8 mm). The other curve in Figure 12 is for outlet flow coefficient fi,2, which conveniently equals 0.1715 for IMPELLER OUTLET 2.60 CHAPTER 2 FIGURE 25 Impeller hub and shroud profiles (design example) s  1. This leads to the exit width b2 after adding in the leakage and the blockage of blades and boundary layers per the computations of Tables 9 and 10. However, Anderson6 points out that what matters for centrifugal pump performance is neither the blade angle nor the exit width individually, but the impeller outlet relative area 2pr2b2 sin bb,2. Choosing a higher blade angle is possible if b2 is correspondingly reduced (and fi,2 increased) so as to maintain this area and therefore the relative velocity W. Figure 15 shows that Vu,2 is thereby essentially unchanged; this in turn preserves the impeller head. Anderson’s overall pump efficiency correlation (Figure 10 and Eq. 44) and the component efficiency expressions of Table 3 lead to the results of Table 9. These give an indication of the relative magnitudes of the losses and are as follows: EFFICIENCIES Overall efficiency hp  0.8550 Mechanical efficiency hm  0.9814 Volumetric efficiency hv  0.9833 (leakage across front and back rings) Hydraulic efficiency hHY  0.8860 hHY is at this point simply deduced from the others, beginning with Anderson’s correlation. Although it is confirmed by Jekat’s correlation in the table, it can be found in a 2.1 CENTRIFUGAL PUMP THEORY 2.61 TABLE 8 Impeller outlet (design example) detailed computation of the hydraulic losses via one-dimensional methods. This will be carried out further on to obtain the performance characteristic curves. Meanwhile, this initial computation enables the determination of Vu,2 at the end of Table 9, which, along with Vm,2 from Table 8, is a major element of the outlet velocity diagram of Figure 26. With the leakage and exit blade angle information, Table 10 contains the computations of the blockage and the exit width b2. This entails the choice of the number of blades, the blade thickness t (2% of the impeller diameter and typically assumed to exist at the exit and elsewhere on the blades except near the leading edges where typically half that value is chosen), and the approximate blade length / (assuming the mean-streamline blade angle to be constant at 2212 deg). The boundary layer blockage is computed from the following approximations: BLOCKAGE AND WIDTH AT IMPELLER EXIT • Adverse pressure gradients on the blades lead to a boundary layer displacement thickness * of twice the zero-pressure gradient value 0* on each blade surface. • Secondary flows scrub the boundary layers from the hub and shroud surfaces; so, * is assumed to be equal to 0* on those surfaces. • 0*  0.0462 /0.8 n0.2/W0.2 for flat-plate, turbulent flow56, and is approximated in this example for low viscosity by a linear growth with length along the blade. The resulting thickness of the boundary blockage is 0.0732 in (1.86 mm) on the blades, which themselves have a thickness of 0.24 in (6.1 mm). Because these thicknesses are inclined at the 2212-deg outlet angle, the actual circumferential blockage is (1 e2,b)  (1 0.870) or 13 percent of 2pr2. In particular, (0.24  0.0732)/sin (2221 deg)  0.82 in or (6.1 2.62 CHAPTER 2 TABLE 9 Component efficiencies (design example)  1.86)/sin(2212 deg)  20.8 mm per blade, which for all six blades is 0.130 times the circumference of 37.7 in (958 mm). The width b2 is computed to maintain Vm,2 at the chosen value of 0.1715 U2, while accommodating this blockage and that of the sidewall boundary layers. Altogether, it can be computed from these data that 85 percent of the meridional exit area 2pr2b2 is estimated to remain open for the one-dimensional flow of (Q  QL) at velocity Vm,2, the boundary layers causing 13 of the blockage. This is quite typical. An openness of 90 percent is possible for larger impellers. HUB-SHROUD PROFILES It is now possible, through the guidelines outlined in the discussion of Figure 13, to finish plotting the hub and shroud profiles in Figure 25, which are also seen in Figure 26c. At this point, the envelope of the leading edges of the blades is approximated by a circular arc—later modified somewhat in the construction of the blades. The arithmetic average radius of the meridional passage at the leading edges is found with the circle of diameter 2.85 in (72.4 mm) to be 2.72 in (69.1 mm). The resulting line passing through the center of the circle is normal to both hub and shroud and is approximated by the dashed straight-line leading-edge quasi-normal shown in the Figure 26c. The rms radial point on this quasi-normal line of Figure 26c crosses the leading edges of the blades at r1, mean K r1, rms  2.91 in (73.9 mm) and is slightly larger than the arithmetic average radius of 2.72 in (69.1 mm). It is this latter radius that INLET VELOCITY DIAGRAMS 2.1 CENTRIFUGAL PUMP THEORY 2.63 TABLE 10 Impeller blade blockage and width at exit (design example) FIGURE 26A through C Impeller velocity diagrams for design example: a) inlet; b) outlet; c) meridional view 2.64 CHAPTER 2 must be used in computing the one-dimensional meridional velocity. After obtained, it is applicable to the rms location (the location of the “mean” or “rms” inlet velocity diagram). This diagram is one of three triangles shown for the inlet in Figure 26a, the other two being located at the hub and shroud locations of the blade leading edges. Notice that Vm,1 for the rms triangle is 16.3 ft/sec (5.0 m/s), which is slightly less than the eye velocity Ve of 17.3 (5.3). Allowing for blade blockage, this would bring the blocked meridional velocity Vm within the blading closer to Ve, the objective being to keep Vm constant in the inlet region and turn into the radial direction. The other triangles correspond to the radial locations of the blade at hub and shroud, as illustrated in Figures 26c and 27, and assume that Vm,1,sh  1.25 times Vm,1, and Vm,1,h  0.75 Vm,1. A full Q3D solution would determine these velocities more accurately; however, the design usually proceeds in this way—largely because the hub blade angle is usually a good deal larger than hub flow angle. Efforts to match the hub flow angle more closely entail special blading that is beneficial for highenergy pumps but has little effect otherwise. The blade angle at the shroud is slightly lower than the flow angle (by about 1 deg). This slightly negative incidence is actually ideal for efficient flow and minimum cavitation. The largest values of U and W exist at the shroud, as can be seen for the shroud inlet triangle, making it important to have the best match at that point. Two deg positive incidence is quite common at the mean or rms radial location and allows for blockage by the blades that does not increase the relative velocity W as the fluid enters the impeller. The outlet velocity components having been found in Tables 8 and 9, the slip velocity Vs must still be found in order to obtain the complete outlet velocity diagram shown in Figure 26b. This slip is computed by Pfleiderer’s method (Table 2), which utilizes the r(m) shape of the mean meridional streamline illustrated in Figure 27, Vs emerging as 15% of U2. The “a” factor for influence of the collector geometry was taken in the middle of the range for volutes at 0.75; (see Table 2). Wiesner’s Eq. 52 yields 17.65%. This would mean 6% less Vu,2 and head. However, this discrepancy is not unexpected, and in view of the earlier discussions on slip, the Pfleiderer result is chosen as more realistic. Nevertheless, uncertainty in the slip is the Achilles heel of the one-dimensional analysis method. For this reason, most analysts “calibrate” their codes by deducing the slip from test results and applying it to impellers of similar geometry. For Pfleiderer’s method, this would be done by calibrating the “a” factor. CFD solutions now appear to be the best approach to overcoming this difficulty. As has been emphasized heretofore, this outlet velocity diagram contains the basic information about the performance and design of the pump. It supplies the boundary conditions for the volute design, but first the impeller blading that must produce it will be established and evaluated. OUTLET VELOCITY DIAGRAM As indicated in Figure 26, the blade angles at both inlet and outlet have been chosen at hub, mean, and shroud (with the same 2212 deg all across the trailing edge being assumed, although some would specify a little variation). Fitting a reasonable blade shape between these end conditions can be done in the conformal plane (illustrated in Figure 19). These shapes, when transformed as described earlier, yield the hub, mean, and shroud blade shapes identified in the polar view of Figure 28. In actuality, the inverse “point-by-point” method was used, specifying Wg as indicated in Figure 29 and, at each of the 21 stations along the mean line of Figure 27, developing the velocity diagrams in the manner employed to arrive at Figure 26b; obtaining bb at each station; and developing the mean-streamline blade shape of the polar view with Eq. 60. Involved in this procedure— which was computerized—was the calculation at each station of the blockage and of the local slip velocity, the latter being estimated as a fraction of the discharge slip velocity Vs that increases rapidly to unity as the exit is approached. Note that the meridional area is needed at each station in order to compute Vm. As indicated in Figure 27, this is approximated here as the area of the frustum of a cone defined by the dashed lines or “quasi-normals,” which are as nearly perpendicular to both hub and shroud as possible. Except for some machines with meridionally curved passages and a large passage width-to-length ratio causing the true normals to be strongly curved, the quasi-normal approach works well. IMPELLER BLADING 2.1 CENTRIFUGAL PUMP THEORY FIGURE 27 2.65 Impeller blading–meridional view (design example) To establish the blade shapes along the hub and shroud, the corresponding “grid lines” of Figure 19 were found from polynomials satisfying the end conditions of blade angle and location, the y-position of the inlet end of the grid line being iterated until the desired blade wrap was obtained. In this way, different wrap angles u ( u2 u1) can be imposed, but a tolerable resulting blade shape requires that these us be not much different from the u ( 119.1 deg) obtained on the mean line from the method just described. In this example, it was possible to maintain u the same on all three construction lines. Alternatively and perhaps more consistently, the inverse approach could be applied at hub and shroud as well as on the mean line. Regardless of which procedure is used to generate the hub, mean and shroud blade shapes—the conformal transformation method, the point-by-point method, or some combination—the shape of the blade everywhere else still needs to be established. To do this, the shapes of the constant-u lines in Figure 27—called “blade elements”—must be specified. This can be done mathematically as indicated on the figure or by eye (the latter approach is widely used). Note that each constant-u line actually lies in a different meridional plane. Thus, Figure 27 depicts a superposition of all 21 meridional planes, each containing an intersection of the blade surface and appropriately identified as having the u-value noted on the figure. Figure 29 is a presentation of the free-stream relative velocity distributions at the blade surfaces and halfway between. The rapid approximate method of DIFFUSION ASSESSMENT 2.66 CHAPTER 2 FIGURE 28 Impeller blading—polar view, showing pressure sides of the blades (design example) Stanitz (Eq. 54) is utilized to obtain the surface W’s, the mean values of W on the mean (rms) streamline coming from the local velocity diagrams developed at each station22. The mean W’s at hub and shroud are found at the ends of the respective constant-u lines under the assumption that the ideal head or UVu-product is constant along each of these lines from hub to shroud. This is usually a fair approximation; however, a Q3D analysis (Figure 14) would yield a closer estimate of these mean W-distributions16,41. Nevertheless, the diffusion factors computed from the plotted surface velocities by means of Eqs. 59 are shown on the figure—these values are well below the 0.6 limit. This would indicate that the resulting solidity of 1.48 (originally estimated at 1.46 in Table 10) is adequate, and that the correct number of blades was chosen. Notice also on Figure 29 that the exit value of W is significantly larger than Wg, which is consistent with the outlet velocity diagram of Figure 26b. Further, a slight jump in the mean W at the inlet is indicated; however, as the higher value of W is not sustained, it is 2.1 CENTRIFUGAL PUMP THEORY FIGURE 29 2.67 Blade surface relative velocity distributions (design example) unlikely that it actually occurs in the 3D velocity field. Moreover, details near leading and trailing edges are not well handled by Eq. 54—a 2D blade-to-blade solution is more desirable for this type of diffusion assessment42. Volute Design With the exit velocity V2 of 43.6 ft/sec (13.3 m/s) and the tangential velocity component Vu,2 of 40.5 ft/sec (12.3 m/s) from Figure 26b, the volute design process is started in Table 11 using Eq. 61. This entails a choice of the radial “Gap B” of 6% of the impeller radius r2 and a tongue leading edge thickness tt equal to 70 percent of the impeller blade thickness. These choices are not critical for a pump of low energy level such as this is, as the stresses imposed by the flow on structural elements are small. The throat area AT is then found to be 32.60 in2 (21,032 mm2) and the throat velocity VT computes to 24.60 ft/sec (7.50 m/s), which is 56% of V2 and so represents considerable diffusion from impeller exit to throat. (The throat area AT is the fundamental feature of a volute and acts in concert with the impeller exit area, as emphasized by Anderson6 and Worster25 and described in the earlier part of this section on Designing the Collector in terms of the intersection of the casing and impeller lines26.) 2.68 CHAPTER 2 The circumferential distribution of the volute cross-sectional area Av versus polar angle uv from the tongue (Figure 20a) is developed in the latter part of Table 11, a final listing being presented for a) linear Av versus uv and b) constant one-dimensional angular momentum rvVv. This listing illustrates an earlier statement that the latter approach pro- TABLE 11 Volute casing (design example) 2.1 CENTRIFUGAL PUMP THEORY 2.69 TABLE 11 Continued. duces smaller areas upstream of the throat than does the former. Frictional effects in the smaller-Av portion of the volute are less prominent using the linear approach. A CFD assessment at all flow rates can be made to guide the design choice here. Finally, the onedimensional constant rvVv method can be improved upon by integrating a constant rVu distribution over each cross section, and the proper design must satisfy also continuity26. Estimated Performance Characteristics Detailed computation of hydraulic losses, together with leakage, disk friction, and other mechanical drags was done as described previously, and the results are presented in Figure 30. The method predicts the pump efficiency to be 85.5% at the design point of 2500 gpm (0.1577 m3/s), peaking at 86 percent at a 5% lower flow rate. The power consumption peaks at the design point at 77 hp (57 kW), indicating that this is a “non-overloading” design; that is, shaft power does not increase beyond this flow rate. The empirical methods reviewed earlier for shut-off head and power were applied here and blended into the one-dimensionally computed curves at half of the BEP flow rate. The shut-off head coefficient value of 0.596 is close to the 0.585 value of Stepanoff 4 and is admittedly open to alteration. The rise of head from design point to shut-off is from 104 up to 161 ft (31.7 up to 49.1 m) or 55%. This percentage could be smaller and still ensure stable operation in any typical system; however, the non-overloading feature could change, the power peaking at a higher flow rate. In retrospect, the design-point head coefficient c of 0.385 could be larger without the diffusion-factor results of Figure 29 becoming excessive. One would conclude from this exercise that the c-curve of Figure 12 is conservative and could be higher. Of more significance, however, is the demonstration in this design example of the utilization of easily applied fundamental fluid dynamical analyses such as the diffusion assessment illustrated in Figure 29 as the basic arbiters of design choices such as the head coefficient, number of blades, and so on. Furthermore, the shapes of the performance characteristics are revealed. 2.70 CHAPTER 2 FIGURE 30 Estimated performance characteristics (design example) The shut-off power is 34 hp (25 kW), arising from a power coefficient of 0.047 from Mockridge’s correlation36 for a b2/D2 of 1.59/12 ( 40.4/304.8)  0.1325—as found from Figure 23. This is 44% of the design-point power, a typical result. As seen in Section 2.3.1, this percentage increases with specific speed, where, of course, b2/D2 is also larger36. The NPSHR ( NPSH3% for this example) at off-design conditions is estimated empirically. At the higher, non-recirculating flow rates, NPSHR is related to the head required to accelerate the relative velocity within the blades to values higher than at the design point14. As stated in Section 2.3.1, operation at NPSH3% involves performance in the presence of extensive internal two-phase behavior. This is complicated by recirculation at the lower flow rates. Therefore, the full NPSHR curve in such a case usually has to be established experimentally. 2.1 CENTRIFUGAL PUMP THEORY 2.71 HIGH-ENERGY PUMPS ________________________________________________ Over the past few decades, there has been a trend toward pumping machinery that concentrates more power within a given volume. This trend is driven by cost and technology improvements. The basic energy transfer relationships show that smaller size demands higher rotative speed. Thus, over the same time period, speeds of high-power pumps have been increasing. Moreover, the number of stages in multistage pumps has been decreasing. There have been spikes in these trends; the resulting pumps suffering from excessive vibration, rotor and hydraulic instabilities, component failure, and cavitation damage57. The term “high energy” has been applied to these machines, and this label can be quantified in terms of the stresses arising in critical pump components and the likelihood of an adverse mechanical response that such stress levels imply. Research has led to technical solutions for effectively controlling rotordynamic behavior and reducing unsteady hydraulic thrust and surge as well as cavitation erosion53,58,59. The resulting pump reliability improvements and life extension should enable the previous trends to continue. Being aware of the energy level enables the pump user to assess whether operation and maintenance difficulties are likely to occur after the pump is installed and running, and it enables the designer to take the appropriate measures to ensure the technical integrity of the product. Pressure Pulsations Measured at the inlet or outlet port, the amplitude of the pressure pulsations can be a significant fraction of the pressure rise of the pump—especially at flow rates well below that of the BEP for the speed involved. Sources are a) the interaction of the pressure fields of the impeller and diffuser or volute, b) unsteady separated and reversed flows at impeller inlet and discharge and in the diffuser, c) cavitating flows, and d) combinations of these phenomena. Pressure pulsations presumed to exist at the impeller OD from the interaction of impeller blade-to-blade and diffuser vane-to-vane variations of pressure have been calculated by inviscid flow analysis to have a peak-topeak amplitude that is of the same order as the static pressure rise of the impeller60. Moreover, the viscous, thicker wakes existing at lower-than-BEP flow rates (here called “low flows”) and separated recirculating fluid from both impeller and diffuser that participate in these interactions can be expected to increase the pressure pulsation amplitude at such conditions. Figure 31 confirms these ideas, showing a bronze impeller that operated extensively at low flow. Cavitation pitting can be observed near the OD of the impeller, which means that the rarefactions of the pressure waves were below the vapor pressure of the liquid—these pressure minima therefore being below the inlet pressure to the impeller. Moreover, the bulged-out shrouds can be assumed to be the result of the repeated occurrence of the associated pressure spikes (that is, the maxima of the pressure waves) within the radial gap (“Gap B”) between the impeller blades and the diffuser vanes, the sidewall pressures on the outsides of the shrouds remaining comparatively constant. At greater values of design pressure rise than was the case for this impeller, this phenomenon creates correspondingly greater forces that have led to actual breakage of the impeller shrouds and diffuser vanes61. The cavitation seen in Figure 31, can also be observed on the leading edges of diffuser vanes, as in Figure 21 of Section 9.5, and this raises the possibility of diffuser vane breakage. Energy Level: Stage Pressure Rise Even in the absence of the weakening effect of cavitation erosion, the leading edge of a diffuser vane or volute tongue is a representative, highly stressed zone within a pump that is subject to failure if the magnitude of the pressure pulsations arising from the impeller-diffuser interactions just described is sufficiently large. Thus, the hydraulically induced stresses in these vanes can be the basis for quantifying the energy level of a pump stage. In Table 12, this concept is developed into an expression for the stress in terms of the fluctuating pressure magnitude dp that is assumed to act across the vane leading edge as illustrated in the table. The width b of the vane is close enough to b2 of the impeller exit to utilize the relationships for c and fi of Figure 12 to relate b/D to specific speed in Eq. (c) of the table. For similar velocity fields, the pressure pulsation magnitude dp is a constant multiplied by the stage pressure rise Pstg. Thus, for a limiting value of stress, the concept of a limiting stage pressure rise 2.72 CHAPTER 2 FIGURE 31 Damage to impeller from low-flow operation (Source: E. Makay in Power) emerges in Eq. (e). The constant K is chosen from experience, which leads to the resulting Eq. (f). This relationship is plotted in Figure 32. [It will be observed that this choice for K corresponds to a limiting stress s from Eq. (d) of 6,600 psi (45.5 MPa) that would exist if dp were equal to Pstg —with t/D  0.01 and e  0.85 as in the example.] The inverse variation with specific speed is a consequence of the greater b/D of higher-s pumps (as developed in Table 12), the wider vane introducing more stress at the juncture with the sidewalls for the same pressure loading and so imposing a lower stage pressurerise limit. Conversely, lower-s pumps should have higher limits for Pstg. Figure 32, therefore, illustrates this concept of a limiting stage pressure rise as a measure of the energy level of centrifugal pumps, the basis being a limiting stress level in a critical component of the pump. Starting with stress at other locations in the pump leads to similar results. To provide perspective, specific examples of pumps that by this definition are in the high-energy domain are plotted on the figure. These data points are taken from Table 13, which contains information for several well-known liquid rocket engine turbo pumps62,63,64 and for some representative high-energy electric utility boiler feed pumps53,57. The last column in Table 13 is another, more general way of comparing the energy level of these machines; namely, the torque per unit volume, which also has the dimensions of stress. For fixed ratios of stage width, casing OD, and other dimensions to impeller radius r ( r2), torque per unit volume differs from the listed values of torque/r3 by a factor. The actual torque per unit volume therefore ranges from one-half to one-sixth of the tabulated 2.1 CENTRIFUGAL PUMP THEORY 2.73 TABLE 12 Hydraulically induced stress levels torque/r3, depending on the casing or barrel thickness, and so on. However, as has been demonstrated, the critical stresses are more closely associated with the impeller OD, which makes comparison of the tabulated values more relevant. Thus, a pump with high torque/r3 can be expected to have correspondingly high local internal stresses. The maximum values listed, namely for the high-pressure propellant pumps on the RD-170 (Russian) and SSME (U.S. Space Shuttle) engines, tend to explain the high level of research and development that was necessary to successfully deploy these machines. Illustrations of some of these rocket engine pumps can be found in Section 9.19.2. Similarly, Section 9.5 2.74 CHAPTER 2 FIGURE 32 Pump energy level defined in terms of stage pressure rise provides examples of high-energy boiler feed pumps: the massive, barrel-type construction of these machines is illustrated in Figure 18. Specifically, the first boiler feed pump listed in Table 13 is the sole feed pump supplying the steam generator of a super-critical 1300 MW electric generating unit and consumes nearly 50 MW of shaft power. It can be seen in Figures 7 and 20 of Section 9.5. For pumps in the low-energy domain of Figure 32, normal design and manufacturing practices result in a more benign mechanical response to the abnormal fluid phenomena discussed here. However, for locations other than the diffuser entrance, which was the basis for the development of the figure, limiting stresses could be reached at considerably lower values of stage pressure rise. For this reason, the dashed line is offered as the upper limit of the low-energy domain; however, a thorough analysis of the stresses in any given application is the ultimate determinant of all the limits suggested in Figure 32 and of the acceptability of the design. It can now be seen that the design example treated earlier in this section is of the low-energy variety; therefore, the special design problems treated here and further on are of relatively little concern in such pumps. On the other hand, if the curve in Figure 32 were extended to much higher specific speeds, it would be found that many existing, large, high-s, low-head pumps are high-energy machines by this stressrelated definition. It is therefore not surprising that such pumps generally require full stress and modal analyses to identify possible destructive resonances and stresses. Fluid/Structure Interactions With the dimensions of pump energy level identified, the next step is to continue the examination of the problems mentioned previously and the methods that have become available for solving them. Attention is focused on hydraulic phenomena because, as the previous discussion of pressure pulsations implies, most of the adverse mechanical behavior exhibited by pumps originates from the behavior of the internal flow field. Excessive measured vibrations, material erosion, and component failures are often the external symptoms of fluid/structure interaction phenomena that are fundamentally explained from a hydraulics perspective. In addition to the hydraulically 2.1 CENTRIFUGAL PUMP THEORY 2.75 TABLE 13 Data on high-energy pumps* *Notes to this table: Flow rates apply to the pump inlet. RD-170 pump data derived from Sutton62 and Advanced Class boiler feed pump data from Ref. 57. Remaining data derived from Table 1 of Section 9.19.2, NASA reports63,64 and from information supplied by Flowserve Corporation.53 induced stresses in the pump structure that culminated in the domain definition of Figure 32, these phenomena encompass a) blade-vane interactions, b) recirculation, c) anomalous axial thrust behavior, and d) cavitation. In order to illustrate the methods for dealing with the problems created by these fluid/structure interactions, a sample suction stage of a high-energy multistage pump is utilized as the quantifying focus in each case. Table 14 contains the conditions and essential features of this machine, a meridional view of which is shown in Figure 33. As with the earlier design example, this stage has been designed in accordance with the procedures outlined for that example. This includes the velocity diagrams shown in the figure, as well 2.76 CHAPTER 2 TABLE 14 Data for high-energy pump suction stage as the design of the eye, the hub and shroud profiles, and the blading. A preliminary design was also made of the vaned diffuser and return vanes, following the guidelines presented in the foregoing Design Procedures subsection. The result is a representative machine to which the following paragraphs continually refer. Blade-Vane Combinations Certain numerical combinations of impeller blades and diffuser vanes (or inlet guide vanes, where they are employed) have been shown to have acoustic consequences that can exacerbate the pressure pulsations arising from the interaction of impeller and diffuser flow fields. Bolleter reviewed the types of interactions that can occur and the consequences with regard to pressure pulsations and resonance33. These 2.1 CENTRIFUGAL PUMP THEORY FIGURE 33 2.77 Sample suction stage of high-energy multistage pump types are associated with the integer difference m between the multiples of the number of impeller blades and the number of diffuser vanes, as explained in Table 15. To check the blade-vane combination, one simply forms a matrix of the multiples as shown and lists the difference in each cell. To illustrate the method, rather than use the pump of Table 14, a different example is chosen in order to dispel the notion that one can always use two different prime numbers for blade-vane combinations. This is an existing case of a 60,000 hp (45 MW) single-stage pump of s  0.6 (Ns  1640) and p  357 psi (2.46 MPa), which has 7 impeller blades and 13 diffuser vanes65. Notice that m  1 in the second order of impeller blade number, which means there should be pressure pulsations at 2 7 rpm/60 Hz. In fact, this pump has exactly this vibration and pressure pulsation frequency in the field, not least because the pump is operating at just the right speed for the acoustic waves emanating from successive interaction points to reinforce each other in producing unacceptable pressure pulsations. Even with this simple method, the final choice of the blade-vane combination is usually a compromise. For example, the method shows that all double-volute pumps have m  0 or 1 in the first order of impeller blade number, depending on whether this number is even or odd. m  0 means that the all the blades or vanes are interacting at the same instant, a consequence of which is torque ripple. High-energy volute pumps exist for both even and odd cases—quite large gaps between blades and vanes ( “Gap B”) are employed to mitigate these effects. The reader who checks the pump of Table 14 by this method will find that m  2 in the first order of both blades (7) and vanes (9) but that m  0 everywhere else in the matrix. 2.78 CHAPTER 2 TABLE 15 Choosing vane combinations to minimize pressure pulsations33 m  1 occurs only in the highest orders of both and should therefore be of little consequence. Checking whether 10 or 11 vanes would be better yields m  1 in the third order of impeller blades and second order of diffuser vanes, which is probably less desirable than m  2 in the first orders. Therefore, Figure 33 shows a value of “Gap B” that is 12 percent of the impeller radius, which should provide adequate protection from pressure pulsations and excitations of resonance. (The gaps at the impeller OD will be discussed further on.) Recirculation Separation and stall of the fluid flowing in the passages of impellers and diffusers occurs at low flow because of the incidence and large reduction in the onedimensional velocity relative to the passage that happens at low flow. This and the con- 2.1 CENTRIFUGAL PUMP THEORY 2.79 sequent recirculation patterns in the impeller were discussed and illustrated in Figure 6. Fischer and Thoma66 visually observed and recorded the flow patterns, finding that as flow rate is reduced, wakes on the suction side of all blades thicken until they occupy half the passage width at half the BEP flow rate. At lesser flow rates, the wakes continue to thicken but become irregular, stalling in one passage and not the others—the stall pattern moving into and out of adjacent passages and so rotating relative to the impeller. As shut-off is approached, this rotating pattern is accompanied by reversed flow emerging from the inlet of the stalled passage. Fraser, working with typical impeller geometries, formulated rules for computing the flow QSR at which this reversal occurs as Q is reduced at constant speed67. His expressions, found further on, include the effect of impeller eye size on QSR. As one might expect, a pump with an eye diameter approaching that of the impeller OD will have QSR approaching QBEP. At Q  QSR, the impeller flow patterns are highly unsteady—as is usually the case with massively separated flows—creating nonsynchronous, low-frequency or random pressure pulsations, the resulting shear layers between the reverse-flowing and in-flowing fluid having vortices with locally low pressures so cavitation can also exist. Fraser also quantified the flow rate QDR below which impeller discharge recirculation exists. Forces from such motion can cause fatigue failure of the impeller blades, diffuser vanes or volute tongue, cavitation erosion also playing a part as in Figure 31. In Section 2.3.2, Fraser describes the identification and consequences of recirculation in detail, the more general designation QR referring to either QSR or QDR, depending on whether Q is between or below both. The ability for pumps to operate with any form of separation; stall; or, worse, flow reversal (recirculation) depends on the energy level. This can be approximately quantified, as outlined under the subject of Minimum Flow Limits further on, which include consideration of accompanying cavitation activity. Axial Thrust Response to Recirculation Discharge recirculation usually involves backflow from the diffuser, itself containing oscillating flow patterns and rotating stall. Fluid emerging from the diffuser will be spinning opposite to the direction of rotation, such fluid having a major effect on the sidewall gap flows as it joins the leakage flows described under Predicting Axial Thrust. As this fluid invades the sidewall gaps, it can slow or virtually cancel the usual positive swirling of the gap fluid. Iino, Sato, and Miyashiro experimentally observed and recorded this behavior, which was exaggerated by shifting the impeller axially and by changing the ring clearances68. An added, not unexpected effect is that as Q is reduced below QDR, the invading flow from the diffuser can favor the front or back side of the impeller and then switch sides upon further reduction of Q. This effect is clearly seen in the experimental thrust-versusQ plots of Figure 34, the impellers having been shifted as just described. Depicted there is the resulting net load on the axial thrust bearing of an eight-stage, 3600-rpm diffuser pump that had a cylindrical balancing drum (not a self-compensating balancing disk). The drum was sized so as not to completely eliminate the thrust—in order to avoid thrust reversals. The solid lines are the predicted net thrust according to the methods outlined in Table 4 for three axial positions of the impeller. The large excursions in net thrust were eliminated by restricting the entry of the invading diffuser backflow into the sidewall gaps —through a tightening of the gap between the shrouds of impeller and diffuser (Gap “A”) in Figure 35. Gap “A” is not effective unless the “overlap” of the two mating shrouds is from four to six times the gap dimension61. Moreover, if Gap “A” is minimized, this can exaggerate the blade-vane interactions, making it necessary to open up Gap “B” more than would be necessary were Gap “A” not minimized69. A further possibility that has been observed in a single-stage double-suction pump is the unsteadiness of impeller discharge recirculation and, most likely, of the diffuser or volute backflow. The side-to-side switching just mentioned appears in Figure 36 to be happening as a function of time as well as of flow rate Q, as evidenced by the axial motion, which is accompanied by discharge pressure pulsations. The “fix” mentioned in the figure was, again, mainly minimizing Gap “A.” Closing Gap “A” and opening Gap “B” are procedures that have been widely and successfully applied in high-energy pumps, which usually work well at BEP but run into difficulties at low flow69. The procedures have proven to cure the thrust and pressure 2.80 CHAPTER 2 FIGURE 34 Axial thrust response to recirculation. (Source: Flowserve Corporation) pulsation behavior just described and also improve the low-flow performance curve shape65. The results are high-energy pumps that can operate smoothly over a wider range of flow rate than—in many cases—was originally expected or specified. Nevertheless, the higher the energy level, the more intolerable is any unsteadiness and pressure pulsation activity. Minimum Flow Limits Because the intensity of pressure pulsations and the accompanying vibrations can increase beyond acceptable limits as flow rate is reduced at constant speed below the BEP flow, or more specifically, below QR, expressions for how low Q/QBEP can be without exceeding these limits have been developed. Manufacturer and user groups such as the Hydraulic Institute (HI), the International Standards Organization (ISO), and the American Petroleum Institute (API) have specified vibration limits that must be met at what is usually called at the minimum continuous stable flow (MCSF) or simply Qmin. In an attempt to answer the question of how far into the recirculation zone (where Q  QR) an operator can take a pump before reaching the MCSF, Gopalakrishnan proposed a general rationale for computing Qmin that takes the energy level into account—through the flow rate and rotative speed of the machine70. These two quantities imply the blade tip speed at the inlet of the impeller. (Through typical ratios of impeller OD to eye diameter, this also implies the OD tip speed and therefore the head, the energy level having been defined in Figure 32 in terms of stage pressure rise pstg.) The theory for this method is defined in Part A of Table 16, in which Qmin is computed as the product of a series of Kfactors multiplying the value of QR, which is computed according to Fraser’s earlier development67. Conceptually, the product of these factors approaches unity in the maximum-energy case, where the instabilities accompanying any recirculation at all are 2.1 CENTRIFUGAL PUMP THEORY FIGURE 35 FIGURE 36 2.81 Gaps at impeller periphery61 Eliminating unsteady thrust and pressure pulsations (Source: E. Makay in Power.) significant. Conceivably, this product could exceed unity at the highest energy levels, as separation and stall and the attendant unsteadiness must occur—as Q is reduced—before the backflows that characterize QR, which was observed by Fraser as the value of flow rate for the onset of recirculation (see Section 2.3.2). K1 and K3 are given in Table 16 as equations that have been curve-fitted to speed/flow rate- and NPSH-effect charts that appeared in Gopalakrishnan’s presentation70. This includes the ability to enter any speed (rpm) into the computation for K1, only 3500 and 2.82 CHAPTER 2 1800 rpm having explicitly appeared in the charts. NPSH plays a role in the determination of Qmin —through the factor K3 —because of the exacerbation of the unsteadiness and pressure pulsations due to dilation of the cavities in the two-phase internal flows that exist for any value of NPSH that is less than the inception NPSHi. (See the ensuing cavitation discussion and the description in Section 2.3.1 that accompanies the definition of NPSH-limits.) As an example, the minimum flow of the high-energy pump suction stage of Table 14 and Figure 33 is computed in Part B of Table 16. (The value of R for the calculation of K3 is taken from Table 18.) This machine has the relatively high energy level of some of the boiler feed pumps shown in Figure 32 and listed in Table 13. Thus it is not surprising that the resulting value of Qmin is 90% of QR (which equals QSR, as is typical) and 48.5% of the BEP flow rate. This is indicated on Figure 37, in which the computed performance curves TABLE 16 General method for computing minimum flow: quantifying the energylevel effect on Qmin relative to QR70 2.1 CENTRIFUGAL PUMP THEORY 2.83 TABLE 16 Continued. FIGURE 37 Estimated performance of high-energy pump suction stage of this pump are displayed. (See Part B of Table 14 for the main elements of these performance predictions.) However, some of the remedies that have been discussed in this section to counter the adverse effects of high-energy pump phenomena can reduce Qmin to lower values than this. In fact, the method contains the flexibility to take into account such improvements through the factor K5. 2.84 CHAPTER 2 In computing QSR, the angle b1 is called for. This is approximated by the inlet flow coefficient fe  0.3 from Table 14, which corresponds to a nominal inlet tip flow angle of 16.7 deg. The actual flow angle (and blade angle) at that location is slightly larger due to the hub-toshroud variation of the incoming meridional velocity as assumed in the development of the inlet velocity diagrams of Figure 33. This will have a small effect on the results, depending on how one interprets “b1”. In this regard, as can also be appreciated from the choices that must be made for the Ks, the method of Table 16 is not precise; however, it is a useful indicator of what the user and designer can expect in determining the operating range of a pump. The above general theory for computing minimum flow is inclusive of all types of centrifugal pumps and has found application especially for high-energy pumps, which can be difficult to evaluate precisely in the varying circumstances of installation and operation in which they are usually applied. The judgments that must be made in order to apply the method of Table 16 can be largely avoided if actual data for Qmin are available. It would in fact be a monumental task to establish precise limits for all pump types and operational envelopes. Nevertheless, the effect of energy level on MCSF has been found experimentally for several classes of API process pumps through extensive testing. From this, Heald and Palgrave have developed a method for computing the minimum flow of these pumps as follows71: Qmin  MSCF  K7 K 100 M QBEP K7  K7 1rpm, configuration, NSS 2 where KM  KM 3 1NPSHA>NPSHR 2 BEP, fluid 4 and (68) (69) (70) Obtained from Figure 38, the factor K7 accounts for the effects of speed and configuration. As used in that chart—and in Eq. 69, the term suction-specific speed Nss (or ss) FIGURE 38 Minimum continuous stable flow (MCSF) for process pumps71 2.1 CENTRIFUGAL PUMP THEORY 2.85 does not mean that NPSH-effects are also included; rather, this expresses the fact that the impeller design is affected by the designer’s choice of Nss as can be seen in the case of the earlier Design Example, beginning in Table 6. This leads immediately in Table 7 to the inlet flow coefficient fe (through the NPSH-correlations of Table 1) and the impeller eye diameter De or radius re. Following on from the discussion of Recirculation, the ratio QSR/QBEP can be expected to increase with eye size or De/D2. Being strictly based on experiments, the Heald and Palgrave method does not deal explicitly with QSR or QDR, but one can see from Eq. 38 that this principle is operative: Higher-Nss pumps require greater MCSF. Before this fact was clearly understood, the trend was to design for greater Nss —in order to increase suction capability of these pumps. When operators ran them back to the same low flows as they had done with earlier pumps designed for lowerNss capability, failures became epidemic. This led to a call for “lower-suction specific speed pumps,” or, more accurately, pumps designed for lower-Nss capability; specifically 11,000 (ss  4) or less72. The NPSH-effect does come into play—as with K3 in the previous general method — through the factor KM in Figure 39a. In this case, the term NPSHR means NPSH3%. Here again, it is the ratio R  NPSHA/NPSHR that is the determining factor because it is a measure of the cavitation activity that is invariably present in the first stage of a pump, and FIGURE 39 NPSH-effects on: a) MCSF and b) impeller life71 2.86 CHAPTER 2 therefore in all single-stage pumps. The lower the value of R, the more violent the pressure pulsations accompanying the recirculating fluid and the consequent vibration. The situation is mitigated considerably at greater R, and KM operates through Eq. 68 to decrease the MCSF. Of course, if R were great enough—often 5 or more—NPSHA would exceed NPSHi (as discussed in Section 2.3.1) and there would then finally really be no cavitation in the pump. (That high a value of NPSHA is rare; for, if it were supplied, there would hardly be a need for a pump in the first place.) Figure 39a also brings in the effect of the liquid being pumped. When room temperature water boils (cavitates), the mass boiled off by the local drop below the vapor pressure makes considerably more cavity or bubble volume than some other liquids, namely hydrocarbons and hot water17. (See Section 2.3.1.) The table refers to room temperature or cold water. By way of illustration, computing the minimum flow of the end-suction volute pump of the Design Example begins with Curve B in Figure 38 (for pumps with 6-inch discharge and larger and for 1800 rpm and lower as stated on the figure). [This pump has a discharge port of about 9 inches (229 mm) as would be the case if the velocity in this port were half of the throat velocity VT in Part A of Table 11, as suggested previously in the paragraphs on Volutes under Designing the Collector.] Nss for this pump was chosen as 12,300 (ss  4.5), which yields 41.5% for K7. If it were decided to provide this cold-water pump with 16.4 ft (5 m) of NPSHA, R would be 1.17, and the figure yields 0.97 for KM. [NPSH3%  14 ft. (4.27 m) for this pump, as seen in Table 6.] Thus, from Eq. 68, this pump has an MCSF of 0.415 0.97 or 40% of QBEP. Had it been designed for Nss  11,000 (ss  4.025), and if R still were 1.17, the NPSHA would have been 19 ft (5.8 m) and MCFS would have been 0.36 0.97 or 35% of QBEP. Moreover, if it were pumping hydrocarbons at this same NPSHA, MCFS would have been even lower, namely 0.36 0.78 or 28% of QBEP. It would appear, though, that for many applications, the pump as designed (at ss  4.5) has a low enough energy level to allow for an adequate range of flow-rate capability, and that therefore, the value Nss  12,300 is in this case not excessive. Cavitation Considerations Having alluded to and treated the subject of cavitation a number of times in this section, we should expand on the role that this ever-present phenomenon plays in the operation and durability of centrifugal pumps, particularly those having a high energy level. The manifestations of cavitation that are encountered and become issues for the operability and life of a pump are a) cavitation accompanying backflow from the impeller eye, b) cavitation-generated instabilities and pressure pulsations, and c) erosion, which involves the prediction of the NPSHR-versus-flow rate characteristic curve to maintain life and, conversely (d) the prediction of life for a given NPSHA. The range of NPSH over which cavitation occurs within a pump extends from the point where pump head or pressure rise undergoes an identifiable drop—usually 3% for repeatable results for the NPSH-value involved—namely the “performance-NPSH” or NPSH3% — upwards. As NPSH increases from this point, there is an extensive range over which no observable performance loss is detected, yet erosive damage is progressing at a sometimes excessive rate. Finally, at the upper end of the range, all two-phase activity ceases—all bubbles and cavities are suppressed—namely, at the inception-NPSH value or NPSHi. As stated previously, NPSHi has been observed to be typically about five times NPSH3%. This is clearly described in Section 2.3.1, in which the different NPSH-limits are distinguished. The following cavitation considerations pertain to this range. a) Cavitation and backflow. The minimum flow limits imposed by the R-value or NPSHeffect arise partly because a lower flow rate at a low NPSH—although it may be in excess of NPSH3% —involves a strong interaction between suction recirculation and cavitation that can be intense, especially for inducers and large-eye (large De/D2) impellers. At flow rates Q V QSR, there exists upstream of the impeller an annulus of back-flowing fluid that emerges from the impeller. Drawing the velocity diagram at the impeller leading edge at the shroud for the case of reversed flow reveals that the absolute velocity component V i) is mostly circumferential and ii) is greater than the impeller tip speed Ut,1 or Ue. Thus the fluid leaving the impeller hugs the outer wall of the approach passage. If this passage is an axial pipe supplying an end-suction impeller or inducer, the pressure along the centerline can be below the vapor pressure, thus creating a vapor core that extends many 2.1 CENTRIFUGAL PUMP THEORY 2.87 diameters upstream, as was shown in a photograph that is part of an article by Yedidiah73. Obviously the energetic backflow has to be balanced by an equivalent inflow that enters the impeller from the interior of the pipe and therefore along the hub streamline of the impeller. Vapor from the core is drawn in also and tends to fill the impeller and vapor-lock it. At this point, the pressure to drive the highly spinning liquid upstream is non-existent and the backflow ceases. The vapor core disappears and the impeller once more begins ingesting liquid, the process just described repeating itself at a very low frequency (as low as 1 to 6 Hz) and called cavitation surge74. It has been found possible to passively divert the backflowing liquid outward from the inlet passage at the impeller or inducer eye into a series of vaned passages surrounding the inlet pipe, the vanes deswirling the backflowing liquid and returning it to a point or annular port upstream. Properly designed, this “backflow recirculator” has completely eliminated all cavitation instabilities in inducers—over the full range of flow rate from shut-off to run-out. It was shown to work for some high-Nss impellers to which it was applied74. Cavitation surge, however, is rarely seen in low-Nss impellers and is usually completely avoided by running the pump at flow rates greater than minimum flow Qmin as established by one of the previous methods or by test. b) Cavitation-related instabilities and pressure pulsations. At flow rates greater than Qmin, cavitation can cause or intensify pressure pulsations. Such instabilities are connected with the variety of cavity and bubble configurations that can exist over the range of flow rates and NPSH- or R-values. The nature and extent of cavitating flow within a pump has been studied extensively by visual observation. Figures 40 and 41 are laboratory photographs61 of a sector of a boiler feed pump impeller eye, in which the suction sides of some of the blades are visible. These were taken with the aid of a bright flash that lasted for 1 ms. They can also be found in Ref. 24 of Section 2.3.1. Figure 40a shows the sheet cavity that exists at BEP flow rate and at an R-value of 2. Also observed at BEP—but at R  1 (that is, NPSH  NPSH3%)—is the thick, extensive cavity of Figure 40b. This cavity extends from the blade leading edge to the throat formed by the leading edge of the following blade, and this same pattern exists on every blade. Most observers note that when this suction-side cavity reaches the next throat, the pressure rise or liquid head starts breaking down—which is what is recorded at this “3-percent-NPSH” point. Instabilities tend to be at a minimum at both of these relatively steady-flow conditions. Figure 41, on the other hand, displays highly unsteady cavity flows. Both photographs (A and B) were taken at the same half-flow condition at different instants—for R  2. A similar sequence of three photographs at this condition appears in Section 9.5, and none of the three possesses a cavity pattern that resembles either of the others. The position of the somewhat smaller cloud in Figure 41b that has broken off the main cavity appears to be traveling toward the pressure-side (out of sight) of the next blade. Pressure-side cavitation erosion would be the result, and this accords with the findings in Section 2.3.2. Pressure pulsations are associated with unsteady cavitation patterns, and plots of the amplitude and frequency of suction pressure pulsations show increasing frequency as R is a. b. FIGURE 40 Cavities on impeller blades at BEP flow rate: a) NPSH = 2 Flowserve Corporation)61 NPSH3%; b) NPSH = NPSH3%. (Source: 2.88 a. CHAPTER 2 b. FIGURE 41A and B Cavities on impeller blades at half the BEP flow rate: NPSH = 2 Flowserve Corporation)61 NPSH3%. (Source: increased—as would be expected as bubble size is reduced. The amplitude peaks at R greater than 1—usually closer to 2. Elimination of these instabilities is best done by avoiding cavitation. This can be done by increasing NPSHA and optimizing the blade shape for minimum cavity activity75. To avoid the excessive pump and system oscillations that can occur, some users specify that a cavitation-flow visualization test be conducted and require that minimal or no cavities shall be observed. c) Erosion due to cavitation and the prediction of the Life-NPSH. In the majority of cases cavitation is unavoidable, and the issue becomes the life of component (usually the impeller) in resistance to attack by collapsing bubbles and larger cavities as they are swept out of the low-pressure regions near the blade leading edges. The pressure created at the point and instant of collapse is immense. Photographs of the erosive damage resulting from this activity can be found in Sections 2.3.1 and 9.5. The surface failure mechanism is one of fatigue due to repeated collapse of bubbles adjacent to the blade. This happens at a large number of closely spaced sites, the resulting erosion having a strongly pitted texture. The rate of erosive depth penetration into the surface of a venturi subject to bubble collapse was found by Knapp76 to increase as the sixth power of velocity V at a constant cavitation number k  2(p1 pv)/rV2. In other words, the erosion rate increased as the third power of (p1 pv) or NPSH at the inlet of the venturi, which means that the collapse pressure rises with NPSH, as is known from the Rayleigh bubble collapse theory. The velocity V in Knapp’s venturi corresponds in a pump to the maximum relative velocity W at inlet—which is conveniently represented by the impeller inlet tip speed Ut,1 as a criterion for damage rate—and t corresponds to the cavitation number k. For a given impeller at constant Q/N and a constant value of t, (or at a constant available suction specific speed,) the higher Ut,1 is, the greater the NPSH and the greater the damage rate. Also, as mentioned earlier, a higher value of Ut,1 implies a correspondingly greater U2 and, therefore, greater pump pressure rise or head. Therefore, high-head pumps are more likely to suffer from cavitation erosion, making cavitation a “high-energy” pump phenomenon. High pressure-rise has already been shown to be a feature of high energy pumps through the definition of energy level in terms of classical stress loading (Figure 32). Facing the reality of high-energy pump destruction due to cavitation erosion, Vlaming redefined the term “NPSHR” to mean NPSH40,000 hrs.; that is, the NPSH needed to limit the damage sufficiently so as to ensure an impeller life of 40,000 hours. He developed an empirical method for predicting the curve of this “damage-NPSH” versus flow rate for conventionally-designed impellers77. This method is defined in Table 17 and includes the inlet tip speed effect on the erosion rate. A value of the vaporization factor Cb that is less than unity applies to hot water and to other liquids such as hydrocarbons, which generate far less vapor volume when they cavitate than does cold water. (See the earlier discussion under NPSH Effects in the subsection on Specific Speed and Optimum Geometry.) (Rather than depend on theory for the API process pumps men- 2.1 CENTRIFUGAL PUMP THEORY 2.89 TABLE 17 NPSH required for 40,000 hours life77 tioned earlier in connection with their minimum flow data, Heald and Palgrave quantified NPSH40,000 hrs. in Figure 39b71.) Applied to the high-energy multistage pump suction stage of Figure 33 and Table 14, Vlaming’s method yields the NPSHR-values shown in Table 18 under the heading fe0.3. These are plotted as the NPSH40,000 hrs.-curve on Figure 37. The blades of the pump are assumed to be set for zero incidence of the incoming flow to their camber lines at the BEP —called “shockless entry” and denoted by “SE.” (Many designers make QSE somewhat larger than QBEP to achieve lower NPSHR at Q  QBEP.) The column to the right in the table contains the results for the same method applied to the more common case for highenergy pump suction stages, namely fe  0.25 (and lower). For the same shaft diameter, flow rate and speed, fe  0.25 yields a larger eye diameter through Eq. 49; namely, 13.92 in. (353.6 mm) versus 13.37 in (339.6 mm) for the fe  0.3 case as shown in Figure 33. Moreover, that figure shows the inlet tip speed Ut,1 (Ue) to be 274 ft/sec (84 m/s), whereas for the larger-eye case the value is computed to be 286 ft/sec (87 m/s). As Table 18 reveals, this larger-eye impeller has a lower value of NPSH3% (as computed from the correlations of Table 1), which is the reason for sizing large eyes. But Vlaming’s method indicates that the smaller eye requires less “damage-NPSH” (R  1.69) than does the larger eye (R  2.06), even though the latter requires less “performance-NPSH” or NPSH3%. So, from both a minimum-flow and a cavitation damage standpoint, the advantage of a smaller eye for high energy levels is evident. d) Prediction of life for a given NPSHA. The inverse of the foregoing problem is how to determine the life under cavitating conditions at a given value of available NPSH. Gülich found a connection between an observed length Lcav of the cavity trailing off the leading 2.90 CHAPTER 2 TABLE 18 NPSHR of high-energy pump suctions stages edge of the blade at BEP (Figure 40a) and the life of an impeller operating at this condition78. The resulting procedure is outlined in Table 19, beginning with the definition that the life is the time it takes for the erosion to penetrate through 75% of the blade (or wall) thickness. In the absence of a cavitation-visualizing test, Lcav can be estimated as suggested by the last two formulas in Table 1975. Critical to this estimate is the assumed value of the “inception-NPSH” or ti. For conventionally designed impeller blades, ti  1 at the BEP, whereas aerodynamically shaped blades that minimize the local reduction of static pressure have been produced with ti  0.5 at the BEP75,79. The life computations of Table 20 follow from application of the method of Table 19 to the sample suction stage of Figure 33 for ti  1 and 0.5. Higher values of ti apply for Q  QBEP —as might be expected from the shape of Vlaming’s NPSH40,000 hrs.-curve in Figure 37 and the fact that the local pressure reduction in the leading-edge region increases with incidence. From this and the results of Table 20, it is evident that improvements to conventional blade-design practice are essential if life is to exceed half a year. Further, this life calculation method is based on the existence of a sheet cavity like that of Figure 40a; and the disordered cavity structures of Figure 41, which happen in the presence of recirculation, are not addressed. Lowflow cavitation damage is generally more severe, and is described in Section 2.3.2. 2.91 2.1 CENTRIFUGAL PUMP THEORY TABLE 19 Prediction of life under cavitating conditions dimension less (mm) Moreover, corrosion can play a role in cavitation-related erosive activity, an effect that was also addressed in Gülich’s work.78 Nevertheless, the ability to compute erosive behavior— even at the BEP—allows one to evaluate design improvements and provides a good idea of the life that can be expected under normal operating conditions. Obviously, if the available NPSH is greater than the inception-NPSH (that is, tA  ti), there is no bubble activity of any kind, and at QBEP the cavity length Lcav  0. An illustration of what can be achieved in this regard is presented in Figures 24 and 25 of Section 9.5, which are the “before” and “after” photographs of the model impeller blades for the first stage of a 24,000 hp (18 MW) pipeline pump. Quasi-three dimensional analysis of the blade pressure loading led to changes in shape that eliminated the cavities53,75. This complete absence of cavitation is becoming the desired objective for the design and application of new and upgraded multistage high-energy pump suction stages. This approach eliminates both the erratic mechanical behavior that occurs in response to unsteady cavity patterns and the erosion due to bubble collapse; thereby substantially increasing the life and reliability of such machines80. 2.92 CHAPTER 2 TABLE 20 Life of high-energy pump suction stage 2 REFERENCES _______________________________________________________ 1. Jennings, G. P., and Meade, L. P. “Determination of Pump Efficiencies from Fluid Temperature Rise.” Presented at the 7th Annual Spring Pipeline Conference, Houston, Texas, May 14, 1956, under the auspices of the American Petroleum Institute’s Division of Transportation. Published in the Division’s Vol. 36 [5] 1956. 2. Cooper, P., and Reshotko, E. “Turbulent Flow Between a Rotating Disk and a Parallel Wall.” AIAA Journal. Volume 13, No. 5, May 1975, pp. 573—578. 2.1 CENTRIFUGAL PUMP THEORY 2.93 3. Sabersky, R., and Acosta, A. J. Fluid Flow, Macmillan, p. 76 (1966). 4. Stepanoff, A. J. Centrifugal and Axial Flow Pumps. 2nd ed. Krieger Publishing, Malabar, FL, 1957. 5. Iversen, H. W. “Performance of the Periphery Pump.” Transactions of the ASME. Vol. 77, Jan. 1955, pp. 15—28. 6. Anderson, H. H. “Prediction of Head, Quantity and Efficiency in Pumps—The Area Ratio Principle.” Performance Prediction of Centrifugal Pumps and Compressors. ASME, 1980, pp. 201—211. 7. American National Standard for Centrifugal Pumps for Design and Application, ANSI/HI 1.3-2000, Section 1.3.4.1.11-14, Hydraulic Institute, Parsippany, NJ, www. pumps.org. 8. Cooper, P. “Analysis of Single- and Two-Phase Flows in Turbopump Inducers.” Transactions of the ASME, Journal of Engineering for Power. Vol. 89, Series A, Oct. 1967, pp. 577—588. 9. Turpin, J. L., Lea, J. F., and Bearden J. L. “Gas-Liquid Flow Through Centrifugal Pumps—Correlation of Data.” Proceedings of the Third International Pump Symposium, Texas A&M University, 1986, pp. 13—20. 10. Runstadler, P. W., Jr., and Dolan, F. X. “Two-Phase Flow, Pump Data for a Scale Model NSSS Pump.” Polyphase Flow in Turbomachinery, ASME, 1978, pp. 65—77. 11. Cooper, P., and others. “Tutorial on Multiphase Gas-Liquid Pumping.” Proceedings of the Thirteenth International Pump Users Symposium, Texas A&M University, Mar. 1996, pp. 159—174. 12. Kasztejna, P. J., and Cooper, P. “Hydraulic Development of Centrifugal Pumps for Coal Slurry Service.” Proceedings of the Eighth International Symposium on Coal Slurry Fuels Preparation and Utilization. U.S. Dept. of Energy, May 1986. 13. Gongwer, C. A. Transactions of the ASME. Vol. 63, Jan. 1941, pp. 29—40. 14. Stripling, L. B., and Acosta, A. J. “Cavitation in Turbopumps.” Parts 1 and 2, Transactions of the ASME. Series D, 1961. 15. Lazarkiewicz, S., and Troskolanski, A. T. Impeller Pumps. Pergamon Press. 16. Katsanis, T., and McNally, W. D. “Revised Fortran Program for Calculating Velocities and Streamlines on the Hub-Shroud Midchannel Stream Surface of an Axial-, Radial-, or Mixed-Flow Turbomachine or Annular Duct.” I, User’s Manual, NASA TN D-8430, and II, Programmer’s Manual, NASA TN D-8431, 1977. 17. Cooper, P. “Application of Pressure and Velocity Criteria to the Design of a CentrifugalPump Impeller and Inlet.” Transactions of the ASME. Series A, Apr. 1964, pp. 181—190. 18. Busemann, A. “Das Förderhohenverhaltniss Radialer Kreiselpumpen mit Logarithmischspiraligen Schaufeln.” Zeitschrift für Angewandte Mathematik und Mechanik. Vol. 8, Oct. 1928, pp. 372—384. 19. Wiesner, F. J. “A Review of Slip Factors for Centrifugal Pumps.” Transactions of the ASME. Series A, Oct. 1967, pp. 558—572. 20. Pfleiderer, C. Die Kreiselpumpen für Flüssigkeiten und Gase. 5te Auflage, SpringerVerlag, 1961. 21. Dicmas, J. L. Vertical Turbine, Mixed Flow, and Propeller Pumps. McGraw-Hill, 1987. 22. Stanitz, J. D., and Prian, V. D. “A Rapid Approximate Method for Determining the Velocity Distribution on Impeller Blades of Centrifugal Compressors.” TN 2421, NACA, July 1951. 23. Lieblein, S. “Experimental Flow in Two-Dimensional Cascades.” Chap. VI in Aerodynamic Design of Axial-Flow Compressors. SP-36, NASA, 1965, pp. 202–205. 24. Guinzburg, A., and others. “Emerging Sewage Pump Design Requirements.” 1997 ASME Fluids Engineering Division Summer Meeting. Paper FEDSM97-3325, June 1997. 2.94 CHAPTER 2 25. Worster, R. C. “The Flow in Volutes and Its Effect on Centrifugal Pump Performance.” Proceedings of the Institution of Mechanical Engineers. Vol. 177, No. 31, 1963. 26. Lorett, J. A., and Gopalakrishnan, S. “Interaction Between Impeller and Volute of Pumps at Off-Design Conditions.” Transactions of the ASME, Journal of Fluids Engineering. Vol. 108, Mar. 1986, pp. 12—18. 27. Miller, D. S. Internal Flow Systems, BHRA Fluid Engineering, 1978. 28. Japikse, D. “A Critical Evaluation of Stall Concepts for Centrifugal Compressors and Pumps—Studies in Component Performance, Part 7.” In Stability, Stall and Surge in Compressors and Pumps. FED—Vol. 19, ASME, Dec. 1984, pp. 1—10. 29. Kovats, A. “Diffusers of Multistage Centrifugal Pumps.” In Return Passages of MultiStage Turbomachinery. FED—Vol. 3, ASME, June 1983, pp. 61—66. 30. Reneau, L. R., and others. “Performance of Straight, Two-Dimensional Diffusers.” Transactions of the ASME. Series D, Vol. 89, 1967, pp. 141—150. 31. Fox, R. W., and Kline, S. J. “Flow Regimes in Curved, Subsonic Diffusers.” Transactions of the ASME. Series D, Vol. 84, 1962, pp. 303—316. 32. Sagi, C. J, and Johnston, J. P. “The Design and Performance of Two-Dimensional, Curved Diffusers.” Transactions of the ASME. Series D, Vol. 89, 1967, pp. 715—731. 33. Bolleter, U. “Blade Passage Tones of Centrifugal Pumps.” Vibrations. Vol. 4, No. 3, Sep. 1988, pp. 8—13. 34. Nykorowytsch, P., Ed. Return Passages of Multi-Stage Turbomachinery. FED—Vol. 3, ASME, June 1983. 35. Hill, P. G., and Peterson, C. R. Mechanics and Thermodynamics of Propulsion. AddisonWesley, 1965, pp. 238–280. 36. Stepanoff, A. J. “Centrifugal Pump Performance as a Function of Specific Speed.” Transactions of the ASME. Aug. 1943, pp. 629—647. 37. Weissgerber, C., and Carter, A. F. “Comparison of Hydraulic Performance Predictions and Test Data for a Range of Pumps.” Performance Prediction of Centrifugal Pumps and Compressors. ASME, Mar. 1980, pp. 219—226. 38. Streeter, V. L., ed. Handbook of Fluid Dynamics. McGraw-Hill, 1961, p. 3–23. 39. Wood, G. M., Welna H., and Lamers, R. P. “Tip—Clearance Effects in Centrifugal Pumps.” Paper No. 64—WA/FE—17, ASME, Nov. 1964. 40. Stepanoff, A. J. “Pumping Solid—Liquid Mixtures.” Paper No. 63—WA—102, ASME, Nov. 1963. 41. Katsanis, T. “Quasi-Three-Dimensional Full Analysis in Turbomachines: A Tool for Blade Design.” In Numerical Simulations in Turbomachinery, FED—Vol. 120, ASME, 1991, pp. 57—64. 42. Katsanis, T. “Fortran Program for Calculating Transonic Velocities on a Blade—to— Blade Stream Surface of a Turbomachine.” NASA TN D—5427, 1969. 43. Spring, H. “Affordable Quasi-Three-Dimensional Inverse Design Method for Pump Impellers.” Proceedings of the Ninth International Pump Users Symposium, Texas A&M University, College Station Texas, Mar. 1992, pp. 97—110. 44. Daily, J. W., and Nece, R. E. “Chamber Dimension Effects on Induced Flow and Frictional Resistance of Enclosed Rotating Disks.” Transactions of the ASME, Series D, Vol. 82, Mar. 1960, pp. 217—232. 45. Nece, R. E., and Daily, J. W. “Roughness Effects on Frictional Resistance of Enclosed Rotating Disks.” Transactions of the ASME. Series D, Vol. 82, Sep. 1960, pp. 553—562. 46. Graf, E. “Analysis of Centrifugal Impeller BEP and Recirculating Flows: Comparison of Quasi—3D and Navier—Stokes Solutions.” Pumping Machinery—1993. FED—Vol. 154, ASME, 1993, pp. 235—245. 2.1 CENTRIFUGAL PUMP THEORY 2.95 47. Gülich, J. F., Favre, J. N., and Denus, K. “An Assessment of Pump Impeller Performance Predictions by 3D—Navier Stokes Calculations.” 1997 ASME Fluids Engineering Division Summer Meeting. Paper FEDSM97—3341, June 1997. 48. Japikse, D., Marscher, W. D., and Furst, R. B. Centrifugal Pump Design and Performance. Concepts, ETI, Inc., Wilder, Vermont, 1997. 49. Denus, K., and others. “A Study in Design and CFD Analysis of a Mixed Flow Pump Impeller.” 1999 ASME Fluids Engineering Division Summer Meeting, Paper FEDSM99—6858, July 1999. 50. Jimbo, H. “Investigation of the Interaction of Windage and Leakage Phenomena in a Centrifugal Compressor.” Paper No. 56—A—47, ASME, Nov. 1956. 51. Keathly, W. C., Due, H. F., and Comolli, C. R. “A Study of Pressure Prediction Methods for Radial Flow Impellers.” Final Report on NASA Contract NAS8—5442, PWA FR— 952, Pratt & Whitney Aircraft, W. Palm Beach, Florida, April 1964. 52. Greitzer, E. M. “The Stability of Pumping Systems.” Transactions of the ASME, Journal of Fluids Engineering, Vol. 103, June, 1981, pp. 193—242. 53. Cooper, P. “Perspective: The New Face of R&D—A Case Study of the Pump Industry.” Transactions of the ASME, Journal of Fluids Engineering, Vol. 118, Dec. 1996, pp. 654—664. 54. Rothe, P. H., and Runstadler, P. W., Jr. “First—Order Pump Surge Behavior.” Paper No. 77—WA/FE—12, ASME, Nov. 1977. 55. Brennen, C., and Acosta, A. J. “The Dynamic Transfer Function for a Cavitating Inducer.” Transactions of the ASME, Journal of Fluids Engineering. Vol. 98, 1976, pp. 182—191. 56. Schlichting, H. Boundary Layer Theory. 4th ed. McGraw-Hill, 1960. 57. “Advanced Class Boiler Feed Pumps.” Proceedings of the Institution of Mechanical Engineers. Vol. 184, Part 3N, 1970. 58. Gopalakrishnan, S. “Pump Research & Development—Past, Present, and Future: An American Perspective.” 1997 ASME Fluids Engineering Division Summer Meeting. Paper FEDSM97—3387, June 1997. 59. Bolleter, U., Frei, A., and Florjancic, D. “Predicting and Improving the Dynamic Behavior of Multistage High Performance Pumps.” Proceedings of the First International Pump Symposium. Texas A&M University, College Station Texas, May 1984, pp. 1—8. 60. Iino, T. “Potential Interaction Between a Centrifugal Pump Impeller and a Vaned Diffuser.” Fluid/Structure Interactions in Turbomachinery. ASME, Nov. 1981, pp. 63—69. 61. Cooper, P. “Hydraulics and Cavitation.” Symposium Proceedings: Power Plant Pumps. M. L. Adams, ed. EPRI CS—5857, June 1988, pp. 4—109 to 4—149. 62. Sutton, G. P. Rocket Propulsion Elements. 6th ed., Wiley, 1992. 63. “Liquid Rocket Engine Centrifugal Flow Turbopumps.” Report SP—8109 of the series entitled NASA Space Vehicle Design Criteria (Chemical Propulsion). NASA, 1973. 64. “Turbopump Systems for Liquid Rocket Engines.” Report SP—8107 of the series entitled NASA Space Vehicle Design Criteria (Chemical Propulsion). NASA, 1975. 65. Makay, E., Cooper, P., Sloteman, D. P., and Gibson, R. “Investigation of Pressure Pulsations Arising from Impeller/Diffuser Interaction in a Large Centrifugal Pump.” Proceedings: Rotating Machinery Conference and Exposition, ASME, Vol. 1, 1993. 66. Fischer, K., and Thoma, D. “Investigation of the Flow Conditions in a Centrifugal Pump.” Transactions of the ASME. Vol. 54, 1932, pp. 143—155. 67. Fraser, W. H. “Recirculation in Centrifugal Pumps.” Materials of Construction of Fluid Machinery and Their Relationship to Design and Performance. ASME, Nov. 1981, pp. 65—86. 2.96 CHAPTER 2 68. Iino, T., Sato, H., and Miyashiro, H. “Hydraulic Axial Thrust in Multistage Centrifugal Pumps.” Transactions of the ASME, Journal of Fluids Engineering. Vol. 102, Mar. 1980, pp. 64—69. 69. Makay, E., and Barrett, J. A. “Changes in Hydraulic Component Geometries Greatly Increased Power Plant Availability and Reduced Maintenance Cost: Case Histories.” Proceedings of the First International Pump Symposium. Texas A&M University, College Station Texas, May 1984, pp. 85—97. 70. Gopalakrishnan, S. “A New Method for Computing Minimum Flow.” Proceedings of the Fifth International Pump Users Symposium. Texas A&M University, College Station Texas, May 1988, pp. 41—47. 71. Heald, C. C., and Palgrave, R. “Backflow Control Improves Pump Performance.” Oil and Gas Journal. Feb. 25, 1985, pp. 96—105. 72. Hallam, J. L. “Centrifugal Pumps: Which Suction—Specific Speeds Are Acceptable.” Hydrocarbon Processing, Apr. 1982. 73. Yedidiah, S. “Performance Curves: Key to Centrifugal Pump Selection.” Machine Design. Apr. 10, 1980, pp. 117—122. 74. Sloteman, D. P., Cooper, P., and Dussourd, J. L. “Control of Backflow at the Inlets of Centrifugal Pumps and Inducers.” Proceedings of the First International Pump Symposium. Texas A&M University, College Station Texas, May 1984, pp. 9—22. 75. Cooper, P., Sloteman, D. P., Graf, E., and Vlaming, D. J. “Elimination of Cavitation— Related Instabilities and Damage in High—Energy Pump Impellers.” Proceedings of the Eighth International Pump Users Symposium. Texas A&M University, 1991, pp. 3—19. 76. Knapp, R. T. “Recent Investigations of the Mechanics of Cavitation and Cavitation Damage.” Transactions of the ASME, Oct. 1955, pp. 1045—1054. 77. Vlaming, D. J. “Optimum Impeller Inlet Geometry for Minimum NPSH Requirements for Centrifugal Pumps.” Pumping Machinery—1989. ASME, July 1989, pp. 25—29. 78. Gülich, J. F. Guidelines for Prevention of Cavitation in Centrifugal Feedpumps. EPRI CS—6398, 1989. 79. Sloteman, D. P., Cooper, P., and Graf, E. “Design of High—Energy Pump Impellers to Avoid Cavitation Instabilities and Damage.” EPRI Power Plant Pumps Symposium, Tampa FL, June 1991. 80. Sloteman, D. P., and others. “Experimental Evaluation of High—Energy Pump Improvements Including Effects of Upstream Piping.” Proceedings of the Twelfth International Pump Users Symposium. Texas A&M University, Mar. 1995, pp. 97—110. SECTION 2.2 CENTRIFUGAL PUMP CONSTRUCTION 2.2.1 CENTRIFUGAL PUMPS: MAJOR COMPONENTS IGOR J. KARASSIK C. C. HEALD CLASSIFICATION AND NOMENCLATURE ________________________________ A centrifugal pump consists of a set of rotating vanes enclosed within a housing or casing that is used to impart energy to a fluid through centrifugal force. Thus, stripped of all refinements, a centrifugal pump has two main parts: (1) a rotating element, including an impeller and a shaft, and (2) a stationary element made up of a casing, casing cover, and bearings. In a centrifugal pump, the liquid is forced by atmospheric or other pressure into a set of rotating vanes. These vanes constitute an impeller that discharges the liquid at its periphery at a higher velocity. This velocity is converted to pressure energy by means of a volute (see Figure 1) or by a set of stationary diffusion vanes (see Figure 2) surrounding the impeller periphery. Pumps with volute casings are generally called volute pumps, while those with diffusion vanes are called diffuser pumps. Diffuser pumps were once quite commonly called turbine pumps, but this term has become more selectively applied to the vertical deep-well centrifugal diffuser pumps usually referred to as vertical turbine pumps. Figure 1 shows the path of the liquid passing through an end-suction volute pump operating at rated capacity (the capacity at which best efficiency is obtained). Impellers are classified according to the major direction of flow in reference to the axis of rotation. Thus, centrifugal pumps may have the following: • Radial-flow impellers (see Figures 25, 34, 35, 36, and 37) • Axial-flow impellers (see Figure 29) • Mixed-flow impellers, which combine radial- and axial-flow principles (see Figures 27 and 28) 2.97 2.98 CHAPTER TWO FIGURE 1 A typical single-stage, end-suction volute pump (Flowserve Corporation) FIGURE 2 A typical diffuser-type pump Impellers are further classified in one of two categories: • Single-suction, with a single inlet on one side (see Figures 25, 33, and 37) • Double-suction, with liquid flowing to the impeller symmetrically from both sides (see Figures 26, 27, and 38) The mechanical construction of the impellers gives a still further subdivision into • Enclosed, with shrouds or side walls enclosing the waterways (see Figures 25, 26, and 27) • Open, with no shrouds (see Figures 29, 33, and 34) • Semiopen or semi-enclosed (see Figure 36) 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.99 A pump in which the head is developed by a single impeller is called a single-stage pump. Often the total head to be developed requires the use of two or more impellers operating in a series, each taking its suction from the discharge of the preceding impeller. For this purpose, two or more single-stage pumps can be connected in a series or all the impellers can be incorporated in a single casing. The unit is then called a multistage pump. The mechanical design of the casing provides the added pump classification of axially split or radially split, and the axis of rotation determines whether the pump is a horizontal or vertical unit. Horizontal-shaft centrifugal pumps are classified still further according to the location of the suction nozzle: • • • • End-suction (see Figures 1, 8, 11, and 13) Side-suction (see Figures 7 and 10) Bottom-suction (see Figure 15) Top-suction (see Figure 22) Some pumps operate in air with the liquid coming to and being conducted away from the pumps by piping. Other pumps, most often vertical types, are submerged in their suction supply. Vertical-shaft pumps are therefore called either dry-pit or wet-pit types. If the wet-pit pumps are axial-flow, mixed-flow, or vertical-turbine types, the liquid is discharged up through the supporting drop or column pipe to a discharge point above or below the supporting floor. These pumps are consequently designated as aboveground discharge or belowground discharge units. Figures 3, 4, and 8 show typical constructions of a horizontal double-suction volute pump, the bowl section of a single-stage axial-flow propeller pump, and a vertical dry-pit single-suction volute pump, respectively. The names recommended by the Hydraulic Institute for various pump parts are given in Table 1. CASINGS AND DIFFUSERS ____________________________________________ The Volute Casing Pump This pump (refer to Figure 1) derives its name from the spiral-shaped casing surrounding the impeller. This casing section collects the liquid discharged by the impeller and converts velocity energy to pressure energy. A centrifugal pump volute increases in area from its initial point until it encompasses the full 360° around the impeller and then flares out to the final discharge opening. The FIGURE 3 Horizontal single-stage double-suction volute pump (the numbers refer to parts listed in Table 1) (Flowserve Corporation) 2.100 CHAPTER TWO TABLE 1 Recommended names of centrifugal pump parts Item no. Name of part Item no. Name of part 1 1A 1B 2 4 6 7 8 9 11 13 14 15 16 17 18 19 20 22 24 25 27 29 31 32 Gasing Gasing (lower half) Gasing (upper half) Impeller Propeller Pump shaft Gasing ring Impeller ring Suction cover Stuffing box cover Packing Shaft sleeve Discharge bowl Bearing (inboard) Gland Bearing (outboard) Frame Shaft sleeve nut Bearing lock nut Impeller nut Suction head ring Stuffing box cover ring Lantern ring (seal cage) Bearing housing (inboard) Impeller key 33 35 36 37 39 40 42 44 46 48 50 52 59 68 72 78 85 89 91 101 103 123 125 127 Bearing housing (outboard) Bearing cover (inboard) Propeller key Bearing cover (outboard) Bearing bushing Deflector Coupling (driver half) Coupling (pump half) Coupling key Coupling bushing Coupling lock nut Coupling pin Handhole cover Shaft collar Thrust collar Bearing spacer Shaft enclosing tube Seal Suction bowl Column pipe Connector bearing Bearing end cover Grease (oil) cup Seal pipe (tubing) a These parts are called out in Figures 3, 4, and 8. wall dividing the initial section and the discharge nozzle portion of the casing is called the tongue of the volute or the “cutwater.” The diffusion vanes and concentric casing of a diffuser pump fulfill the same function as the volute casing in energy conversion. In propeller and other pumps in which axial-flow impellers are used, it is not practical to use a volute casing. Instead, the impeller is enclosed in a pipe-like casing. Generally, diffusion vanes are used following the impeller proper, but in certain extremely low-head units these vanes may be omitted. A diffuser is seldom applied to a single-stage, radial-flow pump, except in special instances where volute passages become so small that machined or precision-cast volute or diffuser-like pieces are utilized for precise flow control. Conventional diffusers are often applied to multistage pump designs in conjunction with guide vanes to direct the flow efficiently from one impeller (stage) to another in a minimum radial and axial space. Diffuser vanes are used as the primary construction method for vertical turbine pumps and singlestage, low-head propeller pumps (see Figure 4). Radial Thrust In a single-volute pump casing design (see Figure 5), uniform or near uniform pressures act on the impeller when the pump operates at design capacity (which coincides with the best efficiency). At other capacities, the pressures around the impeller are not uniform (see Figure 6) and there is a resultant radial reaction (F). A detailed discussion of the radial thrust and of its magnitude is presented in Subsection 2.3.1. Note that the unbalanced radial thrust increases as capacity decreases from that at the design flow. For any percentage of capacity, this radial reaction is a function of total head and of the width and diameter of the impeller. Thus, a high-head pump with a large impeller diameter will have a much greater radial reaction force at partial capacities than a low-head 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 4 Vertical wet-pit diffuser pump bowl (the number refers to parts listed in Table 1) (Flowserve Corporation) 2.101 FIGURE 5 Uniform casing pressures exist at design capacity, resulting in zero radial reaction. pump with a small impeller diameter. A zero radial reaction is not often realized; the minimum reaction occurs at design capacity. Although the same tendency for unbalance exists in the diffuser-type pump, the reaction is limited to a small arc repeated all around the impeller. As a result, the individual reactions cancel each other out as long as flow is constantly removed from around the periphery of the diffuser discharge. If flow is not removed uniformly around its periphery, a pressure imbalance may occur around the diffuser discharge that will be transmitted back through the diffuser to the impeller, resulting in a radial reaction on the shaft and bearing system. In a centrifugal pump design, shaft diameter as well as bearing size can be affected by the allowable deflection as determined by the shaft span, impeller weight, radial reaction forces, and torque to be transmitted. Formerly, standard designs compensated for radial reaction forces encountered at capacities in excess of 50 percent of the design capacity for the maximum-diameter impeller of the pump. For sustained operations at lower capacities, the pump manufacturer, if properly advised, would supply a heavier shaft, usually at a much higher cost. Sustained operations at extremely low flows without the manufacturer being informed at the time of purchase are a much more common practice today. This can result in broken shafts, especially older designs, on high-head units. Because of the increasing application of pumps that must operate at reduced capacities, it has become desirable to design standard units to accommodate such conditions. One solution is to use heavier shafts and bearings. Except for low-head pumps in which only a small additional load is involved, this solution is not economical. The only practical answer is a casing design that develops a much smaller radial reaction force at partial capacities. One of these is the double-volute casing design, also called the twin-volute or dual-volute design. The application of the double-volute design principle to neutralize radial reaction forces at reduced capacity is illustrated in Figure 7. Basically, this design consists of two 180° volutes, and a passage external to the second joins the two into a common discharge. Although a pressure unbalance exists at partial capacity through each 180° arc, the two 2.102 FIGURE 6 At reduced capacities, uniform pressures do not exist in a single-volute casing, resulting in a radial reaction F. CHAPTER TWO FIGURE 7 Transverse view of a double-volute casing pump. forces are approximately equal and opposite. Thus, little if any radial force acts on the shaft and bearings. Subsection 2.3.1 also covers this topic. The double-volute design has many hidden advantages. For example, in large-capacity medium- and high-head single-stage vertical pumps, the rib forming the second volute that separates it from the discharge waterway of the first volute strengthens the casing (see Figure 8). The individual stages of a multistage pump can be made double volute, as illustrated in Figure 9. The kinetic energy of the pumped liquid discharged from the impeller must be transformed into pressure energy and then must be turned 180° to enter the impeller of the next stage. The double volute therefore also acts as a return channel. The back view in Figure 9 shows this as well as the guide vanes used to straighten the flow into the next stage. An alternative to the volute design for multistage pumps is the diffuser and its return vanes that channel the flow from the discharge of the diffuser vanes back into the impeller of the next stage. Solid and Split Casings Solid casing implies a design in which the discharge waterways leading to the discharge nozzle are all contained in one casting or fabricated piece. The casing must have one side open so that the impeller can be introduced into it. Because the sidewalls surrounding the impeller are actually part of the casing, a solid casing, strictly speaking, cannot be used, and designs normally called solid casing are really radially split (refer to Figure 1 and see Figures 11, 12, and 13). A split casing is made of two or more parts fastened together. The term horizontally split had regularly been used to describe pumps with a casing divided by a horizontal plane through the shaft centerline or axis (see Figure 10). The term axially split is now preferred. Because both the suction and discharge nozzles are usually in the same half of the casing, the other half may be removed for inspection of the interior without disturbing the bearings or the piping. Like its counterpart horizontally split, the term vertically split is poor terminology. It refers to a casing split in a plane perpendicular to the axis of rotation. The term radially split is now preferred. End-Suction Pumps Most end-suction, single-stage pumps are made of one-piece solid casings. At least one side of the casing must be open so that the impeller can be assembled in the pump. Thus, a cover is required for that side. If the cover is on the suction side, it becomes the casing sidewall and contains the suction opening (refer to Figure 1). This is called the suction cover or casing suction head. Other designs are made with casing covers (see Figure 12) and still others have both casing suction covers and casing covers (refer to Figure 8 and see Figure 13). 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.103 FIGURE 8 The sectional view of a vertical-shaft, end-suction pump with a double-volute casing (the numbers refer to parts listed in Table 1) (Flowserve Corporation) FIGURE 9 The double volute of a multistage pump, front view (left) and back view (right) (Flowserve Corporation) For general service, the end-suction, single-stage pump design is extensively used for small pumps with a 4- or 6-in (102 or 152 mm) discharge size for both motor-mounted and coupled types. In these pumps, the small size makes it feasible to cast the volute and one side integrally. Whether or not the seal chamber side or the suction side is made integrally with the casing is usually determined by the most economical pump design. For larger pumps, especially those for special services such as sewage handling, there is a demand for pumps of both rotations. A design with separate suction and seal chamber heads permits the use of the same casing for either rotation if the flanges on the two sides 2.104 CHAPTER TWO FIGURE 10 An axially split casing, horizontal-shaft, double-suction volute pump (Flowserve Corporation) FIGURE 11 End-suction pump with semi-open impeller (Flowserve Corporation) are made identical. There is also a demand for vertical pumps that can be disassembled by removing the rotor and bearing assembly from the top of the casing. Many horizontal applications of the pumps of the same line, however, require partial dismantling from the suction side. Such lines are most adaptable when they have separate suction and casing covers. Casing Construction for Open- and Semiopen-Impeller Pumps In the open- or semiopen impeller pump, the impeller rotates within close clearance of the pump casing or suction cover (refer to Figure 11). If the intended service is abrasive, a side plate is mounted within the casing to provide a renewable close-clearance guide to the liquid flow- 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.105 FIGURE 12 End-suction pump with semi-open impeller, inducer and renewable side plate (Flowserve Corporation) FIGURE 13 End-suction pump with removable suction and stuffing box heads (Flowserve Corporation) ing through the impeller (refer to Figure 12). One of the advantages of using side plates is that abrasion-resistant material, such as stainless steel, can be used for the impeller and side plate, while the casing itself may be of a less costly material. Although doublesuction, semiopen-impeller pumps are seldom used today, they were common in the past and were generally made with side plates. In order to maintain pump efficiency, a close running clearance is required between the front unshrouded face of the open or semiopen impeller and the casing, suction cover, or side plate. Pump designs provide either jackscrews or shims to adjust the position of the thrust bearing housing (and, as a result, the axial position of the shaft and impeller) relative to the bearing frame. Pre-rotation and Stop Pieces Improper entrance conditions and inadequate suction approach shapes may cause the liquid column in the suction pipe to spiral for some distance ahead of the impeller entrance. This phenomenon is called pre-rotation, and it is attributed to various operational and design factors in both vertical and horizontal pumps. Pre-rotation is usually harmful to pump operation because the liquid enters between the impeller vanes at an angle other than that allowed in the design. This frequently lowers the net effective suction head and the pump efficiency. Various means are used to avoid pre-rotation both in the construction of the pump and in the design of the suction approaches. 2.106 CHAPTER TWO FIGURE 14 Possible positions of discharge nozzles for a specific design of an end-suction, solid-casing, horizontalshaft pump. The rotation illustrated is counterclockwise from suction end. FIGURE 15 A bottom-suction, axially split casing single-stage pump (Flowserve Corporation) Practically all horizontal, single-stage, double-suction pumps and most multistage pumps have a suction volute that guides the liquid in a streamline flow to the impeller eye. The flow comes to the eye at right angles to the shaft and separates unequally on the two sides of the shaft. Moving from the suction nozzle to the impeller eye, the suction waterways are reduced in area, meeting in a projecting section of the sidewall dividing the two sections. This dividing projection is called a stop piece. To minimize pre-rotation in endsuction pumps, a radial-fin stop piece projecting toward the center is sometimes cast into the suction nozzle wall. Nozzle Locations The discharge nozzle of end-suction, single-stage horizontal pumps is usually in a top-vertical position (refer to Figures 1, 11, and 12). However, other nozzle positions can be obtained, such as top-horizontal, bottom-horizontal, or bottom-vertical. Figure 14 illustrates the flexibility available in discharge nozzle locations. Sometimes the pump frame, bearing bracket, or baseplate may interfere with the discharge flange, prohibiting a bottom-horizontal or bottom-vertical discharge nozzle position. In other instances, solid casings cannot be rotated for various nozzle positions because the seal chamber connection would become inaccessible. Practically all double-suction, axially split casing pumps have a side discharge nozzle and either a side- or a bottom-suction nozzle. If the suction nozzle is placed on the side of the pump casing with its axial centerline (refer to Figure 10), the pump is classified as a side-suction pump. If its suction nozzle points vertically downward (see Figure 15), the pump is called a bottom-suction pump. Single-stage, bottom-suction pumps are rarely made in sizes below a 10-in (254 mm) discharge nozzle diameter. Special nozzle positions can sometimes be provided for double-suction, axially split casing pumps to meet special piping arrangements, such as a radically split casing with bottom suction and top discharge in the same half of the casing. Such special designs are generally costly and should be avoided. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.107 Centrifugal Pump Rotation Because suction and discharge nozzle locations are affected by pump rotation, it is important to understand how the direction of rotation is defined. According to Hydraulic Institute standards, rotation is defined as clockwise or counterclockwise by looking at the driven end of a horizontal pump or looking down on a vertical pump. To avoid misunderstanding, clockwise or counterclockwise rotation should always be qualified by including the direction from which one looks at the pump. The terms inboard end or drive end (the end of the pump closest to the driver) and outboard end or nondrive end (the end of the pump farthest from the driver) are used only with horizontal pumps. The terms lose their significance with dual-driven pumps. Any centrifugal pump casing pattern can be arranged for either clockwise or counterclockwise rotation, except for end-suction pumps, which have integral heads on one side. These require separate directional patterns. Casing Handholes Casing handholes are furnished primarily on pumps handling sewage and stringy materials that may become lodged on the impeller suction vane edges or on the tongue of the volute. The holes permit removal of this material without completely dismantling the pump. End-suction pumps used primarily for liquids of this type are provided with handholes or access to the suction side of the impellers. These access points are located on the suction head or in the suction elbow. Handholes are also provided in drainage, irrigation, circulating, and supply pumps if foreign matter may become lodged in the waterways. On very large pumps, manholes provide access to the interior for both cleaning and inspection. Mechanical Features of Casings Most single-stage centrifugal pumps are intended for service at moderate pressures and temperatures. As a result, pump manufacturers usually design a special line of pumps for high operating pressures and temperatures rather than make their standard line unduly expensive by making it suitable for too wide a range of operating conditions. If axially split casings are subject to high pressure, they tend to “breathe” at the split joint, leading to misalignment of the rotor and, even worse, leakage. For such conditions, internal and external ribbing is applied to casings at the points subject to the greatest stress. In addition, whereas most pumps are supported by feet at the bottom of the casing, high temperatures require centerline support so that, as the pump becomes heated, expansion will not cause misalignment. Series Units For large-capacity medium-high-head conditions, two single-stage, doublesuction pumps can be connected in a series on one baseplate with a single driver. Such an arrangement was at one time very common in waterworks applications for heads of 250 to 400 ft (76 to 122 m). One series arrangement uses a double-extended shaft motor in the middle, driving two pumps connected in a series by external piping. In a second type, a standard motor is used with one pump having a double-extended shaft. This latter arrangement may have limited applications because the shaft of the pump next to the motor must be strong enough to transmit the total pumping horsepower. If the total pressure generated by such a series unit is relatively high, the casing of the second pump could require ribbing. Higher heads per stage are becoming more and more common, and series units are generally used in only very high ranges of total head. MULTISTAGE PUMP CASINGS__________________________________________ Although the majority of single-stage pumps are of the volute-casing type, both volute and diffuser casings are used in multistage pump construction. Because a volute casing gives rise to radial thrust, axially split multistage casings generally have staggered volutes so that the resultant of the individual radial thrusts is balanced out (see Figures 16 and 17). Both axially and radially split casings are used for multistage pumps. The choice between the two designs is dictated by the design pressure, with 2000 lb/in2 (138 bar)° being typical for 3600-rpm axially split casing pumps. °1 bar = 105 Pa 2.108 CHAPTER TWO FIGURE 16 FIGURE 17 An arrangement of a multistage volute pump for radial-thrust balance The horizontal flange of an axially split casing, six-stage pump (Flowserve Corporation) Axially Split Casings Regardless of the arrangement of the stages in the casing, it is necessary to connect the successive stages of a multistage pump. In the low and medium pressure and capacity ranges, these interstage passages are cast integrally with the casing proper (see Figures 18 and 19). As the pressures and capacities increase, the desire to maintain as small a casing diameter as possible, coupled with the necessity of avoiding sudden changes in the velocity or the direction of the flow, leads to the use of external interstage passages cast separately from the pump casing. They are formed in the shape of a loop, bolted or welded to the casing proper (see Figure 20). Interstage Construction for Axially Split Casing Pumps A multistage pump inherently has adjoining chambers subjected to different pressures, and means must be made available to isolate these chambers from one another so that the leakage from high to low pressure will take place only at the clearance joints formed between the stationary and rotating elements of the pump. Thus, the leakage will be kept to a minimum. The isolating wall used to separate two adjacent chambers of a multistage pump is called a stage piece, diaphragm, or inter-stage diaphragm. The stage piece may be formed of a single piece, or it may be fitted with a renewable stage piece bushing at the clearance joint between the stationary stage piece and the part of the rotor immediately inside the former. The stage pieces, which are usually solid, are assembled on the rotor along with impellers, sleeves, bearings, 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.109 FIGURE 18 Two-stage axially split casing volute pump for small capacities and pressures up to 450 lb/in2 (31 bar) (Flowserve Corporation) FIGURE 19 Corporation Two-stage axially split casing volute pump for pressures up to 500 lb/in2 (34 bar) (Flowserve FIGURE 20 A six-stage, axially split casing volute pump for pressures up to 1,300 lb/in2 (90 bar) (Flowserve Corporation) 2.110 CHAPTER TWO and similar components. To prevent the stage pieces from rotating, a locked tongue-andgroove joint is provided in the lower half of the casing. Clamping the upper casing half to the lower half securely holds the stage piece and prevents rotation. The problem of seating a solid stage piece against an axially split casing is one that has given designers much trouble. First, there is a three-way joint and, second, this seating must make the joint tight and leakproof under a pressure differential without resorting to bolting the stage piece directly to the casing. To overcome this problem, it is wise to make a pump that has a small-diameter casing so that when the casing bolting is pulled tight, there is a seal fitting of the two casing halves adjacent to the stage piece. The small diameter likewise helps to eliminate the possibility of a stage piece cocking and thereby leaving a clearance on the upper-half casing when it is pulled down. No matter how rigidly the stage piece is located in the lower-half casing, there must be a sliding fit between the seat face of the stage piece and that in the upper-half casing so that the upper-half casing can be pulled down. Each stage piece, furthermore, must be arranged so that the pressure differential developed by the pump will tend to seat the piece tightly against the casing rather than open up the joint. We have said that axially split casing pumps are typically used for working pressures of up to 2,000 lb/in2 (138 bar). High-pressure piping systems, of which these pumps form a part, are inevitably made of steel because this material has the valuable property of yielding without breaking. Considerable piping strains are unavoidable, and these strains, or at least a part thereof, are transmitted to the pump casing. The latter consists essentially of a barrel that is split axially, flanged at the split, and fitted with two necks that serve as inlet and discharge openings. When piping stresses exist, these necks, being the weakest part of the casing, are in danger of breaking off if they cannot yield. Steel is therefore the safest material for pump casings whenever the working pressures in the pump are in excess of 1,000 lb/in2 (70 bar). This brings up an important feature in the design of the suction and discharge flanges. Although raised-face flanges are perfectly satisfactory for steel-casing pumps, their use is extremely dangerous with cast-iron pumps. This danger arises from the lack of elasticity in cast iron, which leads to flange breakage when the bolts are being tightened, the fulcrum of the bending moment being located inward from the bolt circle. As a result, it is essential to avoid raised-face flanges with cast-iron casings as well as the use of a raisedface flange pipe directly against a flat-face cast-iron pump flange. Suction flanges should obviously be suitable for whatever hydrostatic test pressure is applied to the pump casing. The location of the pump casing feet is not critical in smaller pumps operating at discharge pressures below 275 lb/in2 (19 bar) and at moderate temperatures of up to 300°F (150°C). Since the unit is relatively small, very little distortion is likely to occur. However, for larger units operating at higher pressures and temperatures, it is important that the casing be supported at the horizontal centerline or immediately below the bearings (refer to Figures 19 and 20). Radially Split Double-Casing Pumps The oldest form of radially split casing multistage pump is commonly called the ring-section, ring-casing, or the doughnut type. When more than one stage was found necessary to generate higher pressures, two or more single-stage units of the prevalent radially split casing type were assembled and bolted together. In later designs, the individual stage sections and separate suction and discharge heads were held together with large throughbolts. These pumps, still an assembly of bolted-up sections, can present serious dismantling and reassembly problems because suction and discharge connections have to be broken each time the pump is serviced. The double-casing pump retains the advantages of the radially split casing design and minimizes the dismantling problem. The basic principle consists of enclosing the working parts of a multistage centrifugal pump in an inner casing and building a second casing around this inner casing. The space between the two casings is maintained at the discharge pressure of the last pump stage. The construction of the inner casing follows one of two basic principles: (1) axial splitting (see Figure 21) or (2) radial splitting (see Figure 22). 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 21 FIGURE 22 2.111 Double-casing multistage pump with axially split inner casing (Allis-Chalmers) Double-casing multistage pump with radially split inner casing (Flowserve Corporation) The double-casing pump with radially split inner casing is an evolution of the ringcasing pump with added provisions to ease dismantling. The inner unit is generally constructed exactly as a ring-casing pump. After assembly, it is inserted and bolted inside a cylindrical casing that supports it and leaves it free to expand under temperature changes. In Figure 23, the inner assembly of such a pump is being inserted into the outer casing. Figure 24 shows the external appearance of this type of pump. The suction and discharge nozzles form an integral part of the outer casing, and the internal assembly of the pump can be withdrawn without disturbing the piping connections. Hydrostatic Pressure Tests It is standard practice for the manufacturer to conduct hydrostatic tests for the parts of a pump that contain fluid under pressure. This means that the pump casing and, where applicable, parts like suction or casing covers are assembled with the internal parts, removed, and are then subjected to a hydrostatic test, generally for a minimum of 30 minutes. Such a test demonstrates that the casing containment is sound and that there is no leakage of fluid to the exterior. Additional tests may be conducted by the pump manufacturer to determine the soundness of internal partitions separating areas of the pump operating under different pressures. 2.112 CHAPTER TWO FIGURE 23 Rotor assembly of a radially split, double-casing pump being inserted into its outer casing (Flowserve Corporation) FIGURE 24 Multistage radially split, double-casing pump (Flowserve Corporation) The definition of the applicable hydrostatic pressure for these tests varies. The most generally accepted definition is that given by the Hydraulic Institute Standards. Each part of the pump that contains fluid under pressure shall be capable of withstanding a hydrostatic test at not less than the greatest of the following: • 150 percent of the pressure that will occur in that part when the pump is operated at rated conditions for the given application of the pump, except thermoset parts • 125 percent of the pressure that would occur in that part when the pump is operating at rated speed for a given application, but with the pump discharge valve closed 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.113 IMPELLERS _________________________________________________________ In a single-suction impeller, the liquid enters the suction eye on one side only. A doublesuction impeller is, in effect, two single-suction impellers arranged back to back in a single casing. The liquid enters the impeller simultaneously from both sides, while the two casing suction passageways are connected to a common suction passage and a single suction nozzle. For the general service single-stage, axially split casing design, a double-suction impeller is favored because it is theoretically in an axial hydraulic balance and because the greater suction area of a double-suction impeller permits the pump to operate with less net absolute suction head. For small units, the single-suction impeller is more practical for manufacturing reasons, as the waterways are not divided into two very narrow passages. It is also sometimes preferred for structural reasons. End-suction pumps with single-suction overhung impellers have both first-cost and maintenance advantages unobtainable with double-suction impellers. Most radially split casing pumps therefore use single-suction impellers. Because an overhung impeller does not require the extension of a shaft into the impeller suction eye, single-suction impellers are preferred for pumps handling suspended matter, such as sewage. In multistage pumps, single-suction impellers are almost universally used because of the design and first-cost complexity that double-suction staging introduces. Impellers are called radial vane or radial flow when the liquid pumped is made to discharge radially to the periphery. Impellers of this type usually have a specific speed below 4200 (2600) if single-suction and below 6000 (3700) if double-suction. Specific speed is discussed in detail in Subsection 2.3.1. The units for specific speed used here are in USCS rpm, gallons per minute, and feet. In SI, they are rpm, liters per second, and meters. Impellers can also be classified by the shape and form of their vanes: • • • • The straight-vane impeller (see Figures 25, 34, 35, 36, and 37) The Francis-vane or screw-vane impeller (see Figures 26 and 27) The mixed-flow impeller (see Figure 28) The propeller or axial-flow impeller (see Figure 29) In a straight-vane radial impeller, the vane surfaces are generated by straight lines parallel to the axis of rotation. These are also called single-curvature vanes. The vane surfaces of a Francis-vane radial impeller have a double curvature. An impeller design that has both a radial-flow and an axial-flow component is called a mixed-flow impeller. It is generally restricted to single-suction designs with a specific speed above 4200 (2600). Types with lower specific speeds are called Francis-vane impellers. Mixed-flow impellers with a small radial-flow component are usually referred to as propellers. In a true propeller, or axial-flow impeller, the flow strictly parallels the axis of rotation. In other words, it moves only axially. An inducer is a low-head, axial-flow impeller with few blades that is placed in front of a conventional impeller. The hydraulic characteristics of an inducer are such that it requires considerably less NPSH than a conventional impeller. Both inducer and impeller FIGURE 25 Straight-vane, radial, single-suction closed impeller (Flowserve Corporation) 2.114 CHAPTER TWO FIGURE 26 Francis-vane, radial, double-suction closed impeller (Flowserve Corporation) FIGURE 27 High-specific-speed, Francis-vane, radial, double-suction closed impeller (Flowserve Corporation) FIGURE 28 Open mixed-flow impeller (Flowserve Corporation) FIGURE 29 Axial-flow impeller (Flowserve Corporation) are mounted on the same shaft and rotate at the same speed (see Figure 30). The main purpose of the inducer is not to generate an appreciable portion of the total pump head, but to increase the suction pressure to a conventional impeller. Inducers are therefore used to reduce the NPSH requirements of a given pump or to permit the pump to operate at higher speeds with a given available NPSH. For a further discussion of inducers, see Section 2.1 and Subsection 2.3.1. The relation of single-suction impeller profiles to specific speed is shown in Figure 31. The classification of impellers according to their vane shape is naturally arbitrary inasmuch as there is much overlapping in the types of impellers used in the different types of pumps. For example, impellers in single- and double-suction pumps of low specific speeds have vanes extending across the suction eye. This provides a mixed flow at the impeller entrance for low pickup losses at high rotative speeds but enables the discharge portion of the impeller to use the straight-vane principle. In pumps of higher specific speed operating against low beads, impellers have double-curvature vanes extending over the full vane surface. They are therefore full Francis-type impellers. The mixed-flow impeller, usually a single-suction type, is essentially one-half of a double-suction, high-specific-speed, Francisvane impeller. In addition, many impellers are designed for specific applications. For instance, the conventional impeller design with sharp vane edges and restricted areas is not suitable for handling liquids containing rags, stringy materials, and solids like sewage because it will become clogged. Special nonclogging impellers with blunt edges and large waterways have been developed for such services (see Figure 32). For pumps up to the 12- to 16-in (305- to 406-mm) discharge size, these impellers have only two vanes. Larger pumps normally use three or four vanes. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 30 2.115 A centrifugal pump with a conventional impeller preceded by an inducer (Flowserve Corporation) Another impeller design used for paper pulp pumps (see Figure 33) is fully open and nonclogging; it has screw and radial streamlined vanes. The screw-conveyor end projects far into the suction nozzle, permitting the pump to handle high-consistency paper pulp stock. Impeller Mechanical Types Mechanical design also determines impeller classification. Accordingly, impellers may be completely open, semiopen, or closed. Strictly speaking, an open impeller (see Figures 34 and 35) consists of nothing but vanes attached to a central hub for mounting on the shaft without any form of sidewall or shroud. The disadvantage of this impeller is structural weakness. If the vanes are long, they must be strengthened by ribs or a partial shroud. Generally, open impellers are used in small, low energy pumps. One advantage of open impellers is that they are better suited for handling liquids containing stringy materials. It is also sometimes claimed that they are better suited for handling liquids containing suspended matter because the solids in such matter are more likely to be clogged in the space between the rotating shrouds of a closed impeller and the stationary casing walls. It has been demonstrated, however, that closed impellers do not clog easily, thus disproving the claim for the superiority of the open-impeller design. In addition, the open impeller is much more sensitive to wear than the closed impeller and therefore its efficiency may deteriorate rather rapidly. The open impeller rotates between two side plates, between the casing walls of the volute, or between the casing cover and the suction head. The clearance between the impeller vanes and the sidewalls enables a certain amount of water slippage. This slippage increases as wear increases. To restore the original efficiency, both the impeller and the side plate(s) must be replaced. This, incidentally, involves a much larger expense than would be entailed in closed impeller pumps where simple rings form the leakage joint. The semiopen impeller (see Figure 36) incorporates a single shroud, usually at the back of the impeller. This shroud may or may not have pump-out vanes, which are vanes located at the back of the impeller shroud (see Figure 37). This function reduces the pressure at the back hub of the impeller and prevents foreign matter from lodging in back of the impeller that would interfere with the proper operation of the pump and the seal chamber. The closed impeller (refer to Figures 25 through 27), which is almost universally used in centrifugal pumps handling clear liquids, incorporates shrouds or sidewalls that totally enclose the impeller waterways from the suction eye to the periphery. Although this design prevents the liquid slippage that occurs between an open or semiopen impeller and its side plates, a running joint must be provided between the impeller and the casing to separate the discharge and suction chambers of the pump. This running joint is usually formed by a relatively short cylindrical surface on the impeller shroud that rotates within a slightly larger stationary cylindrical surface. If one or both surfaces are made renewable, the leakage joint can be repaired when wear causes excessive leakage. 2.116 FIGURE 31 Variations in impeller profiles with specific speeds and approximate ranges of specific speeds for the various types. [Universal specific speed s = Ns/2733. Nq (in rpm, m3/s, m) = Ns/51.65]. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 32 Phantom view of a radial-vane nonclogging impeller (Flowserve Corporation) 2.117 FIGURE 33 Paper plup impeller (Flowserve Corporation) FIGURE 34 Open impellers. Notice that the impellers at left and right are strengthened by a partial shroud (Flowserve Corporation). FIGURE 35 An open impeller with partial shroud (Flowserve Corporation) FIGURE 36 Semiopen impeller (Flowserve Corporation) FIGURE 37 The front and back views of an open impeller with a partial shroud and pump-out vanes on the back side (Flowserve Corporation) 2.118 CHAPTER TWO FIGURE 38 Parts of a double-suction impeller If the pump shaft terminates at the impeller so that the latter is supported by bearings on one side, the impeller is called an overhung impeller. This type of construction is the best for end-suction pumps with single-suction impellers. Impeller Nomenclature The inlet of an impeller just before the section where the vanes begin is called the suction eye (see Figure 38). In a closed-impeller pump, the suction eye diameter is taken as the smallest inside diameter of the shroud. In determining the area of the suction eye, the area occupied by the impeller shaft hub is deducted. The hub is the central part of the impeller that is bored out to receive the pump shaft. The term, however, is also frequently used for the part of the impeller that rotates in the casing fit or in the casing wearing ring. It is then referred to as the outer impeller hub or the wearing-ring hub of the shroud. WEARING RINGS ____________________________________________________ Wearing rings provide an easily and economically renewable leakage joint between the impeller and the casing. A leakage joint without renewable parts is illustrated in Figure 39. To restore the original clearances of such a joint after wear occurs, the user must either (1) build up the worn surfaces by welding or metal spraying, or (2) buy new parts. The new parts are not very costly in small pumps, especially if the stationary casing element is a simple suction cover. This is not true for larger pumps or where the stationary element of the leakage joint is part of a complicated casting. If the first cost of a pump is of prime importance, it is more economical to provide for remachining both the stationary parts and the impeller. Renewable casing and impeller rings can then be installed (see Figures 40, 41, and 42). The nomenclature for the casing or stationary part forming the leakage joint surface is as follows: (1) casing ring (if mounted in the casing), (2) casing ring or suction head ring (if mounted in a suction cover or head), and (3) casing cover ring or head ring (if mounted in the casing cover or head). Some engineers like to identify the part 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 39 A plain flat leakage joint with no rings 2.119 FIGURE 40 A single flat casing ring construction. further by adding the word wearing, such as casing wearing ring. A renewable part for the impeller wearing surface is called the impeller ring. Pumps with both stationary and rotating rings are said to have double-ring construction. Wearing Ring Types Various types of wearing ring designs, and the selection of the most desirable type depends on the liquid being handled, the pressure differential across the leakage joint, the rubbing speed, and the particular pump design. In general, centrifugal pump designers use the ring construction that they have found to be most suitable for each particular pump service. The most common ring constructions are the flat type (see Figures 40 and 41) and the L type. The leakage joint in the former is a straight, annular clearance. In the L-type ring (see Figure 43), the axial clearance between the impeller and the casing ring is large, so the velocity of the liquid flowing into the stream entering the suction eye of the impeller is low. The L-type casing rings shown in Figures 43 and 44 have the additional function of guiding the liquid into the impeller eye; they are called nozzle rings. Impeller rings of the L type shown in Figure 44 also furnish protection for the face of the impeller wearing ring hub. Some designers favor labyrinth-type rings (see Figures 45 and 46) that have two or more annular leakage joints connected by relief chambers. In leakage joints involving a single unbroken path, the flow is a function of both the area and the length of the joint as well as of the pressure differential across the joint. If the path is broken by relief chambers (see to Figure 42, 45, and 46), the velocity energy in the jet is dissipated in each relief chamber, increasing the resistance. As a result, with several relief chambers and several leakage joints for the same actual flow through the joint, the area and hence the clearance between the rings can be greater than for an unbroken, shorter leakage joint. The single labyrinth ring with only one relief chamber (refer to Figure 45) is often called an intermeshing ring. The step-ring type (refer to Figure 42) utilizes two flat-ring elements of slightly different diameters over the total leakage joint width with a relief chamber between the two elements. Other ring designs also use some form of relief chamber. For example, one commonly used in small pumps has a flat joint similar to that in Figure 40, but with one surface broken by a number of grooves. These act as relief chambers to dissipate the jet velocity head, thereby increasing the resistance through the joint and decreasing the leakage. For raw water pumps in waterworks service and for larger pumps in sewage services in which the liquid contains sand and grit, water-flushed rings have been used (see Figure 47). Clear water under a pressure greater than that on the discharge side of the rings is piped to the inlet and distributed by the cored passage, the holes through the stationary ring, and the groove to the leakage joint. Ideally, the clear water should fill the leakage joint with some flow to the suction and discharge sides to prevent any sand or grit from getting into the clearance space. Wearing-ring flush is also employed in some process pumps when pumping solids or abrasives to minimize ring wear by injecting a compatible clean liquid between the rings. 2.120 FIGURE 41 CHAPTER TWO A double flat ring construction FIGURE 43 An L-type nozzle casing ring FIGURE 42 A step-type leakage joint with double rings FIGURE 44 Double rings, both of L type In large pumps (with roughly a 36-in [900-mm] or larger discharge size), particularly vertical, end-suction, single-stage volute pumps, size alone permits some refinements not found in smaller pumps. One example is the inclusion of inspection ports for measuring ring clearance (see Figure 48). These ports can be used to check the impeller centering after the original installation as well as to observe ring wear without dismantling the pump. The lower rings of large vertical pumps handling liquids containing sand and grit in intermittent services are highly susceptible to wear. During shutdown periods, the grit and sand settle out and naturally accumulate in the region where these rings are installed, as it is the lowest point on the discharge side of the pump. When the pump is started again, this foreign matter is washed into the joint all at once and causes wear. To prevent this action in medium and large pumps, a dam-type ring is often used (see Figure 49). Periodically, the pocket on the discharge side of the dam can be flushed out. One problem with the simple water-flushed ring is the failure to get uniform pressure in the stationary ring groove. If pump size and design permit, two sets of wearing rings arranged in tandem and separated by a large water space (see Figure 50) provide the best solution. The large water space enables a uniform distribution of the flushing water to the full 360° degrees of each leakage joint. Because ring 2 is shorter and because a greater clearance is used there than at ring 1, equal flow can take place to the discharge pressure side and to the suction pressure side. This design also makes it easier to harden or coat the surfaces. For pumps handling gritty or sandy water, the ring construction should provide an apron on which the stream leaving the leakage joint can impinge, as sand or grit in the jet will erode any surface it hits. Thus, a form of L-type casing ring similar to that shown in Figure 49 should be used. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 45 A single-labyrinth, intermeshing type. It is a double-ring construction with a nozzle-type casing ring. 2.121 FIGURE 46 Labyrinth-type rings in a double-ring construction FIGURE 47 Water-flushed wearing ring FIGURE 48 Wearing ring with an inspection port for checking clearance FIGURE 49 Dam-type ring construction Wearing-Ring Location In some designs, leakage is controlled by an axial clearance (see Figure 51). Generally, this design requires a means of adjusting the shaft position for proper clearance. Then, if uniform wear occurs over the two surfaces, the original clearance can be restored by adjusting the position of the impeller. A limit exists to the amount of wear that can be compensated, because the impeller must be nearly central in the casing waterways. 2.122 CHAPTER TWO FIGURE 50 Two sets of rings with space between for flushing water (Flowserve Corporation) FIGURE 51 A leakage joint with axial clearance Leakage joints with axial clearance are not popular for double-suction pumps because a very close tolerance is required in machining the fit of the rings in reference to the centerline of the volute waterways. Joints with radial clearances, however, enable some shifting of the impeller for centering. The only adverse effect is a slight inequality in the lengths of the leakage paths on the two sides. So far, this discussion has treated only those leakage joints located adjacent to the impeller eye or at the smallest outside shroud diameter. Designs have been made where the leakage joint is at the periphery of the impeller. In a vertical pump, this design is advantageous because the space between the joint and the suction waterways is open and so sand or grit cannot collect. Because of rubbing speed and because the impeller diameters used in the same casing vary over a wide range, the design is impractical in regular pump lines. Mounting Stationary Wearing Rings In small single-suction pumps with suction heads, a stationary wearing ring is usually pressed into a bore in the head and may or may not be further locked by several set screws located half in the head and half in the ring (refer to Figure 41). Larger pumps often use an L-type ring with the flange held against a face on the head. In axially split casing pumps, the cylindrical casing bore (in which the casing ring will be mounted) should be slightly larger than the outside diameter of the ring. Unless some clearance is provided, distortion of the ring may occur when the two casing halves are assembled. However, the joint between the casing ring and the 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.123 casing must be tight enough to prevent leakage. This is usually provided by a radial metalto-metal joint (refer to Figure 43) arranged so that the discharge pressure presses the ring against the casing surface. As it is not desirable for the casing ring of an axially split casing pump to be pinched by the casing, the ring will not be held tightly enough to prevent its rotation unless special provisions are made to keep it in place. One means of accomplishing this is to place a pin in the casing that will project into a hole bored in the ring or, conversely, to provide a pin in the ring that will fit into a hole bored in the casing or into a recess at the casing split joint. Another method is to have a tongue on the casing ring that extends around 180° and engages a corresponding groove in one-half of the casing. This method can be used with casing rings having a central flange by making the diameter of the flange larger for 180° and cutting a deeper groove in that half of the casing. Many methods are used for holding impeller rings on the impeller. Probably the simplest is to rely on a press fit of the ring on the impeller or, if the ring is of proper material, on a shrink fit. Designers do not usually feel that a press fit is sufficient and often add several machine screws or set screws located half in the ring and half in the impeller, as in Figure 41. An alternative to axial machine screws being located half in the ring and half in the impeller is the radial pin, which goes through the center of the ring into the impeller (or from the inside of the impeller eye outward into the ring). This pinning method avoids having to drill and tap holes half in a hardened wearing ring and half in a softer impeller hub. For higher speeds, installing the radial pins from the inside of the impeller eye outward into the wearing ring captures the pins and protects against pin loss, especially at higher operating speeds. Generally, some additional locking method is used rather than relying solely on friction between the ring and the impeller. In the design of impeller rings, consideration has to be given to the stretch of the ring caused by centrifugal force, especially if the pump is of a high-speed design for the capacity involved. For example, some pumps operate at speeds that would cause the rings to become loose if only a press fit is used. For such pumps, shrink fits should be used or, preferably, impeller rings should be eliminated. Wearing-Ring Clearances Typical clearance and tolerance standards for nongalling wearing-joint metals in general service pumps are shown in Figure 52. They apply to the FIGURE 52 Wearing-ring clearances for single-stage pumps using nongalling materials 2.124 CHAPTER TWO following combinations: (1) bronze with a dissimilar bronze, (2) cast-iron with bronze, (3) steel with bronze, (4) Monel metal with bronze, and (5) cast-iron with cast iron. If the metals gall easily (like the chrome steels), the values given should be increased by 0.002 to 0.004 in. The tolerance indicated is positive for the casing ring and negative for the impeller hub or impeller ring. In a single-stage pump with a joint of nongalling components, the correct machining dimension for a casing-ring diameter of 9.000 in would be 9.000 plus 0.003 and minus 0.000 in. For the impeller hub or ring, the values would be 9.000 minus 0.018 (or 8.982) plus 0.000 and minus 0.003 in. Diametral clearances would be between 0.018 and 0.024 in. Obviously, clearances and tolerances are not translated into SI units by merely using conversion multipliers; they are rounded off as they are in the USCS system. The SI values for the previous USCS examples given are outlined here. For a single-stage pump with a casing ring diameter of 230 millimeters, the machining dimensions would be as follows: Casing ring Impeller hub Diametral clearances 230 plus 0.08 and minus 0.00 mm 230 minus 0.50 (229.5) plus 0.00 and minus 0.08 mm Between 0.50 and 0.66 mm Naturally, the manufacturer’s recommendation for ring clearances and tolerances should be followed. AXIAL THRUST ______________________________________________________ Axial Thrust in Single-Stage Pumps with Closed Impellers The pressures generated by a centrifugal pump exert forces on both stationary and rotating parts. The design of these parts balances some of these forces, but separate means may be required to counterbalance others. An axial hydraulic thrust on an impeller is the sum of the unbalanced forces acting in the axial direction. Because reliable, large-capacity thrust bearings are readily available, an axial thrust in single-stage pumps remains a problem only in larger, higher speed units. Theoretically, a double-suction impeller is in hydraulic axial balance, with the pressures on one side equal to and counterbalancing the pressures on the other (see Figure 53). In practice, this balance may not be achieved for the following reasons: 1. The suction passages to the two suction eyes may not provide equal or uniform flows to the two sides. 2. External conditions, such as an elbow located too close to the pump suction nozzle, may cause unequal flow to the two suction eyes. FIGURE 53 The origin of pressures acting on impeller shrouds to produce an axial thrust 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.125 3. The two sides of the discharge casing waterways may not be symmetrical, or the impeller may be located off-center. These conditions will alter the flow characteristics between the impeller shrouds and the casing, causing unequal pressures on the shrouds. 4. Unequal leakage through the two leakage joints can upset the balance. Combined, these factors can create an axial unbalance. To compensate for this, all centrifugal pumps, even those with double-suction impellers, incorporate thrust bearings. The ordinary single-suction, closed, radial-flow impeller with the shaft passing through the impeller eye (refer to Figure 53) is subject to an axial thrust because a portion of the front wall is exposed to suction pressure and thus relatively more backwall surface is exposed to discharge pressure. If the discharge chamber pressure is uniform over the entire impeller surface, the axial force acting toward the suction would be equal to the product of the net pressure generated by the impeller and the unbalanced annular area. Actually, pressure on the two single-suction, closed impeller walls is not uniform. The liquid trapped between the impeller shrouds and casing walls is in rotation, and the pressure at the impeller periphery is appreciably higher than the impeller hub. Although we need not be concerned with the theoretical calculations for this pressure variation, Figure 54 describes it qualitatively. Generally speaking, the axial thrust toward the impeller suction may be about 20 to 30 percent less than the product of the net pressure and the unbalanced area. To eliminate the axial thrust of a single-suction impeller, a pump can be provided with both front and back wearing rings. To equalize the thrust areas, the inner diameter of both rings is made the same (see Figure 55). Pressure approximately equal to the suction pressure is maintained in a chamber located on the impeller side of the back wearing ring by FIGURE 54 Pressure distribution on the front and back shrouds of the single-suction impeller with a shaft through the impeller eye FIGURE 55 Balancing the axial thrust of a single-suction impeller by means of wearing rings on the back side and balancing holes 2.126 CHAPTER TWO FIGURE 56 FIGURE 57 chamber Pump-out vanes used in a single-suction impeller to reduce axial thrust An axial thrust problem with a single-suction, overhung impeller and a single stuffing box or seal drilling balancing holes through the impeller. Leakage past the back wearing ring is returned to the suction area through these holes. However, with large single-stage, singlesuction pumps, balancing holes are considered undesirable because leakage back to the impeller suction opposes the main flow, creating disturbances. In such pumps, a piped connection to the pump suction replaces the balancing holes. Another way to eliminate or reduce an axial thrust in single-suction impellers is to use pump-out vanes on the back shroud. The effect of these vanes is to reduce the pressure acting on the back shroud of the impeller (see Figure 56). This design, however, is generally used only in pumps handling dirty liquids where it keeps the clearance space between the impeller back shroud and the casing free of foreign matter. So far, our discussion of axial thrust has been limited to single-suction, closed impellers with a shaft passing through the impeller eye that are located in pumps with two seal chambers, one on either side of the impeller. In these pumps, suction-pressure magnitude does not affect the resulting axial thrust. On the other hand, axial forces acting on an overhung impeller with a single seal chamber (see Figure 57) are definitely affected by suction pressure. In addition to the unbalanced force found in a single-suction, two-seal chamber design (refer to Figure 54), there is an axial force equivalent to the product of the shaft area through the seal chamber and the difference between suction and atmospheric pressure. This force acts toward the impeller suction when the suction pressure is less than atmospheric or in the opposite direction when it is higher than atmospheric. When an overhung impeller pump handles a suction lift, the additional axial force is very low. For example, if the shaft diameter through the stuffing box is 2 in (50.8 mm) 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.127 (area  3.14 in2 or 20.26 cm2), and if the suction lift is 20 ft (6.1 m) of water equivalent to an absolute pressure of 6.06 lb/in2 (0.42 bar abs), the axial force caused by the overhung impeller and acting toward the suction will be only 27 lb (121 N). On the other hand, if the suction pressure is 100 lb/in2 (6.89 bar), the force will be 314 lb (1405 N) and will act in the opposite direction. Therefore, because the same pump may be used under many conditions over a wide range of suction pressures, the thrust bearing of pumps with single-suction, overhung impellers must be arranged to take the thrust in either direction. They must also be selected with a sufficient thrust capacity to counteract forces set up under the maximum suction pressure established for that particular pump. This extra thrust capacity can become quite significant in certain special cases, such as with boiler circulating pumps. These are usually of the single-suction, single-stage, overhung impeller type and may be exposed to suction pressures as high as 2800 lb/in2 gage (193 bar). This is approximately the vapor pressure of 685°F (363°C) boiler water. If such a pump has a shaft diameter of 6-in (15.24 cm), the unbalanced thrust would be as much as 77,500 lb (346,770 N), and the thrust bearing has to be capable of counteracting this. Axial Thrust of Single-Suction Semiopen Radial-Flow Impellers The axial thrust generated in semiopen impellers is higher than that in closed impellers. This is illustrated in Figure 58, which shows that the pressure on the open side of the impeller varies from essentially the discharge pressure at the periphery (at diameter D2) to the suction pressure at the impeller eye (at diameter D1). The pressure distribution on the back shroud is essentially the same as that illustrated in Figure 54, varying from discharge pressure at the periphery to some portion of this pressure at the impeller hub. This latter pressure is, of course, substantially higher than the suction pressure. The unbalanced portion of the axial thrust on the impeller is represented by the cross-hatched area in Figure 58. One of the means available for partially balancing this increased axial thrust is to provide the back shroud with pump-out vanes, as in Figures 37 and 56. Fully open impellers or semiopen impellers with a portion of the back shroud removed produce an axial thrust somewhat higher than closed impellers and somewhat lower than semiopen impellers. Axial Thrust of Mixed-Flow and AxiaL-Flow Impellers The axial thrust in radial impellers is produced by the static pressures on the impeller shrouds. Axial-flow impellers have no shrouds, and the axial thrust is created strictly by the difference in pressure on the two faces of the impeller vanes. In addition, a difference may exist in pressure acting on the two shaft hub ends, one generally subject to discharge pressure and the other to suction pressure. With mixed-flow impellers, axial thrust is a combination of forces caused by the action of the vanes on the liquid and those arising from the difference in the pressures acting on the various surfaces. Wearing rings are often provided on the back of mixed-flow impellers, with either balancing holes through the impeller hub or an external balancing pipe leading back to the suction. FIGURE 58 Axial thrust in a semiopen single-suction impeller 2.128 CHAPTER TWO FIGURE 59 Multistage pump with single-suction impellers facing in one direction and hydraulic balancing device (Flowserve Corporation) FIGURE 60 A four-stage pump with opposed impellers (Flowserve Corporation) Except for very large units and in certain special applications, the axial thrust developed by mixed-flow and axial-flow impellers is carried by thrust bearings with the necessary load capacity. Axial-Thrust in Multistage Pumps Most multistage pumps are built with single-suction impellers in order to simplify the design of the interstage connections. Two obvious arrangements are possible for the single-suction impellers: 1. Several single-suction impellers can be mounted on one shaft, each having its suction inlet facing in the same direction and its stages following one another in ascending order of pressure (see Figure 59). The axial thrust is then balanced by a hydraulic balancing device. 2. An even number of single-suction impellers can be used, one-half facing in one direction and the other half facing in the opposite direction. With this arrangement, an axial thrust on the first half is compensated by the thrust in the opposite direction on the other half (see Figure 60). This mounting of single-suction impellers back to back is frequently called opposed impellers. An uneven number of single-suction impellers can be used with this arrangement, provided the correct shaft and interstage bushing diameters are used to give the effect of a hydraulic balancing device that will compensate for the hydraulic thrust on one of the stages. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.129 It is important to note that the opposed impeller arrangement completely balances an axial thrust only under the following conditions: • The pump must be provided with two seal chambers. • The shaft must have a constant diameter. • The impeller hubs must not extend through the interstage portion of the casing separating adjacent stages. Except for some special pumps that have an internal and enclosed bearing at one end, and therefore only one seal chamber, most multistage pumps fulfill the first condition. Because of structural requirements, however, the last two conditions are not practical. A slight residual thrust is usually present in multistage opposed impeller pumps and is carried on the thrust bearing. HYDRAULIC BALANCING DEVICES _____________________________________ If all the single-suction impellers of a multistage pump face the same direction, the total theoretical hydraulic axial thrust acting toward the suction end of the pump will be the sum of the individual impeller thrusts. The thrust magnitude will be approximately equal to the product of the net pump pressure and the annular unbalanced area. Actually, the axial thrust turns out to be about 70 to 80 percent of this theoretical value. Some form of hydraulic balancing device must be used to balance this axial thrust and to reduce the pressure on the seal chamber adjacent to the last-stage impeller. This hydraulic balancing device may be a balancing drum, a balancing disk, or a combination of the two. Balancing Drums The balancing drum is illustrated in Figure 61. The balancing chamber at the back of the last-stage impeller is separated from the pump interior by a drum that is usually keyed to the shaft and rotates with it. The drum is separated by a small radial clearance from the stationary portion of the balancing device, called the balancingdrum head, or balancing sleeve, which is fixed to the pump casing. The balancing chamber is connected either to the pump suction or to the vessel from which the pump takes its suction. Thus, the back pressure in the balancing chamber is only slightly higher than the suction pressure, the difference between the two being equal to the friction losses between this chamber and the point of return. The leakage between FIGURE 61 Balancing drum 2.130 CHAPTER TWO the drum and the drum head is, of course, a function of the differential pressure across the drum and of the clearance area. The forces acting on the balancing drum in Figure 61 are the following: • Toward the discharge end: the discharge pressure multiplied by the front balancing area (area B) of the drum • Toward the suction end: the back pressure in the balancing chamber multiplied by the back balancing area (area C) of the drum The first force is greater than the second, thereby counterbalancing the axial thrust exerted upon the single-suction impellers. The drum diameter can be selected to balance the axial thrust completely or within 90 to 95 percent, depending on the desirability of carrying any thrust-bearing loads. It has been assumed in the preceding simplified description that the pressure acting on the impeller walls is constant over their entire surface and that the axial thrust is equal to the product of the total net pressure generated and the unbalanced area. Actually, this pressure varies somewhat in the radial direction because of the centrifugal force exerted upon the liquid by the outer impeller shroud (refer to Figure 54). Furthermore, the pressures at two corresponding points on the opposite impeller faces (D and E in Figure 61) may not be equal because of a variation in clearance between the impeller wall and the casing section separating successive stages. Finally, a pressure distribution over the impeller wall surface may vary with head and capacity operating conditions. This pressure distribution and design data can be determined quite accurately for any one fixed operating condition, and an effective balancing drum could be designed on the basis of the forces resulting from this pressure distribution. Unfortunately, varying head and capacity conditions change the pressure distribution, and as the area of the balancing drum is necessarily fixed, the equilibrium of the axial forces can be destroyed. The objection to this is not primarily the amount of the thrust, but rather that the direction of the thrust cannot be predetermined because of the uncertainty about internal pressures. Still it is advisable to predetermine normal thrust direction, as this can influence external mechanical thrust-bearing design. Because 100 percent balance is unattainable in practice and because the slight but predictable unbalance can be carried on a thrust bearing, the balancing drum is often designed to balance only 90 to 95 percent of the total impeller thrust. The balancing drum satisfactorily balances the axial thrust of single-suction impellers and reduces pressure on the discharge-side stuffing box. It lacks, however, the virtue of automatic compensation for any changes in axial thrust caused by varying impeller reaction characteristics. In effect, if the axial thrust and balancing drum forces become unequal, the rotating element will tend to move in the direction of the greater force. The thrust bearing must then prevent excessive movement of the rotating element. The balancing drum performs no restoring function until such time as the drum force again equals the axial thrust. This automatic compensation is the major feature that differentiates the balancing disk from the balancing drum. Balancing Disks The operation of the simple balancing disk is illustrated in Figure 62. The disk is fixed to and rotates with the shaft. It is separated by a small axial clearance from the balancing disk head, or balancing sleeve, which is fixed to the casing. The leakage through this clearance flows into the balancing chamber and from there either to the pump suction or to the vessel from which the pump takes its suction. The back of the balancing disk is subject to the balancing chamber back pressure, whereas the disk face experiences a range of pressures. These vary from discharge pressure at its smallest diameter to back pressure at its periphery. The inner and outer disk diameters are chosen so that the difference between the total force acting on the disk face and that acting on its back will balance the impeller axial thrust. If the axial thrust of the impellers should exceed the thrust acting on the disk during operation, the latter is moved toward the disk head, reducing the axial clearance between 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 62 2.131 A simple balancing disk the disk and the disk head. The amount of leakage through the clearance is reduced so that the friction losses in the leakage return line are also reduced, lowering the back pressure in the balancing chamber. This lowering of pressure automatically increases the pressure difference acting on the disk and moves it away from the disk head, increasing the clearance. Now the pressure builds up in the balancing chamber, and the disk is again moved toward the disk head until an equilibrium is reached. To assure proper balancing in disk operation, the change in back pressure in the balancing chamber must be of an appreciable magnitude. Thus, with the balancing disk wide open with respect to the disk head, the back pressure must be substantially higher than the suction pressure to give a resultant force that restores the normal disk position. This can be accomplished by introducing a restricting orifice in the leakage return line that increases back pressure when leakage past the disk increases beyond normal. The disadvantage of this arrangement is that the pressure on the seal chamber is variable, a condition that may be injurious to the life of the seal and therefore should avoided. Combination Balancing Disk and Drum For the reasons just described, the simple balancing disk is seldom used. The combination balancing disk and drum (see Figure 63) was developed to obviate the shortcomings of the disk while retaining the advantage of automatic compensation for axial thrust changes. The rotating portion of this balancing device consists of a long cylindrical body that turns within a drum portion of the disk head. This rotating part incorporates a disk similar to the one previously described. In this design, radial clearance remains constant regardless of disk position, whereas the axial clearance varies with the pump rotor position. The following forces act on this device: • Toward the discharge end: the sum of the discharge pressure multiplied by area A, plus the average intermediate pressure multiplied by area B • Toward the suction end: the back pressure multiplied by area C Whereas the position-restoring feature of the simple balancing disk required an undesirably wide variation of the back pressure, it is now possible to depend upon a variation of the intermediate pressure to achieve the same effect. Here is how it works: When the pump rotor moves toward the suction end (to the left in Figure 63) because of increased axial thrust, the axial clearance is reduced and pressure builds up in the intermediate 2.132 CHAPTER TWO FIGURE 63 A combination balancing disk and drum relief chamber, increasing the average value of the intermediate pressure acting on area B. In other words, with reduced leakage, the pressure drop across the radial clearance decreases, increasing the pressure drop across the axial clearance. The increase in intermediate pressure forces the balancing disk toward the discharge end until equilibrium is reached. Movement of the pump rotor toward the discharge end would have the opposite effect, increasing the axial clearance and the leakage and decreasing the intermediate pressure acting on area B. Now numerous hydraulic balancing device modifications are in use. One typical design separates the drum portion of a combination device into two halves, one preceding and the second following the disk (see Figure 64). The virtue of this arrangement is a definite cushioning effect at the intermediate relief chamber, thus avoiding too positive a restoring action, which might result in the contacting and scoring of the disk faces. SHAFTS AND SHAFT SLEEVES ________________________________________ The basic function of a centrifugal pump shaft is to transmit the torques encountered when starting and during operation while supporting the impeller and other rotating parts. It must do this job with a deflection less than the minimum clearance between rotating and stationary parts. The loads involved are (1) the torques, (2) the weight of the parts, and (3) both radial and axial hydraulic forces. In designing a shaft, the maximum allowable deflection, the span or overhang, and the location of the loads all have to be considered, as does the critical speed of the resulting design. Shafts are usually proportioned to withstand the stress set up when a pump is started quickly, such as when the driving motor is energized directly across the line. If the pump handles hot liquids, the shaft is designed to withstand the stress set up when the unit is started cold without any preliminary warmup. Critical Speeds Any object made of an elastic material has a natural period of vibration. When a pump rotor or shaft rotates at any speed corresponding to its natural frequency, minor unbalances will be magnified. These speeds are called the critical speeds. In conventional pump designs, the rotating assembly is theoretically uniform around the shaft axis and the center of mass should coincide with the axis of rotation. This theory does not hold for two reasons. First, minor machining or casting irregularities always occur. Second, variations exist in the metal density of each part. Thus, even in verticalshaft machines having no radial deflection caused by the weight of the parts, this eccentricity of the center of mass produces a centrifugal force and therefore a deflection when 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.133 FIGURE 64 A combination balancing disk and drum with a disk located in the center portion of the drum (Flowserve Corporation) the assembly rotates. At the speed where the centrifugal force exceeds the elastic restoring force, the rotor will vibrate as though it were seriously unbalanced. If it is run at that speed without restraining forces, the deflection will increase until the shaft fails. Rigid and Flexible Shaft Designs The lowest critical speed is called the first critical speed, the next highest is called the second, and so forth. In centrifugal pump nomenclature, a rigid shaft means one with an operating speed lower than its first critical speed. A flexible shaft is one with an operating speed higher than its first critical speed. Once an operating speed has been selected, relative shaft dimensions must still be determined. In other words, it must be decided whether the pump will operate above or below the first critical speed. Actually, the shaft critical speed can be reached and passed without danger because frictional forces tend to restrain the deflection. These forces are exerted by the surrounding liquid, and the various internal leakage joints acting as internal liquid-lubricated bearings. Once the critical speed is passed, the pump will run smoothly again up to the second speed corresponding to the natural rotor frequency, and so on to the third, fourth, and all higher critical speeds. Designs rated for 1,750 rpm (or lower) are usually of the rigid-shaft type. On the other band, high-head 3,600 rpm (or higher) multistage pumps, such as those in a boiler-feed service, are frequently of the flexible-shaft type. It is possible to operate centrifugal pumps above their critical speeds for the following two reasons: (1) very little time is required to attain full speed from rest (the time required to pass through the critical speed must therefore be extremely short) and (2) the pumped liquid in the internal leakage joints acts as a restraining force on the vibration. Experience has proved that, although it was usually assumed necessary to use shafts of such rigidity that the first critical speed is at least 20 percent above the operating speed, equally satisfactorily results can be obtained with lighter shafts with a first critical speed of about 60 to 75 percent of the operating speed. This, it is felt, is a sufficient margin to avoid any danger caused by an operation close to the critical speed. 2.134 CHAPTER TWO Influence of Shaft Deflection To understand the effect of critical speed on the selection of shaft size, consider the fact that the first critical speed of a shaft is linked to its static deflection. Shaft deflection depends upon the weight of the rotating element (w), the shaft span (l), and the shaft diameter (d). The basic formula is as follows: f where f w l C E I wl3 CEI  deflection, in (m)  weight of the rotating element, lb (N)  shaft span, in (m)  coefficient depending on shaft-support method and load distribution  modulus of elasticity of shaft materials, lb/in2 (N/m2)  moment of inertia (p d4/64) in4 (m4) This formula is given in its most simplified form, that is, for a shaft of constant diameter. If the shaft is of varying diameter (the usual situation), deflection calculations are much more complex. A graphical deflection analysis may then be the most practical answer. This formula works only for static deflection, the only variable that affects critical speed calculations. The actual shaft deflection, which must be determined to establish minimum permissible internal clearances, must take into account all transverse hydraulic reactions on the rotor, the weights of the rotating element, and other external loads. It is not necessary to calculate the exact deflection to make a relative shaft comparison. Instead, a factor can be developed that will be representative of relative shaft deflections. As a significant portion of rotor weight is in the shaft, and as methods of bearing support and the modulus of elasticity are common to similar designs, deflection f can be shown as follows: f  function of 1ld2 21l3 2 d4 f  function of l4 d2 In other words, pump deflection varies approximately as the fourth power of the shaft span and inversely as the square of shaft diameter. Therefore, the lower the l4/d2 factor for a given pump, the lower the unsupported shaft deflection, essentially in proportion to this factor. For practical purposes, the first critical speed Nc can be calculated as in USCS units in SI units Nc  Nc  187.7 2f 1in2 946 rpm 2f 1mm2 rpm To maintain internal clearances at the wearing rings, it is usually desirable to limit the shaft deflection under the most adverse conditions to between 0.005 and 0.006 in (0.127 and 0.152 mm). It follows that a shaft design with a deflection of 0.005 to 0.006 in will have a first critical speed of 2,400 to 2,650 rpm. This is the reason for using rigid shafts for pumps that operate at 1,750 rpm or lower. Multistage pumps operating at 3,600 rpm or higher use shafts of equal stiffness (for the same purpose of avoiding wearing ring contact). However, their corresponding critical speed is about 25 to 40 percent less than their operating speed. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.135 Lomakin Effect All the previous material refers to the behavior of a rotor and its shaft operating in air. In reality, the rotor operates immersed in the liquid being pumped, and this liquid flows through one or more of the small annular areas created by clearances separating regions in the pump under different pressures, such as at the wearing rings, interstage bushings, or balancing devices. This flow of liquid creates what is called a hydrodynamic bearing effect and essentially transforms the rotor from one supported at two bearings external to the pump to one with several additional internal bearings lubricated by the liquid pumped. This phenomenon is generally called the Lomakin effect. The result of the Lomakin effect is that the deflection of the shaft when a pump is running is reduced somewhat from the value calculated for the shaft operating in air and the critical speed is increased. The advantage of this effect, particularly in the design of some multistage pumps, is that it permits the use of longer and more slender shafts. Whether this is sound practice remains a controversial subject. The supportive effect of the hydrodynamic hearings depends on (a) the pressure differential, which disappears completely when the pump is at rest, and (b) the clearance, which decreases substantially as the internal clearances increase with erosive or contact wear. Thus, contact between rotating and stationary parts will take place every time a pump is started if the internal clearances are initially less than the shaft deflection in air. This contact will again take place as the pump coasts down after being stopped. Furthermore, as wear takes place at the running joints, the shaft assumes a deflection closer and closer to its deflection in air, unsupported by the Lomakin effect. In view of all these facts, it is recommended that pump users acquaint themselves not only with the calculated shaft deflections with a pump running in new condition, but also with the shaft deflections in air. This way they can compare these with the internal clearances. Shaft Sizing Shaft diameters are usually larger than what is actually needed to transmit the torque. A factor that assures this conservative design is a requirement for ease of rotor assembly. The shaft diameter must be stepped up several times from the end of the coupling to its center to facilitate impeller mounting (see Figure 65). Starting with the maximum diameter at the impeller mounting, there is a step down for the shaft sleeve and another for the external shaft nut, followed by several more for the bearings and the coupling. Therefore, the shaft diameter at the impellers exceeds that required for torsional strength at the coupling by at least an amount sufficient to provide all intervening step downs. One frequent exception to shaft oversizing at the impeller occurs in units consisting of two double-suction, single-stage pumps operating in a series, one of which is fitted with a FIGURE 65 Rotor assembly of a single-stage, double-suction pump (Flowserve Corporation) 2.136 CHAPTER TWO double-extended shaft. As this pump must transmit the total horsepower for the entire series unit, the shaft diameter at its inboard bearing may have to be larger than normal. The shaft design of end-suction, overhung impeller pumps presents a somewhat different problem. One method for reducing shaft deflection at the impeller and seal chamber, where the concentricity of running fits is extremely important, is to considerably increase the shaft diameter between the bearings. Except in certain smaller sizes, centrifugal pump shafts are protected against wear, erosion, and corrosion by renewable shaft sleeves. In small pumps, however, shaft sleeves present a certain disadvantage. As the sleeve cannot appreciably contribute to shaft strength, the shaft itself must be designed for the full maximum stress. Shaft diameter is then materially increased by the addition of the sleeve, as the sleeve thickness cannot be decreased beyond a certain safe minimum. The impeller suction area may therefore become dangerously reduced, and if the eye diameter is increased to maintain a constant eye area, the liquid pickup speed must be increased unfavorably. Other disadvantages accrue from greater hydraulic and seal losses caused by increasing the effective shaft diameter out of proportion to the pump size. To eliminate these shortcomings, very small pumps frequently use shafts of stainless steel or some other material that is sufficiently resistant to corrosion and wear that it does not need shaft sleeves. One such pump is illustrated in Figure 66. Manufacturing costs, of course, are much less for this type of design, and the cost of replacing the shaft is about the same as the cost of new sleeves (including installation). Shaft Sleeves Pump shafts are usually protected from erosion, corrosion, and wear at seal chambers, leakage joints, internal bearings, and in the waterways by renewable sleeves. The most common shaft sleeve function is that of protecting the shaft from wear at packing and mechanical seals. Shaft sleeves serving other functions are given specific names to indicate their purpose. For example, a shaft sleeve used between two multistage pump impellers in conjunction with the interstage bushing to form an interstage leakage joint is called an interstage or distance sleeve. In medium-size centrifugal pumps with two external bearings on opposite sides of the casing (the common double-suction and multistage varieties), the favored shaft sleeve construction uses an external shaft nut to hold the sleeve in an axial position against the impeller hub. Sleeve rotation is prevented by a key, usually an extension of the impeller key (see Figure 67). The axial thrust of the impeller is transmitted through the sleeve to the external shaft nut. In larger high-head pumps, a high axial load on the sleeve is possible and a design similar to that shown in Figure 68 may be preferred. This design has the advantages of simplicity and ease of assembly and maintenance. It also provides space for a large seal chamber FIGURE 66 Section of a small centrifugal pump with no shaft sleeves (Flowserve Corporation) 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 67 FIGURE 68 2.137 A sleeve with external locknut and impeller key extending into the sleeve to prevent rotation A sleeve with an internal impeller nut, external shaft-sleeve nut, and a separate key for the sleeve FIGURE 69 A sleeve threaded onto a shaft with no external locknut and cartridge-type mechanical seals. When shaft sleeve nuts are used to retain the sleeves and impellers axially, they are usually manufactured with right- and left-hand threads. The friction of the pumpage and inadvertent contact with stationary parts or bushings will tend to tighten the nuts against the sleeve and impeller hub (rather than loosen them). Usually, the shaft sleeves utilize extended impeller keys to prevent rotation. Some manufacturers favor the sleeve shown in Figure 69, in which the impeller end of the sleeve is threaded and screwed to a matching thread on the shaft. A key cannot 2.138 CHAPTER TWO be used with this type of sleeve, and right- and left-hand threads are substituted so that the frictional grip of the packing on the sleeve will tighten it against the impeller hub. As a safety precaution, the external shaft nuts and the sleeve itself use set screws for a locking device. In pumps with overhung impellers, various types of sleeves are used. Most pumps use mechanical seals, and the shaft sleeve is usually a part of the mechanical seal package supplied by the seal manufacturer. Many mechanical seals are of the cartridge design, which is set and may be bench-tested for leakage prior to installation in the pump. (For a further discussion of mechanical seals, see Subsection 2.2.3.) For overhung impeller pumps that utilize packing for sealing, the packing sleeves generally extend from the impeller hub through the seal chambers (or stuffing boxes) to protect the pump shaft from wear (see Figure 70). The sleeves are usually keyed to the shaft to prevent rotation. If a hook-type sleeve is used, the hook part of the sleeve is clamped between the impeller and a shaft shoulder to maintain the axial position of the sleeve. A hook-type sleeve used to be popular for overhung impeller pumps that operate at high temperatures because it is clamped at the impeller end and the rest of the sleeve is free to expand axially with temperature changes. But with the increased use of cartridge-type seals, the use of hook-type sleeves is diminishing. In designs with a metal-to-metal joint between the sleeve and the impeller hub or shaft nut, a sealing device is required between the sleeve and the shaft to prevent leakage. Pumped liquid can leak into the clearance between the shaft and the sleeve when operating under a positive suction head and air can leak into the pump when operating under a negative suction head. This seal can be accomplished by means of an O-ring, as shown in Figure 71, or a flat gasket. For high temperature services, the sealing device must be either acceptable for the temperature to which it will be exposed, or it must be located outside the high temperature liquid environment. An alternative design used for some hightemperature process pumps is shown in Figure 72. In this arrangement, the contact surface of the hook-type sleeve and the shaft is ground at a 45-degree angle to form a metal-to-metal seal. That end of the sleeve is locked, but the other is free to expand with temperature changes. When O-rings are used, any sealing surfaces must be properly finished to ensure a positive seal is achieved. All bores and changes in diameter over which O-rings must be passed should be properly radiused and chamfered to protect against damage during assembly. Guidelines for assembly dimensions and surface finish criteria are listed in Oring manufacturers’ catalogs. Material for Packing Sleeves Packing sleeves are surrounded in the stuffing box by packing. The sleeve must be smooth so that it can turn without generating too much friction and heat. Thus, the sleeve materials must be capable of taking a very fine finish, preferably a polish. Cast-iron is therefore not suitable. A hard bronze is generally used for pumps handling clear water, but chrome or other stainless steels are sometimes pre- FIGURE 70 A sleeve for pumps with an overhung impeller 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 71 2.139 A seal arrangement for the shaft sleeve to prevent leakage along the shaft CONTACTING SURFACE FIGURE 72 A sleeve with a 45º bevel contacting surface ferred. For pumps subject to abrasives, hardened chrome or other stainless steels give good results. In most applications, a hardened chromium steel sleeve will be technically adequate, and the most economical choice. For severe or unusual conditions, coated sleeves are used. Ceramic coatings, applied using a plasma spray process, have also been used. Chromium oxide and aluminum oxide are the most common ceramic coatings. Both are extremely hard and resist abrasive wear well. Ceramic coatings have been replaced in some applications by tungsten carbide coatings applied using a high-velocity oxyfuel (HVOF) process. The superior impact resistance and bond strength of these coatings is well documented. Another coating that is widely used on pump sleeves is a nickel-chromium-silicon-boron self-fluxing coating. This coating has a good resistance to galling and moderate resistance to abrasive wear. Sleeves intended for coating should be machined with an undercut, so that the coating does not extend to the edge of the sleeve. This will prevent chipping at the edge, especially with the more brittle ceramic coatings. SEAL CHAMBERS AND STUFFING BOXES _______________________________ Seal chambers have the primary function of protecting the pump against leakage at the point where the shaft passes out through the pump pressure casing. If the pump handles a suction lift and the pressure at the bottom of the seal chamber (the point closest to the inside of the pump) is below atmospheric, the seal chamber function is to prevent air leakage into the pump. If this pressure is above atmospheric, the function is to prevent liquid leakage out of the pump. 2.140 CHAPTER TWO FIGURE 73 FIGURE 74 A conventional stuffing box with throat bushing A conventional stuffing box with a bottoming ring FIGURE 75 A lantern ring (also called a seal cage) When sealing is accomplished by means of a mechanical seal, the seals are installed in a seal chamber. When sealing is accomplished by means of packing, the seal chamber is commonly referred to as a stuffing box. For general service pumps, a stuffing box usually takes the form of a cylindrical recess that accommodates a number of rings of packing around the shaft or shaft sleeve (see Figures 73 and 74). If sealing the box is desired, a lantern ring or seal cage (see Figure 75) is used to separate the rings of packing into approximately equal sections. The packing is compressed to give the desired fit on the 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.141 shaft or sleeve by a gland that can be adjusted in an axial direction. The bottom or inside end of the box can be formed by the pump casing (refer to Figure 70), a throat bushing (see Figure 78), or a bottoming ring (see Figure 74). For manufacturing reasons, throat bushings are widely used on smaller pumps with axially split casings. Throat bushings are always solid rather than split. The bushing is usually held from rotation by a tongue-and-groove joint locked in the lower half of the casing. Packing Lantern Rings (Seal Cages) When a pump operates with negative suction head, the inner end of the stuffing box is under vacuum and air tends to leak into the pump. For this type of service, packing is usually separated into two sections by a lantern ring or seal cage (refer to Figure 73). Water or some other sealing fluid is introduced under pressure into the lantern ring connection, causing a flow of sealing fluid in both axial directions. This construction is useful for pumps handling chemically active or dangerous liquids since it prevents an outflow of the pumped liquid. Lantern rings are usually axially split for ease of assembly. Some installations involve variable suction conditions, the pump operating part of the time with suction head and part of the time with suction lift. When the operating pressure inside the pump exceeds atmospheric pressure, the liquid lantern ring becomes inoperative (except for lubrication). However, it is maintained in services so that when the pump is primed at starting, all air can be excluded. Sealing Liquid Arrangements When a pump handles clean, cool water, sealing liquid connections are usually to the pump discharge or, in multistage pumps, to an intermediate stage. An independent supply of sealing water should be provided if any of the following conditions exist: • A suction lift in excess of 15 ft (4.5 m) • A discharge pressure under 10 lb/in2 (0.7 bar) • Hot water (over 250°F or l20°C) being handled without adequate cooling (except for boiler-feed pumps, in which lantern rings are not used) • Muddy, sandy, or gritty water being handled • The pump is a hot-well pump. • The liquid being handled is other than water, such as acid, juice, molasses, or sticky liquids, without special provision in the stuffing box design for the nature of the liquid If the suction lift exceeds 15 ft (4.5 m), excessive air infiltration through the stuffing boxes may make priming difficult unless an independent seal is provided. A discharge pressure under 10 lb/in2 (0.7 bar) may not provide sufficient sealing pressure. Hot-well (or condensate) pumps operate with as much as a 28-in Hg (710 mm Hg) vacuum, and air infiltration would take place when the pumps are standing idle in standby service. When sealing water is taken from the pump discharge, an external connection may be made through small-diameter piping (see Figure 76) or internal passages. In some pumps, these connections are arranged so that a sealing liquid can be introduced into the packing space through an internal drilled passage either from the pump casing or from an external source (see Figure 77). When the liquid pumped is used for sealing, the external connection is plugged. If an external sealing liquid source is required, it is connected to the external pipe tap with a socket-head pipe plug inserted at the internal pipe tap. It is sometimes desirable to locate the lantern ring with more packing on one side than on the other. For example, in gritty-water services, a lantern ring location closer to the inner portion of the pump would divert a greater proportion of sealing liquid into the pump, thereby keeping grit from working into the box. An arrangement with most of the packing rings between the lantern ring and the inner end of the stuffing box would be applied to reduce dilution of the pumped liquid. Some pumps handle water in which there are small, even microscopic, solids. Using water of this kind as a sealing liquid introduces the solids into the leakage path, shortening the life of the packing and sleeves. It is sometimes possible to remove these solids 2.142 CHAPTER TWO FIGURE 76 Piping connections from the pump discharge to seal cages FIGURE 77 An end-suction pump with provisions for an internal or external sealing-liquid supply (Flowserve Corporation) by installing small pressure filters in the sealing water piping from the casing to the stuffing box. Filters ultimately get clogged, though, unless they are frequently backwashed or otherwise cleaned out. This disadvantage can be overcome by using a cyclone (or centrifugal) separator. The operating principle of the cyclone separator is based on the fact that if a liquid under pressure is introduced tangentially into a vortexing chamber, a centrifugal force will make it rotate in the chamber, creating a vortex. Particles heavier than the liquid in which they are carried will hug the outside wall of the vortexing chamber and the liquid in the center of the chamber will be relatively free of foreign matter. The action of such a separator is illustrated in Figure 78. Liquid piped from the pump discharge or from an intermediate stage of a multistage pump is piped to inlet tap A, which is drilled tangentially to the cyclone bore. The liquid containing solids is directed downward to the apex of the cone at outlet tap B and is piped to the suction or to a low-pressure point in the system. The cleaned liquid is taken off at the center of the cyclone at outlet tap C and is piped to the stuffing box. Sand that will pass through a No. 40 sieve will be completely eliminated in a cyclone separator with supply pressures as low as 20 lb/in2 (1.4 bar). With l00 lb/in2 (7-bar) supply pressure, 95 percent of the particles of 5-micron size will be eliminated. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 78 2.143 An illustration of the priniciple of cyclone separators (Flowserve Corp. ) FIGURE 79 Backflow preventer (Hersey Products) Most city ordinances require that some form of backflow preventer be interposed between city water supply lines and connections to equipment where backflow or siphoning could contaminate a drinking water supply. This is the case, for instance, with an independent sealing supply used for stuffing boxes of sewage pumps. Quite a variety of backflow preventers are available. In most cases, the device consists of two spring-loaded check valves in a series and a spring-loaded, diaphragm-actuated, differential-pressure relief valve located in the zone between the check valves (see Figure 79). In a normal operation, both check valves remain open as long as there is a demand for sealing water. The differential-pressure relief valve remains closed because of the pressure drop past the first check valve. If the pressure downstream of the device increases, tending to reverse the direction of the flow, both check valves close and prevent backflow. If the second check valve is prevented from closing tightly, the leakage past it increases the pressure between the two check valves, the relief valve opens, and water is discharged to the 2.144 CHAPTER TWO FIGURE 80 FIGURE 81 A water seal unit A weighted grease sealer atmosphere. Thus, the relief valve operates automatically to maintain the pressure between the two check valves lower than the supply pressure. Some local ordinances prohibit any connections between city water lines and a sewage or process liquid line. In such cases, an open tank under atmospheric pressure is installed into which city water can be admitted and from which a small pump can deliver the required quantity of sealing water. Such a water-sealing supply unit (see Figure 80) can be installed in a location where it can serve a number of pumps. The tank is equipped with a float valve to feed and regulate the water level so that contamination of the city water supply is prevented. A small close-coupled pump is mounted directly on the tank and maintains a constant pressure of clear water at the stuffing box seals of the battery of pumps it serves. A small recirculation line is provided from the closecoupled pump discharge back to the tank to prevent operation at shutoff. The discharge pressure of the small supply pump is set by the maximum sealing pressure required at any of the pumps served. The supply at the individual stuffing boxes is then regulated by setting small control valves in each individual line. If clean, cool water is not available (as with some drainage, irrigation, or sewage pumps), grease or oil seals are often used. Most pumps for sewage service have a single stuffing box subject to discharge pressure that operates with a flooded suction. It is therefore not necessary to seal these pumps against air leakage, but forcing grease or oil into the sealing space at the packing helps to exclude grit. Figure 81 shows a typical weighted grease sealer. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 82 2.145 An automatic grease sealer mounted on a vertical pump (Zimmer & Francescon) Automatic grease or oil sealers that exert pump discharge pressure in a cylinder on one side of a plunger, with light grease or oil on the other side, are available for sewage service. The oil or grease line is connected to the stuffing box seal, which is at about 80 percent of the discharge pressure. As a result, there is a slow flow of grease or oil into the pump when the unit is in operation. No flow takes place when the pump is out of service. Figure 82 shows an automatic grease sealer mounted on a vertical sewage pump. Water-Cooled Stuffing Boxes High temperatures or pressures complicate the problem of maintaining stuffing box packing. Pumps in these more difficult services are usually provided with mechanical seals and seal support systems. When it is necessary or desirable to use packing, however, the pumps are usually equipped with jacketed, water-cooled stuffing boxes. The cooling water removes heat from the liquid leaking through the stuffing box and heat generated by friction in the box, thus improving packing service conditions. In some special cases, liquid other than water can be used in the cooling jackets. Two water-cooled stuffing box designs are commonly used. The first, shown in Figure 83, provides cored passes in the casing casting. These passages that surround the stuffing box are arranged with inand-out connections. The second type uses a separate cooling chamber combined with the stuffing box proper, with the whole assembly inserted into and bolted to the pump casing (see Figure 84). The choice between the two is based on manufacturing preferences. Caution is required when depending on water cooling to provide proper operation because of the danger of passage fouling and a loss of cooling effectiveness during operation. It is important that any such cooling passages be accessible for periodic inspection and cleaning to ensure that effective cooling is maintained. Stuffing box pressure and temperature limitations vary with the pump type because it is generally not economical to use expensive stuffing box construction for infrequent hightemperature or high-pressure applications. Therefore, whenever the manufacturer’s stuffing box limitations for a given pump are exceeded, the application of pressure-reducing devices ahead of the stuffing box is recommended. Pressure-Reducing Devices Essentially, pressure-reducing devices consist of a bushing or meshing labyrinth ending in a relief chamber located between the pump interior and the stuffing box or seal chamber. The relief chamber is connected to some suitable 2.146 CHAPTER TWO FIGURE 83 FIGURE 84 A water-cooled stuffing box with a cored water passage cast in casing A separate water-cooled stuffing box with pressure-reducing stuffing box bushing low-pressure point in the installation, and the leakage past the pressure-reducing device is returned to this point. If the pumped liquid must be salvaged, as with treated feedwater, it is returned to the pumping cycle. If the liquid is expendable, the relief chamber can be connected to a drain. Many different pressure-reducing device designs exist. Figure 84 illustrates a design for limited pressures. A short serrated bushing is inserted at the bottom of the stuffing box or seal chamber, followed by a relief chamber. The leakage past the serrated bushing is bled off to a low-pressure point. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 85 2.147 Split stuffing box gland With relatively high-pressure units, intermeshing labyrinths can be located following the balancing device and ahead of the stuffing box or seal chamber. Piping from the chamber following pressure-reducing devices should be amply sized so that as wear increases leakage, piping friction will not increase seal chamber pressure. Stuffing Box Packing Glands Stuffing box packing glands may assume several forms, but basically they can be classified into two groups: solid glands and split glands (see Figure 85). Split glands are made in halves so that they can be removed from the shaft without dismantling the pump, thus providing more working space when the stuffing boxes are being repacked. Split glands are desirable for pumps that have to be repacked frequently, especially if the space between the box and the bearing is restricted. The two halves are generally held together by bolts, although other methods are also used. Split glands are generally a construction refinement rather than a necessity, and they are rarely used in smaller pumps. They are commonly furnished for large single-stage pumps, for some multistage pumps, and for certain refinery pumps. Another common refinement is the use of swing bolts in stuffing box packing glands. Such bolts may be swung to the side, out of the way, when the stuffing box is being repacked. Stuffing box leakage into the atmosphere might, in some services, seriously inconvenience or even endanger the operating personnel. An example would be when volatile liquids are being pumped at vaporizing temperatures or temperatures above their flash point. As this leakage cannot always be cooled sufficiently by a water-cooled stuffing box, smothering glands are used (refer to Figure 83). Provisions are made in the gland to introduce a liquid, either water or another compatible liquid at a low temperature, that mixes intimately with the leakage, lowering its temperature or, if the liquid is volatile, absorbing it. Stuffing box packing glands are usually made of bronze, although cast-iron or steel may be used for all iron-fitted pumps. Iron or steel glands are generally bushed with a nonsparking material like bronze in hazardous process services to prevent the ignition of flammable vapors by the glands sparking against a ferrous metal shaft or sleeve. Stuffing Box Packing See Subsection 2.2.2. MECHANICAL SEALS VERSUS PACKING ________________________________ Sealing with a packed stuffing box is impractical for many conditions of service. In an ordinary packed stuffing box, the sealing between the rotating shaft or shaft sleeve and the stationary portion of the stuffing box (or seal chamber) is accomplished by means of rings of packing forced between the two surfaces and held tightly in place by a packing gland. The leakage around the shaft is controlled by merely tightening or loosening the packing gland nuts (or bolts). The actual sealing surfaces consist of the axial rotating surface of the shaft or shaft sleeve and the stationary packing. Attempts to reduce or eliminate all leakage from a conventional packed stuffing box increase the compressive load of the packing 2.148 CHAPTER TWO gland on the packing. The packing, being semiplastic, forms more closely to the shaft or shaft sleeve and tends to reduce the leakage. After a certain point, however, the leakage between the packing and the rotating shaft or shaft sleeve becomes inadequate to carry away the heat generated by the packing rubbing on the rotating surface, and the packing fails to function. This failure can result in burned packing, packing “blow out,” and severely damaged shaft or shaft sleeve surfaces. If the sealing surface is coated, this coating may be destroyed. Even before this condition is reached, the shaft or shaft sleeve may be severely worn and scored by the packing, so that it becomes impossible to pack the stuffing box satisfactorily. These undesirable characteristics prohibit the use of packing as a sealing method if some leakage of the pumpage to the atmosphere is not acceptable. Packing is limited in its application pressure and temperature range (see Section 2.2.2), and it is usually not acceptable for any flammable or hazardous pumping services. To address these limitations, the mechanical seal was developed (see Section 2.2.3). The mechanical seal has found general acceptance in nearly all pumping applications. Packing is still used in certain low-pressure, low-temperature applications where leakage of the pumpage is not a problem and a history of satisfactory, economical service exists. Mechanical seals are not always the solution to every sealing situation. Seals are still subject to failure, and their failure may be more rapid and abrupt than that of packing. If packing fails, the pump can many times be kept running by temporary adjustments until it is convenient to shut it down. If a mechanical seal fails, most often the pump must be shut down immediately. As both packed stuffing boxes and conventional mechanical face seals are subject to wear, both are subject to failure. Whether one or the other should be used depends on the specific application and the experience of the user. In some cases, both give good service and the choice becomes a matter of personal preference or cost. Table 2 TABLE 2 Comparison of packing and mechanical seals Advantages Disadvantages Packing 1. Lower initial cost 2. Easily installed as rings and glands are split 3. Good reliability to medium pressures and shaft speeds 4 Can handle large axial movements (thermal expansion of stuffing box versus shaft) 5. Can be used in rotating or reciprocating applications 6. Leakage tends to increase gradually, giving warning of impending breakdown 1. Relatively high leakage 2. Requires regular maintenance 3. Wear of shaft of shaft sleeve can be relatively high 4. Power losses may be high Mechanical seals 1. Very low leakage/no leakage 2. Require no maintenance 3. 4. 5. 6. Eliminate sleeve wear/shaft wear Very good reliability Can handle higher pressures and speeds Easily applied to carcinogenic, toxic, flammable, or radioactive liquids 7. Low power loss Source: John Crane Inc. 1. Higher initial cost 2. Easily installed but may require some disassembly of pump (couplings and so on) 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.149 adds other comments and summarizes some advantages and disadvantages of packing and mechanical seals. Principles and Construction of Mechanical Seals See Subsection 2.2.3. Injection-Type Shaft Seals See Subsection 2.2.4. BEARINGS __________________________________________________________ The function of bearings in centrifugal pumps is to keep the shaft or rotor in correct alignment with the stationary parts under the action of radial and transverse loads. Bearings that give radial positioning to the rotor are known as radial or line bearings, and those that locate the rotor axially are called thrust bearings. In most applications, the thrust bearings actually serve both as thrust and radial bearings. Types of Bearings Used All types of bearings have been used in centrifugal pumps. Even the same basic design of pump is often made with two or more different bearings, required either by varying service conditions or by the preference of the purchaser. In most pumps, however, either rolling element or oil film (sleeve-type) bearings are used today. In horizontal pumps with bearings on each end, the bearings are usually designated by their location as inboard, or drive end, and outboard, or non-drive end. Inboard (drive end) bearings are located between the casing and the coupling. Pumps with overhung impellers have both bearings on the same side of the casing so that the bearing nearest the impeller is called inboard and the one farthest away outboard. In a pump provided with bearings at both ends, the thrust bearing is usually placed at the outboard end and the line bearing at the inboard end. The bearings are mounted in a housing that is usually supported by brackets attached or integral to the pump casing. The housing also serves the function of containing the lubricant necessary for proper operation of the bearing. Occasionally, the bearings of very large pumps are supported in housings that form the top of pedestals mounted on soleplates or on the pump bedplate. These are called pedestal bearings. Because of the heat generated by the bearing or the heat in the liquid being pumped, some means other than radiation to the surrounding air must occasionally be used to keep the bearing temperature within proper limits. If the bearings have a force-fed lubrication system, cooling is usually accomplished by circulating the oil through a separate water-tooil or air-to-oil cooler. Otherwise, a jacket through which a cooling liquid is circulated is usually incorporated as part of the housing. Pump bearings may be rigid or self-aligning. A self-aligning bearing will automatically adjust itself to a change in the angular position of the shaft. In babbitted or sleeve bearings, the name self-aligning is applied to bearings that have a spherical fit of the sleeve in the housing. In rolling element bearings, the name is applied to bearings, the outer race of which is spherically ground or the housing of which provides a spherical fit. Although double-suction pumps are theoretically in hydraulic balance, this balance is rarely realized in practice, and so even these pumps are provided with thrust bearings. A centrifugal pump, being a product of the foundry, is subject to minor irregularities that may cause differences in the eddy currents set up on the two sides of the impeller. As this disturbance can create an axial hydraulic thrust, some form of thrust bearing that is capable of taking a thrust in either direction is necessary to maintain the rotor in its proper position. The thrust capacity of the bearing of a double-suction pump is usually far in excess of the probable imbalance caused by irregularities. This provision is made because (1) unequal wear of the rings and other parts may cause an imbalance and (2) the flow of the liquid into the two suction eyes may be unequal and cause an imbalance because of an improper suction-piping arrangement. Rolling Element Bearings The most common rolling element bearings used on centrifugal pumps are the various types of ball bearings. Roller bearings are used less often, although the spherical roller bearing (see Figure 86) is used frequently for large shaft 2.150 CHAPTER TWO FIGURE 86 Self-aligning spherical roller bearing (SKF USA, Inc.) sizes, for which there is a limited choice of ball bearings. As most roller bearings are suitable only for radial loads, their use on centrifugal pumps tends to be limited to applications in which they are not required to carry a combined radial and thrust load. Ball Bearings As the coefficient of rolling friction is less than that of sliding friction, one must not consider a ball bearing in the same light as a sleeve bearing. In the former, the load is carried on a point contact of the ball with the race, but the point of contact does not rub or slide over the race and no appreciable heat is generated. Furthermore, the point of contact is constantly changing as the ball rolls in the race, and the operation is practically frictionless. In the sleeve bearing, a constant rubbing of one surface over another occurs, and the friction must be reduced by the use of a lubricant. Ball bearings that operate at an absolutely constant speed theoretically require no lubricant. No speed can be called absolutely constant, however, for the conditions affecting the speed always vary slightly. For instance, a motor with a full-load speed rated at 3,510 rpm might vary in speed over the course of a minute from 3,505 to 3,515 rpm. Each variation in speed causes the balls in a ball bearing to lag or lead the race because of their inertia. Consequently, a very slight, almost immeasurable sliding action takes place. Another limiting condition is that the hardest of metals suffers minute deformations on carrying loads, thus upsetting perfect point contacts and adding another slight sliding action. For these reasons, ball bearings must be given some lubrication. Ball thrust bearings are built to carry heavy loads by pure rolling motion on an angular contact. As a thrust load is axial, it is equally distributed to all the balls around the race, and the individual load on each ball is only a small fraction of the total thrust load. In such bearings, it is essential that the balls be equally spaced, and for this purpose, a retaining cage is used between the balls and between the inner and outer races. This cage carries no load, but the contact between it and the ball produces sliding friction that requires lubrication. Types and Applications Pump designers have a wide variety of rolling element bearings and arrangements to choose from. Ball bearings with their high-speed capabilities and low friction make them ideal for small and medium-size pumps, while roller bearings are more common in larger, slower speed pumps where a heavy capacity is required. Depending upon the specific bearing type, optional characteristics such as seals, shields, various cage materials and designs, and special internal clearances and preloads are available. Although several might be dimensionally acceptable, it is best for users to adhere to manufacturer recommendations to ensure optimum reliability. The most common ball bearings used in centrifugal pumps are 1) single-row, deepgroove, 2) single-row, angular contact, and 3) double-row, angular contact ball bearings. Sealed ball bearings are used in special applications such as vertical in-line pumps. Sealed prelubricated bearings require special attention if the unit in which they are 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 87 Inc.) Self-aligning ball bearing (SKF USA, 2.151 FIGURE 88 Single-row, deep-groove ball bearing (SKF USA, Inc.) installed is not operated for long periods of time (such as stand-by units or units kept in stock or storage). The shaft should be rotated occasionally (see specific instruction manual directions) to agitate the lubricant and maintain a film coating on the bearing elements. Self-aligning ball bearings (see Figure 87) are sometimes used for heavy loads, high speeds, long-bearing spans (large deflection angles at the bearings) and no axial thrust requirements. This bearing design acts as a pivot that compensates for misalignment and shaft deflection. For large shafts, the self-aligning spherical roller bearing (refer to Figure 86) is used instead of the self-aligning ball bearing, and it can carry both radial loads and axial thrust loads. The single-row, deep-groove ball bearing (see Figure 88), sometimes referred to as a Conrad-type bearing, is the most commonly used bearing in centrifugal pumps, except for the larger size pumps. The Conrad-type design is recommended for use in centrifugal pumps because it can support either radial, axial, or a combination of radial and axial loads. This makes it ideal for the radial bearing in end-suction centrifugal pumps or as both the radial and thrust bearings in small pumps. The bearing design requires a careful alignment between the shaft and the housing. It is often used with seals or shields in greaselubricated applications to help exclude dirt and retain lubricants within the bearing. Angular contact ball bearings are commonly used in centrifugal pump applications to support axial loads or a combination of both axial and radial loads. Their axial stiffness and small operating clearances provide precise position accuracy for the shaft. Angular contact bearings are manufactured in a single-row design (see Figure 89), typically with a 40° contact angle, and also as a double-row bearing (see Figure 90), most commonly with a 30° contact angle. Single-row, angular contact ball bearings support axial loads in only one direction when used singly. To support reversing axial loads or combined loads, single-row bearings must be mounted in a back-to-back or face-to-face arrangement where the contact angles oppose each other. Owing to its more rigid design, the back-to-back arrangement is generally recommended for centrifugal pumps, while the face-to-face arrangement is common when a slight misalignment is expected. When required to support heavy axial loads, single-row, angular contact ball bearings can be mounted in tandem where their contact angles are in the same direction. This arrangement must still be opposed with a third bearing in a back-to-back or face-to-face arrangement with the tandem pair when radial or reversing thrust loads must also be supported (see Figure 91). Depending upon the operating conditions of the pump, single-row, angular contact ball bearings typically operate with either a small clearance or a light preload. Some applications exist where a high axial load occurs predominantly in one direction, but the thrust bearing must be capable of carrying occasional smaller axial loads in the 2.152 CHAPTER TWO FIGURE 89 Single-row, angular contact bearing (SKF USA, Inc.) FIGURE 91 FIGURE 90 Double-row, angular-contact bearing (SKF USA, Inc.) Paired bearing arrangements (SKF USA, Inc.) reversing direction. When this occurs, a typical back-to-back angular contact bearing arrangement can result in one bearing becoming nearly completely unloaded. In the most severe cases of axial unloading of angular contact bearings, skidding of the unloaded balls within the bearing races can occur. This skidding can result in bearing heating and subsequent damage, even failure, with time. To avoid ball skidding under light load or no-load conditions, standard angular contact bearing sets can be arranged for a light preload that will result in a sufficient load on the dynamically unloaded bearing to prevent skidding. Another alternative is to install a matched set of two angular contact bearings with different contact angles (see Figure 92). By utilizing an angular contact bearing with a lower contact angle (say 15 degrees instead of the normal 40 degrees), the unloaded bearing will have a lower requirement for an axial load and be more resistant to ball skidding. This means the bearing will run at a lower temperature. The double-row, angular contact ball bearing (see Figure 93) is similar in design to a back-to-back pair of single-row, angular contact ball bearings, but in a narrower width package. Its ease of mounting, along with its low-friction operation, high-speed capability, and seal or shield availability, make it an ideal bearing for light- to medium-duty end suction centrifugal pumps and submersible pumps. Lubrication of Antifriction Bearing In the layout of a line of centrifugal pumps, the choice of the lubricant for the pump bearings is dictated by application requirements, by cost considerations, and sometimes by the preferences of a group of purchasers committed to the major portion of the output of that line. For example, in vertical wet-pit condenser circulating pumps, water is the lubricant of choice, in preference to grease or oil. If oil or grease is used in such pumps and the lubricant leaks into the pumping system, the condenser operation might be seriously affected because the tubes would become coated with the lubricant. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 92 Angular contact bearings with different contact angles (SKA USA, Inc.) 2.153 FIGURE 93 A double-row, angular-contact ball thrust bearing that is grease-lubricated and watercooled Most centrifugal pumps for refinery services are supplied with oil-lubricated bearings because of the insistence of refinery engineers on this feature. In the marine field, on the other hand, the preference lies with grease-lubricated bearings. For high pump operating speeds (5,000 rpm and above), oil lubrication is found to be the most satisfactory. For highly competitive lines of small pumps, the main consideration is cost, and so the most economical lubricant is chosen, depending upon the type of bearing used. Ball bearings used in small centrifugal pumps are usually grease-lubricated, although some services use oil lubrication. In grease-lubricated bearings, the grease packed into the bearing is thrown out by the rotation of the balls, creating a slight suction at the inner race. (Even if the grade of grease is relatively light, it is still a semisolid and flows slowly. As heat is generated in the bearing, however, the flow of the grease is accelerated until the grease is thrown out at the outer race by the rotation.) As the expelled grease is cooled by contact with the housing and thus is attracted to the inner race, a continuous circulation of grease lubricates and cools the bearing. This method of lubrication requires a minimum amount of attention and has proved itself very satisfactory. A vertically mounted thrust bearing arranged for grease lubrication is shown in Figure 94. A bearing fully packed with grease prevents proper grease circulation in itself and its housing. Therefore, as a general rule, it is recommended that only one-third of the void spaces in the housing be filled. An excess amount of grease will cause the bearing to heat up, and grease will flow out of the seals to relieve the situation. Unless the excess grease can escape through the seal or through the relief cock that is used on many large units, the bearing will probably fail early. In oil-lubricated ball bearings, a suitable oil level must be maintained in the housing. This level should be at about the center of the lowermost ball of a stationary bearing. It can be achieved by a dam and an oil slinger to maintain the level behind the dam and thereby increase the leeway in the amount of oil the operator must keep in the housing. Oil rings are sometimes used to supply oil to the bearings from the bearing housing reservoir (see Figure 95). In other designs, a constant-level oiler is used (see Figure 96). Because of the advantages of interchangeability, some pump lines are built with bearing housings that can be adapted to either oil or grease lubrication with minimum modifications (see Figure 97). Oil Film or Sleeve Bearings See Subsection 2.2.5. 2.154 CHAPTER TWO FIGURE 94 Vertically mounted thrust bearing arranged for grease lubrication (SKF USA, Inc.) FIGURE 95 A ball bearing pump with oil rings COUPLINGS _________________________________________________________ Centrifugal pumps are connected to their drivers through couplings of one sort or another, except for close-coupled units, in which the impeller is mounted on an extension of the shaft of the driver. Because couplings can be used with both centrifugal and positive displacement pumps, they are discussed separately in Section 6.3. BEDPLATE AND OTHER PUMP SUPPORTS_______________________________ For very obvious reasons, it is desirable that pumps and their drivers be removable from their mountings. Consequently, they are usually bolted and doweled to machined surfaces that in turn are firmly connected to a foundation. To simplify the installation of horizontal-shaft units, these machined surfaces are usually part of a common bedplate on which either the pump or the pump and its driver have been prealigned. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 96 FIGURE 97 2.155 A constant-level oiler Ball bearings arranged (left) with oil rings in the housing and (right) for grease lubrication Bedplates The primary function of a pump bedplate is to furnish mounting surfaces for the pump feet that can be rigidly attached to the foundation. Mounting surfaces are also necessary for the feet of the pump driver or drivers and for the feet of any independently mounted power transmission device. Although such surfaces could be provided by separate bedplates or by individually planned surfaces, it would be necessary to align these separate surfaces and fasten them to the foundation with the utmost care. Usually, this method requires in-place mounting in the field as well as drilling and tapping for the bolts after all the parts have been aligned. To minimize such field work, coupled horizontal-shaft pumps are usually purchased with a continuous base extending under the pump and its driver. Ordinarily, both these units are mounted and aligned at the place of manufacture. Although such bases are designed to be quite rigid, they deflect if improperly supported. It is therefore necessary to support them on a foundation that can supply the required rigidity. Furthermore, as the base can be sprung out of shape by improper handling during transit, it is imperative that the alignment be carefully rechecked during erection and prior to starting the unit. 2.156 CHAPTER TWO FIGURE 98 Horizontal shaft overhung pump and driver on a structural steel bedplate with a raised edge around the base and a tapped drainage connection (Flowserve Corporation) FIGURE 99 Horizontal shaft centrifugal pump and internal combustion engine mounted on a structural steel bedplate (Flowserve Corporation) As the unit size increases, so does the size, weight, and cost of the base required. The cost of a prealigned base for most large units exceeds the cost of the field work necessary to align individual bedplates or soleplates and to mount the component parts. Such bases are therefore used only if appearances require them or if their function as a drip collector justifies the additional cost. Even in fairly small units, the height at which the feet of the pump and the other elements are located may differ considerably. A more rigid and nicelooking installation can frequently be obtained by using individual bases or soleplates and building up the foundation to various heights under the separate portions of equipment. Baseplates are usually provided with a raised edge or raised lip around the base to prevent dripping or draining onto the floor (see Figure 98). The base itself is sloped toward 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 100 Horizontal shaft centrifugal pump and driver on a structural steel bedplate made of a simple channel shape (Flowserve Corporation) 2.157 FIGURE 101 Single-stage double-suction pump with centerline support (Flowserve Corporation) one end to collect the drainage for further disposal. A drain pocket is provided near the bottom of the slope, sometimes with a mesh screen. A tapped connection in the pocket permits piping the drainage to a convenient point. Bedplates are usually fabricated from steel plate and structural steel shapes (see Figures 99 and 100). Even though most of these fabrications have a drain capability, because of the popular use of mechanical seals and the containment of stuffing box leakage for pumps that continue to use shaft packing (leakage is usually collected in the bearing bracket and piped to a common collecting point), the bedplate surfaces actually are seldom used to collect leakage from the pumping equipment during operation. Bedplate drain surfaces are usually employed to contain the leakage of pumpage and other liquids during pump maintenance and removal or in the event of a seal or packing failure. Soleplates Soleplates are cast-iron or steel pads located under the feet of the pump or its driver and are embedded in the foundation. The pump or its driver is doweled and bolted to them. Soleplates are customarily used for vertical dry-pit pumps and also for some of the larger horizontal units to save the cost of the large bedplates otherwise required. Centerline Support For operation at high temperatures, the pump casing must be supported as near to its horizontal centerline as possible in order to prevent excessive strains caused by temperature differences. Such strains might seriously disturb the alignment of the unit and eventually damage it. Centerline construction is usually employed in boilerfeed, refinery, and hot-water circulating pumps (see Figure 101). The exact temperature at which centerline support construction becomes mandatory varies from 250 to 350°F (121 to 177°C). Horizontal Units Using Flexible Pipe Connections The previous discussion of bedplates and supports for horizontal-shaft units assumed their application would be to pumps with piping setups that do not impose hydraulic thrusts on the pumps. If flexible pipe connections or expansion joints are desirable in the suction or discharge piping of a pump (or in both), the pump manufacturer should be so advised for several reasons. First, the pump casing will be required to withstand various stresses caused by the resultant hydraulic thrust load. Although this is rarely a limiting or dangerous factor, it is best that the manufacturer have the opportunity to check the strength of the pump casing. Second, the resulting hydraulic thrust has to be transmitted from the pump casing through the casing feet to the bedplate or soleplate and then to the foundation. Usually, horizontalshaft pumps are merely bolted to their bases or soleplates, and so any tendency to displacement is resisted only by the frictional grip of the casing feet on the base and by relatively small dowels. If flexible pipe joints are used, this attachment may not be sufficient to withstand the hydraulic thrust. If high hydraulic thrust loads are to be encountered, therefore the pump feet must be keyed to the base or supports. Similarly, the 2.158 CHAPTER TWO bedplate or supporting soleplates must be of a design that will permit transmission of the load to the foundation. Each of these design elements must be checked to confirm their capability to withstand hyraulic thrust loads from flexible pipe connections. Preferably expansion joints will be restrained to avoid transmitting loads to the pump nozzles. VERTICAL PUMPS____________________________________________________ Vertical-shaft pumps fall into two classifications: dry-pit and wet-pit. Dry-pit pumps are surrounded by air, and the wet-pit types are either fully or partially submerged in the liquid handled. Vertical Dry-Pit Pumps Dry-pit pumps with external bearings include most small, medium, and large vertical sewage pumps, most medium and large drainage and irrigation pumps for medium and high heads, many large condenser circulating and water supply pumps, and many marine pumps. Sometimes the vertical design is preferred (especially for marine pumps) because it saves floor space. At other times, it is desirable to mount a pump at a low elevation because of suction conditions, and it is then also preferable or necessary to have the pump driver at a high elevation. The vertical pump is normally used for large capacity applications because it is more economical than the horizontal type, all factors considered. Many vertical dry-pit pumps are basically horizontal designs with minor modifications (usually in the bearings) to adapt them for vertical-shaft drive. This is not true of smalland medium-sized sewage pumps, however. In these units, a purely vertical design is the most popular. Most of these sewage pumps have elbow suction nozzles (see Figures 102 through 104) because their suction supply is usually taken from a wet well adjacent to the FIGURE 102 Corporation) Section of a vertical sewage pump with end-suction (elbow) and side discharge (Flowserve 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 103 2.159 An installed vertical sewage pump similar to that shown in Figure 102 (Flowserve Corporation) FIGURE 104 Vertical sewage pump with a direct-mounted motor (Flowserve Corporation) pit in which the pump is installed. The suction elbow usually contains a handhole with a removable cover to provide easy access to the impeller. To dismantle one of these pumps, the stuffing box head must be unbolted from the casing after the intermediate shaft or the motor and motor stand have been removed. The rotor assembly is drawn out upward, complete with the stuffing box head, the bearing 2.160 FIGURE 105 Corporation) CHAPTER TWO Vertical bottom-suction volute pumps installed in a sewage pumping station (Flowserve housing, and the like. This rotor assembly can then be completely dismantled at a convenient location. Vertical-shaft installations of single-suction pumps with a suction elbow are commonly furnished with either a pedestal or a base elbow (refer to Figure 102), both of which can be bolted to soleplates or even grouted in. The grouting arrangement is not desirable unless there is full assurance that the pedestal or elbow will never be disturbed or that the grouted space is reasonably regular and the grout will separate from the pump without excessive difficulty. Vertical single-suction pumps with bottom suction are commonly used for larger sewage, water supply, or condenser circulating applications. Such pumps are provided with wing feet that are bolted to soleplates grouted in concrete pedestals or piers (see Figure 105). Sometimes the wing feet may be grouted right in the pedestals. These must be suitably arranged to provide proper access to any handholes in the pump and to allow clearance for the elbow suction nozzles if these are used. If a vertical pump is applied to a condensate service or some other service for which the eye of the impeller must be vented to prevent vapor binding, a pump with a bottom singleinlet impeller is not desirable because it does not permit effective venting. Neither does a vertical pump employing a double-suction impeller (see Figure 106). The most suitable design for such applications incorporates a top single-inlet impeller (see Figure 107). If the driver of a vertical dry-pit pump can be located immediately above the pump, it is often supported on the pump itself (refer to Figure 104). The shafts of the pump and driver may be connected by a flexible coupling, which requires that each have its own thrust bearing. If the pump shaft is rigidly coupled to the driver shaft or is an extension of the driver shaft, a common thrust bearing is used, normally in the driver. Although the driving motors are frequently mounted on top of the pump casing, one important reason for the use of the vertical shaft design is the possibility of locating the motors at an elevation sufficiently above the pumps to prevent the accidental flooding of the motors. The pump and its driver may be separated by an appreciable length of shafting, which may require steady bearings between the two units. Subsection 6.3.1 discusses the construction and arrangement of the shafting used to connect vertical pumps to drivers located some distance above the pump elevation. Bearings for vertical dry-pit pumps and for intermediate guide bearings are usually antifriction grease-lubricated types to simplify the problem of retaining a lubricant in a housing with a shaft projecting vertically through it. Larger units, for which antifriction bearings are not available or desirable, use self-oiling, babbitt steady bearings with spiral 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.161 FIGURE 106 A vertical double-suction volute pump with a direct-mounted motor (Flowserve Corporation) FIGURE 107 A section of a vertical pump with a top-suction inlet impeller (Flowserve Corporation) 2.162 CHAPTER TWO FIGURE 108 FIGURE 109 Corporation) A self-oiling, babbitt steady bearing for large vertical shafting (Flowserve Corporation) A section of a large, vertical, bottom-suction volute pump with a single sleeve bearing (Flowserve oil grooves (see Figures 108 and l09). Figure l09 illustrates a vertical dry-pit pump design with a single-sleeve line bearing. The pump is connected by a rigid coupling to its motor (not shown), which is provided with a line and a thrust bearing. Vertical dry-pit centrifugal pumps are structurally similar to horizontal-shaft pumps. It is to be noted, however, that many of the large, vertical, single-stage, single-suction (usually bottom) volute pumps that are preferred for large storm water pumpage, drainage, irrigation, sewage, and water supply projects have no comparable counterpart among horizontal-shaft units. The basic U-section casing of these pumps, which is structurally weak, often requires the use of heavy ribbing to provide sufficient rigidity. In the comparable water turbine practice, a set of vanes (called a speed ring) is employed between the casing and the runner to act 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.163 as a strut. Although the speed ring does not adversely affect the operation of a water turbine, it would function basically as a diffuser in a pump because of the inherent hydraulic limitations of that construction. Some high-head pumps of this type have been made in the twinvolute design. The wall separating the two volutes acts as a stengthening rib for the casing, thus making it easier to design a casting strong enough for the pressure involved. Bases and Supports for Vertical Pumping Equipment Vertical-shaft pumps, like horizontal-shaft units, must be firmly supported. Depending upon the installation, the unit can be supported at one or several elevations. Vertical units are seldom supported from walls, but even that type of support is sometimes encountered. Occasionally, a nominal horizontal-shaft pump design is arranged with a vertical shaft and a wall used as the supporting foundation. Regular horizontal-shaft units can be used for this purpose without modification, except that the bedplate is attached to a wall. Careful attention must be given to the arrangement of the pump bearings to prevent the escape of the lubricant. Installations of double-suction, single-stage pumps with the shaft in the vertical position are relatively rare, except in some marine or navy applications. Hence, manufacturers have few standard pumps of this kind arranged so that a portion of the casing forms the support (to be mounted on soleplates). Figure 106 shows such a pump, which also has a casing extension to support the driving motor. Vertical Wet-Pit Pumps Vertical pumps intended for submerged operations are manufactured in a great number of designs, depending mainly upon the service for which they are intended. Small pumps of this type are often referred to as sump pumps. Wet-pit centrifugal pumps can be classified in the following manner: • Vertical turbine pumps • Propeller or modified propeller pumps • Volute pumps Vertical Turbine Pumps Vertical turbine pumps were originally developed for pumping water from wells and have been called deep-well pumps, turbine well pumps, and borehole pumps. As their application to other fields has increased, the name vertical turbine pumps has been generally adopted by manufacturers. This is not too specific a designation because the term turbine pump has been applied in the past to any pump employing a diffuser. There is now a tendency to designate pumps using diffusion vanes as diffuser pumps to distinguish them from volute pumps. As that designation becomes more universal, applying the term vertical turbine pumps to the construction formerly called turbine well pumps will become more specific. The largest fields of application for the vertical turbine pump are pumping from wells for irrigation and other agricultural purposes, for a municipal water supply, and for industrial water supplies, as well as for processing, circulating, refrigerating, and air conditioning. This type of pump has also been utilized for brine pumping, mine dewatering, oil field repressuring, and other purposes. These pumps have been made for capacities as low as 10 or 15 gpm (2 or 3 m3/h) and as high as 25,000 gpm (5700 m3/h) or more and for heads up to 1,000 feet (300 m). Most applications naturally involve the smaller capacities. The capacity of the pumps used for bored wells is naturally limited by the size of the well as well as by the rate at which water can be drawn without lowering its level to a point of insufficient pump submergence. Vertical turbine pumps should be designed with a shaft that can be readily raised or lowered from the top to permit proper positioning of the impeller in the bowl. An adequate thrust bearing is also necessary to support the vertical shafting, the impeller, and the hydraulic thrust developed when the pump is in service. As the driving mechanism must also have a thrust bearing to support its vertical shaft, it is usually provided with one large enough to carry the pump parts as well. For these two reasons, the hollow-shaft motor or gear is most commonly used for vertical turbine pump drives. In addition, these pumps are sometimes made with their own thrust bearings to allow for a belt drive or for a drive through a flexible coupling by a solid-shaft motor, gear, or turbine. Dual-driven pumps usually employ an angle gear with a vertical motor mounted on its top. 2.164 (a)1 CHAPTER TWO (b) FIGURE 110 A vertical turbine pump design with enclosed impellers and: a) enclosed line shafting and b) openline shafting (Flowserve Corporation) The design of vertical pumps illustrates how a centrifugal pump can be specialized to meet a specific application. Figure 110 illustrates a turbine design with closed impellers and enclosed-line shafting and another turbine design with closed impellers and open-line shafting. The bowl assembly, or section, consists of the suction case (also called suction head or inlet vane), the impeller or impellers, the discharge bowl, the intermediate bowl or bowls (if more than one stage is involved), the discharge case, the various bearings, the shaft, and the miscellaneous parts, such as keys, impeller-locking devices, and the like. The column pipe assembly consists of the column pipe, the shafting above the bowl assembly, the shaft bearings, and the cover pipe or bearing retainers. The pump is suspended from the driving head, which consists of the discharge elbow (for above ground discharge), the motor or driver and support, and either the stuffing box (in an open-shaft construction) or the assembly for providing tension on the cover pipe and introducing a lubricant into it. Below ground discharge is taken from a tee in the column pipe, and the driving head functions principally as a stand for the driver and a support for the column pipe. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.165 Liquid in a vertical turbine pump is guided into the impeller by the suction case or head. This may be a tapered section for the attachment of a conical strainer or suction pipe, or it may be a bell mouth. Semiopen and enclosed impellers are both commonly used. For proper clearances in the various stages, the semiopen impeller requires more care in assembly on the impeller shaft and more accurate field adjustments of the vertical shaft position in order to obtain the best efficiency. Enclosed impellers are favored over semiopen ones because wear on the latter reduces capacity, which cannot be restored unless new impellers are installed. Normal wear on enclosed impellers does not affect impeller vanes, and worn clearances may be restored by replacing wearing rings. The thrust produced by semiopen impellers may be as much as 150 percent of that by enclosed impellers. Occasionally, in power plants, the maximum water level that can be carried in the condenser’s hot well will not give adequate NPSH for a conventional horizontal condensate pump mounted on the basement floor, especially if the unit has been installed in a space originally allotted for a smaller pump. Building a pit for a conventional horizontal condensate pump or a vertical dry-pit pump that will provide sufficient submergence involves considerable expense. Pumps of the design shown in Figure 111 have become quite popular in such applications. This is basically a vertical turbine pump mounted in a tank (often called a can) that is sunk into the floor. The length of the pump has to be such that sufficient NPSH will be available for the first-stage impeller design, and the diameter and length of the tank have FIGURE 111 Vertical turbine can pump for condesate service (Flowserve Corporation) 2.166 CHAPTER TWO FIGURE 112 Vertical propeller pump installed (Flowserve Corporation) to allow for proper flow through the space between the pump and tank and then for a turn and flow into the bell mouth. Installing this design in an existing plant is naturally much less expensive than making a pit because the size of the hole necessary to install the tank is much smaller. The same basic design has also been applied to pumps handling volatile liquids that are mounted on the operating floor and not provided with sufficient NPSH. Propeller Pumps Originally, the term vertical propeller pump was applied to vertical wet-pit diffuser or turbine pumps with a propeller or axial-flow impellers, usually for installation in an open sump with a relatively short setting (see Figures 112 and 113). Operating heads exceeding the capacity of a single-stage axial-flow impeller might call for a pump of two or more stages or a single-stage pump with a lower specific speed and a mixed-flow impeller. High enough operating heads might demand a pump with mixedflow impellers and two or more stages. For lack of a more suitable name, such high-head designs have usually been classified as propeller pumps also. Although vertical turbine pumps and vertical modified propeller pumps are basically the same mechanically and could even be of the same specific speed hydraulically, a basic turbine pump design is suitable for a large number of stages. A modified propeller pump design, however, is basically intended for a maximum of two or three stages. Most wet-pit drainage, low-head irrigation, and storm water installations employ conventional propeller or modified propeller pumps. These pumps have also been used for condenser circulating services, but a specialized design dominates this field. As large power plants are usually located in heavily populated areas, they frequently have to use badly contaminated water (both fresh and salt) as a cooling medium. Such water quickly short- 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 113 Vertical propeller pump with above ground discharge (Flowserve Corporation) 2.167 FIGURE 114 Vertical pull-out design allows the rotating element and critical non-rotating wear components to be removed for inspection and replacement without removing the complete pump (Flowserve Corporation) ens the life of fabricated steel. Cast iron, bronze, or an even more corrosion-resistant cast metal must therefore be used for the column pipe assembly. This requirement means a very heavy pump if large capacities are involved. To avoid the necessity of lifting this large mass for maintenance of the rotating parts, some designs (one of which is illustrated in Figure 114) are built so that the impeller, diffuser, and shaft assembly can be removed from the top without disturbing the column pipe assembly. These designs are commonly designated as pullout designs. Like vertical turbine pumps, propeller and modified propeller pumps have been made with both open- and enclosed-line shafting. Except for condenser circulating services, enclosed shafting, using oil as a lubricant but with a grease-lubricated tail bearing below the impeller, seems to be favored. Some pumps handling condenser 2.168 CHAPTER TWO circulating water use enclosed shafting but with water (often from another source) as the lubricant, thus eliminating any possibility of oil getting into the circulating water and coating the condenser tubes. Propeller pumps have open propellers. Modified propeller pumps with mixed-flow impellers are made with both open and closed impellers. Volute Pumps A variety of wet-pit pumps are available. The liquid pumped, be it clean water, sewage, abrasive liquids or slurries, dictates whether a semiopen or an enclosed impeller will be used, whether the shafting will be open or closed to the liquid pumped, and whether the bearings will be submerged or located above the liquid. Figure 115 illustrates a single-volute pump with a single-suction enclosed nonclog impeller, no pump-out vanes or wearing-ring joints on the hack side of the impeller, and enclosed shafting. The pump is designed to be suspended from an upper floor by means of a drop pipe and for pumping sewage or other solid-laden liquids. To seal against leakage along the shaft at the point where it passes through the casing, a seal chamber or a stuffing box is provided. The design of a stuffing box can be either like the one shown in Figure 116, which uses rings of packing and a spring-loaded gland, or the one shown in Figure 117, which uses U-cup packing requiring no gland. The pump shown in Figure 115 uses two sleeve bearings above the impeller. The bottom bearing is grease-lubricated, and a seal is provided to prevent grease leakage as well as to keep out any grit. The upper bearing connects the shaft cover pipe to the pump-bearing bracket and the upper end of the cover pipe to the floorplate. This bearing is gravity-feed oil-lubricated. If intermediate bearings are required, they are also supported by the cover pipe and are oil-lubricated. The pump thrust is carried by the motor, which can be either a hollow-shaft or a solid-shaft construction. The latter type requires the use of a rigid coupling between the pump and motor shafts. In most applications, these volute-type pumps have been replaced with vertical wet-pit pumps with the stuffing box/seal chamber in the discharge head (see Section 9.2, Figure 4b). A design that uses open shafting and no seal chamber or stuffing box at the pump casing, incorporating its own thrust bearing, is shown in Figure 118. The impeller shown in Figure 118 is of the vortex type (sometimes called a recessed impeller, as shown in Figure 119), which is suitable for pumping heavy concentrations of solid material (such as sludges or slurries) or in certain food-processing applications, but other types of impellers can be substituted. Pumped liquid leakage from the casing is relieved back to the suction through holes in the support pipe. The seal chamber or stuffing box at the driver floor elevation is used only when gas tight construction is desired. The lower and any intermediate sleeve bearings are grease-lubricated as shown, but gravity-feed oil lubrication is also available in other designs. The upper antifriction thrust bearing is grease-lubricated. A solid shaft motor and a flexible shaft are used. Figure 120 illustrates what is called a cantilever-shaft pump, which has the unique feature of having no bearings below the liquid surface. The shaft is exposed to the liquid pumped. External antifriction grease-lubricated bearings are provided above the floor and are properly spaced to support the rigid shaft. They carry both the thrust and the radial load. A flexible coupling is used between the pump and the solid-shaft motor. The stuffing box at the floorplate may be eliminated if holes are provided in the drop pipe to maintain the liquid level in the pipe even with the sump liquid level. Either semiopen or enclosed impellers may be used. An interesting design of the wet-pit pump is shown in Figure 121. It uses a singlestage, double-suction impeller in a twin-volute casing. Because the axial thrust is balanced, the thrust bearing need carry only the weight of the rotating element. The pump requires no stuffing box or mechanical seal. The shaft is entirely enclosed, and the bearings are externally lubricated, either with oil or with water. The lower bearing receives its lubrication from an external pipe connection. The term sump pump ordinarily conveys the idea of a vertical wet-pit pump that is suspended from a floorplate or sump cover. It could be supported by a foot on the bottom of a well, be motor-driven and automatically controlled by a float switch, and be used to remove drainage collected in a sump. The term does not indicate a specific construction, for both diffuser and volute designs are used. These may be single-stage or multistage and have open or closed impellers of a wide range of specific speeds. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.169 FIGURE 115 A section of a vertical wet-pit, nonclogging pump (Flowserve Corporation) For small capacities driven by fractional-horsepower motors, cellar drainers can be used. These are small and usually single-stage volute pumps with single-suction impellers (either top or bottom suction) supported by a foot on the casing. The motor is supported well above the impeller by some form of column enclosing the shaft. These drainers are made as complete units, including float, float switch, motor, and strainers (see Figure 122). The larger sump pumps are usually standardized but obtainable in any length, with covers of various sizes (on which a float switch may be mounted) and the like. Duplex 2.170 CHAPTER TWO FIGURE 116 FIGURE 117 A typical stuffing box arrangement A stuffing box arrangement with U-cup packing (Flowserve Corporation) units, that is, two pumps on a common sump cover (sometimes with a hole for access to the sump), are often used (see Figure 123). Such units may operate their pumps in a fixed order, or a mechanical or electric alternator may be used to equalize their operation. The Application of Vertical Wet-Pit Pumps Like all pumps, the vertical wet-pit pump has advantages and disadvantages. One advantage is that installation does not require a separate dry pit to collect the pumped liquid. If the impeller (first-stage impeller in multistage pumps) is submerged, no priming problem exists and the pump can be automatically controlled without fear of its ever running dry. Moreover, the available NPSH is greater (except in closed tanks) and often permits a higher rotative speed for the same service conditions. A second advantage is that the motor or driver can be located at any desired height above any flood level. It has the following mechanical disadvantages: (1) the possibility of freezing when idle, (2) the possibility of damage by floating objects if the unit is installed in an open ditch or similar installation, (3) the inconvenience of lifting out and dismantling for inspection and repairs, no matter how small, and (4) the pump bearings have a relative short life unless the water and bearing design are ideal. In summary, the vertical wet-pit pump is the best pump available for some applications. It’s not ideal but can be the most economical for certain installations, a poor choice for some, and the least desirable for still others. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.171 FIGURE 118 A vertical wet-pit volute pump with open shafting, no stuffing box, and its own thrust bearing (Aurora Pump) FIGURE 119 The principle operation of the vortex pump in Figure 118 (Aurora Pump) Axial Thrust in Vertical Pumps with Single Suction Impellers The subject of axial thrust in horizontal pumps has already been discussed in this section. When pumps are installed in a vertical position, additional factors need be taken into account when determining the amount and direction of the thrust to be absorbed in the thrust bearings. The first and most significant of these factors is the weight of the rotating parts, which is a constant downward force for any given pump, completely independent of the pump operating capacity or total head. Since in most cases the single-suction impellers of vertical pumps are mounted with the suction eye facing downward, the normal hydraulic axial thrust is exerted downward and the weight of the rotating parts is additive to this axial thrust. The second factor involves the dynamic force (or change in momentum) caused by the change in the direction of flow, from vertical to either horizontal or partly horizontal, as the pumped liquid flows through the impeller. This force acts upward and balances a small portion of the hydraulic downthrust and of the rotor weight. The magnitude of this force for water having a specific weight (force) of 62.34 lb/ft3 (9.79 kN/m3) is 2.172 CHAPTER TWO FIGURE 120 Vertical wet-pit cantilever-type volute pump: (1) impeller nut, (2) open impeller, (3) closed impeller, (4) open casing, (5) closed casing, (6) casing wearing ring, (7) casing gasket, (8) discharge flange gasket, (9) back cover, (10) packing or mechanical seal, (11) packing or seal gland, (12) radial bearing, (13) shaft, (14) supporting pipe, (15) separate stuffing box, (16) sump cover, (17) discharge pipe, (18) upper motor pedestal, (19) lower motor pedestal, (20) lower thrust bearing cover, (21) thrust bearing, (22) upper thrust bearing cover, (23) coupling, (24) motor, (25) bearing locknut and washer (Laurence Pump) 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 121 Double-suction, wet-pit pump (Flowserve Corporation) FIGURE 122 A cellar-drainer sump pump (Sta-Rite Products) in USCS units Fu  KV2e Ae 2g12.312 in SI units Fu  KV2e Ae 2g11.022 where Fu  upthrust, lb (N) K  constant related to the impeller type Ve  velocity in the impeller eye ft/s (m/s) 2.173 2.174 CHAPTER TWO FIGURE 123 A duplex sump pump (Economy Pump) Ae  net eye area, in2 (cm2) g  32.2 ft/s2 (9.81 m/s2) The constant K is 1.0 for a fully radial-flow impeller, less than 1.0 for a mixed-flow impeller, and essentially zero for a fully axial-flow impeller. Under normal operating conditions, the upward thrust caused by the change of momentum is hardly significant in comparison with the downward thrust caused by the unbalanced pressures acting on the single-suction impeller. Consider the example of an impeller with the following characteristics: Capacity Total head Net pressure Unbalanced eye area 2,500 gpm (568 m3/h) 231 ft (70.4 m) 100 lb/in2 (6.89 bar) 40 in2 (258 cm2) If we neglect the effect of the pressure distribution on the shrouds of the impeller, the downward thrust is in USCS units in SI units 100  40  4000 lb 6.89  105  258  17,800 N 10,000 The upward force caused by the change of momentum is 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS in USCS units in SI units 1.0  a 1 a 2.175 2500  0.32 2 b  40 40  107.5 lb 64.4  2.31 568  10,000 2 b  258 258  3600  481 N 2  9.81  1.02 This is certainly negligible relative to the downward hydraulic axial thrust. The situation is quite different, however, during the startup of a vertical pump. Although the motor may get up to full speed in just a few seconds, it takes a certain amount of time for the total head to increase from zero that corresponding to the normal operating capacity. Consequently, the pump will be operating in a very high capacity range. Since the upward force caused by the change of momentum varies as the square of the capacity while the downward axial thrust caused by pressure differences is very low, there can be a momentary net upward force, or upthrust. This means that thrust bearings intended to accommodate the axial thrust of vertical pumps must be capable of accommodating some thrust in the upward direction in addition to the normal downward thrust. This is particularly true for pumps with relatively low heads per stage and with short settings because in these units the rotor weight does not at all times compensate for the upthrust from the change in momentum. Particular care must be exercised in defining the range of vertical movement allowed for the thrust bearings because any such movement must remain within the displacement limits of any mechanical seal used in the pump. In rather rare cases, a close-coupled vertical pump is operated at such a high capacity that a continuous net upthrust is generated. Such operations can damage the pump because the line shaft is operated in compression and may buckle, causing vibration and bearing wear. The manufacturer’s comments should be invited if such an operation is contemplated. Shaft Elongation in Vertical Pumps The elongation of a vertical pump shaft is caused by three separate phenomena: (1) the tensile stress caused by the weight of the rotor, (2) the tensile stress caused by the axial thrust, and (3) the thermal expansion of the shaft. In most cases, the tensile stress created by the axial thrust is several times greater than that created by the weight. In a typical example of a 16,000-gpm (3636-m3/h) pump designed for a 175-ft (53.3 m) head and 50 ft (15.25 m) long, the elongation caused by the weight of a 1600-pound (726 kg) impeller will be of the order of 0.0033 in (0.084 mm). The elongation caused by the axial thrust will be approximately 0.0315 in (0.8 mm). The elongation caused by thermal expansion has to be considered from two angles. First, if the shaft and the stationary parts are built of materials that have essentially the same coefficient of expansion, both will expand equally and no significant relative elongation will take place. Second, if the pumps are operated in essentially the same range of temperatures in which they were assembled, no significant relative expansion will take place, even if dissimilar coefficients of expansion are involved. Whatever the case, the pump manufacturers take these factors into consideration by providing the necessary vertical end-play between the stationary and rotating pump components. Loads on Foundations of Vertical Pumps If the motor support is integral with the pump discharge column, as in Figures 110, 112, and 113, and if the hydraulic thrust is carried by the motor thrust bearing, this thrust is not additive to the deadweight of the pump and of its motor plus the weight of the water contained in the pump, insofar as the load on the foundations is concerned. This is because the pump and motor mounted in this fashion form a self-contained entity and all internal forces and stresses are balanced within this entity. If the pump and motor are supported separately, however, as in Figure 124, and are joined by a rigid coupling that transmits the pump hydraulic thrust to the motor thrust bearing, the foundations will carry the following loads when the pump is running: 2.176 CHAPTER TWO FIGURE 124 The vertical pump and driving motor supported separately (Flowserve Corporation) • The foundation supporting the motor: the weight of the motor plus the weight of the pump rotor plus the axial hydraulic thrust developed by the pump • The foundation supporting the pump: the weight of the pump stationary parts plus the weight of the water in the pump, including the water in the discharge pipe supported by the foundation, less the axial hydraulic thrust developed by the pump Since when the pump is idle the foundation supporting the pump is not benefited by the reduction in load equivalent to the axial hydraulic thrust, both running and idle operating conditions must be considered in this case. Typical Arrangements of Vertical Pumps A pump is only part of a pumping system. The hydraulic design of the system external to the pump will affect the overall economy of the installation and can easily have an adverse effect upon the performance of the pump itself. Vertical pumps are particularly susceptible because the small floor space occupied 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS 2.177 by each unit offers the temptation to reduce the size of the station by placing the units closer together. If the size is reduced, the suction arrangement may not permit the proper flow of water to the pump suction intake. Recommended arrangements for vertical pumps are discussed in Section 10.1. SPECIAL PURPOSE PUMPS ___________________________________________ In addition to the more or less general purpose pumps described on the preceding pages, literally hundreds of centrifugal pumps are intended for very specific applications. Although it is impossible to describe every one of these special purpose designs, many of them are discussed and illustrated in Sections 9.1 through 9.22, covering a variety of pump services. However, several specific designs are seeing rapidly increasing usage and so are discussed in greater detail here. Submersible Motor-Driven Wet-Pit Pumps The installation of conventional vertical wet-pit pumps with the motor located above the liquid level may require a considerable length of drive shafting, particularly in the case of deep settings. The addition of this shafting, of the many line bearings, and possibly of an external lubrication system may represent a major portion of the total installed cost of the pumping unit. Furthermore, shaft alignment becomes more critical, and shaft elongation and power losses increase rapidly as the setting is increased, especially for deep-well pumps. A great variety of submersible motors have been developed to obviate these shortcomings. They are described in Subsection 6.1.1, and a classification of the various types of such motors is presented in Figure 26 of that subsection. Submersible wet-pit pumps eliminate the need for extended shafting, shaft couplings, a mechanical seal or stuffing box, a subsurface motor stand, and, in some cases, an expensive pump house. Both vertical turbine and volute-type wet-pit pumps may be so driven. Figures 125 and 126 illustrate, respectively, the external appearance and a cross section of a vertical turbine pump driven by a submersible motor located at the bottom of the pump. The pump suction is through a perforated strainer located between the motor and the first-stage impeller bowl. There is, of course, no shafting above the pump and the pump-and-motor unit is supported by the discharge pipe only. No external lubrication is required. The motor is completely enclosed and oil-filled and is provided with a thrust bearing to carry the pump downthrust. A mechanical seal is provided at the motor shaft extension, which is connected to the pump shaft with a rigid coupling. Only a discharge elbow and the electric cable connection are seen above the surface support plate. On occasion, this type of pump is used horizontally as a booster pump in a pipeline, and in such cases the elbow at the discharge is eliminated. Vertical wet-pit volute-type sump pumps can be obtained with close-coupled submersible motors for drainage, sewage, process, and slurry services. Figure 127 illustrates dual submersible sewage pumps in a below ground collecting tank. The pumps (see Figure 128) are supported by guide rails that make it possible to lower and raise the pumps by means of a chain hoist. During this operation, the discharge pipe is connected and disconnected without dewatering the tank. Other arrangements use foot-supported pumps with rigid discharge piping. Motors used for this type of construction are usually hermetically sealed, employing a double mechanically sealed oil chamber with a moisture-sensing probe to detect any influx of conductive liquid past the outer seal. Controls to start and stop the pump motors can be either an air compressor bubbler system or level-sensing switches that tilt when floated (refer to Figure 127). Small portable pumps are available with flexible discharge hoses and built-in waterlevel motor control switches activated by trapped air pressure. Motors for these pumps are usually oil-filled and have a single mechanical shaft seal but are also available in a hermetically sealed design. The submersible motors are cooled by the liquid in which they are immersed and therefore should not be run dewatered, although some motors can operate for short periods (10 to 15 min.) this way. 2.178 CHAPTER TWO FIGURE 125 Vertical turbine pump driven by a submersible motor (Flowserve Corporation) FIGURE 127 FIGURE 126 Cross-sectional view of a submersible pump (Flowserve Corporation) Dual submersible sewage pumps in a below ground collecting tank (Flowserve Corporation) 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 128 2.179 A section of a submersible sewage pump shown in Figure 127 (Flowserve Corporation) In a pump-motor combination (see Figure 129), the motor is cooled by the pumped liquid as it moves through a passage around the sides of the motor. This design also uses a pressurized oil-seal chamber to assure positive sealing. Sealless Pumps Completely leakproof pumps are available for pumping corrosive, volatile, radioactive, and otherwise hazardous liquids. Canned motor pumps (see Figure 130) are an assembly of a standard centrifugal pump and a squirrel-cage induction motor in a hermetically sealed unit. Modifying a recirculating flow system in a canned motor pump can allow it to be used in applications at up to 1000°F (538°C). Magnetic drive pumps (see Figure 131) are an assembly of a rotor, an impeller, product lubricated bearings, and a magnetic carrier inside an isolation shell or diaphragm. This rotor is driven by magnets outside the shell or diaphragm. No mechanical connection exists between the driven magnets and the driving magnets. No seals exist and thus we have the term “sealless.” Sealless pumps are described in detail in Subsection 2.2.7. Straight-Radial-Vane High-Speed Pumps For handling volatile liquids at low flow rates and high heads, the straight-radial-vane impeller in a diffuser casing offers several advantages. Volatile, low-specific-gravity, poor-lubricity liquids require larger running clearances, which is not possible with conventional high-head multistage centrifugal or positive displacement pumps. High-head pumping of these liquids can be handled by operating this completely open impeller at very high speeds through an integral gear increaser and with very large impeller-to-casing clearance, typically 0.030 to 0.070 in (0.76 to 1.8 mm). Tests of one manufacturer’s design have shown this clearance can increase to 0.125 in (3.2 mm) with virtually no change in performance, and consequently there is no need to provide adjustment for impeller axial clearance. A pump of this design can also run “dry,” as liquid lubricity is not required to lubricate the bearings, which can be separately lubricated (providing the mechanical seal is lubricated). Figure 132 illustrates one manufacturer’s design of a radial-vane high-speed pump. The impeller rotates in a circular casing that has a single emission point leading to a conical diffusion section. The advantage of this type of casing is that the conversion to pressure occurs outside the circular housing, thus eliminating any recirculation forces that 2.180 FIGURE 129 CHAPTER TWO A portable submersible pump (Peabody Barnes) FIGURE 130 Typical canned motor pump (Ref. Subsection 2.2.7) (Crane Chempump) would require a close clearance between the impeller and casing. For higher flow ranges, a double emission point design, as shown in Figure 133, is used. This additional emission acts like a double-volute casing in conventional centrifugal pumps. 2.2.1 CENTRIFUGAL PUMP: MAJOR COMPONENTS FIGURE 131 FIGURE 132 Corporation) 2.181 Typical magnetic drive sealless pump (Ref. Subsection 2.2.7) (Flowserve Corporation) A straight-radial-vane impeller in a single emission-point, conical-diffuser circular casing (Sundyne Figure 134 is a sectional view of this pump with an integral speed increaser. Because of the high rotative speed, a single-stage pump can achieve the head normally associated with multistage centrifugal or positive displacement pumps. With this smaller liquid end and an integral gear box, cost and space savings can be appreciable. 2.182 CHAPTER TWO FIGURE 133 A straight-radial-vane impeller in a double-emission-point, conical-diffuser circular casing (Sundyne Corporation) FIGURE 134 A section of a straight-radial-vane impeller and inducer in a conical-diffuser circular casing with an integral gear increaser with the following: (1) pump casing, (2) impeller, (3) impeller bolt, (4) impeller tab washer, (5) inducer, (6) diffuser, (7) diffuser cover, (8) upper throttle bushing, (9) seal housing, (10) single shaft sleeve, (11) pump seal rotating face, (12) gearbox seal rotating face, (11) pump mechanical seal, (14) gearbox mechanical seal, (15) separator orifice, (16) separator fitting, (17) gearbox output housing, (18) gearbox input housing, (19) bearing plate, (20) interconnecting shaft, (21) low-speed shaft, (22) input gear, (23) lower idler gear, (24) high-speed shaft, (25) pinion gear, (26) idler shaft, (27) O-ring packings (Sundyne Corporation). 2.2.2 CENTRIFUGAL PUMP PACKING JAMES P. NETZEL Packing is used in the stuffing box of a centrifugal pump to control the leakage of the pumped liquid out, or the leakage of air in, where the shaft passes through the casing. This basic form of a seal can be applied in light- to medium-duty services and to those liquids that prove difficult for mechanical seals. THE DESIGN OF PACKING RINGS_______________________________________ Packing may be referred to as compression, automatic, or floating. Each term describes the type of operation in which the packing will be used. Automatic and floating packings require no gland adjustments in controlling leakage. Automatic packings are confined to a given space and are activated by the operating pressure. Automatic packing rings are designed in the form of V rings, U cups, and O rings. Floating packing includes piston rings and segmental rings that may be energized by a spring. These types of packing are commonly used in reciprocating applications. Compression packing is most commonly used on rotating equipment. The seal is formed by the packing being squeezed between the inboard end of the stuffing box and the gland (see Figure 1a). A static seal is formed at the ends of the packing ring and at the inside diameter of the stuffing box. The dynamic seal is formed between the packing and shaft or shaft sleeve. Under a load, the packing deforms down against the shaft, controlling leakage. Some leakage along the shaft is necessary to cool and lubricate the packing. The amount of leakage will depend on the materials of construction for the packing, the operating conditions of the application, and the condition of the equipment. Packing must be able to withstand equipment variables (see Figure 1b). The design of the packing ring and the materials of construction must be resilient to follow shaft runout and misalignments, as well as to compensate for thermal growth of the equipment without an appreciable increase in leakage. 2.183 2.184 CHAPTER TWO FIGURE 1 Compression packing; (a) new packing installation, (b) installation variables, (c) installation after many adjustments As rotating equipment is operated, the load on the packing must be adjusted to control leakage (see Figure 1c). Care should be taken not to overtighten the packing. Most compression packings have a lubricant designed into them to prevent overheating of the packing or scoring of the shaft. Repeated adjustments will drive some of the lubricant from the packing, which will result in reduced operating time. Compression packings are made of twisted, braided, woven, or wrapped elements formed into square or round cross-sections or other configurations. Square cross-sections are more common for rotating equipment. A selection of the proper materials for the packing must include the chemical resistance to the product being sealed as well as the temperature, pressure, and shaft speed. Complete lists of the construction materials, packing lubricants and binders are given in Tables 1 and 2. OPERATING FUNDAMENTALS__________________________________________ The Size and Number of Packing Rings The number of packing rings may vary, depending on the objective of the sealing system or the requirements of the rotating equipment. The most common packing arrangement for rotating equipment is illustrated in Figure 2. Three rings of packing are used to seal the process liquid from the packing lubricant. Two rings between the lantern and gland are used to restrict the leakage of the lubricant to the atmosphere. The size of the packing depends on the size of the equipment. Typically, for rotating shafts, the standard square size packings shown in Table 3 may be considered. 2.185 2.2.2 CENTRIFUGAL PUMP PACKING TABLE 1 Common materials of construction for packing Fibers Mineral Metal Graphite Metals Animal Wool Hair Leather Vegetable Flax Ramie Jute Cotton Paper Synthetic Nylon Rayon TFE Carbon Aramid Polyamide Lead Copper Brass P-bronze Aluminum Iron Stainless Steel Nickel Monel Inconel Zinc TABLE 2 Common lubricants and binders for packing Lubricants Mineral Lube Oil Paraffin Petrolatum Waxes Greases Vegetable Caster Oil Palm Oil Cottonseed Oil Linseed Oil Carnauba Wax Dry Lubricants Animal Tallow Glycerol Beeswax Lard Oil Fish Oil Soap Synthetic Oils Waxes Fluorolubes Silicones FIGURE 2 Graphite Moly Mica Talc Teflon Carbon Tungsten Disulfide Binders Grease Waxes Elastomers TFE Other Resins Common packing arrangement The packing size for an existing piece of equipment can be found by using the formula: Packing size  box ID ⫺ shaft or sleeve OD 2 2.186 CHAPTER TWO TABLE 3 Packing sizes for rotating shafts Shaft (or sleeve) Diameter, in (mm) 5 8 1 81 to (15 to 30) to 1 78 (30 to 50) to 3 (50 to 75) 3 to 4 43 (75 to 120) 4 34 to 12 (120 to 305) 1 81 1 78 Packing size, in (mm) 1 4 5 16 3 8 1 2 5 8 (6.3) (8) (10) (12.5) (16) FIGURE 3 A smothering gland and water-cooled stuffing box For packing to operate properly, the finish on the shaft sleeve must be at least 16 min (0.4mm) centerline average (CLA) and the finish in the bore should be 63 min (1.65 mm) CLA. The sleeve must be harder than the packing and chemically resistant to the liquid being sealed. If the sleeve has a coated material for a hard-wear surface, the sleeve must also have good thermal shock resistance. Lantern Rings (Seal Cages) When an application requires that a lubricant be introduced to the packing, a lantern ring is used to distribute the flow (refer to Figure 2). This ring is used at or near the center of the packing installation. For ease of assembly, most lantern rings are axially split. The construction materials range from metal to TFE (tetrafluoroethlyene). TFE lantern rings are usually filled with glass or with glass and molybdenum disulfide. They are inherently self-lubricating and will not score the shaft. A throat bushing at the bottom of the stuffing box can be used to provide a closer clearance with the shaft to prevent packing extrusion. Stuffing Box Gland Plates All mechanical packings are mechanically loaded in the axial direction by the stuffing box gland (refer to Figure 2). In cases where leakage of the process liquid is dangerous or can vaporize and create a hazard to operating personnel, a smothering gland is used to introduce a neutral liquid at lower temperatures (see Figure 3). A sufficient quantity of quenching liquid should be used to eliminate the danger from the liquid being pumped. The neutral liquid circulated in the gland mixes with the leakage and carries it to a safe place for disposal. Close clearances in the gland control the leakage of the combined liquids to the atmosphere. This quench can also be used to protect the packing from any wear through abrasion, because the leakage cannot vaporize and leave behind abrasive crystals. 2.2.2 CENTRIFUGAL PUMP PACKING 2.187 TABLE 4 Leakage to prevent packing burning and sleeve scoring Pressure lb/in2 (bar) Leakage, drops/min cc/min 60 4 190 470 0–60 (0–4.0) 61–100 (4.1–6.8) 101–250 (6.9–17) TABLE 5 Typically coefficients of friction Material f Plain cotton TFE impregnated fiber Grease-lube fiber Flexible graphite 0.22 0.17 0.10 0.05 Glands are usually made of bronze, but cast iron or steel can be used for all-iron pumps. When iron or steel glands are used, they are normally bushed with a non-sparking material like bronze. The basic operating parameters for compression packing are the PV (pressure  velocity) factor and the projected bearing area of the assembly. Together they determine the rate of heat generation for the system. Some leakage of the product being sealed or of the packing lubricant is necessary to keep the packing from burning up or scoring the shaft sleeve. The minimum values for leakage at different packing pressures are given in Table 4. Flexible graphite and carbon filament compression packings can be used with reduced leakage rates, or in dry, gas-tight pump stuffing boxes, as the developed heat is dissipated through the packing and pump housing. The exact pressure P between the shaft sleeve and the packing is a function of the pressure distribution over the length of the packing and the axial loading from the gland. For ease of calculation when determining the PV value, the gage pressure of the liquid at the packing is multiplied by p  the packing ID  rpm. The heat generation at the packing can then be estimated as fp2PND2L Q CJ Leakage and Power Consumption where Q f P N D L C J  heat generated, Btu/min (W)  coefficient of friction  liquid pressure at packing, lb/in2 gage (bar)  shaft speed, rpm  sleeve OD or packing ID, in (m)  sleeve length covered by packing, in (m)  12 (60 for SI units)  mechanical equivalent of heat  778 ft  lb/Btu (1 N  m/s  W) The coefficient of friction for various packings at a pressure of 100 lb/in2 (6.8 bar) is given in Table 5. The heat generated by the packing must be removed by the leakage through the packing. APPLICATION INFORMATION AND SEALING ARRANGEMENTS _____________ Materials of Construction Basically, stuffing box packing is a pressure breakdown device. In order for a packed stuffing box to operate properly, the correct packing must be 2.188 CHAPTER TWO FIGURE 4 Graphite acrylic packing in continuous form (John Crane Inc.) FIGURE 5 Non-asbestos packing (John Crane Inc.) FIGURE 6 Flexible graphite packing rings (John Crane, Inc.) applied and the appropriate design features must be included in the system. Numerous types of packing materials are available, each is suited to a particular service. These may be grouped into three categories: 1. Non-asbestos packing Since asbestos is no longer available as a packing material, other types of packing material are being used. These include cotton, TFE filament, Aramid fiber, aromatic polyamides, graphite/carbon yarn, and flexible graphite. All of these materials, with the exception of flexible graphite, can be impregnated with various lubricants. An example of graphited acrylic packing is shown in Figure 4. With the exception of flexible graphite, all the materials are of an interlaced construction for greater flexibility (see Figure 5). Packing made with this type of construction remains intact even if individual yarns wear away at the inside diameter of the packing. 2. Flexible graphite These types of packing rings were originally available cut from laminated sheet stock that required the rings to be made with an interference fit. Flexible graphite rings are made today from ribbon tape that is easily compressed to the shaft and bore to affect a better seal. Flexible graphite rings are available in split or endless rings (see Figure 6) or as the ribbon tape itself for ease of maintenance and installation. Operating limits for non-asbestos packings are found in Table 6. 3. Metallic packing The basic materials of construction are lead or babbit, aluminum, and copper in either wire or foil form. Metallic packing rings have flexible cores of non-asbestos materials such as twisted glass fiber. The packing is impregnated with graphite grease and/or oil lubricants (see Figures 7 and 8). Babbit is used in water and oil services at temperatures up to 450°F (229°C) and pressures up to 250 lb/in2 (17 bar). Copper foil is used with water and low sulfur oils. Aluminum TABLE 6 Service limitations of common packing materialsa Packing Material Pressure (max.)b lb/in2 gage (bar) PV rating (max.)c lb/in2 gage # fpm (bar # m/s) Temp. (max)d °F (°C) pH Range Comments 2.189 Cotton 100 (6.8) 150(65.6) 5–7 Non-abrasive material; for cold water and dilute salt solutions Flax/ramie 100 (6.8) 188,000 (65.8) 150 (65.6) 5–7 High, wet strength and excellent resistance to fungi and rotting; for cold water and dilute salt solutions Plastic 100 (6.8) 250 (17) 188,000 (65.8) 471,000 (165) 600 (315.5) 150 (65.6) 4–8 Excellent sealing qualities, reacts well to gland adjustments, can extrude at higher pressure if not backed up by braided or metallic packing Graphited acrylic 250 (17) 471,000 (165) 350 (175) 4–8 For mild chemicals and solvents Acrylic TFEimpregnated 250 (17) 471,000 (165) 350 (175) 2–10 For mild chemicals and solvents Babbitt (lead) 250 (17) 471,000 (165) 450( 232.2) 2–10 Shaft sleeve must have a Brinell hardness of 500 or more; for hot oils and boiler feed water Aluminum or copper 250 (17) 471,000 (165) 750 (398.8) 3–10 Shaft sleeve must have a Brinell hardness of 500 or more; for hot oils and boiler feed water TFE filament 250 (17) 471,000 (165) 500 (260) 0–14 For corrosive liquids and food service; usually requires slightly higher break-in leakage Aramid fiber 250 (17) 471,000 (165) 500 (260) 3–11 Strong resilient packing; maximum speed 1,900 fpm (9.6 m/s); good in abrasives and chemicals Graphite/carbon filamente 250 (17) 471,000 (165) 750 (398.8) 0–14 For corrosive liquids and high- temperature applications Flexible Graphitef 250 (17) 471,000 (165) 1000 (540) 0–14 Excellent conductor of heat from the sealing surfaces; operates with minimum leakage; excellent radiation resistance (a) Continuous lubrication introduced at the lantern ring. This table is only a guide. Consult the packing manufacturer with complete operating conditions for exact recommendations. (b) Pressure relates to the operating pressure at the stuffing box. (c) PV data based on a 2 in (5.08 cm) shaft at 1,750 and 3,600 rpm. (d) Temperature is the product temperature. (e) Functional temperature. Graphite can be used up to 3000°F (1650°C) in non-oxidizing atmospheres. (f) Functional temperature. Flexible graphite can be used to 5300°F (2970°C) in non-oxidizing atmospheres. 2.190 CHAPTER TWO FIGURE 7 FIGURE 8 Metallic packing in spiral form (John Crane Inc.) Metallic packing in ring form (John Crane Inc.) is used in oil service and heat transfer fluids. Both copper and aluminum have a temperature range up to 1000°F (537°C) and pressures of 250 lb/in2 (17 bar). Babbit is not suitable for running against brass or bronze shaft sleeves, and where copper or aluminum is used, the sleeves should be 550 Brinell (55–60 Rockwell C) or harder. Additional service limitations for common packing materials can be found in Table 6. ENVIRONMENTAL LIMITATIONS ________________________________________ Pressure Every pumping application results in either positive or negative pressure at the throat of the pump stuffing box. A positive pressure will force the liquid pumped through the packing to the atmosphere side of the pump. Higher pressure will result in greater leakage from the pump. This results in excessive tightening of the gland, which causes accelerated wear of the shaft or shaft sleeve and packing. For pressures at the stuffing box greater than 75 lb/in2 (5.1 bar), some means of throttling the pressure should be considered. A combination of hard and soft rings die-formed to the exact stuffing box bore 2.2.2 CENTRIFUGAL PUMP PACKING 2.191 FIGURE 9 Combination of hard and soft packing (John Crane Inc.) FIGURE 10 Compression packing with throttle bushing for pressure breakdown and sleeve dimensions can be used (see Figure 9). Harder rings at the inboard end of the box and at the lantern ring and gland break down the pressure and prevent the extrusion of the packing. If the packing itself cannot be used to break down the pressure, then a throttle bushing must be used (see Figure 10). This is a typical arrangement for a vertical turbine pump where the stuffing box is subject to the discharge pressure of the pump. Here the throttle bushing is used to bring the liquid to almost suction pressure, and most of the leakage through the bushing is bled back to the suction. When the suction pressure is less than atmospheric, as in condensate pump service, an orifice is used in the piping back to the suction in order to maintain pressure above atmospheric pressure in the stuffing box. This prevents air leakage into the pump and excessive flow and wear at the bushing. When a pump is fitted with a bypass line from the discharge, a valve can also be used to reduce the pressure at the stuffing box. This is another method for reducing pressure for the benefit of the installation. 2.192 CHAPTER TWO Air Leakage When the pressure at the stuffing box is atmospheric or just below during normal operation, a bypass from the pump casing discharge through an orifice can be used to inject liquid into the lantern ring. The sealing liquid will flow partly into the pump and partly out to the atmosphere, thereby preventing air from entering at the stuffing box. This arrangement is commonly used to handle clean, cool water. In some pumps, these connections are arranged so that liquid can be introduced into the lantern ring through internally drilled passages. When the negative pressure is very low, such as when the suction lift is in excess of 15 ft (4.6 m), an independent injection of 10 to 25 lb/in2 (4.8 to 11.7 bar) higher than the atmospheric pressure must be used on the pump. Otherwise, priming may be difficult. Hot well pumps or condensate pumps operate with as much as 28 in (0.7 m) of vacuum, and air leakage into the pump would occur even on standby service. Here a continuous injection of clean, cool water is required, or alternatively, a cross connection of sealing water to another operating pump will provide sealing pressure as long as one pump is operating. A lantern ring is provided for this, as shown in Figure 10. Temperature The control of temperature at the stuffing box is an important factor in promoting the life of the packing. Even though packings are rated for high product temperatures, cooling in most cases is desirable. The heat developed at the packing must also be removed. The rules of thumb are as follows: • For light service conditions with pressures at 15 lb/in2 (1 bar) and temperatures at 200°F (90°C), cooling is desirable. • For medium service conditions with pressures at 50 lb/in2 (3.4 bar) and temperatures at 250°F (118°C), lantern ring cooling is desirable, such as one gpm (3.78 l/min) at 5 lb/in2 (0.34 bar) above process pressure. • For high service conditions with pressures at 100 lb/in2 (6.8 bar) and temperatures at 300°F (131°C), lantern ring cooling is desirable, such as one gpm (3.78 l/min) at 5 lb/in2 (0.34 bar) above process pressure plus a water-cooled stuffing box (refer to Figure 3). When cooling water is required, it is circulated through the stuffing box at the lantern ring. For horizontal-shaft pumps, the inlet should be at the bottom, with the outlet at the top of the stuffing box. Some of the coolant may flow to the process liquid and some to the atmosphere. The outboard packing rings seal only the cool liquid. If the product to be sealed will solidify, then the packing box will have to be heated before the pump is started. This can be accomplished by steam or electric tracing of the pump stuffing box. Abrasives Liquids that contain abrasives in the form of suspended solids such as sand and dirt will shorten the life of the packing. Particles will imbed themselves in the packing and will begin to wear the shaft or shaft sleeve. Abrasives can be eliminated at the sealing surfaces by injecting a clean liquid into the lantern ring. The injection may take the form of a bypass line from the pump discharge through a filter or centrifugal separator. Where necessary, the clean injection may also be from an external source. To keep the abrasives from the packing in horizontal-shaft pumps, the injection can be made directly to the inboard end of the stuffing box through the lantern ring (see Figure 11). A soft rubber gasket between the lantern ring and the box shoulder can be used to limit the flow of clean liquid and the ingress of abrasives. When separators and filters cannot be used, an injection from an external source must be considered. Two rings of packing are located between the lantern ring and the inboard end of the stuffing box to keep the product dilution to a minimum. External flushing should be injected into the stuffing box at a pressure 10 to 25 lb/in2 (1.7 bar) greater than the pressure at the inboard end of the box from the liquid being pumped. A regulating valve, illustrated in Figure 12, can be used to control the pressure and flow to the packing installation. Flow to the packing can be regulated to ensure the best operating environment for this seal, while conserving water used for injection. 2.2.2 CENTRIFUGAL PUMP PACKING 2.193 FIGURE 11 A clean injection through a centrifugal separator to keep abrasives out of the stuffing box FIGURE 12 A regulation valve that controls and monitors the flow of water to the packing (Safematic) Dissolved solids in the liquid being pumped can also create a wear problem. Here an increase or decrease in stuffing box temperatures may be necessary to keep solids in solution. INSTALLING CONTINUOUS COIL PACKING _______________________________ To install continuous coil packing, perform the following steps: 1. Loosen and remove the gland from the stuffing box. 2. Using a packing puller, begin to remove the old packing rings. 3. Remove the split lantern ring (if present) and then continue removing the packing with the puller. 4. After the packing has been removed, check the sleeve for scoring and nicks. If the shaft sleeve or shaft cannot be cleaned up, it must be replaced. Check the size of the 2.194 5. 6. 7. 8. CHAPTER TWO stuffing box bore and the shaft sleeve or shaft diameter to determine which size packing should be used. After the size of the packing has been determined, wrap the packing tightly around a mandrel, which should be the same size as the pump shaft or sleeve. The number of coils should be sufficient to fill the stuffing box. Cut the packing along one side to form the individual rings. Before beginning the assembly of any packing material, be sure to read all the instructions from the manufacturer. Assemble the split packing rings on the pump. Each ring should be sealed individually with the split ends staggered 90° and the gland tightened to seal and fully compress the ring. Be sure the lantern ring is reinstalled correctly at the flush connection. Then back off the gland and retighten it, but only finger-tight. The exception to this procedure is that TFE packing should be installed one ring at a time, but not seated because TFE packings have high thermal expansion. Allow excess leakage during break-in to avoid the possibility of rapid expansion of the packing, which could score the shaft sleeve or shaft so that leakage could not be controlled. Leakage should be generous upon startup. If the packing begins to overheat at startup, stop the pump and loosen the packing until leakage is obtained. Restart only if the packing is leaking. CAUSES FOR A SHORT PACKING LIFE __________________________________ In order for packing to operate properly, the equipment must be in good condition. Shafts should be checked for runout and eccentricity to be sure they are within the manufacturer’s recommended tolerances. Surfaces in contact with the packing should be finished to the correct smoothness and tolerance. Table 7 lists common troubles that affect the packing life. Causes and possible cures are also given. 2.2.2 CENTRIFUGAL PUMP PACKING 2.195 TABLE 7 Packing troubles, causes, and cures Trouble Cause Cure No liquid delivered by pump Lack of prime (packing loose or defective, to allowing air leak into suction) Tighten or replace packing and prime pump. Not enough liquid delivered by pump Air leaking into stuffing box Check for leakage through stuffing box while operating. If no leakage occurs after after reasonable gland adjustment, new packing may be needed. or Lantern ring may be clogged or displaced and may need centering in line with sealing liquid connection. or Sealing liquid line may be clogged. or Shaft or shaft sleeve beneath packing may be badly scored, allowing air to be sucked into pump. Defective packing Replace packing and check the smoothness of the shaft or shaft sleeve. Not enough pump pressure Defective packing As per preceding Pump works for a while and then quits Air leaks into stuffing box As per preceding Pump takes too much power Packing too tight Release gland pressure and retighten reasonably. Keep leakage flowing. If none, check packing, sleeve, or shaft. Pump leaks excessively at stuffing Defective packing Replace worn packing or replace packing damaged by lack of lubrication. Wrong type of packing Replace packing not properly installed or run in. Replace improper packing with correct grade for liquid being handled. Shaft or shaft sleeve scored Put in lathe and machine-true and smooth or replace. Recheck dimensions for correct packing size. Packing too tight Release gland pressure and retighten. Packing not lubricated Release gland pressure and replace all packing if any burnt or damaged. Wrong grade of packing Check with pump or packing manufacturer for correct grade. Insufficient cooling water to jacket Check for open supply line valve or clogged line. Stuffing box improperly packed Repack. Shaft or shaft sleeve worn or scored Remachine or replace. Insufficient or no lubrication Repack, making sure packing is loose enough to allow some leakage. Packing packed improperly Repack properly, making sure all old packing is removed and the box is clean. Stuffing box overheats Packing wears too fast Wrong grade of packing Check with pump or packing manufacturer. Pulsating pressure on external seal liquid line Remove cause of pulsation. 2.196 CHAPTER TWO FURTHER READING __________________________________________________ Berzinsa, A. “Finer Points of Compression Packing Help with Centrifugal Pumps,” Power, June 1974, pp. 58–59. Ciffone, J. G. “Packing,” John Crane Mechanical Maintenance Training Center, Morton Grove, Illinois, November 1994. Crane Packing Company. Engineering Fluids Sealing, Materials, Design and Applications: An Information Compendium of Technical Papers. Morton Grove, IL 1979. Elonka, S. “Packing: Power Practical Manual,” Power, March 1955, pp. 107–130. Krisle, O.M. “How to Get Longer Life out of Pump Packing and Shaft Sleeve,” Water and Sewage Works, June 1967, pp. 199–201. Neale, M.J., “Packing Glands,” Tribology Handbook (Sec. A36), Wiley, New York; 1973. 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS JAMES P. NETZEL Mechanical seals have been used for many years to seal any number of liquids at various speeds, pressures, and temperatures. Today plant operators are benefiting from improved seal technologies driven by the U.S. Clean Air Act of 1990, and the American Petroleum Institute (API) Standard 682. These new seal technologies are based on advanced computer programs used to optimize seal designs, which are then verified through performance testing at simulated refinery conditions required by the API. The results to date indicate not only an improvement in emissions control, but also a major increase in equipment reliability. CLASSES OF SEAL TECHNOLOGY _____________________________________ Emerging seal technologies are providing clear choices for sealing. Various plant services require the application of these new technologies for emissions control, safety, and reliability. Sealing systems are now available that are based on the preferred method of lubrication to be used. These classes of seals are as follows: 1. Contacting liquid lubricated seals: • Normally, a single seal arrangement is cooled and lubricated by the liquid being sealed. This is the most cost-effective seal installation available to the industry. • Dual seals are arranged to contain a pressurized or non-pressurized barrier or buffer liquid. Normally, this arrangement will be used on applications where the liquid being sealed is not a good lubricating fluid for a seal and for emissions containment. These arrangements require a lubrication system for the circulation of barrier or buffer liquids. 2.197 2.198 CHAPTER TWO 2. Non-contacting gas lubricated seals: • Dual non-contacting, gas-lubricated seals are pressurized with an inert gas such as nitrogen. • Dual non-contacting, gas-lubricated seals are used in a tandem arrangement and pressurized by the process liquid being sealed, which is allowed to flash to a gas at the seal. A tandem seal arrangement is used on those liquids that represent a danger to the plant environment. For non-hazardous liquids, a single seal can be used. Each of these solutions has been used on difficult applications to increase the mean time between maintenance (MTBM). SEAL DESIGN _______________________________________________________ Advancing the state-of-the-art sealing systems are new suites of computer programs such as C’StedySM(1) used to analyze the performance of both contacting and non-contacting seal designs during steady-state and transient conditions. This type of finite analysis considers all of the operating conditions, the fluid sealed, the materials of construction, and seal geometry. The outputs from the program are seal distortion, temperature distribution, friction power, actual PV (pressure ⫻ velocity), leakage, the percentage of face in liquid or vapor, and fluid film stability (see Figure 1). This type of analysis requires accurate fluid and material properties. The results from the program can predict the success or failure of a given installation. For example, a mixture of liquid hydrocarbon made up of ethane, propane, butane, and hexane has to be sealed. This is a new application. The operating condition is 1,300 psig at 70°F. The shaft speed is 3,600 rpm. To determine the performance of this seal prior to installation, an analysis must be made. The results of this study indicate stable operations for a contacting seal, as shown in Figure 2. This study was used to predict seal performance. Actual field results from this difficult service were excellent at startup and during equipment operation. FIGURE 1 1 C’StedySM fluid film model for a mechanical seal (John Crane Inc.) Service Mark of John Crane Inc. 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS 2.199 FIGURE 2 C’Stedy output for a successful seal on high pressure light hydrocarbon service (John Crane Inc.) FIGURE 3 The basic components of a mechanical seal These state-of-the-art computer tools not only predict performance, but they also can be used to determine any short seal life. By using a series of calculations per second, this type of analysis can be used to create an animation that will visibly show changes to a seal at startup and during fluctuations in operating conditions. A seal can be examined for stable and unstable operations. This is a useful analysis tool for critical applications. DESIGN FUNDAMENTALS _____________________________________________ Contacting Liquid Lubricated Seals The basic components of a mechanical seal are the primary and mating rings. Together they form the dynamic sealing surfaces, which are perpendicular to the shaft. The primary ring is part of the seal head assembly, while the mating ring and static seal form a second assembly, making a complete installation for a pump. These basic seal parts are shown in Figure 3. For slower and normal shaft speeds, the seal head assembly will rotate with the shaft, while on high shaft speeds, the seal head assembly will be held stationary to the equipment. The only difference between contacting and non-contacting seal technologies is found in the design of the seal faces. Each system has the same type and number of parts. Each has its own area of application for maximum sealing efficiency. Non-contacting seal technology will be discussed later. 2.200 CHAPTER TWO FIGURE 4 Processes involved at contacting seal faces In a contacting seal, as the shaft begins to rotate, a small fluid film develops, along with frictional heat from the surfaces in sliding contact. These processes occurring at the seal faces are shown in Figure 4. The amount of heat developed at the seal faces must be removed to prevent the liquid being sealed from flashing or beginning to carbonize. Seal heat can be removed with a seal flush located at the seal faces. To analyze the performance of a seal and determine amount of cooling, the following calculations can be made. Seal Balance The greatest concern to the seal user is the dynamic contact between the mating seal surfaces. The performance of this contact determines the effectiveness of the seal. If the load at the seal faces is too high, the liquid at the seal faces will vaporize or carbonize and the seal faces can wear out. Damage to the seal faces can occur due to unstable conditions. A high wear rate from solid contact and leakage can occur if the bearing limits of the materials are exceeded. Seal balancing is a feature that is used to avoid these conditions and provide for a more efficient installation. The pressure in any seal chamber acts equally in all directions and forces the primary ring against the mating ring. Pressure acts only on the annular area ac (see Figure 5a), so that the force in pounds (Newtons) on the seal face is as follows: Fc  pac where p  seal chamber pressure, lb/in2 (N/m2) and ac  hydraulic closing area, in2 (m2) The pressure in lb/in2 (N/m2) between the primary ring and mating ring is 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS Pfœ  2.201 pac Fc  ao ao where ao  hydraulic opening area (seal face area), in2(m2). To relieve the pressure at the seal faces, the relationship between the opening and closing forces can be controlled. If ao is held constant and ac is decreased by a shoulder on a sleeve or seal hardware, the seal face pressure can be lowered (see Figure 5b). This is called seal balancing. A seal without a shoulder in the design is referred to as an unbalanced seal. A balanced seal is designed to operate with a shoulder. The ratio of the hydraulic closing area to the face area is defined as seal balance b: b ac ao Seals can be balanced for pressure at the outside diameter of the seal faces, as shown in Figure 5b. This is typical for a seal mounted inside the seal chamber. Seals installed outside the seal chamber can be balanced for pressure at the inside diameter of the seal faces. In special cases, seals can be double-balanced for pressure at both the outside and inside diameters of the seal. Seal balances can range from 0.65 to 1.35, depending on operating conditions. Face Pressure As relative motion takes place between the seal planes, a liquid film develops. The generation of this film is believed to be the result of surface waviness in the individual sealing planes. Pressure and thermal distortion, as well as anti-rotation devices such as drive pins, keys, or dents used in the seal design, have an influence on surface waviness and on how the film develops between the sliding surfaces. Hydraulic pressure develops in the seal face, which tends to separate the sealing planes. The pressure distribution, referred to as a pressure wedge, shown in Figure 6, can be considered as linear, concave, or convex. The actual face pressure pf in lb/in2 (N/m2) is the sum of the hydraulic pressure ph and the spring pressure Psp designed into the mechanical seal. The face pressure Pf is a further refinement of Pfœ , which does not take into account the liquid film pressure or the mechanical load of the seal: FIGURE 5 Hydraulic pressure acting on the primary ring: a) unbalanced, b) balanced FIGURE 6 The pressure distribution can be considered linear, concave, or convex. 2.202 CHAPTER TWO Pf  Ph  Psp where Ph p k b  p(b—k), lb/in2(N/m2) and  pressure differential across seal face, lb/in2 (N/m2)  pressure gradient factor  seal balance The mechanical pressure for a seal design is Fsp , lb>in2 1N>m2 2 Psp  ao where Fsp  seal spring load, lb (N) and ao  seal face area, in2 (m2) Then the actual face pressure can be expressed as Pf  ¢p1b  k2  Psp The actual face pressure is used in the estimate of the operating pressure and velocity for a given seal installation. Pressure-Velocity As the sealing planes move relative to each other, they are affected by the actual face pressure and rotational speed. The product of the two, pressure times velocity, is referred to as PV and is defined as the power Nf per unit area with a coefficient of friction of unity: Nf PV  ao For seals, the equation for PV can be written as follows: PV  PfVm  3 ¢p1b  k2  Psp 4Vm where Vm  velocity at the mean face diameter dm, ft/min (m/s). The PV for a given seal installation can be compared with values developed by seal manufacturers as a measure of adhesive wear. Power Consumption The PV value also enables the seal user to estimate the power loss at the seal with the following equation: Nf  1PV2fao, ft  lb>min 1N  m>s2 where f is the coefficient of friction. As a rule of thumb, the power to start a seal is generally five times the running value. The coefficients of friction for various common seal face materials are given in Table 1. These coefficients were developed with water as a lubricant at an operating PV value of 100,000 lb/in2  ft/min (35.03 bar  m/s). The coefficient of friction is a function of the tribological properties of the mating pairs of seal face materials and the fluid being sealed. Values in oil would be slightly higher because of the viscous shear of the fluid film at the seal faces. For a double or tandem seal, the barrier/buffer oil should have a low viscosity and be a good lubricant. The values given are suitable for estimating the power loss in a seal. For example, let’s say we have a pump having a 2-in (50.8-mm) diameter sleeve at the seal chamber is fitted with a balanced seal of this size and mean diameter. The seal operates in water at 300 lb/in2 (20.68 bar), 3,600 rpm, and ambient temperatures. The materials of construction are carbon and tungsten carbide. Determine the PV value and power loss of the seal, given the following: 2.203 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS TABLE 1 Coefficient of friction for various seal face materials (John Crane Inc.) Sliding Materials Rotating Stationary Coefficient of friction Carbon-graphite (resin filled) Cast iron Ceramic Tungsten carbide Silicon carbide Silicon carbide converted carbon 0.07 0.07 0.07 0.02 0.015 Silicon carbide Tungsten carbide 0.05 Silicon carbide Silicon carbide converted carbon 0.04 Silicon carbide converted carbon 0.05 Silicon carbide 0.05 Tungsten carbide 0.01 ¢p  300 lb>in2 120.68 bar2  0.75  0.5  2 in (50.8 min)  25 lb/in2(1.72 bar) p  50.8  3600  9.57 m>s p  2  3600  1885 ft>min a b Vm = 12 1000  60 ao  0.4 in2 (0.000258 m2) f  0.07 (Table 1) b k dm Psp In USCS units: PV  330010.75  0.52  254 118852  188,400 lb>in2 # ft>min Nf  1188,400210.072 10.42  5275 ft # lb>min  0.16 hp In SI units: PV  3 20,6810.75  0.52  1.724 19.572  66 bar # m>s  66  105 N>m2 # m>s Nf  166  105 210.072 12.58  10 4 2  119 N # m>s  119 W Temperature Control Controlling the temperature at the seal faces is desirable because wear is a direct function of temperature. Heat at the seal faces also causes thermal distortion, which will contribute to increased seal leakage. Many applications require some type of cooling. The temperature of the sealing surfaces is a function of the heat generated by the seal, plus the heat gained or lost to the pumpage. The heat generated at the faces from sliding contact is the mechanical power consumption of the seal being transferred into heat. Therefore, Qs  CaNf  C1 1PVfao 2 where Qs  heat input from the seal, Btu/h(W) and C1  0.077 for USCS units and 1 for SI units If the heat is removed at the same rate it is produced, the temperature will not increase. If the amount of heat removed is less than that generated, the seal face temperature will 2.204 CHAPTER TWO FIGURE 7 Unbalanced seal heat generation (John Crane Inc.) increase to a point where seal face damage will occur. The estimated values for heat input are given in Figures 7 and 8. Heat removal from a single seal is accomplished by a seal flush. The seal flush is usually a bypass from the discharge line on the pump or an injection from an external source. The flow rate for cooling can be found by calculating the following: Qs gpm 1m3>h2  C2 1sp. ht.21sp. gr.2¢T where Qs C2 sp. ht. sp. gr. T  seal heat, Btu/h(W)  500 in USCS units and 1,000 in SI units  specific heat of coolant, Btu>lb. # °F 1J>kg # K2  specific gravity of coolant  temperature rise, °F (K) When handling liquids at elevated temperatures, the heat input from the process must be considered in the calculation of coolant flow. Thus, Qnet  Qp  Qs The heat load Qp from the process can be determined from Figure 9. As an example, let’s determine the net heat input for a 4-in (102-mm) diameter balanced seal in water at 1,800 rpm. The pressure and temperature are 400 lb/in2 (27.6 bar) and 170°F (76.7°C). From Figure 8, we have the following: In USCS units: Qs  13500 Btu>h>1000 rpm2  118002  6300 Btu>h 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS 2.205 FIGURE 8 Balanced seal heat generation (John Crane Inc.) In SI units: Qs  11025 W>1000 rpm2  118002  1845 W From Figure 9, assuming that the seal chamber will be cooled to 70°F (21°C) and that the temperature difference between the seal chamber and pumpage is 100°F (37.8°C), we have the following: In USCS units: In USCS units: ⬖ Qp  255 Btu>h Qp  75 W Qnet  6555 Btu>h 11920 W2 The total heat input can be used to estimate the required flow to the seal. When multiple seals are used in a pump seal chamber, the heat load from each seal must be considered as well as any heat soak from the process. Different methods are used to supply cool liquid to the seal chamber (see Figure 10). When the liquid is clean, an internal flush connection at (A) can be used to cool the seal. When the liquid is dirty, an external flush at (B) can be used. This will allow the flush, a bypass from the discharge line, to pass through a filter or centrifugal separator. The seal faces will be flushed with clean, cool liquid. Increased pressure from the flush provides positive circulation and prevents flashing at the seal faces caused by the heat generation. When a pump handles liquids near their boiling point, additional cooling of the seal chamber is required. A typical arrangement to accomplish this is shown in Figure 11. This seal is equipped with a pumping ring and a heat exchanger. The pumping ring acts as a miniature pump, causing the liquid to flow through the outlet piping at the top of the seal 2.206 CHAPTER TWO FIGURE 9 Heat soak from process when water is used for lubrication (John Crane Inc.) FIGURE 10 plug port Cooling circulation to mechanical seal: (A) internal circulation plug port, (B) external circulation 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS FIGURE 11 2.207 High performance boiler feed pump seal with external cooling chamber. The liquid passes through the heat exchanger and returns directly to the faces at the bottom inlet in the end plate. As the liquid is circulated, heat is removed from the seal and seal chamber. A closed loop system is commonly used on hot water pumps. This method is extremely efficient since the coolant is circulated only in the seal chamber and does not reduce the temperature of the liquid in the pump. It should be noted that during periods of shutdown different wear problems might exist because the seal faces may be too cold. The product being pumped may be a solution that can crystallize or solidify at ambient temperatures. For these applications, the seal faces may have to be preheated before starting to avoid damage to the seal. Leakage Leakage is affected by the parallelism of the sealing planes, angular misalignment, coning (negative face rotation), thermal distortion (positive face rotation), shaft runout, axial vibration, and fluctuating pressure. For parallel faces only, which take into account seal geometry only, the theoretical leakage in cubic centimeters per hour can be estimated from the following: Q3  C3  h3 1P2  P1 2>u ln 1R2>R1 2 where C3 h P2 P1 u R2 R1  2.13  1010 in USCS units and 1.88  109 in SI units  face gap, in (m)  pressure at face ID, lb/in2 (N/m2)  pressure at face OD, lb/in2 (N/m2)  dynamic viscosity, Cp (N  s/m2)  outer face radius, in (m)  inner face radius, in (m) Negative leakage indicates flow from the face’s outer diameter to the inner diameter. The effect of centrifugal force from one of the rotating sealing planes is very small and can be neglected in normal pump applications. The gap between the seal face is a function of the materials of construction, flatness, and the liquid being sealed. The face gap can range from 20  106 to 50  106 in (0.508  106 to 1.27  106 m). Contacting Seal Operating Envelope Every seal has an operating envelope. The basic envelope for a contacting seal is shown in Figure 12. The upper limits are defined by wear of the seal faces, usually defined by a pressure-velocity limit. The fluid being sealed should be cooled so the liquid at the seal faces does not flash. Operating within the envelope will result in excellent seal performance. 2.208 CHAPTER TWO FIGURE 12 Operating envelope for a contacting seal CLASSIFICATION OF SEALS BY ARRANGEMENT _________________________ Sealing arrangements can be classified into two groups: 1. Single seal installations a. Internally mounted b. Externally mounted 2. Multiple seal installations a. Double seals b. Tandem seals Single seals are used in most applications. This is the simplest seal arrangement with the least number of parts. An installation can be referred to as inside-mounted or outsidemounted, depending on whether the seal is mounted inside or outside the seal chamber (see Figure 13). The most common installation is an inside-mounted seal. Here the liquid under pressure acts with the spring load to keep the seal faces in contact. Outside-mounted seals are considered to be used for low-pressure applications since both seal faces, the primary ring and mating ring, are put in tension. This limits the pressure capability of the seal. An external seal installation is used to minimize corrosion that might occur if the metal parts of the seal were directly exposed to the liquid being sealed. Multiple seals are used in applications requiring • A neutral liquid for lubrication • Improved corrosion resistance • A buffered area for plant safety Double seals consist of two single seals back to back, with the primary rings facing in opposite directions in the seal chamber. The neutral liquid, at a pressure higher than that of the liquid being pumped, lubricates the seal faces (see Figure 14). The inboard seal keeps the liquid being pumped from entering the seal chamber. Both inboard and outboard seals prevent the loss of neutral lubricating liquid. 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS FIGURE 13 2.209 Single seal installations: a) outside mounted, b) inside mounted FIGURE 14 Double Seals Double seals can be used in an opposed arrangement. Two seals are mounted face to face, with the primary sealing rings rotating on a common mating ring (see Figure 15). In this case, the neutral liquid is circulated between the seals at a pressure lower than that of the process fluid. This pressure is limited since the outboard seal faces are in tension. The inboard seal is similar to a single inside-mounted seal and carries the full differential pressure of the seal chamber to the neutral liquid. The outboard seal carries only the pressure of the neutral liquid to the atmosphere. The purpose of this arrangement is to fit a seal installation having a shorter axial length than is possible with back-to-back double seals and still form a buffered area for plant safety. Tandem seals are arranged with two single seals mounted in the same direction (see Figure 16). The outboard seal and neutral liquid create a buffer zone between the liquid being pumped and the atmosphere. Normally, the pressure differential from the liquid 2.210 CHAPTER TWO FIGURE 15 Opposed double seals FIGURE 16 Tandem seals being sealed and atmosphere is taken across the inboard seal, while the neutral lubricating liquid is at atmosphere pressure. This arrangement can also be used as a method to break down the pressure on high-pressure applications. For example, the pressure difference across each seal can be half the fluid pressure being sealed. The liquid in the outboard seal chamber may be circulated to remove seal heat. Tandem seals are used on toxic or flammable liquids, which require a buffered or safety zone. Package or cartridge seals are an extension of other seal arrangements. A package seal requires no special measurements prior to seal installation. For a single seal, the seal package consists of the gland plate, sleeve, and drive collar (see Figure 17). A spacer is provided on most package seals to properly set the seal faces. The spacer is removed after the drive collar has been locked to the shaft and the gland plate bolted to the pump. CLASSIFICATION OF SEALS BY DESIGN ________________________________ There are four seal classification groups: • Unbalanced or balanced • Rotating or stationary seal head 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS FIGURE 17 2.211 A single package (cartridge) seal assembly (John Crane Inc.) • Single-spring or multiple-spring construction • Pusher or nonpusher secondary seal design The selection of an unbalanced or balanced seal is determined by the pressure in the seal chamber. Balance is a way of controlling the contact pressure between the seal faces and power generated by the seal. When the percentage of balance b (the ratio of hydraulic closing area to seal face area) is 100 percent or greater, the seal is referred to as unbalanced. When the percentage of balance for a seal is less than 100 (1.0), the seal is balanced. Figure 18 illustrates common unbalanced and balanced seals. The selection of a rotating or stationary seal is determined by the speed of the pump shaft. A seal that rotates with the shaft is a rotating seal assembly. Typical rotating seals are shown in Figures 17, 21 and 22. When the mating ring rotates with the shaft, the seal is stationary (see Figure 19). Rotating seal heads are common in the industry for normal pump shaft speeds. As a rule of thumb, when the shaft speed exceeds 5,000 ft/min (25.4 m/s), stationary seals are required. Higher speed applications require a rotating mating ring to keep unbalanced forces, which may result in seal vibration, to a minimum. A stationary seal should be considered for all split case pumps. This will eliminate seal problems that occur when the top and bottom halves of the pump casing do not line up. The pressure in the pump can cause a misalignment of these parts that creates an out-of-square condition at the seal faces. The selection of a single-spring or multiple-spring seal head construction is determined by the space limits and the liquid sealed. Single-spring seals are most often used with bellows seals to load the seal faces (see Figure 20a). The advantage of this type of construction is that the openness of design makes the spring a nonclogging component of the seal assembly. The coils are made of a large diameter spring wire and therefore can withstand a great deal of corrosion. Multiple-spring seals require a shorter axial space. Face loading is accomplished by a combination of springs placed about the circumference of the shaft (refer to Figure 1 and see Figure 20b). Most multiple-spring designs are used with assemblies having O-rings or wedges as secondary seals. Pusher-type seals are defined as seal assemblies in which the secondary seal is moved along the shaft by the mechanical load of the seal and the hydraulic pressure in the seal chamber. The designation applies to seals that use an O-ring, wedge, or V-ring. A typical construction is illustrated in Figure 21. The primary ring, with a hardened metal surface, rotates with the shaft and is held against the stationary ring by the compression ring through loading of the O-ring. The 2.212 CHAPTER TWO FIGURE 18 Common unbalanced and balanced seals compression ring supports a nest of springs that is connected at the opposite end by a collar, which is fixed to the shaft. The primary ring is flexibly mounted to take up any shaft deflection or equipment vibration. The collar is fixed to the shaft by setscrews. Another pusher-type seal is illustrated in Figure 22. When elastomers cannot be used in the product, a wedge made of TFE must be considered. A metal retainer locked to the shaft by (A) provides a positive drive through the shaft and to the primary ring (F) through drive dents (D), which fit corresponding grooves. The seal between the primary ring and shaft or sleeve is made by a wedge (E), which is preloaded by multiple springs (B). The spring load is distributed uniformly by a metal disc (C). The primary ring (G) contacts the mating ring (H) to form the dynamic seal. Pusher seals also come in split designs. Illustrated in Figure 23 is a split seal design for an ANSI pump. This is a fully split seal design with all of the basic parts fitted outside the seal chamber. The gland plate is fully split and provides easy access to other seal components. A finger spring located on the atmospheric side of the seal provides an axial load and drive to the stationary primary ring. Since it is located on the atmospheric side of the seal, it will not be clogged from material in the pumpage. This is suited for a variety of applications, including paper stock, sewage, slurries, and river water. Two flush ports in the gland plate provide for a seal flush for cooling. 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS FIGURE 19 FIGURE 20 Stationary seal with a rotating mating ring (John Crane Inc.) Comparison of a) single-spring and b) multiple-spring seals FIGURE 21 O-ring type mechanical seal (Flowserve Corp.) 2.213 2.214 CHAPTER TWO FIGURE 22 Wedge-type mechanical seal (John Crane Inc.) FIGURE 23 A split seal design for an ANSI pump (John Crane Inc.) FIGURE 24 A half-convolution bellows seal (John Crane Inc.) Nonpusher seals are defined as seal assemblies in which the secondary seal is not forced along the shaft by the mechanical load or hydraulic pressure in the seal chamber. Instead, all movement is taken up by the bellows convolution. This definition applies to those seals that use half-, full-, and multiple-convolution bellows as a secondary seal. The half-convolution bellows seals are always made of an elastomer (see Figure 24). The tail of the bellows is held to the shaft by a drive band. This squeeze fit seals the shaft and enables the unit to rotate with the shaft. Positive drive is accomplished through the drive band, retainer, and primary ring by a series of slots and dents. A static seal is created at the back of the primary ring and at the front of the bellows. This type of seal is used for light-duty service conditions. The amount of axial travel along the shaft is half that of a full convolution bellows. The full-convolution bellows seal is illustrated in Figure 25. The tail of the bellows is held to the shaft by a drive band. The squeeze fit seals the shaft and enables the unit to rotate with the shaft. The drive for the seal assembly is similar to that of the halfconvolution seal. Static sealing is accomplished at the front of the bellows and the back of the primary ring. The heavier full-convolution bellows design can tolerate greater shaft motion and runout to pressures of 1200 lb/in2 (8.3 bar). Multiple-convolution bellows seals are necessary to add flexibility to those secondary seal materials that cannot be used in any other shape. The mechanical characteristics of TFE and metals require multiple-convolution designs. 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS FIGURE 25 2.215 A full-convolution bellows seal (John Crane Inc.) A TFE bellows assembly is illustrated in Figure 26. Because of the large crosssectional area, these types of seals are mounted outside the seal chamber. Pressure at the inside diameter of the seal helps keep the faces closed. Small springs on the atmosphere side of the seal supply the mechanical load to keep the seal faces closed initially. Multiple convolution metal bellows seals come in various designs and are discussed in the following section. SEALING REQUIREMENTS IN THE PETROLEUM REFINING INDUSTRY _______ The API Standard 682 is an industrial standard developed by users with input from equipment and seal manufacturers. The goal of the standard was to create a specification for seals that would have a good probability of meeting emission standards defined by the U.S. Clean Air Act of 1990 and have a life of at least three years. The implementation of this specification indicates not only an increase in emissions control, but also a major increase in equipment reliability. The following contacting, liquid-lubricated seal designs were identified by API 682, 1st edition (October 1994) as solutions to sealing refinery services. These have been verified by seal manufacturer tests under simulated refinery conditions. These are as follows: • Type A A single, pusher-type seal mounted inside the seal chamber with a rotating flexible element. This is a balanced cartridge design with multiple springs and an Oring as a secondary seal (see Figure 27). This seal is preferred for all refinery services except non-flashing hydrocarbons above 300°F (150°C). It is considered to be the standard for temperatures up to 500°F (265°C). • Type B A single, low-temperature, non-pusher, inside-mounted seal, with a rotating metal bellows flexible element. The secondary static seals for this nickel allow metal bellows design are fluorocarbon elastomer O-rings. This low-temperature seal design is a standard optional selection for non-flashing hydrocarbon services up to 300°F (150°C). • Type C A single, high-temperature, non-pusher, inside-mounted seal with a stationary metal bellows flexible element. The secondary static seals for this hightemperature bellows design are flexible graphite. This seal is the standard selection for non-flashing hydrocarbon applications when temperatures are above 300°F (150°C) and pressures are less than 250 lb/in2 absolute (17 bar). Each of the previous seal types is also available as a dual seal arrangement (see Figure 30). When the space between the inboard and outboard seals is pressurized with a barrier fluid, the seal arrangement is referred to as a pressurized dual seal. When the 2.216 CHAPTER TWO FIGURE 26 A Teflon bellows seal (John Crane Inc.) FIGURE 27 Type A Single mounted pusher seal (John Crane Inc.) FIGURE 28 Type B Single inside mounted nonpusher seal (John Crane Inc.) FIGURE 29 Type C Single inside mounted high temperature seal (John Crane Inc.) FIGURE 30 Pressurized dual API 682 seal for HF alkylatin service (John Crane Inc.) space between the inboard and outboard seals is unpressurized with a buffer fluid, the seal arrangement is referred to as a nonpressurized dual seal. This is the only seal cartridge that can function as a double or tandem seal with the same individual seal parts. The terms barrier fluid and buffer fluid refer to the same fluid lubricating the seal. When the fluid is pressurized, it is a barrier fluid. When the fluid is non-pressurized, it is a buffer fluid. 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS FIGURE 31 2.217 API 682 qualification test rig (John Crane Inc.) The dual seal design shown in Figure 30 is without a pumping ring on the outside diameter of the outboard seal. This figure also represents a successful installation on an HF alkylation unit. In this design, isobutane is circulated at a pressure greater than at the pressure at the outside diameter of the inboard seal. The isobutane is then flushed over the inboard seal to keep the hydrofluoric acid away from the inboard seal. Seal life has been significantly increased with this improved sealing technology. API 682 requires qualfication testing for all seal designs by the seal manufacturer. To meet these requirements, seal manufacturers constructed new testing facilities that allow testing at simulated refinery conditions for common process fluids. Figure 31 shows an API qualification test rig with instrumentation installed. Each seal type from each seal application group is required to be tested in four different test fluids that model fluids that model fluids from the application groups. These fluids include water, propane, 20 percent NaOH solution, and mineral oil. Each qualification test for each test fluid consists of three phases: a. the dynamic phase at constant temperature, pressure, and speed b. the static phase at 0 rpm using the same temperature and pressure as the dynamic phase c. the cyclic phase at varying temperatures and pressures, including start-ups and shutdowns. For flashing hydrocarbons, the cyclic test phase includes excursions into vapor and back to liquid. The seal is expected to perform within the regulated emissions limits after being exposed to qualification testing and upset conditions and demonstrate a capability of at least three years life in service. The result of this effort is not only an improvement in emssions control but also a major improvement in seal reliability. This naturally results in substantially lower life-cycle cost for the user. The success or failure of a seal installation can often be traced to the selection of the proper piping arrangement. A piping arrangement or plan defines how a seal installation will be cooled or, in some cases, heated. Commonly used systems have been defined by API and are shown in Figure 32. Performance testing to qualify seal designs to API 682 has resulted in an improved seal flush required for cooling. The mating ring, chamfered at the outside diameter, enables the flow of the flush liquid not only around the circumferential groove, but also directly to the 2.218 CHAPTER TWO PLAN 01 Fluid being pumped is circulated internally from discharge to seal chamber. Internal recirculation must be sufficient to maintain stable conditions at the seal face. Recommended for clean pumpage only and horizontal pumps. Not recommended for vertical pumps. PLAN 02 Plugged connections for possible future circulating fluid Dead-ended seal chamber with no circulation of a seal flush fluid. Used on special applications with horizontal pumps. Not recommended for vertical pumps. PLAN 11 Fluid pumped is circulated externally from discharge to seal chamber. An orifice may be used to control flow. The flow enters the seal chamber adjacent to the mechanical seal faces. This flow must be sufficient to maintain stable conditions at the seal faces. Not recommended on vertical pumps. PLAN 13 Fluid pumped is circulated from the seal chamber back to pump suction. An orifice may be used to control flow. PLAN 14 Fluid pumped is circulated from discharge to the seal chamber and back to the suction nozzle. An orifice, as shown, may be used to control flow, and must be sized in accordance with the throat bushing and the return line. Similar to Plan 11, flow back to suction side will evacuate vapor that may collect in the seal chamber. Recommended for vaporizing liquid such as those found in light hydrocarbon services. PLAN 21 TI PLAN 23 TI By vendor By purchaser PI TI FIGURE 32 Fluid pumped is moved from the seal chamber by a pumping ring through a heat exchanger and back to the seal chamber. This plan can be used on hot applications to minimize heat load on the heat exchanger by cooling only the small amount of liquid that is recirculated. a dial thermometer (*) may be used in the recirculation line. Plan 21 Fluid pumped is circulated from discharge through a heat exchanger and into the seal chamber. An orifice, as shown, may be used to control flow. A dial thermometer (*) may be used in the recirculation line. Fluid pumped is circulated from discharge through a heat exchanger and into the seal chamber. An orifice, as shown, may be used to control flow. A dial thermometer (*) may be used in the recirculation line. PLAN 31 PLAN 32 Fluid pumped is circulated from discharge through a heat exchanger and into the seal chamber. An orifice, as shown, may be used to control flow. A dial thermometer (*) may be used in the recirculation line. Piping plans for mechanical seals A fluid separate from the pumpage is injected into the seal chamber from an external source. Care must be exercised in selecting an external fluid for injection to provide good lubrication to the seal and eliminate the potential for vaporization and also to avoid contamination of the pumpage with the injected flush. A dial thermometer (*) and flow indicator (*) are optional. 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS PLAN 41 When a hot fluid is pumped which contains suspended abrasive particles, flow from the discharge to a cyclone separator delivers clean flow to the seal chamber through a heat exchanger. An orifice, as shown, may be used to control flow. Solids are delivered to pump suction. Clean discharge to the seal chamber and dirty discharge to pump suction must be at equal pressures. A dial thermometer (*) may be used in the flush line to seals. TI Vent PLAN 52 PI PS Normally open FI PLAN 53 PS Reservoir External Pressure source PI Normally open FI 2.219 Reservoir Applies to an outer seal of an unpressurized dual seal arrangement. An external reservoir provides a buffer fluid which is circulated by an internal pumping ring in the outboard seal cavity during normal operation. The reservoir is usually continuously vented to a vapor recovery system which is maintained at a pressure less than the pressure in the seal chamber. A pressure switch (*) and heat exchanger (*) are optional. Applies to an outer seal of a pressurized dual seal arrangement. An external reservoir provides a barrier fluid under pressure which is circulated by an internal pumping ring in the outboard seal cavity during normal operation. Reservoir pressure is greater than the process pressure. A pressure switch (*) and heat exchanger (*) are optional. PLAN 54 An outboard seal chamber is pressurized by a barrier fluid from an external reservoir. Circulation is by an external pressure system or pump. Reservoir pressure is greater than the process pressure being sealed. PLAN 61 Tapped connections are plugged. When used, the purchaser provides quench fluid (steam, gas, water, etc.) to an auxiliary sealing device. Plugged Inlet Plugged outlet An external source is used to provide a quench which is required to prevent solids from building up on the atmospheric side of the seal. Typically used with a close clearance throttle bushing. PLAN 62 NOTE: This table provides a quick reference to piping plans described in API Standard 610, 8th edition August 1995, for centrifugal pumps for petroleum, heavy duty chemical and gas industry. The reader is encouraged to consult this specific standard for more detailed information. LEGEND Heat exchanger Cyclone separator Check valve Flow indicator Orifice PI Pressure gauge with block valve TI Dial thermometer Flow regulating valve PS Pressure switch with block valve Block valve FIGURE 32 Continued. FI LI Level indicator 2.220 CHAPTER TWO seal faces for cooling. This design also helps to eliminate any trapped vapor at the seal faces (see Figure 33). GLAND PLATE CONSTRUCTION ________________________________________ An essential component of any seal installation is the gland plate. The purpose of this part is to hold either the mating ring assembly or the seal head assembly, depending on whether the seal head is rotating with the shaft or stationary to the pump casing. It is also a pressure-containing component of the installation. The alignment of one of the sealing surfaces, particularly the mating ring used with a rotating seal assembly and a gland plate bushing, is dependent on the fit of the gland plate to the pump. To ensure the proper installation, the API specification requires a register fit with the inside or outside diameter of the seal chamber. The static seal on the face of the seal chamber must be completely confined. Three basic gland plate constructions are shown in Figure 34: • A plain gland plate is used where seal cooling is provided internally through the pump stuffing box and where the liquid to be sealed is not considered hazardous to the plant environment and will not crystallize or carbonize at the atmospheric side of the seal. • A flush gland plate is used where internal cooling is not available. Here coolant (liquid sealed or liquid from an external source) is directed to the seal faces where the seal heat is generated. • A flush-and-quench gland plate is required on those applications that need direct cooling as well as a quench fluid at the atmospheric side of the seal. The purpose of the quench fluid, which may be a liquid, gas, or steam, is to prevent the buildup of any carbonized or crystallized material along the shaft. When properly applied, a seal quench can increase the life of a seal installation by eliminating the loss of seal flexibility due to hangup. This gland plate can also be used for flush, vent, and drain where seal leakage needs to be controlled. Flammable vapors leaking from the seal can be vented to a flare and burned off, while nonflammable liquid leakage can be directed to a safe sump. Figure 35 illustrates some restrictive devices used in the gland plate when quench or vent-and-drain connections are used. These bushings can be pressed in place, as in Figure 35a, or allowed to float as in Figures 35b, c, and d. Floating bushings enable closer running clearances with the shaft because such bushings are not restricted at their outside diameter. The bushing shown in Figure 35d is also split to enable the thermal expansion of the shaft. This restrictive bushing is preferred on refinery applications. Small packing rings can also be used for a seal quench, as shown in Figure 35e. SEAL CHAMBER DESIGN _____________________________________________ A critical part of the sealing system is the seal chamber design. The selection of the proper seal chamber can increase the life of the mechanical seal. Changes in pump design have resulted in three chambers, as shown in Figure 36. These are the conventional seal chambers, referred to as the standard bore, the enlarged bore, and the tapered bore. Seal chambers can influence the seal environment through pressure, solids handling, vapor removal, and temperature. A standard bore seal chamber was originally designed for packing and has a restriction at the bottom of the chamber that limits the interchange of fluids between the chamber and pumpage. This seal chamber is dependent on the application of the proper piping plan to remove heat or abrasives. The enlarged bore is similar to a standard bore, enabling the installation of large seal cross-sections and dual seals. The increased cross-section increases the volume of the liquid for cooling. This seal is also dependent on the proper piping plan to remove heat. The tapered bore seal chamber has increased radial clearance and the bottom of the chamber is exposed to the impeller. The walls of the chamber promote self-venting during shutdown and self-draining during disassembly. Internal flow within the chamber elimi- 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS FIGURE 33 Improved seal flush for refinery services (John Crane Inc.) 2.221 FIGURE 34 Basic gland plate designs: (a) plain gland plate, (b) flush gland plate, (c) flush and quench, or flush vent and drain gland plate nates the need for external piping for cooling. For some applications, operating at a higher pressure will require an external flush. NON-CONTACTING GAS LUBRICATED SEALS ____________________________ The evolution of this sealing concept for pumps has its origin in gas sealing technology developed for gas compressors in the mid-1970s. The development of this technology for 2.222 FIGURE 35 CHAPTER TWO Common restrictive devices used with quench, or vent and drain gland plates. pumps was accelerated in the early 1990s as a way to control emissions. Eliminating the contact between the seal faces while the pump shaft is turning also eliminates the tribological problems of frictional heat and wear. This is no easy task with a liquid present in the pump. Deflections of the seal faces from temperature and pressure must be controlled to very precise levels. Gas, rather than a liquid, must be used as a barrier fluid and one of the seal faces must be designed with a lift mechanism. A lift mechanism can take the shape of a spiral groove, an L-shaped slot, or a controlled wavy surface. When the shaft turns, pressure builds up in the seal faces, which causes the face separation. The basic construction of a spiral groove face and the pressure buildup is shown in Figure 37. The separation of the face is very small and could be measured in nanometers, which enables a small amount of gas to flow across the seal face. Since this type of seal is non-contacting, the only heat that is developed is from the shearing of gas at the seal faces. The small amount of gas flow helps cool the seal faces. The temperature rise at the seal faces is just a few degrees, making this a preferred seal for heat-sensitive liquids. The processes occurring at the seal faces are shown in Figure 38. The operating envelope for a non-contacting, gas-lubricated seal for liquid pumping services is shown in Figure 39. Since the rubbing contact at the seal faces has been elim- 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS FIGURE 36 Standard bore, enlarged bore, and tapered bore seal chamber arrangements. 2.223 FIGURE 37 Spiral groove seal face and pressure build up in the grooves (John Crane Inc.) inated, the seal can be operated at the vapor pressure of the liquid being sealed. In addition, no limiting factor exists due to the pressure-velocity relationship and wear at the seal faces. The limiting factor in the application of the seal is the pressure that has an effect on seal face deflection. Dual pressurized, gas-lubricated seals have been designed to fit oversized and small bore seal chambers. An oversized seal chamber that enables a larger cross-section seal and that can handle pressures up to 600 psig (40 bar) is illustrated in Figure 40.Many existing pumps in the field have small bore seal chambers and do not require the same pressure capability as a larger cross-section seal. The seal to fit these units is illustrated in Figure 41. This type of seal is being used to pressures of 230 psig (16 bar). Dual seals are pressurized with an inert gas. Gas pressure is normally 20 to 30 psig (1.4 to 2 bar) above the process liquid being sealed. The amount of gas consumed through a dual seal can be estimated from Figure 42. The gas consumption is the sum of the flow across the inboard and outboard seals. The amount of gas consumed on an annual basis is very small. This makes this type of installation very economical when compared to a fully pressurized liquid barrier/buffer system. Gas pressure and flow are monitored by a safety gas panel like that shown in Figure 43. This technology represents a solution for those sealing problems that affect the cooling and lubrication of a contacting seal by the liquid being sealed. These problems, identified 2.224 CHAPTER TWO FIGURE 38 FIGURE 39 Crane Inc.) The processes occurring at the seal faces The operating envelope for a non-contacting, gas-lubricated seal for liquid pumping services. (John by users as reasons for a short seal life, include a loss of seal flush, dry running, startup without venting, low net positive suction head (NPSH), and cavitation. Cavitation may also result in vibration of the equipment. The vibration limit for a non-contacting gas lubricated seal is 0.4 in/s (10 mm/s). The benefits of this technology are increased performance, emissions control, safety, reliability, and efficiency for a conventional pump fitted 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS FIGURE 40 Non-contacting gas-lubricated seal for pumps with a large bore seal chamber (John Crane Inc.) FIGURE 42 2.225 FIGURE 41 Non-contacting gas-lubricated seal for pumps with a small cross-section seal chamber (John Crane Inc.) Gas consumption through a seal face. (John Crane Inc.) with non-contacting gas lubricated seals. This translates into increased mean time between maintenance (MTBM) and reduced costs of ownership of the equipment. An inert gas, such as nitrogen, is a preferred barrier fluid in refinery and petrochemical industries, while purified air is used in the pharmaceutical and biotech industries. Protecting the environment is the main reason for the development of this technology, but, it is also a primary sealing system used to maintain product purity in the pharmaceutical and biotech industries. Pumping liquid near its vapor pressure represents a challenge to equipment manufacturers and plant operators. Trying to seal this type of application with a contacting seal will result in an inefficient installation. The amount of heat generated would require too much cooling to prevent the liquid from flashing. The only solution is to eliminate the heat generated at the seal faces, allowing the liquid being sealed to flash to a gas and use a noncontacting gas lubricated seal. For those liquids that are dangerous to the environment, a tandem seal arrangement would be used. The space between the seals would be vented to a flair or vapor disposal 2.226 CHAPTER TWO FIGURE 43 FIGURE 44 A safety gas panel for monitoring gas pressure and flow (John Crane Inc.) Early non-contacting gas-lubricated seal design for vaporizing hydrocarbon service (John Crane Inc.) TABLE 2 Typical vaporizing hydrocarbon services (John Crane Inc.) Operating Condition/ Seal Size (in) Liquid Pressure (lb/in2) Temperature (Degree F) Min/Max Speed (rpm) 1.250 1.875 2.875 3.375 5.250 Ethane LNG LNG Ethane LNG 554 327 400 392 850 45 / 125 37 / 125 30 / 125 49 / 125 30 / 125 3,560 3,560 3,560 1,750 3,560 area. Figure 44 represents an early tandem seal arrangement for this type of service. Applying this technology to pumps has resulted in a significant increase in equipment reliability. A list of typical applications for non-contacting gas seals to be applied to vaporizing hydrocarbon liquid services is shown in Table 2. 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS FIGURE 45 2.227 A non-contacting gas-lubricated seal for cryogenic service (John Crane Inc.) Cryogenic liquids represent a similar design challenge in sealing technology. Traditionally, pumps that are used to pump these liquids relied on contacting seal designs. Although these fluids were at cryogenic temperatures, the seals were operating near the boiling point of the liquid. Frictional heat was enough to vaporize or flash the liquid to a gas. This resulted in short seal life. By allowing the liquid to flash to a gas and by using non-contacting gas lubricated seals, seal life has been extended from weeks to years. A non-contacting gas lubricated cryogenic seal is illustrated in Figure 45. Due to the low temperatures involved, a metal bellows is required in the seal design. MATERIALS OF CONSTRUCTION _______________________________________ All component parts of a seal are selected based on their corrosion resistance to the liquid being sealed. The National Association of Corrosion Engineers (NACE) Corrosion Handbook provides corrosion rates for many materials of construction for mechanical seals used with a variety of liquids and gases. When the corrosion rate is greater than two mils (0.05 mm) per year, double seals that keep the hardware items of the seal in a neutral liquid should be selected to reduce corrosion. In this design, only the inside diameter of the mating ring, the primary ring, and the secondary seal are exposed to the corrosive liquid and should be constructed of corrosion-resistant materials, such as ceramic, carbon, and Teflon. Common materials of construction are given in Table 3. Table 4 lists the properties of common seal face materials. The operating temperature is a primary consideration in the design of the secondary and static seals in the assembly. These parts must retain their flexibility throughout the life of the seal, as flexibility is necessary to retain the liquid at the secondary seal as well as to enable a degree of freedom for the primary ring to follow the mating ring. The usable temperature limits for common secondary and static seal materials are given in Table 5. An additional consideration in the selection of the primary and mating ring materials in sliding contact is their PV limitation. This value is an indication of how well the material combination will resist adhesive wear, which is the dominant wear in mechanical seals. Limiting PV values for various face combinations are given in Table 6. Each limiting value has been developed for a wear rate that provides an equivalent seal life of two years. A PV value for an individual application can be compared with the limiting PV 2.228 CHAPTER TWO TABLE 3 Common materials of construction for mechanical seals Components Secondary Seals: O-rings Bellows Materials of Construction Nitrile, Ethylene Propylene, Chloroprene, Fluoroelastomer, Perfluoroelastomer Nitrile, Ethylene Propylene, Chloroprene, Fluoroelastomer Wedge or U Cups Fluorocarbon Metal Bellows Stainless steel, Nickel-base Alloy Primary Ring Carbon, Metal-filled Carbon, Tungsten Carbide, Silicon Carbide, Siliconized Carbon, Bronze Hardware (retainer, disc, snap rings, set screws, springs) Stainless Steel, Nickel-base Alloy Mating Ring Ceramic, Cast Iron, Tungsten Carbide, Silicon Carbide value for the materials used to determine satisfactory service. These values apply to aqueous solutions at 120°F (49°C). For lubricating liquids such as oil, values of 60 percent or higher can be used. Higher or lower values of PV may apply, depending on the seal face design. INSTALLING THE SEAL AND IDENTIFYING CAUSES OF SEAL LEAKAGE _____ A successful seal installation requires operation of the pump within the manufacturer’s specification. Relative movement between the seal parts or shaft sleeve usually indicates that mechanical motion has been transmitted to the seal parts from misalignment (angular or parallel), endplay, or radial runout of the pump (see Figure 46). Angular misalignment results when the mating ring is not square with the shaft and will cause excessive movement of internal seal parts as the primary ring follows the outof-square mating ring. This movement will fret the sleeve or seal hardware on pusher type seal designs. Angular misalignment may also occur from a seal chamber that has been distorted by piping strain developed at operating temperatures. Damage in the wearing rings can also be found here if the pump seal chamber has been distorted. Parallel misalignment results when the seal chamber is not properly aligned with the rest of the pump. No seal problems will occur unless the shaft strikes the inside diameter of the mating ring. If damage has occurred, there will also be damage to the bushing at the bottom of the seal chamber at the same location as the mating ring. Excessive axial endplay can damage the seal surfaces and cause fretting. If the seal is continually being loaded and unloaded, abrasives can penetrate the seal faces and cause premature wear of the primary and mating rings. Thermal damage in the form of heat checking in the seal faces because of excessive endplay can occur if the seal is operated below working height. Radial runout in excess of limits established by the pump manufacturer could cause excessive vibration at the seal. This vibration, coupled with small amounts of the other types of motion that have been defined, will shorten seal life. Instructions and seal drawings should be reviewed to determine the installation dimension or spacing required to ensure that the seal is at its proper working height (see Figure 47). The installation reference can be determined by locating the face of the seal chamber on the surface of the sleeve and then measuring along the sleeve after it has been TABLE 4 Typical properties of common seal face material CERAMIC CARBIDES CARBON Cast Iron Ni-resist 85% (AL2O3) 99% (AL2O3) Tungsten (6% Co) Silicon (SiC) Resin Antimony Bronze SiC Conv. Modulus of Elasticity  106 lb/in2 ( 103 Mpa) 13–15.95 (90–110) 10.5–16.9 72–117) 32 (221) 50 (245) 90 (621) 48–57 (331–393) 2.5–4.0 (17.2–27.6) 3.8–4.8 (26.2–33.1) 2.9–4.4 (20–30) 2-2. (13.8–15.9) Tensile Strength  103 lb/in2 (Mpa) 65–120 (448–827) 20–45 (138–310) 20 (138) 39 (269) 123.25 (8;50) 20.65 (142) 4.5–9 (31–62) 7.5–9.0 (52–62) 7.5–9 (52–62) 2 (14) Coefficient of Thermal Expansion  106 in/in F (cm/cm K) 6.6 (11.88) 6.5–6.8 (11.7–12.24) 3.9 (7.02) 4.3 (7.74) 2.53 (4.55) 1.88 (4.55) 2.3.–3.4 (4.14–6.12) 2.3.–4.7 (4.14–8.46) 2.4–3.1 (4.32–5.58) 2.4–3.2 (4.32–5.76) Thermal Conductivity Btu  ft/h  ft2 ºF (w/m  k) 23–29 (39.79–50.17) 25–28 (43.25–48.44) 8.5 (14.70) 14.5 (25.08) 41–48 (70.93–83.04) 41–60 (70.93–103.8) 3.8–12 (6.57–20.76) 5.8–9.0 (10.0–15.6) 8–8.5 (13.84–14.70) 30 (51.9) Density: lb/in3 (kg/m3) 0.259–0.268 (7169–7418) 0.264–0.268 (7307–7418) 0.123 (3405) 0.137 (3792) 0.50 (16.331) 0.104 (2879) 0.064–0.069 (1771–1910) 0.083–0.112 (2297–3100) 0.083–0.097 (2297–2685) 0.067–0.070 (1854–1938) 80–105 75–100 Property 2.229 Brinell Hardness 217–269 Rockwell A 131–183 87 87 Rockwell 45N 92 86–88 Shore Rockwell 15T 70–92 90 2.230 CHAPTER TWO TABLE 5 Temperature limits of secondary seal materials removed from the unit. It is not necessary to use this procedure if a step in the sleeve or collar has been designed into the assembly to provide for proper seal setting. Assembling other parts of the seal will bring the unit to its correct working height. All package or cartridge shaft seals can be assembled with relative ease because just the bolts at the gland plate and set screws on the drive collar need to be fastened to the seal chamber and shaft. After the seal spacer is removed, the unit is ready to operate. To assemble a mechanical seal to a pump, a spacer coupling is required. If the pump is packed but may later be converted to mechanical seals, a spacer coupling should be included in the pump design. Since a seal has precision-lapped faces and because secondary seal surfaces are critical in the assembly, installations to the equipment should be kept as clean as possible. All lead edges on sleeves and glands should have sufficient chamfers to facilitate installation. When mechanical seals are properly applied, there should be no static leakage and, under normal conditions, the amount of dynamic leakage should range from none to just a few drops per minute. Under a full vacuum, a mechanical seal is used to prevent air from leaking into the pump. If excessive leakage occurs, the cause must be identified and corrected. Causes for seal leakage with possible corrections are listed in Table 7. In addition, Figure 48 illustrates the most common causes for mechanical seal leakage. Further information on seal leakage and the related condition of seal parts can be found in the works listed in the “Further Reading” section. 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS 2.231 TABLE 6 Frequently used seal face materials and their PV limitations Sliding Materials PV limit, lb/in2  ft/min (bar  m/s) Rotating Stationary Comments Carbon-graphite Ni-resist 100,000 (35.03) Better thermal shock resistance than ceramic Ceramic (85% Al2O3) 100,000 (35.03) Poor thermal shock resistance and much better corrosion resistance than Ni-resist Ceramic (99% Al2O3) 100,000 (35.03) Better corrosion resistance than 85% Al2O3 Ceramic Tungsten Carbide (6% Co) 500,000 (175.15) With bronze-filled carbon graphite, PV is up to 100,000 lb/in2 ft/min (35.02 bar  m/s) Tungsten Carbide (6% Ni) 500,000 (175.15) Ni binder for better corrosion resistance Silicon Carbide (converted Carbon) 500,000 (175.15) Good wear resistance; thin layer of SiC makes relapping questionable Silicon Carbide (solid) 500,000 (175.15) Better corrosion resistance than Tungsten Carbide but poorer thermal shock resistance Carbon-graphite 500,000 (17.51) Low PV, but very good against face blistering Ceramic 10,000 (3.50) Good service on sealing paint pigments Tungsten Carbide 120,000 (42.04) PV is up to 185,000 lb/in2 ft/min (64.8 bar  m/s) with two grades that have different % of binder Tungsten Carbide/ Silicon Carbide (solid) 300,000 (105.1) Excellent abrasion resistance. Commonly used on high temperature applications Silicon Carbide (converted carbon) 500,000 (175.15) Excellent abrasion resistance, more economical than solid Silicon Carbide Silicon Carbide (solid) 350,000 (122.6) Excellent abrasion resistance, good corrosion resistance and moderate thermal shock resistance 2.232 CHAPTER TWO FIGURE 46 Common types of motion that influence seal performance FIGURE 47 Typical installation reference dimensions 2.233 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS TABLE 7 Checklist for identifying causes of seal leakage Symptom Possible Causes Corrective procedures Seal spits and sputters (“face popping”) in operation Seal fluid vaporizing at seal interfaces Increase cooling of seal faces. Check for proper seal balance with seal manufacturer Add bypass flush line if not in use Enlarge bypass flush line and/or orifices in gland plate Check for seal interface cooling with seal manufacturer Seal drips steadily Faces not flat Carbon graphite seal faces blistered Seal faces thermally distorted Check for incorrect installation dimensions Check for improper materials or seals for the application Improve cooling flush lines Check for gland plate distortion due to overtorquing of gland bolts Check gland gasket for proper compression Clean out foreign particles between seal faces; relap faces if necessary Check for cracks and chips at seal faces; replace primary and mating rings Secondary seals nicked or scratched during installation O-rings overaged Secondary seals hard and brittle from compression set Secondary seals soft and sticky from chemical attack Spring failure Hardware damaged by erosion Drive mechanisms corroded Replace secondary seals Check for proper lead-in chamfers, burrs, and so on Check for proper seals with seal manufacturer Check with seal manufacturer for other material Replace parts Check with seal manufacturer for other material (continues) 2.234 CHAPTER TWO TABLE 7 Continued. Symptom Possible Causes Corrective procedures Seal squeals during operation Amount of liquid inadequate to lubricate seal faces Add bypass flush line if not in use Carbon dust accumulates on outside of gland ring. Amount of liquid inadequate to lubricate seal faces Liquid film evaporating between seal faces Enlarge bypass flush line and/or orifices in gland plate Add bypass flush line if not in use Enlarge bypass flush line and/or orifices in gland plate Check for proper seal design with seal manufacturer if pressure in stuffing box is excessively high Seal leaks Nothing appears to be wrong Refer to list under “Seal drips steadily” Check for squareness of stuffing box to shaft Align shaft, impeller, bearing, and so on to prevent shaft vibration and/ or distortion of gland plate and/or mating ring Seal life is short. Abrasive fluid Prevent abrasives from accumulating at seal faces Add bypass flush line if not in use Use abrasive separator or filter Seal running too hot Increase cooling of seal faces Increase bypass flush line flow Check for obstructed flow in cooling lines Equipment mechanically out of line Align Check for rubbing of seal on shaft 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS FIGURE 48 Identifying causes of seal leakage (John Crane Inc.) 2.235 2.236 CHAPTER TWO FIGURE 48 Continued. FURTHER READING __________________________________________________ Abar, J. W. “Failures of Mechanical Face Seals.” in Metals Handbook, Vol. 10, 8th ed., American Society for Metals. Metals Park, OH, 1975. American Petroleum Institute. “Centrifugal Pumps for General Refinery Services.” API Standard 610, 8th ed., Washington, DC, 1995. Crane Packing Company. Engineered Fluid Sealing: Materials, Design and Application. Morton Grove, IL, 1979. Crane Packing Company. Identifying Causes of Seal Leakage, S-2031 and Bulletin. Morton Grove, IL, 1979. 2.2.3 CENTRIFUGAL PUMP MECHANICAL SEALS 2.237 Gabriel, R. P., and Niamathullah, S. K. “Design and Testing of Seals to Meet API 682 Requirements.” Proceedings of the 13th International Pump Users Symposium, Turbomachinery Laboratory, Texas A&M University. College Station, TX, Mar. 1996. Ganzon, N. “Seal Chamber Design Affects Reliability, Emissions.” Pumps and Systems. Nov. 1995. Hamaker, J. B. “Mechanical Seal Lubrication Systems.” 1977 ASLE Education Program Fluid Sealing Course, American Society of Lubrication Engineers and Crane Packing Company. Morton Grove, IL, May 1977. Hamner, N. E. (compiler). Corrosion Data Survey, 5th ed. National Association of Corrosion Engineers. Houston, TX, 1975. Morrissey, C. P. “A New Shaft Sealing Solution for Small Cryogenic Pumps,” 51st Annual Meeting Society of Tribologists and Lubrication Engineers. Cincinnati, OH, May 1996. Netzel, J. P. “Mechanical Seals for Biochemical and Sterile Processes.” American Society of Mechanical Engineers, Bioprocessing Equipment Design Conference. Charlottesville, VA, October 1993. Netzel, J. P. “Sealing Solutions.” Plant Engineering and Maintenance. Feb. 1991. Netzel, J. P. “Sealing Technology, A Control for Industrial Pollution,” Lubrication Engineering. pp. 483–493, 1990. Netzel, J. P. “Surface Disturbances in Mechanical Face Seals From Thermoelastic Instability.” American Society of Lubrication Engineers, 35th Annual Meeting. Anaheim, CA, May 5, 1980. Netzel, J., and Wray, M. “Improving Equipment Reliability in the Petroleum Refining Industry.” Conference Proceedings, Methods, Strategies, and Technologies to Reduce Total Equipment Ownership Costs, Aramco Service Company. Houston, TX, Oct. 1997. O’Brien, A. and Wasser, J. R. “Design and Application of Dual Gas Seals for Small Bore Seal Chambers.” 14th International Pump Users Symposium, Turbomachinery Laboratory, Texas A&M University. College Station, TX. Mar. 1997. Schoenherr, K. “Design Terminology for Mechanical Face Seals.” SAE Transactions 74 (650301), 1966. Schoenherr, K. and Johnson, R. L. “Seal Wear.” Wear Control Handbook. (M. Peterson and W. Winer, eds.) American Society of Mechanical Engineers. New York, 1980. Snapp, R. B. “Theoretical Analysis of Face Type Seals with Varying Radial Face Profiles.” 64-WA/LUB 6, American Society of Mechanical Engineers. New York, 1964. STLE. “Guidelines for Meeting Emission Regulations for Rotating Machinery with Mechanical Seals.” SP-30, Society of Tribololgists and Lubrication Engineers. Park Ridge, IL, Apr. 1994. Wasser, J. R. “Dry Seal Technology for Rotating Equipment.” 48th Annual Meeting Society of Tribologists and Lubrication Engineers. Calgary, Alberta, Canada, May 1993. Wasser, J. R., Sailer, R., and Warner, G. “Design and Development of Gas Lubricated Seals for Pumps.” Proceeding of the Eleventh International Pump Users Symposium, Turbomachinery Laboratory, Texas A&M University. College Station, TX, March 1994. 2.2.4 CENTRIFUGAL PUMP INJECTION-TYPE SHAFT SEALS ROBERT D. ASHTON TIMOTHY L. WOTRING Injection-type shaft seals (sometimes called packless stuffing boxes) are designed to control leakage from hot-water pumps. Cool water is injected into each seal to either suppress or regulate the hot leakage, which would otherwise flash upon reaching the outside of the pump. Injection-type shaft seals provide high reliability and yet require little maintenance. They are used primarily in power plant boiler-feed and reactor-feed pump applications where shaft peripheral speeds are high (3600 rpm and up) and pumping temperatures are greater than 250°F (120°C). Under these conditions, conventional packing or mechanicalseal-type stuffing boxes may not be suitable or desirable. Injection shaft seals are either serrated throttle bushings or floating ring seal designs that regulate the flow, temperature, and pressure of the controlled leakage. The flow of the cool injection and of any hot water in the seals depends on operating pressures, but it is restricted by close seal clearances and is regulated by injection control valves. The operating temperature of the seals is controlled by allowing cool injection water to surround the outside of the seal. Ports in the seal enable the cool injection water access to the shaft to either overcome or mix with hot water in the seal so that the resulting seal leakage is cool. The seal designs must provide sufficient pressure breakdown between the pump suction or balance device chamber pressures inside the pump and the atmospheric conditions outside the pump. Upon reaching the outside of the seal, the cool leakage is piped away by gravity drain for eventual return to the power plant feedwater system. SERRATED THROTTLE BUSHINGS ______________________________________ The construction of a serrated (sometimes called labyrinth or grooved) bushing (see Figure 1) varies from one pump manufacturer to another. However, the serrated designs basically involve a rotating shaft running with a reasonably small clearance, such as 0.002 to 2.239 2.240 CHAPTER 2 FIGURE 1 ***Author: Replacement caption was not provided*** 0.003 in per inch (0.002 to 0.003 mm per millimeter) of the shaft diameter within a solid stainless steel (typically 12 percent chrome) stationary bushing installed in the pump casing end cover. Grooved serrations are applied to the hardened stainless steel rotating surface (usually a shaft sleeve) or to the stationary bushing or to both rotating and stationary surfaces to effect a high throttling action or stuffing box leakage reduction. The serrations create a more effective pressure breakdown than a smooth axial surface and enable increased running clearances to let foreign particles, such as grit, pass through or flush out in the grooves. The increased clearances also enable more tolerance for any displacement between rotating and stationary components that might result from assembly misalignment or from distortions induced by transient temperatures. The grooved running surface area at clearances is greatly reduced relative to that of a smooth axial surface, and therefore possible metal surface contact between rotating and stationary parts is reduced during transient operations. Seal leakage could be reduced by decreasing the running clearance in the injection seal, but the smaller clearance would limit the capability of the serrated seal to conform to radial shaft movement. Reduced clearances in any injection seal design may lead to problems with thermal distortions caused by a loss of available cool injection water and may also increase galling by particles trapped in the seal running fits. FLOATING RING DESIGN ______________________________________________ A segmented throttle bushing made of many floating rings will enable conformity with the radial shaft movement and smaller running clearances to reduce seal leakage. The construction of a floating ring stuffing box (see Figure 2) varies from one pump manufacturer 2.2.4 CENTRIFUGAL PUMP INJECTION-TYPE SHAFT SEALS 2.241 FIGURE 2 Floating ring seal design with shaft sleeve. Pegs prevent rotation of spring-loaded rings but let them float radially. (From Power, August 1980, McGraw-Hill, New York, copyright 1980) to another, but the unit basically consists of a stack of hardened, martensitic stainless steel rings (typically 17 percent chrome), instead of a stationary bushing. The rings run against a smooth, hardened, martensitic stainless steel (typically 12 to 17 percent chrome) rotating shaft sleeve surface. The separate rings are all contained within a housing installed in the pump casing end cover. Each ring is loaded against an adjacent stainless steel ring spacer so that a stationary seal is produced in the axial direction, but the ring is still able to move radially with the shaft. An axial loading of the rings is provided by hydraulic pressure during operation and by springs during idle pump periods. The rings are locked against rotation usually by pegs and/or a slot arrangement. A small radial clearance, usually 0.001 to 0.0015 in per inch (0.001 0.0015 mm per mm) of the shaft diameter, is provided between the rings and the shaft sleeve to provide an adequate throttling of seal leakage across the limited number of seal rings. The length of each ring varies with the diameter of the injection-type seal but is generally about 0.50 in (13 mm). The radial clearance enables a reduced injection flow but increases the sensitivity of the seal to foreign particles in the seal leakage. The individual seal rings are permitted to move radially (float), finding an equilibrium running position relative to the shaft. The capability to float up to .016 in (.4 mm) radially increases the tolerance for shaft misalignment as the shaft passes through the seal area. The multiplicity of seal rings further reduces the effect of any angular displacement between the rotating and stationary components that might arise from errors in the original assembly or from distortions caused by temperature changes during pump load variations. Some injection-type shaft seal designs use a combination of serrations and floating rings in the same stuffing box. For either type of seal, careful maintenance during the assembly or disassembly requires accurate alignment of the rotating and stationary components as well as careful handling to avoid scratching the components. Cleanliness is of utmost importance with shaft seals. CONDENSATE INJECTION REQUIREMENTS ______________________________ Because boiler-feed and reactor-feed pumps normally have high feedwater temperatures, 250 to 500°F (120 to 260°C), with consequent vapor pressures higher than external atmospheric conditions, the seal leakage must be cooled to avoid flashing in the stuffing box or outside the pump. This cooling is usually accomplished by injecting cold condensate from a power plant condensate pump directly into the seal to cool the stuffing box components as well as the seal leakage. The amount of condensate injection water required will depend on several factors, including (1) whether the design is a serrated or floating ring, (2) the type of seal control 2.242 CHAPTER 2 system, (3) the diameter, clearance, and rotating speed of the running fit, and (4) the internal pump pressure and temperature. A typical example of stuffing box flows would be a 5500-rpm, boiler-feed pump with a temperature-controlled seal system (described later) and a serrated 5-in (127-mm) diameter running seal with about 0.015 inches (0.38 mm) of diametrical clearance. The approximate flows would be as follows: 1. Total injection per seal: 5 to 15 gpm (0.3 to 1.0 l/s) 2. Leakage from internal pump to seal: 0 to 5 gpm (0 to 0.3 l/s) 3. Drainage from each seal: 8 to 20 gpm (0.5 to 1.3 l/s) Given a seal with a floating ring design (having typically half the clearance of seals with a serrated bushing design), the previous conditions would require only 50 to 60 percent of the injection water needed by the serrated bushing design with the larger clearance. During pump standby periods, higher injection and leakage flows are required because of reduced seal throttling caused by low or zero speed, by high internal pump pressures, or by a pump balance device that induces higher internal leakage during standby. The injection flows to both stuffing boxes on a multistage pump may not be the same as a result of pump balance action at one end of the pump. Cold condensate, 85 to 110°F (30 to 45°C), is usually available from the power plant condensate pump or secondary condensate booster pump discharge. The variation in condensate supply pressure relative to internal boiler-feed pump pressure makes it necessary to use injection control systems for satisfactory seal operation. Condensate injection systems vary to accommodate the wide range of pressures and temperatures found in different power plant feedwater system layouts. The four basic types of injection flow systems are (1) manual, (2) differential pressure-controlled, (3) differential temperature-controlled, and (4) constant drain temperature control. Manual System Manual flow control requires setting and readjusting a valve by hand for each stuffing box each time a change occurs in pump operating conditions due to varying plant loads. Although it is possible to use such a control system, it is not usually recommended because of the inability to automatically compensate for rapid changing conditions. Differential Pressure-Controlled System The differential pressure-controlled shaft seals operate as cold condensate maintained at a pressure greater than boiler-feed pump suction is injected into the central portion of the seal. By maintaining an injection pressure above the pump seal chamber pressure, a small portion of this injection water flows into the pump proper, and most of the injected water flows out of the seal into a collection chamber adjacent to the pump bearing bracket. The leakage in the collection chamber, which is vented to the atmosphere, is drained by gravity for an eventual return to the main feedwater system. The differential pressure-controlled system has a pneumatic injection control valve. This valve in the condensate injection line is governed by an air signal received from a differential pressure control monitor that maintains the preset pressure differential, 10 to 25 lb/in2 (70 to 170 kN/m2), between the injection pressure and the internal pump pressure. This system is not always favored, however, because of a feedback instability tendency resulting from operating pressure changes affecting the valve position, which changes pressure, and so on. Another unfavorable factor for hot-water service is the introduction of cold condensate into a hot pump at all times, including pump idle periods, affecting pump prewarming conditions. Differential Temperature-Controlled System The differential temperature-controlled system operates by maintaining a seal drain temperature differential of 25°F (14°C) above the injection water temperature. The temperature differential ensures a continuous outleakage of hot boiler feed water. The hot out-leakage is desirable, ensuring that cold water cannot be injected into the pump and thus eliminating thermal distortions that occur with the differential pressure-controlled system. The valve controllers are identical to the controllers used for the differential pressure-controlled system. 2.2.4 CENTRIFUGAL PUMP INJECTION-TYPE SHAFT SEALS 2.243 FIGURE 3 Temperature-controlled system regulating condensate injection to feed pump shaft seals Constant Drain Temperature-Controlled System The constant drain temperaturecontrolled seal system (see Figure 3) controls hot out-leakage by throttling the injection flow to maintain a preset seal drain temperature. This system has a temperature-sensing probe in each seal drain line. Each probe is connected to an indicating temperature controller, which provides an air signal to a pneumatic control valve in the condensate injection line for control of the seal injection flow rate. Electronic control systems are often used where thermocouples or RTD’s sense drain temperatures. The signal output is then processed through an I/P controller that adjusts the position of the pneumatically operated throttle valve as required. The cold condensate is injected into the stuffing box central portion and allowed to mix with hot water entering the seal from inside the pump. The drain temperature is maintained at a preset 140 to 150°F (60 to 66°C) to preclude flashing in the stuffing box or in the drains. Note that with this system some hot water enters the seal. Therefore, cold condensate does not enter the hot pump and does not adversely affect pump warming conditions, especially during extended idle periods. The required condensate injection pressure is at least equal to the internal stuffing box pressure plus interim frictional loss between the condensate supply source and the point of hot-water mixture. Note that this system may enable satisfactory operation even when the condensate supply pressure is nearly equal to boiler-feed pump suction pressure. In addition to providing a rapid response to variations in operating pump conditions, this type of control will always supply just enough injection water to maintain the recommended drainage temperature. Intermediate Leakoff System The intermediate-leakoff shaft seal system (see Figure 4) has many variations but basically uses a bleedoff from a central portion of the stuffing box. This system may be used to reduce internal stuffing box pressure if high boiler-feed pump suction pressure exists. To create a positive leakoff flow, the intermediate bleedoff flow is piped back to a plant feedwater system low-pressure point, such as a plant condenser, heater, or booster pump suction where the pressure is less than the boiler-feed pump suction pressure. However, the back pressure of the leakoff destination must be above the bleedoff vapor pressure to suppress flashing in the leakoff lines. This back pressure may be the leakoff destination pressure or may be created by an orifice or a valve. Cold condensate injection into the stuffing box is controlled by a pressure differential monitor maintaining a preset pressure above the bleedoff pressure (refer to Figure 4). A 2.244 CHAPTER 2 FIGURE 4 Typical intermediate-leakoff shaft seal system. (From Power, September 1980, McGraw-Hill, New York, copyright 1980) stuffing box drain temperature control can also be used. Note that the cold condensate injection pressure need not equal or overcome a high feed pump suction pressure. The condensate injection temperature must still be 85 to 110°F (30 to 45°C) to keep the drain temperature below the flashing condition. Condensate injection shaft seals without an intermediate bleedoff but that are subject to suction pressures in excess of about 250 lb/in2 (1725 kN/m2) must be extremely long for a proper pressure breakdown. Longer shaft seals require thicker pump case end covers, affecting pump cost, and a longer rotating element, which could adversely affect rotor dynamics. The intermediate-leakoff shaft seal is effective where there is high feed pump suction pressure imposed by boiler-feed booster pumps or in a closed feedwater system with no deaerating open heater, wherein a condensate pump discharge can be fully imposed on the feed pump at low plant loads. In plant systems with feedwater heaters between a booster pump and a feed pump, a high pressure condensate from the cooler booster pump is injected into the seal to enable a cooler intermediate leakoff to help prevent flashing (refer to Figure 2). INJECTION SOURCES ________________________________________________ Many power plant feedwater systems with deaerating direct-contact heaters (open cycle) usually have the boiler-feed pump drawing water directly from the deaerator. These open cycles with constant-speed condensate pumps always have cold condensate supply pressures in excess of boiler-feed pump suction pressure because of the increased available condensate pump head at low system flows and the interim system frictional loss at high system flows. If a boiler-feed booster pump or a closed feedwater system with no open heaters and resultant higher boiler-feed pump suction pressures is used, the pump manufacturer may elect to require condensate injection seal water booster pumps or may use an intermediateleakoff packless shaft seal with a condensate injection overcoming only the intermediate leakoff pressure. If the condensate pumps are variable-speed units that enable the condensate injection pressure to drop to an equal feed pump suction pressure at low loads with little or no feedwater system frictional drop, then condensate injection seal water pumps are required. In this situation, at least one pump manufacturer offers an optional pumping 2.2.4 CENTRIFUGAL PUMP INJECTION-TYPE SHAFT SEALS 2.245 ring configuration in their packless seal that can increase the seal water pressure in the stuffing box to overcome any pump suction pressure. AUXILIARY EQUIPMENT _______________________________________________ The condensate injection piping should be conservatively sized based on the maximum injection flow requirements to obtain a low pressure drop between the injection source and the seal injection control valve. These control valves may be equipped with limit stops to prevent full closure and enable a continuous cool injection to the seals under almost all operating conditions. In some installations, isolating lines are furnished around the valves to enable a continued injection flow even during control valve maintenance. The valves can also be designed to remain open during a failure, such as a loss of station air to the pneumatic controls, and to close only with an air supply. Injection control systems that are not properly maintained might result in cold water entering a hot pump. Should this problem occur while a boiler feed pump is in the hot standby mode or when turning gears, thermal gradients will occur, leading to contact among the close running fits within the pump. Proper maintenance and operation of the injection control systems is necessary to ensure reliable operation of the pump itself. The air supply filter regulators for each control must be furnished with relatively dry clean air at station supply pressure. The condensate injection supply to the seals must be clear and free of foreign matter to prevent damage to stuffing box components. It is therefore necessary to install filters in the injection line prior to the control valve. To keep damaging fine mill scale, oxide particles, abrasives, and other materials from entering the small seal clearances, several pump manufacturers recommend 100-mesh (150-micron) dual strainers. If dual strainers with isolating valves are used, each filter can be cleaned without interrupting injection flow during pump operation. Pressure gages should be installed before and after each filter to permit the operator to monitor filter pressure drop. A differential pressure switch and alarm for each filter are preferable to alert the operator to clean the strainer when pressure drop becomes excessive. The condensate injection shaft seals should always be filled with cool water before and during pump operation, even during reverse pump rotation. Some pump manufacturers stipulate that condensate injection must be continuous without any interruption during all operation modes. The clearances in the condensate injection shaft seal may double over the service life of the internal wearing parts. With double clearances, the leakage will approximately double. This factor should be considered when sizing the return drain piping back to the plant condenser if frictional losses are to be kept to a minimum. The drain line should be pitched at least a quarter-inch per foot (20 mm per meter). The collecting chamber at the pump stuffing box is vented to the atmosphere, and the only head available to evacuate the chamber is the static head between the pump and the point of return. This head must always be well in excess of the frictional losses (even after the leakage is doubled). Otherwise, the drains may back up, the collection chambers may overflow, and the adjacent bearing brackets may flood, with subsequent possible intrusion of water into the pump bearings and lubricating oil. The seal collection chambers have especially large connections to assure proper drainage, provided no back pressure exists. Two types of condensate drain systems can be used to dispose of the drain coming from the collecting seal chambers. One system uses traps that are piped directly to the plant condenser if sufficient static head exists for positive drain flow. The second system collects the drain in a condensate storage tank into which various other drains (from other pumps shaft seals and so on) are also directed. As this vented storage tank is under atmospheric pressure, it must be set at a reasonable elevation below the pump centerline so that the static elevation difference will overcome frictional losses in the drain piping. A separate condensate transfer pump, under control of the storage tank liquid level control system, can then pump the condensate drains from the storage tank into the plant condenser. The storage tank should have its own overflow 2.246 CHAPTER 2 protection system that enables outside drainage if, for some reason, proper drainage cannot be achieved. For example, the top of the tank vent pipe should be below the pump centerline to help preclude the possibility of drainage backing up to the level of the pump seal collection chambers. Note that this storage tank should also be large enough for an adequate drainage collection to help prevent backups. PACKLESS SHAFT SEALS WITHOUT INJECTION __________________________ The packless shaft seals that have been so successfully applied to boiler-feed, reactor-feed, and booster pump services are applicable to a number of other services. For instance, they are very suitable for cold-water condensate booster pumps and for high-pressure pumps applied to hydraulic descaling or hydraulic press work. In such services, there is no need to bring in an injection supply water to the breakdown seals (unless the pumped water is not clear and free of gritty material), because the water handled by the pump is already cold with no danger of flashing as it leaves the pump stuffing boxes. 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS WILBUR SHAPIRO PRINCIPLES OF OPERATION___________________________________________ A journal bearing is essentially a viscous pump, and it derives load capacity by pumping the lubricant through a small clearance region. In Figure 1, the fluid is dragged along by FIGURE 1 Two-groove cylindrical bearing. 2.247 2.248 CHAPTER 2 the rotating journal. To generate pressure, the resistance to pumping must increase in the direction of the flow. In the figure, the journal moves to form a converging tapered clearance in the direction of the rotation or flow. The eccentricity e is the total displacement of the journal from its concentric position. The attitude angle g in Figure 1 is the angle between the load direction and the line of centers. Note that, because of the necessity to form a converging wedge, the displacement of the journal is not along a line that is coincident with the load vector. A positive pressure is produced in the converging region of the clearance. Downstream from the minimum film thickness, which occurs along the line of centers, the film becomes divergent. The resistance decreases in the direction of pumping, and either negative pressures occur or the air in the lubricant gasifies or cavitates and a region of atmospheric pressure occurs in the bearing area. This phenomenon is known as fluid film bearing cavitation. It should be clearly distinguished from other forms of cavitation that take place in pumps, such as in the impeller, for example. Here the fluid is traveling at a high velocity and the inertia forces on each fluid element dominate. Implosions occur in the impeller and can cause damage. In a bearing, the viscous forces dominate and each fluid particle moves at a constant velocity in proportion to the net shearing forces on it. Thus, cavitation in a bearing is more of a change of the phase of the lubricant that occurs in a region of lower pressure that permits the release of entrained gases. Generally, bearing cavitation does not cause damage. Regimes of Lubrication Whether or not a fluid film can be formed, journal rotation is dependent on several factors including the surface speed, viscosity, and load capacity. A parameter1 is often used to determine a particular regime of lubrication, ZN>P, where Z  viscosity of lubricant, cP (Pa  s) N  rotating speed, rpm P,  average pressure of the bearing, lb/in2 (bar)* A plot of the coefficient of friction versus ZN>P generally has the form shown in Figure 2. At low values of ZN>P, a combination of viscosity, speed, and load places a bearing in a boundary lubricated regime where typical coefficients of friction are 0.08 to 0.14. Boundary lubrication implies intimate contact between the opposed surfaces. As the value of the parameter increases as a result of the increased speed, increased viscosity, or lowered load, there is a dramatic reduction in the coefficient of friction. In this region, there is a mixed film lubrication and the coefficient of friction varies between 0.02 and 0.08. By mixed film lubrication, it is meant that the journal is partly surrounded by a fluid film and is partly supported by rubbing contact between the opposed members. As ZN>P increases FIGURE 2 Coefficient of friction versus ZN>P (Ref. 1). *1 bar  105 Pa. For a discussion of bar, see “SI Units: A Commentary” in the front matter. 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS 2.249 further, a situation of full fluid film lubrication prevails. A general rule of thumb is that ZN>P should be 30 (0.44) or greater for a fluid film to be generated. Note that as ZN>P continues to increase beyond the full film demarcation, the coefficient of friction rises, but at a relatively low rate and generally remains in regions of low coefficients of friction. EXAMPLE Z ⫽ 30 cP 10.03 Pa # s2 N ⫽ 150 rpm P ⫽ 200 lb>in2 113.6 bar2 ZN 30 ⫻ 150 ⫽ 22.5 ⫽ 200 P The bearing is not fluidborne and is operating in the mixed film regime. At what speed will the bearing become hydrodynamic? For hydrodynamic operations, ZN>P  30. Therefore, N 30P 30 ⫻ 200 ⫽ 200 rpm  30 Z THEORETICAL FOUNDATIONS _________________________________________ The foundation of a fluid film-bearing analysis emanates from the boundary layer theory of fluid mechanics. The governing differential equation was first formulated by Osborne Reynolds in 1886 and is known as Reynolds’ equation in his honor. It has been only in the last 30 years or so that general solutions have been obtained, and this has been primarily due to the use of numerical methods applied to the digital computer. References 2 and 3 go into the details of contemporary numerical solutions and are recommended for those interested in the analytical aspects of lubrication. Principal Assumptions Reynolds’ equation can be derived from the Navier-Stokes equation of fluid mechanics, and a number of textbooks are available that comprehensively describe the derivation.4 The primary assumptions are as follows: • Laminar flow conditions prevail, and the fluids obey a Newtonian shear stress distribution where the shear stress is proportional to the velocity gradient. • Inertial forces, resulting from acceleration of the liquid, are small relative to the viscous shear forces and may be neglected. • The pressure across the film is constant since the fluid films are so thin. • The height of the fluid film is small relative to other geometric dimensions, and so the curvature of the fluid film can be ignored. • The viscosity of the liquid remains constant. In most cases, this is a reasonable assumption since it has been repeatedly demonstrated that, if the average viscosity is used, little error is introduced and the complexity of the analysis is considerably reduced. Derivaton of Reynold’s Equation of Lubrication Assume that the rotating journal has a peripheral velocity U. Consider an elemental volume in the clearance space of the bearing and establish equilibrium (see Figure 3). Note that since inertial forces are neglected, the volume is in equilibrium by the pressure and shear forces acting upon it, so there is no acceleration. As shown in Figure 4, p is the pressure and t is the shear stress acting upon the volume. Summing forces in the x direction 2.250 CHAPTER 2 FIGURE 3 FIGURE 4 Force equilibrium on fluid element. Fluid control volume. FIGURE 5 p dy dz ⫺ a p⫹ Laminar velocity distribution across film. 0p 0t dx dy dz ⫺ t dx dz ⫹ t ⫹ dy dx dz ⫽ 0 0x b a 0y b ⫺ (1) 0p 0t dx dy dz ⫹ dx dy dz ⫽ 0 0x 0y or 0p 0t ⫽ 0x 0y (2) For a Newtonian fluid in a laminar flow, the shear stress is directly related to the velocity gradient with the proportionality constant being the absolute viscosity m (see Figure 5): t⫽m dv dy (3) 0 2v 0t ⫽m 2 0y 0y where n is the fluid velocity. Substituting Equation 3 into Equation 2, we obtain 0p 0 2v ⫽m 2 0x 0y (4) Integrating with respect to y twice produces the following equation: v⫽ 2 1 0p y ⫹ C1y ⫹ C2 m 0x 2 (5) where C1 and C2 are constants of integration. The boundary conditions are v ⫽ 0, y ⫽ 0 and (6) v ⫽ u, y ⫽ h Substituting the boundary conditions of Equation 6 into Equation 5 results in the following expression for y: v uy 1 0p 2 1y ⫺ h2y2 ⫹ 2m 0x 2 (7) 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS 2.251 The velocity in the z direction would be similar, except that the surface velocity term would be omitted because no surface velocity exists in the z direction. In addition, the pressure gradient would be with respect to z. Now let us consider the flow across the film due to this velocity. Note that Equation 7 is the velocity computed in the x direction, which is in the direction of rotation of the journal: h qx ⫽ 冮v x 0 dy ⫽ 冮 h 0 c uy 1 0p 2 d dy 1y ⫺ h2y2 ⫹ 2m 0x 2 (8) After integrating, qx ⫽ ⫺ 1 0p 3 1 h ⫹ uh 12m 0x 2 (9) Note that qx is the flow per unit width across the film. The flow in the axial direction is qz ⫽ ⫺ 1 0p 3 h 12m 0z (10) Now let us consider a flow balance through an elemental volume across the film (see Figure 6). The net outflow through the volume equals the net reduction in volume per unit time: a qx ⫹ 0qx 0qz 0h dxb dz ⫺ qx dz ⫹ a qz ⫹ dzb dx ⫺ qz dx ⫽ ⫺ dx dz 0x 0z 0t (11) Thus, this gives us 0qx 0qz 0h ⫹ ⫽⫺ 0x 0z 0t (12) Substituting Equations 9 and 10 into Equation 12, we obtain ⫺ 0p 0p 0h u 0h 0 0 1 1 a a ⫺ h3 b ⫺ h3 b ⫽ ⫺ 0x 12m 0x 0z 12m 0z 0t 2 0x (13) and the final equation becomes 0h 0h 0 h3 0p 0 h3 0p a b⫹ a b ⫽ 6m ⫹ 12 0x m 0x 0z m 0z 0x 0t (14) Equation 14 is the general form of Reynolds’ equation used for laminar, twodimensional lubrication problems. Reynolds’ equation is a flow balance equation. The left-hand side represents pressureinduced flows in the x and z directions through the differential element. The first term on the right-hand side represents the shear flow of the fluid induced by the surface velocity of the journal u. Note that this term contains the derivative of clearance with respect to distance. If this term is zero, then there is zero pressure produced by hydrodynamic action, FIGURE 6 Flow balance across control volume. 2.252 CHAPTER 2 and the term 0h/0x is the mathematical representation of the tapered wedge. The second term on the right-hand side refers to a time rate of change of the film thickness, which can be translated to a normal velocity of the center of the journal. It produces pressure by a fluid velocity normal to the bearing surfaces that attempts to squeeze fluid out of a restricted clearance space. This phenomenon is called the squeeze film effect in bearing terminology. Since it is proportional to the velocity of the center of the journal, it is the phenomenon that produces viscous damping in a bearing. The solution to Reynolds’ equation (refer to Equation 14) provides the pressure at all points in the bearing. The application of the digital computer has enabled a rapid solution of Reynolds’ equation over a grid network representing the bearing area.2,3 Once the pressures have been obtained, a numerical integration is applied to determine the performance parameters (in other words, the load capacity): w⫽ 冮 冮 prdu dr (15) The flow across any circumferential line is qu ⫽ 1 冯 a ⫺ 12m 0p 3 1 h ⫹ uhb dz r0u 2 (16) The flow across any axial line is qz ⫽ 1 冯 a ⫺ 12m 0p 3 h b rdu 0z (17) The viscous frictional moment is obtained by integrating the shear stress over the area and can be shown to be Mf  冮 冮 c 1r mrv 2 0p h ⫹ d r du dz 0u 2 h (18) where v  journal surface speed, rad/s. Typical computer program output includes the following: • • • • • • • Pressure distribution throughout the grid network Load capacity Side leakage and carryover flows Viscous power losses Righting moments due to misalignments Attitude angles Cross-coupled spring and damping coefficients due to displacements and velocity perturbations of the journal center • Clearance distribution Turbulence Equation 14 is for laminar conditions. For very high speed bearings, operations beyond the turbulent regime may occur and Reynolds’ equation must be modified. The turbulent theory has been developed, and the literature on this topic can enable performance predictions for turbulent bearings.5,6 The onset of turbulence is determined by examining the bearing’s Reynolds number, which is the ratio of inertia to viscous forces and is defined as 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS Re ⫽ where Re r u h m ruh m 2.253 (19)  Reynolds number  fluid density, lb  s2/in4 (kg/m3)  surface velocity, in/s (m/s)  local film thickness, in (m)  viscosity, lb  s/in2 (Pa  s) A reasonable approximation is to use the concentric clearance c for h. In terms of N rpm and journal diameter D, the Reynolds number is Re ⫽ r pDN c 60 m (20) The criterion for turbulence in journal bearings is that Re ⱖ 1000. EXAMPLE As an example, consider the following: Journal diameter D ⫽ 5 in (127 mm) Bearing length L ⫽ 5 in (127 mm) Radial clearance c ⫽ 0.0025 in (0.064 mm) Operating speed N ⫽ 5000 rpm Lubricant viscosity m ⫽ 2 ⫻ 10 ⫺6 lb  s>in2 114 ⫻ 10 ⫺3 Pa  s2 Lubricant density r ⫽ 7.95 ⫻ 10 ⫺5 lb  s2>in4 18.66 ⫻ 10 ⫺11 kg  s2>mm4 2 Re ⫽ r pDN c 60 m ⫽ 7.95 ⫻ 10⫺5 ⫻ p ⫻ 5 ⫻ 5000 0.0025 ⫽ 130 ⫻ 60 2 ⫻ 10⫺6 Thus, the bearing is operating in the laminar regime. To become turbulent (assuming constant viscosity), the operating speed would have to increase to approximately 38,500 rpm. The example cited is for a relatively high-speed, oil-lubricated bearing for pump applications. In general, most pump bearings operate in the laminar regime. Exceptions might occur when water is used as tbe lubricant because it its much less viscous than oil. Evaluation of Frictional Losses It is often desirable to obtain a quick estimate of viscous drag losses that the journal bearings produce. If we consider shear forces again, we return to the laminar flow equation: F ⫽ mA where F m A u h m h (21) ⫽ viscous shear force, lb (N) ⫽ viscosity, lb  s/in2 (Pa  s) ⫽ surface area, in2 (m2) ⫽ journal velocity, in/s (m/s) ⫽ film thickness, in (m) To obtain friction, we multiply both sides of Equation 21 by the journal radius R. Then the viscous frictional moment is M ⫽ mAR u h 2.254 CHAPTER 2 The frictional horsepower loss is FHP ⫽ NM 63,000 where N  rotating speed, rpm M  moment, lb  in (N  m) Also A  surface area of bearing  pDL, in2 (m2) u ⫽ surface speed ⫽ pDN in>s 1m>s2 60 Substituting, we obtain FHP ⫽ m L D3N2 766,000h (22) In obtaining approximate losses for estimation purposes, the concentric clearance c is substituted for the local film thickness h. If we consider the previous example where D  5 in (127 mm), L  5 in (127 mm), c  0.0025 in (0.064 mm), N  5000 rpm, and m  2 ⫻ 10⫺6 lb  s/in2 (14 ⫻ 10⫺3 Pa  s), the horsepower loss is FHP ⫽ 12 ⫻ 10⫺6 2152 152 3 150002 2 1766,0002 10.00252 ⫽ 16.32 hp Note that for thrust bearings, the frictional horsepower loss is FHP ⫽ mN2 OD4 ⫺ ID4 h 6.127 ⫻ 106 (23) where OD ⫽ outside diameter ID ⫽ inside diameter A general rule of thumb is that the frictional horsepower in a thrust bearing is approximately twice that in a journal bearing. BEARING TYPES _____________________________________________________ Cylindrical Bearing The most common type of journal bearing is the plain cylindrical bushing shown schematically in Figure 1. It can be split and have lubricating feed grooves at the parting line. A ramification is to incorporate axial grooves to enable better cooling and to improve whirl stability (described in more detail below in the discussion of cylindrical bearings with axial grooves). The principle advantages of cylindrical bearings are (1) simple construction and (2) a high-load capacity relative to other bearing configurations. This type of bearing also has several disadvantages: • Whirl Instability: This is prone to subsynchronous whirling at high speeds and also at low loads. Whirling is an orbiting of the journal (shaft) center in the bearing, a motion that is superimposed upon the normal journal rotation. The orbital frequency is approximately half the rotating speed of the shaft. The expression half-frequency whirl is commonly used. The reason for the occurrence of this whirl and more details concerning bearing dynamics are presented in the section on bearing dynamics. • Viscous Heat Generation: Because of the generally large and uninterrupted surface area of this bearing, it generates more viscous power loss than some other types. 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS 2.255 • Contamination: The cylindrical bearing is more susceptible to contamination problems than other types because contaminants that are dragged in at the leading edge of the bearing cannot easily dislodge because of the absence of grooves or other escape paths. The advantages of simplicity and load capacity make the plain journal a leading candidate for most applications, but performance should be carefully investigated for whirl instability and potential thermal problems. Cylindrical bearings are generally used for medium-speed (500 in/s [200 mm/sec] surface speed) and medium- to heavy-load applications (250 to 400 lb/in2 [17 to 28 bar] on a projected area). Cylindrical Bearing with Axial Grooves A typical configuration of this type of bearing is a plain cylindrical bearing with four equally spaced longitudinal grooves extending most of the way through the bearing. Usually, a slight land area exists at either end of the groove to force the inlet flow to each groove into the bearing clearance region (see Figure 7), rather than out the groove ends. This configuration is a little less simple than the plain cylindrical bearing, and because the grooves consume some land area, this configuration has less load capacity than the plain bushing. Since oil is fed into each of the axial grooves, this bearing requires more inlet flow but also will run cooler than the plain bushing. The grooves act as convenient outlets for any contaminants in the lubricant, and thus the grooved bearing can tolerate more contamination than the plain cylindrical bearing. In general, this bearing can be considered as an alternate to a plain bearing if the former can correct a whirl or overheating problem. Elliptical and Lobe Bearings Elliptical and lobe bearings have noncircular geometries. Figure 8 shows two types of three-lobe bearings with the clearance distribution exaggerated so that the lobe geometry is easily discernible. An elliptical bearing is simply a twolobe bearing with the major axis along the horizontal axis. The lobe bearing shown in Figure 8a is a symmetric lobe bearing where the minimum concentric clearance occurs in the center of each lobed region. Thus, at the leading edge region, a converging clearance produces positive pressure, but downstream from the minimum film thickness, a divergent film thickness distribution can be found with resulting negative, or cavitation, pressures. The canted lobe in Figure 8b, on the other hand, generally develop positive pressure throughout the lobe because the bearing is constructed with a completely converging film thickness in each lobed region. This design has excellent whirl resistance (superior to that of the symmetric lobe bearing) and a reasonably good load capability. A 2:1 ratio between leading and trailing edge concentric clearance is generally a reasonable compromise with respect to performance. Elliptical and lobe bearings are often used because they provide better resistance to whirls than cylindrical configurations. They do so because they have multiple load-producing pads that assist in preventing large-attitude angles and cross-coupling (see the section on bearing dynamics). Elliptical and lobe bearings are generally used for high-speed, low-load applications where whirls might be a problem. FIGURE 7 Cylindrical bearing with axial grooving. 2.256 CHAPTER 2 FIGURE 8 (a) symmetric lobe bearing and (b) canted lobe bearing. FIGURE 9 Five-pad tilting pad bearing. Elliptical, or two-lobe, bearings generally have poor horizontal stiffness because of the large clearances along the major diameter of the ellipse. The split elliptical configuration, however, is easier to manufacture than the other types because it is two cylindrical bearing halves with material removed along the parting line. Lobe bearings are usually clearance- and tolerance-sensitive. The other types of lobe bearings are complicated to manufacture. Tilting-Pad Bearings Tilting-pad bearings are used extensively, especially in highspeed applications, because of their whirl-free characteristics. They are the most whirlfree of all bearing configurations. An important geometric variable for tilting-pad bearings is the preload ratio, defined as shown in Figure 9. The preload ratio equals Preload ratio  PR  c¿ c ⫺ c¿ ⫽1⫺ c c (24) where c  machined clearance c¿  concentric pivot film thickness The variable c¿ is an installed clearance and is dependent upon the radial position of the pivot. Figure 10 displays two pads. Pad 1 has been installed such that the preload ratio 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS FIGURE 10 2.257 Tilting-pad bearing preload. is less than one. For pad 2, the preload ratio is one. The solid line represents the position of the journal in the concentric position. The dashed portion of the journal represents its position when a load is applied to the bottom pads (not shown). Pad 1 is operating with a good converging wedge, even though the journal is moving away from it. Pad 2, on the other hand, is operating with a completely diverging film, which means that it is totally unloaded. Thus, bearings with installed pad preload ratios of one or greater will operate with unloaded pads, which reduces overall stiffness of the bearing and results in a deterioration of stability because the unloaded pads do not aid in resisting cross-coupling influences. In the unloaded position, they are also subject to flutter instability and to a phenomenon known as leading edge lockup, where the leading edge is forced against the shaft and is maintained in that position by the frictional interaction of the shaft and the pad. This is especially prevalent in bearings that operate with low-viscosity lubricants, such as gas or water bearings. Thus, it is important to design bearings with preload, although for manufacturing reasons it is common practice to produce bearings without preload. Tilting-pad bearings have some other characteristics that are both positive and negative: • They are not as clearance-sensitive as most other bearings. • Because the pads can move, they can operate safely at a lower minimum film thickness than other bearings. • They do not provide as much squeeze film damping as rigid configurations. • Generally, they are more expensive than other bearings. • For high-speed applications, their pivot contacts can be subjected to fretting corrosion. Hybrid Bearings A hybrid bearing, schematically shown in Figure 11, derives a load capacity from two sources: (1) the normal hydrodynamic pressure generation and (2) an external high-pressure supply that introduces oil into recesses machined into the bearing surface via restrictors (orifices or capillaries upstream of the recesses). External pressure significantly enhances load capacity. Also, these bearings have excellent low- or zero-speed load capabilities. They are sometimes used as startup devices to lift off the rotor. When self-sustaining hydrodynamic speeds are attained, the external pressure is shut off. The characteristics of externally pressurized, or hybrid, bearings include the following: • High load and stiffness capabilities • An external flow that assists in cooling 2.258 CHAPTER 2 FIGURE 11 Cross-coupling influences in hybrid bearings. • Their clearances and tolerances are generally more liberal than in hydrodynamic bearings • They require external fluid-supply systems • They are applied when there is not a sufficient generating speed or when a high-load capacity and stiffness are required • They are sometimes applied to prevent a whirl, but rotational speeds can unbalance recess pressures, introduce cross coupling, and promote a whirl. STEADY STATE PERFORMANCE________________________________________ Computer-generated performances have been obtained for most of the bearing types previously discussed. Information has been plotted in a nondimensional format so that no restrictions exist on operating conditions, lubricant properties, and so on. Use of the charts will be subsequently demonstrated by numerical example. Viscosity One of the key parameters in determining the performance of a bearing is the lubricant viscosity. Viscosity characteristics of commonly used Society of Automotive Engineers (SAE) grades of oil are shown in Figure 12. The units of viscosity are microreyns, where the reyn has the units of lb  s/in2 and comes from the ratio of shear stress to the velocity gradient across the film, as indicated by Equation 21. Other units of viscosity are centipoises, Saybolt seconds universal (SSU), and centistokes. The conversion factors are as follows: m1reyns2 ⫽ Z ⫻ 1.45 ⫻ 10 ⫺7 n1centistokes2 ⫽ 0.22 1SSU2 ⫺ 180 SSU Z1centiposes2  n1centistokes2 ⫻ SG 1sp. gr.2 (25) (26) (27) 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS FIGURE 12 2.259 Viscosity characteristics for SAE oil grades. Performance Curves Performance plots have been generated for the following types of bearings: • Two-groove cylindrical bearings • Symmetric three-lobe bearings: Each lobe is offset such that in the concentric position the minimum film thickness in the center of each lobed region is half the machined clearance c (see definition following). The pads are each 110° in the angular extent. • Canted three-lobe bearing: The lobing is canted such that in the concentric position the leading edge clearance was twice the trailing edge and the trailing edge film thickness (minimum) in the concentric position was 0.5c where c equals the machined clearance. The pads are each 110° in the angular extent. • Tilting-pad bearing: The tilting-pad bearing that information is obtained for is a fivepad bearing with a 60° pad and a preload ratio of 30 percent. Two length/diameter ratios are examined for each type of bearing: L/D  0.5 and 1.0. The definition of the nondimensional parameters is as follows: W ⫽ nondimensional load parameter ⫽ wc2>6mvRL3 (28) P ⫽ nondimensional viscous power loss parameter ⫽ 1100cp>m1vRL2 2 (29) Q ⫽ nondimensional flow parameter ⫽ 2q>0.26vRLc (30) HM ⫽ nondimensional minimum film thickness ⫽ hM>c where w  bearing load capacity, lb (N) and c  reference clearance (machined clearance  radius of bearing ⫺ radius of shaft), in (mm) m ⫽ absolute viscosity, reyns (lb  s/in2) (cP) v ⫽ shaft or journal rotational speed rad/s R ⫽ shaft radius, in (mm) L ⫽ bearing length, in (mm) p ⫽ viscous power loss, hp (kW) q ⫽ flow, gpm (m3/h) (31) 2.260 CHAPTER 2 FIGURE 13 Performance characteristics for two-groove cylindrical bearings. FIGURE 14 Performance characteristics for three-lobe bearings. qi  inlet flow to the leading edge of bearing (for multipad bearings, equals the sum of inlet flow to each pad), gpm (m3/h) qs  side leakage flow or flow out of the bearing ends (for multipad bearings, equals the sum of side leakage flow of each pad), gpm (m3/h) hM  minimum film thickness in bearing, in (mm) Performance curves are shown in Figures 13 through 17. At times, the nondimensional data can be confusing and lead to erroneous judgments. For example, the nondimensional power loss P is greater for an L/D equal to 0.5 than for an L/D equal to 1.0. However, when the dimensional value of the power loss is being computed, the nondimensional value is multiplied by L2. Therefore, the power loss for the L/D 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS FIGURE 15 2.261 Performance characteristics for canted three-lobe bearings. FIGURE 16 Performance characteristics for tilting-pad hearings. equal to 1.0 will be, as expected, greater than for the L/D equal to 0.5. If the reader uses the data as presented, the dimensional information will prove consistent. To make comparisons among the bearings, using the nondimensional data is not strictly proper because there may be slight inconsistencies in preloads, the bearings will not be operating at the same average viscosity, and so on. Subsequently, dimensional data derived from the performance curves will be compared, but comparisons of the nondimensional information will provide an indication of performance parameters among the bearings. Comparisons have been made at equal values of the load parameter W and the results are shown in Table 1. Comparisons of the different bearing types should be made only at the same L/D ratio because of the anomalies (discussed above) of nondimensional parameters that occur at 2.262 CHAPTER 2 FIGURE 17 Performance characteristics for tilting-pad hearings. TABLE 1 Comparative Results of Bearing Types at W  0.2 L/D  0.5 Bearing type Two-groove cylindrical Three-lobe Canted three-lobe Tilting-pad L/D  1 P HM QS QI P HM QS QI 1.13 1.80 1.52 1.50 0.36 0.32 0.31 0.30 0.78 0.79 1.42 0.44 1.81 1.85 3.30 2.82 0.72 1.05 0.91 0.90 0.25 0.22 0.22 0.20 0.50 0.22 0.90 0.16 1.50 1.65 2.58 1.50 QI  nondimensional inlet flow  2qi/0.26 wRLc QS  nondimensional side leakage flow  2qs/0.26 wRLc different L/D ratios. If we assume that all the reference variables that go into the nondimensional parameters are identical, we can establish the following conclusions: • The two-groove cylindrical hearing has the highest film thickness and thus the highest load capacity. • The symmetric three-lobe bearing has the highest power loss. • The canted three-lobe bearing has the greatest flow requirements. Note that these comparisons were made on the basis of steady-state performances only. The major reason for applying lobe and tilting-pad bearings is to avoid dynamic instabilities. Heat Balance Performance is based upon the assumption of a uniform viscosity in the fluid film. Since the viscosity is a strong function of temperature and since the temperature rise of the lubricant due to viscous heat generation is not known a priori, an iterative procedure is required to determine the average viscosity in the film. To determine an average viscosity, there must be a simplified heat balance in the film. The assumption is made that all the viscous heat generated in the film is absorbed by the lubricant as it flows through the film and produces a temperature rise. 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS FIGURE 18 2.263 Heat balance in a fluid film. Figure 18 shows a developed view of the bearing surface and the parameters involved with conducting the heat balance. The lubricant enters a pad or bearing at the leading edge with an inlet flow qi and an inlet temperature ti. As the lubricant enters the bearing surface and flows through it, the lubricant is exposed to a viscous shear, which adds heat to the fluid. In Figure 18, the viscous shear is indicated as a power input p. Some of the fluid flows out of the sides of the bearing, represented by qs/2, and some of the fluid, qc, is carried either over to the grooving of the next pad or back to the inlet groove for a single pad bearing. Although the fluid temperature, and thus the viscosity, changes along the length of the pad, it is assumed that the temperature in the film increases to some value to for both the side leakage and carry-over fluids. Then the heat balance is conducted as follows: Heat added to a fluid by a viscous shear equals heat absorbed by the fluid. Heat added to fluid by viscous shear ⫽ heat absorbed by fluid pJ ⫽ qs rCp 1to ⫺ ti 2 ⫽ qcrCp 1to ⫺ ti 2 ⫽ 1qs ⫹ qc 2rCp 1to ⫺ ti 2 (32) Since by continuity of flow, where p J r Cp qi ⌬t qi ⫽ qs ⫹ qc (33) pJ ⫽ qirCp ¢t pJ ¢t ⫽ qirCp (34) (35)  viscous heat generation, hp  heat equivalent constant  0.7069 Btu/hp  s (J/kg  s)  specific weight of flow, lb/in3 (kg/m3)  specific heat of fluid, Btu/lb  °F (J/kg  °C)  bearing inlet flow, in3/s (mm3/s) ⫽ temperature rise, °F (°C) If qi is in gallons per minute [(1 in3/s ⫽ 0.26 gpm) ⫽ 16.39 cm3/s], ¢t ⫽ 0.706910.262p qirCp ⫽ 0.1838p qirCp (36) Before proceeding to a sample problem, a word about the flows qi and qs, and the general heat balance philosophy. The flow required by the bearing to prevent starvation is qi. The minimum make-up flow, or the flow that is lost from the ends of the bearing, is the side leakage qs. Theoretically then, the only flow that need be supplied to the bearing is qs. However, if this were true, then the heat balance formulation would require another balance between the carry-over flow qc at some temperature to and the make-up flow qs coming into the pad at temperature ti. It would be found that, if we supplied only qs to the bearing, excessive temperatures would result. In most instances, the amount of flow supplied to a bearing exceeds both the inlet flow and the side leakage flow by a significant 2.264 CHAPTER 2 margin. Designers often determine the flow to be supplied to a bearing system by a bulk temperature rise of the total flow entering the inlet pipe and exiting the exhaust pipe. Thus, the entire bearing is treated as a black box, and the total flow qT to the bearing system for a given bulk temperature rise ⌬tt is qT ⫽ pJ rCp ¢tt (37) Normally, ⌬tt is selected between 20 and 40°F. Since qT will exceed qi, not all the flow will enter the film. Some will surround the bearing or flow through the feed groove and act as a cooling medium for heat transfer through the bearing walls or for cooling the fluid that does enter the bearing surface qi. This example demonstrates the use of the design curves and how to perform a simplified heat balance in a bearing analysis. In this example, we know these values: EXAMPLE Bearing type: two-groove cylindrical Bearing length L ⫽ 3 in (76 mm) Bearing diameter D ⫽ 6 in (152 mm) L/D ⫽ 0.5 N ⫽ 1800 rpm; v ⫽ 1800 ⫻ p/30 ⫽ 188.5 rad/s Bearing load v ⫽ 6000 lb (2721 kg) Inlet temperature ti ⫽ 120°F (49°C) Lubricant ⫽ SAE 20 Bearing radial clearance c ⫽ 0.003 in (0.076 mm) A general rule of thumb is that c/R ⫽ 0.001. Assume an average lubricant temperature of 130°F ⫺ m (SAE 20 at 130°F) ⫽ 4.5 ⫻ 10⫺6 reyn (lb  s/in2) (31 ⫻ 10⫺3 Pa  s) NOTE: r ⫽ 0.0307 lb>in3 1849.7 kg>m3 2 1average value for most oils2 Cp ⫽ 0.5Btu>lb # F° 12093 J>kg # C°2 1average value for most oils2 Compute the following: W⫽ 160002 10.0032 2 wc2 ⫽ ⫽ 0.1309 3 6mvRL 16214.5 ⫻ 10⫺6 2 1188.52 132133 2 From Figure 13 and Equation 29, P ⫽ 1.0 ⫽ P⫽ 1100cp m1vRL2 2 11214.5 ⫻ 10⫺6 2 1188.5 ⫻ 3 ⫻ 32 2 11100210.0032 ⫽ 3.92 hp 12.95 kW2 From Figure 13 and Equation 30, QI ⫽ 1.85 ⫽ qi ⫽ 2qi 0.26vRLC 11.85210.2621188.52132 13210.0032 ⫽ 1.224 gpm 14.63 l>min2 2 From Equation 36, 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS ¢t ⫽ tav ⫽ 2.265 10.1838213.922 0.1838p ⫽ ⫽ 38.35 F° 121.3 C°2 qirCp 11.2242 10.0307210.52 122 11202 ⫹ 38.85 ti ⫹ 1ti ⫹ ¢t2 2ti ⫹ ¢t ⫽ ⫽ ⫽ 139.2°F 159.6°C2 2 2 2 The assumed tav of 130°F was apparently not high enough and therefore we must repeat the calculations with a higher value of assumed tav. Assume tav  138°F, m  3.8 ⫻ 10⫺6 lb  s/in2 (26.2 ⫻ 10⫺3 Pa  s) and repeat the calculations: W⫽ 5.8945 ⫻ 10⫺7 ⫽ 0.1552 m p P ⫽ 1.08 ⫽ 11.1466 ⫻ 10⫺6 2 a b m p⫽ 11.08213.8 ⫻ 10⫺6 2 1.1466 ⫻ 10⫺6 qi QI ⫽ 1.83 ⫽ 0.6616 ⫽ 3.579 hp 12.67 kW2 qi ⫽ 1.211 gpm 14.58 l>min2 111.974213.5792 p ¢t ⫽ 11.974 ⫽ ⫽ 35.38 F° 119.65 C°2 qi 1.211 122 11202 ⫹ 35.38 ⫽ 137.7°F 158.7°C2 tav ⫽ 2 For all practical purposes, these equal the assumed value of tav ⫽ 138°F (58.9°C). In conducting the iterative procedure for determining the average fluid temperature, it is sometimes helpful to plot points on a viscosity chart. Referring to Figure 19, suppose FIGURE 19 Graphical determination of average viscosity for SAE 20 oil. 2.266 CHAPTER 2 point 1 is the resultant average viscosity and temperature of the initial guess and point 2 represents the result of the second guess. Then, by drawing a straight line between points 1 and 2 and establishing where it intersects the lubricant viscosity curve, we can determine the convergence to the proper result of average viscosity and temperature. Now that we have convergence, we can determine the remaining variables, which are the minimum film thickness HM and side leakage QS from Figure 13: HM ⫽ 0.41 ⫽ hM c hM ⫽ 10.41210.0032 ⫽ 0.00123 in 10.031 mm2 2qs qs QS ⫽ 0.74 ⫽ ⫽ 0.6616 0.26vRLC qs ⫽ 10.6616210.742 ⫽ 0.490 gpm 11.85 l>min2 A summary of the total bearing performance is as follows: Load v ⫽ 6000 lb (2721 kg) Minimum film thickness hM ⫽ 0.00123 in (0.031 mm) Viscous power loss r ⫽ 3.579 (2.669 kW) Inlet flow qi ⫽ 1.211 gpm (4.58 l/min) Side leakage qs ⫽ 0.490 gpm (1.85 l/mm) Fluid temperature rise ⌬t ⫽ 35.38°F (19.65°C) We can repeat the same calculation for the other types of bearings, using Figures 14 through 17. The results are shown in Table 2. The two-groove cylindrical bearing operates with the highest film thickness. The symmetric three-lobe has the lowest film thickness and highest temperature rise. It appears that, on the basis of steady-state performance, manufacture, and cost, the two-groove cylindrical bearing is the best choice. BEARING DYNAMICS7_________________________________________________ Fluid film bearings can significantly influence the dynamics of rotating shafts. They are a primary source of damping and thus can inhibit vibrations. Alternatively, they provide a TABLE 2 Comparative Bearing Performance Bearing type Two-groove cylindrical Symmetric three-lobe Canted three-lobe Tilting-pad ⌬t, °F (°C) tav, °F (°C) 35 (19.44) 49 (27.22) 31 (17.22) 26 (14.44) 138 (58.9) 144 (62.2) 135 (57.2) 134 (56.7) p, hp 3.58 4.97 5.20 4.88 qi, gpm (cm3/s) 1.21 (76.7) 1.22 (77.3) 2.01 (127.4) 2.26 (143.2) qs, gpm (cm3/s) 0.490 (31.05) 0.516 (32.7) 0.940 (59.6) 0.311 (19.71) hM, mils (mm) 1.23 (0.0312) 0.96 (0.0244) 1.05 (0.0267) 1.02 (0.0259) The following values were used: v ⫽ 6000 lb (2721 kg), SAE 20 oil, L ⫽ 3 in (76 mm), D ⫽ 3 in (76 mm), N ⫽ 1800 rpm, ti⫽ 120°F (48.9°C), c ⫽ 0.008 in (0.076 mm). 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS 2.267 mechanism for self-excited rotor whirl. Whirl is manifest as an orbiting of the journal at a subsynchronous frequency, usually close to one-half the rotating speed. Whirl is usually destructive and must be avoided. It is the purpose of this section to provide some insight into bearing dynamics, present some background on analytical methods and representations, and discuss some particular bearings and factors that can influence dynamic characteristics. Dynamic performance data and sample problems are presented for several bearing types. The Concept of Cross Coupling As mentioned in the opening paragraphs of this subsection, a journal bearing derives load capacity from viscous pumping of the lubricant through a small clearance region. To generate pressure, the resistance to pumping must increase in the direction of the fluid flow. This is accomplished by a movement of the journal such that the clearance distribution takes on the form of a tapered wedge in the direction of rotation, as shown in Figure 1. The attitude angle g in Figure 1 is the angle between the load direction and the line of centers. Thus, the displacement of the journal is not along a line that is coincident with the load vector, and a load in one direction causes not only displacements in that direction, but orthogonal displacements as well. Similarly, a displacement of the journal in the bearing will cause a load opposing the displacement and a load orthogonal to it. Thus, strong cross-coupling influences are introduced by the mechanism by which a bearing operates. The concept of cross-coupling is significant in dynamic characteristics. It is the cross-coupling characteristics of a journal bearing that can promote selfexcited instabilities in the form of bearing whirl. Motion in one direction produces orthogonal forces that in turn cause orthogonal motion. The process continues, and an orbital motion of the journal results. This orbital motion is generally in the same direction as shaft rotation and subsynchronous in frequency. Half-frequency whirl is a self-excited phenomenon and does not require external forces to promote it. Cross-Coupled Spring and Damping Coefficients For dynamic considerations, a convenient representation of bearing characteristics is a cross-coupled spring and damping coefficients. These are obtained as follows (refer to Figure 1): 1. The equilibrium position to support the given load is established by computer solution of Reynolds’ equation. 2. A small displacement to the journal is applied in the y direction. A new solution of Reynolds’ equation is obtained, and the resulting forces in the x and y directions are produced. The spring coefficients are as follows: Kxy ⫽ Kyy ⫽ where Kxy ⌬Fx ⌬y Kyy ⌬Fy ¢Fx ¢y ¢Fy ¢y (38) (39)  the stiffness in the x direction due to y displacement ⫽ the difference in x forces between displaced and equilibrium positions ⫽ the displacement from the equilibrium position in the y direction ⫽ the stiffness in the y direction due to y displacement ⫽ the difference in y forces between displaced and equilibrium positions 3. The journal is returned to its equilibrium position and an x displacement is applied. Similar reasoning produces Kxx and Kyx. The cross-coupled damping coefficients are produced in a similiar manner, except, instead of displacements in the x and y direction, velocities in these directions are consecutively applied with the journal in the equilibrium position. The mechanism for increasing the load capacity is squeeze film in which the last term on the right-hand side of 2.268 CHAPTER 2 Equation 14 is actuated. Thus, for most fixed bearing configurations, eight coefficients exist: four spring and four damping. The total force on the journal is # (40) Fi ⫽ Kijxj ⫹ Dijxj where Fi ⫽ force in the ith direction. Repeated subscripts imply the following summation: Kijxj ⫽ Kixx ⫹ Kiyy ⫹ p It should be realized that the cross-coupled spring and damping coefficients represent a linearization of bearing characteristics. When they are used, the equilibrium position should be accurately determined, as the coefficients are valid for only a small displacement region encompassing the equilibrium position of the journal. This is true because the spring and damping coefficients remain constant for only a small region of the equilibrium position. Consider the two-groove cylindrical bearing shown in Figure 1, with the geometric and operating conditions indicated in Table 3. The computer solution (also the performance curves in Figure 13) produces the following results: Bearing load w ⫽ 20,780 lb (9,424 kg) Power loss ⫽ 15.51 hp (11.56 kW) Minimum film thickness hM ⫽ 0.00125 in (0.032 mm) Side leakage qs ⫽ 0.941 gpm (3.56 l/mm) The spring and damping coefficients are Spring coefficients, lb/in (kg/mm): c Kxx Kyx ⫺12.14 ⫻ 106 Kxy d ⫽ c Kyy 28.3 ⫻ 106 ⫺4.64 ⫻ 106 d ⫺20.41 ⫻ 106 Damping coefficients, lb  s/in kg  s/in: c Dxx Dyx Dxy ⫺2.85 ⫻ 104 d ⫽ c Dyy ⫺2.69 ⫻ 104 2.66 ⫻ 104 d ⫺1.11 ⫻ 105 The negative signs imply a positive stiffness because the restoring load is opposite the applied load. Note that for this bearing configuration there is very strong cross-coupling, evidenced by the magnitude of the off-diagonal terms. Critical Mass The cross-coupled spring and damping coefficients provide a convenient way of representing a bearing in a stability analysis. They reduce the fluid film bearing to a spring-mass system (see Figure 20), and consequently stability and dynamics problems are simplified considerably. Consider a journal of mass M operating in a bearing. The journal can be considered to have two degrees of freedom, x and y. The governing equations are TABLE 3 Two-Groove Cylindrical Bearing Geometry and Operating Conditions Journal diameter D  5 in (127 mm) Bearing length L  5 in (127 mm) Active pad angle up,  160° (10° grooves on either side) Radial clearance c  0.0025 in (0.064 mm) Operating speed N  5000 rpm Lubricant viscosity m  2 ⫻ 10⫺6 lb  s/in2 (13.79 ⫻ 10⫺3 Pa  s) Eccentricity ratio e ⫽ 0.5 Load direction is vertical downward 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS FIGURE 20 2.269 Point mass representation of a bearing supported on cross-coupled springs and dampers. Interpretation of x ⫽ xoebt where b ⫽ a ⫹ iv x  xoeateivt  xoeat 1COS vt ⫹ i sin vt2 FIGURE 21 Interpretation of the growth factor a and orbital frequency v. $ # # Mx ⫹ Dxxx ⫹ Dxyy ⫹ Kxxx ⫹ Kxyy ⫽ 0 $ # # My ⫹ Dyyy ⫹ Dyxx ⫹ Kyyy ⫹ Kyxx ⫽ 0 (41) (42) Assume a sinusoidal response to the form x ⫽ x0ebt (43) bt y ⫽ y0e (44) b ⫽ a ⫹ iv (45) where b is a complex variable: By the Euler expansion of e , another way to write Equations 43 and 44 is bt x ⫽ xoeat 1cos vt ⫹ i sin vt2 y ⫽ yoe 1cos vt ⫹ i sin vt2 at (46) (47) An interpretation of a and v is shown in Figure 21. The real part of b ⫽ a is called the growth or attenuation factor. The imaginary part is the frequency of vibration. A positive real part means that the response to a disturbance grows in time. The growth factor is similar to the logarithmic mean decrement, which is common in vibration theory: at ⫽ ln xn⫹1 xn (48) 2.270 where CHAPTER 2 t ⫽ period of vibration xn⫹1 ⫽ amplitude at time n ⫹ 1 xn ⫽ amplitude at time n Thus, the growth factor a is a measure of the growth or decay of the journal to a small disturbance. A positive growth factor implies an instability. The solutions to Equations 41 and 42 are obtained by substituting Equations 43 and 44, which produces the following: c 1Mb2 ⫹ Dxxb ⫹ Kxx 2 1bDyx ⫹ Kyx 2 1Dxyb ⫹ Kxy 2 x d e 0 f ⫽ 506 1Mb2 ⫹ bDyy ⫹ Kyy 2 y0 To obtain a solution, the determinant of the coefficient matrix must vanish. Expansion produces a polynomial in b that can be solved for the roots of b, which in turn provide the growth factors and frequencies of vibration. It is possible to obtain a closed-form solution of Equations 41 and 42 for the critical mass and resulting orbital frequency. The critical mass M is defined as that mass above which an instability will occur. At the threshold of instability, the real part of b ⫽ a goes to zero and b ⫽ iv. Substituting into Equation 49, expanding the determinant and separating real and imaginary components produces the following equations: M1Dyy ⫹ Dxx 2v2 ⫹ 1DxyKyx ⫹ DyxKxy ⫺ DxxKyy ⫺ DyyKxx 2 ⫽ 0 E (51) { { { (50) { { M2v4 ⫺ M1Kyy ⫹ Kxx 2v2 ⫹ 1DyxDxy ⫺ DxxDyy 2v2 ⫹ KxxKyy ⫺ KxyKyx ⫽ 0 A B  C A B C D E The two equations can be solved for M and v: M⫽ v⫽ BED E2 ⫺ AED ⫹ CD2 (52) B (53) ⫺1AED ⫹ E2 ⫹ CD2 2 BD2 Thus, if the cross-coupled coefficients are known, it is possible to determine the critical mass and the orbital frequency. If the mass acting on the bearing exceeds or equals the critical value, then an instability will occur. Dynamic Stability of Various Bearing Types Several parameters can be used to establish the stability characteristics of a particular type of bearing. The most significant is the critical mass, derived above. It is also possible to get some feel for stability from purely steady-state performances by examining the bearing attitude angle. The larger the attitude angle, the greater the cross-coupling influence and the worse the stability characteristics. Interpreting attitude angles, however, can prove misleading. Figure 22 shows plots of the attitude angle g versus the load parameter W for some of the different types of bearings previously described. The contradiction in these results is that the symmetric three-lobe bearing has higher attitude angles than the two-groove cylindrical bearing even though, as will be subsequently demonstrated, it has superior stability characteristics. The reason for the higher attitude angle is that the bearing is cavitating in the diverging region of the loaded pad, which results in a large shift in the journal. Whirl motion, however, is prevented by the accompanying lobes. A more direct and accurate approach to establishing bearing stability is to determine the critical mass acting on the bearing. Figure 23 shows the results obtained for the fixed 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS FIGURE 22 Attitude angle versus load parameter for various bearing types. FIGURE 23 Critical mass versus load parameter for various bearing types. 2.271 bearing geometries previously discussed. For any particular bearing geometry, if the critical mass attributable to the bearing exceeds that of the data plotted, the bearing will be unstable. Thus, if the plotted point falls to the right of the bearing line, the bearing is stable; if it falls to the left, the bearing is unstable. The critical mass is defined as follows: M  mvRc3>24mL5 (54) 2.272 where M m v R c m L CHAPTER 2 ⫽ nondimensional critical mass ⫽ dimensional critical mass, lb  s2/in (kg  s2./mm) ⫽ shaft speed, rad/s ⫽ shaft radius, in (mm) ⫽ bearing machined radial clearance, in (mm) ⫽ lubricant viscosity, lb sec/in2 (Pa  s) ⫽ bearing length, in (mm) An examination of the curves in Figure 23 clearly indicates the superiority of the canted three-lobe bearing and the inferiority of the cylindrical configuration. EXAMPLE Consider a high-speed bearing for which N ⫽ 60,000 rpm v ⫽ 60,000p/30⫽6283 rad/s D ⫽ 1 in (25.4 mm) L ⫽ 0.5 in (12.7 mm) m ⫽ 1⫻10⫺6 lb  s/in5 (6.9 x 10⫺3 Pa  s) C ⫽ 0.0005 in (0.013 mm) w ⫽ 100 lb (45.35 kg) W⫽ 11002 10.00052 2 wC2 ⫽ ⫽ 0.0106 6mvRL3 162 11 ⫻ 10⫺6 216283210.52 10.53 2 The value of M is obtained from Figure 22 and indicated in Table 4 for the various bearing types considered. Dimensional units are also given. Table 4 clearly indicates that the lobe bearings can permit significantly more attributable mass than the cylindrical bearing can. If a symmetric rotor is being supported by two bearings, the cylindrical bearings would be unstable if half the weight of the rotor exceeded 14.7 lb (6.67 kg). The half weights go to 258 lb (117 kg) and 501 lb (227.2 kg) for the symmetric three-lobe and canted three-lobe bearings, respectively. No mention has been made of the tilting-pad bearing because, for all practical purposes, these bearings are always stable. OPERATING CONDITIONS THAT AFFECT BEARING STABILITY ______________ Cavitation Figure 1 shows a pressure distribution in a journal bearing. Positive pressure is generated in the converging wedge because the journal is pumping fluid through a restriction. On the downstream side of the minimum film thickness, the journal is pumping fluid out of a diverging region. In this region, the pressure decreases. Either the pressure becomes negative, which is defined as pressure below the ambient pressure, TABLE 4 Stability Comparison Chart Bearing type Two-groove cylindrical Symmetric three-lobe Canted three-lobe M m, lb  s2/in (kg  s2/mm) w, lb (kg)a 0.02 0.85 0.68 0.038 (6.79 ⫻ 10⫺4) 0.668 (11.9 ⫻ 10⫺3) 1.299 (23.2 7 10⫺3) 14.7 (6.7) 257.9 (117.1) 501.4 (227.6) m⫽ a 24Mmh5 3 vRc ⫽ M1242 110⫺6 2 10.55 2 16283210.52 10.00053 2 ⫽ 1.91M The w⬘ represents the maximum mass in weight units that could act on the bearing. 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS 2.273 or the film cavitates and decreases to atmospheric pressure as the lubricant releases entrapped air. With respect to stability, cavitation is a more desirable condition than the development of negative pressure. From an examination of Figure 1, it can be seen that the negative pressure pulls the journal in an orthogonal direction and increases the cross-coupling. The more eccentric the bearing, the larger the negative pressure or cavitated region. In lightly loaded bearings that are pressure-fed, negative pressures can occur because pressures in the divergent region have not approached atmospheric pressure. Cavitation does not occur, and the bearing is prone to instability. Thus, the feed pressure and load on a bearing are two additional parameters that affect stability. Lobe bearings are often used because of their excellent antiwhirl characteristics. Figure 9 earlier showed two types of lobe bearings: symmetric and canted. The symmetric lobe bearing is designed so that, in the concentric position, the minimum film thickness occurs at the center of each lobe. Note that this permits a region of converging film followed by a region of diverging film. Thus, depending upon the ambient pressure, it is possible to have negative pressures in a symmetric lobe bearing. Under very high ambient conditions, a symmetric lobe bearing can go unstable. The canted lobe bearing is designed to have a completely converging wedge and positive pressure throughout its arc length. Its stability characteristics are superior to those of the symmetric lobe bearing. Its steady-state characteristics are also superior. Hybrid Bearings At times, externally pressurized bearings are resorted to for stability improvements. The philosophy is that externally pressurized bearings are not subject to high attitude angles, as is the case with hydrodynamic journal bearings. Although this is generally true, hybrid hearings can still be subject to considerable cross-coupling. Figure 11 showed a schematic arrangement of a hybrid bearing. Oil is fed through restrictors from an external source into pocket recesses. From there, it exits into the clearance region between recesses. Lubricant is also pumped into and out of recesses by the rotating shaft by a viscous drag in the same manner as with a purely hydrodynamic bearing. Consider recess 2 in Figure 11. Oil is pumped from the shaft via a converging wedge, and it augments the pressure in the recess provided by the external system. The net result is a higher pressure in the bearing domain covered by recess 2 than would occur without rotation. Now consider recess 3. Here the journal is pumping fluid out of the recess into a diverging film, so that the hydrodynamic action tends to reduce the pressure in this recess domain. By similar reasoning, it can be shown that recess 4 operates at a lower pressure than recess 1. The net result of these variations in pressure due to rotation is that crosscoupling forces are introduced and the hybrid bearing may not prevent instability. BEARING MATERIALS AND FAILURE MODES ____________________________ Materials The most common material used for oil-lubricated fluid film bearings is babbitt. Tin- and lead-based babbitts are relatively soft materials and offer the best insurance against shaft damage. They also enable embedded dirt and contaminants without significant damage. Two types of babbitts are in common use. One has a tin base (86 to 88 percent), with about three to eight percent copper and four to 14 percent antimony. The other has a lead base with a maximum of 20 percent tin and about 10 to 15 percent antimony. The remainder is principally lead. The physical properties of babbitt are shown in Table 5. The primary limitations of babbitt are operating temperature (300°F [140°C] max) and fatigue strength. The chemical composition of various babbitt alloys are indicated in Table 6. Tin-based babbitts have better characteristics than lead-based babbitts; they have better corrosion resistance, are less likely to wipe under poor conditions of lubrication, and can be bonded more easily than lead-based materials. Because of cost considerations, however, lead-based babbitts are widely used. The more widely used is the SAE 15 alloy containing one percent arsenic (refer to Table 6). 2.274 CHAPTER 2 TABLE 5 Properties of Bearing Alloys Bearing material Brinell hardness at room temperature Brinell hardness at 30°F (17°F) Minimum Brinell hardness of shaft 20–30 6–12 150 or less 15–20 6–12 150 or less Tinbased babbitt Leadbased babbitt Load carrying capacity, lb/in2 (kg/m2) Max operating temperature, °F (°C) 800–1500 (5.62 ⫻ 105 ⫺ 10.55 ⫻ 105) 800–1200 (5.62 ⫻ 105 ⫺ 8.44 ⫻ 105) 300 (149) 300 (149) SOURCE: Ref. 8 TABLE 6 Compositiona Percentages of Babbitts with SAE Classifications of 11 to 15 SAE No. (Similar ASTM Spec.) Element 11 (None) 12 (B23, alloy 2) 13 (None) 14 (B23, alloy 7) 15 (B23, alloy 15) Tin (mm) Antimony Lead Copper Iron Arsenic Bismuth Zinc Aluminum Cadmium Others 86.0 6.0–7.5 0.5 5.0–6.5 0.08 0.1 0.08 0.005 0.005 — 0.2 88.2 7.0–8.0 0.5 3.0–4.0 0.08 0.1 0.08 0.005 0.005 — 0.2 5.0–7.0 9.0–11.0 Remainder 0.5 — 0.25 — 0.005 0.005 0.05 0.2 9.2–10.8 14.0–16.0 Remainder 0.5 — 0.6 — 0.005 0.005 0.05 0.2 0.9–1.2 14.0–15.5 Remainder 0.5 — 0.8-1.2 — 0.005 0.005 0.05 0.2 a The percentage of minor constituents represents limiting values except as noted. SOURCE: SAE Handbook, Society of Automotive Engineers. New York, 1960, p. 201. Tin-based babbitts do not experience as many corrosion problems as lead-based babbitts. SAE 11 babbitts (containing eight percent antimony and eight percent copper) are used extensively for industrial applications. Babbitts are bonded to a backing shell of another material, such as steel or bronze, because they are not a good structural material. The thinner the babbitt layer, the greater the fatigue resistance. In automotive applications where resistance to fatigue is important, babbitt thickness is from 0.001 to 0.005 in (0.02 to 0.12 mm). For pump applications, the 1 thickness varies from 32 to 81 in (0.8 to 3 mm). The thicker layers provide good conformity and embedability. Failure Modes The failure modes most commonly found are fatigue, wiping, overheating, corrosion, and wear. Fatigue occurs because of cyclic loads normal to the bearing surface. Figure 24 shows babbitt fatigue in a 7-in (178-mm) diameter journal bearing from a steam turbine. Wiping results from surface-to-surface contact and smears the babbitt, as shown in Figure 25. The usual causes of wiping are a bearing overload, insufficient rotational speed to form a film, and loss of lubricant. 2.2.5 CENTRIFUGAL PUMP OIL FILM JOURNAL BEARINGS FIGURE 24 Babbitt fatigue in a 7-in (178-mm) diameter turbine bearing. (Westinghouse Electric Corp. Photo originally reproduced in Ref. 8, p.18–19.) 2.275 FIGURE 25 Bearing wipe on a 3-in (76-mm) diameter bearing due to temporary loss of lubricant. (Westinghouse Electric Corp. Photo originally reproduced in Ref. 8, p.18–25) FIGURE 26 Corrosion on a 5-in (127-mm) diameter, lead-based babbitt bearing. (Westinghouse Electric Corp. Photo originally reproduced in Ref. 8, p. 18–21) Overheating is manifest by discoloration of the surface and cracking of the babbitt material. Corrosion is failure by a chemical action. It is more common with lead-based babbits, which react with acids in the lubricant. Figure 26 shows corrosion damage for a 5-in (127-mm) diameter, lead-based babbitt bearing. Wear results from contaminants in the film and is evidenced by scoring marks that may be localized or persist around a large circumferential region of the bearing. REFERENCES _______________________________________________________ 1 Fuller, D. D. Theory and Practice of Lubrication for Engineers. Wiley, New York, 1956. Castelli, V. and W. Shapiro. “Improved Method of Numerical Solution of the General Incompressible Fluid-Film Lubrication Problem.” Trans. ASME, J. Lub. Technol., April, 1967, pp. 211—218. 2 2.276 CHAPTER 2 3 Castelli, V.,and J. Pirvics. “Review of Methods in Gas-Bearing Film Analysis.” Trans. ASME, J. Lub. Technol., October 1968, pp. 777—792. 4 Pinkus, O. and B. Sternlicht. Theory of Hydrodynamic Lubrication. McGraw-Hill, New York, 1961. 5 Ng, C. W. and C. H. T. Pan. “A Linearized Turbulent Lubrication Theory.” Trans. ASME, J. Basic Eng., Series D, Vol. 87, 1965, p. 675. 6 Elrod, H. G. and C. W. Ng. “A Theory for Turbulent Films and Its Application to Bearings.” Trans. ASME, J. Lub. Technol., July 1967, p. 346. 7 Shapiro, W., and R. Colsher. “Dynamic Characteristics of Fluid Film Bearings.” Proc. Sixth Turbomachinery Symposium, sponsored by the Gas Turbine Laboratories, Department of Mechanical Engineering, Texas A&M University, College Station, Texas, December 1977. 8 O’Connor, J. I., J Boyd, and E. A. Avallone. Standard Handbook of Lubrication Engineering. McGraw-Hill, New York, 1968. 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS GRAHAM JONES LARRY HAWKINS PAUL COOPER Magnetic bearings maintain the rotor of a pump in suspension through the forces of attraction of a magnetic circuit. Thus, although they bear up the weight and hydraulic loads of the impellers and the shaft, they are not really bearings in the traditional sense of the rotating and stationary surfaces bearing on one another. The supporting magnet circuit for each bearing includes stationary magnets in a stator surrounding the shaft, a laminated rotor that fits on the shaft, and the shaft itself. The stator consists of electromagnets in the traditional heteropolar design, and if a homopolar design is employed, permanent magnets can be added. Sensors monitor the position of the shaft and signal a controller to adjust the magnetic loads to keep the shaft to within about 0.001 in (25 mm) of the desired position. Magnetic bearings are found in small, high-speed turbomachinery such as high-speed, multistage, axial-flow turbomolecular vacuum pumps1. They were introduced into large turbomachinery in the early 1980s, mainly in gas compressors and turboexpanders. Their use and acceptance has grown slowly but steadily since then2. Pump applications of a significant size have appeared and have confirmed the general position that magnetic bearings can provide a technically sound bearing with maintenance and operating advantages, including zero wear. However, due to the technical complexity of magnetic bearing systems, the economies of scale associated with production quantities are required to make these systems affordable. Two representative magnetic-bearing-equipped pumps are summarized in Table 1. One is a multistage boiler feedwater pump3–6 and the other a single-stage double-suction hydrocarbon process pump7. The multistage pump was retrofitted with magnetic bearings (as shown in Figure 1) and is shown in Figure 2, together with another identical pump that still contains the oil-lubricated bearings—both installed in an electric generating station. The magnetic-bearing pump is not encumbered with the usual complexity of a bearing lubrication system. 2.277 2.278 CHAPTER 2 TABLE 1 Example of magnetic-bearing-equipped pumps Parameter Multistage Pump Single-Stage Pump Power, hp (MW) Rated speed, rpm Shaft weight, lb (kg) Radial bearing design load, lb (kN) 610 (0.46) 3,580 520 (236) 800 (3.6) Thrust bearing design load, lb (kN) Number of stages 4,000 (17.8) 8 800 (0.6) 1780 732 (332) Thrust end: 930 (4.1) Drive end: 1,415 (6.3) 4,000 (17.8) 1 FIGURE 1 Magnetic bearing configuration in multistage pump6 A B FIGURE 2 Multistage pumps installed in boiler feed service [610 hp (0.46 MW)]: a) pump with magnetic bearings; b) the same pump with oil-lubricated bearings. (Flowserve Corporation) MAGNETIC BEARING PRINCIPLES ______________________________________ How Magnetic Bearings Work In an active magnetic bearing system, a stator composed of an array of stationary magnets, or electromagnetic coils, interacts with a ferrous rotor (or a ferrous sleeve on a non-ferrous rotor) so as to suspend the shaft in a magnetic field (see Figure 3). The position of the shaft is maintained dynamically through a continuous feedback system which comprises a position sensor, a controller, and an amplifier system (see Figure 4). Typically there are two radial bearings and one thrust bearing for a complete sys- 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS 2.279 FIGURE 3 Typical radial electromagnetic bearing. (Axial thrust bearings have the stator coils arranged in a disk configuration, a ferrous rotating disk being supported in the resulting magnetic circuit.) FIGURE 4 Typical system control loop tem. This system is tuned to the required characteristics of the pump, through a digital or analog controller (Figure 5), with the capability of adjusting the bearing stiffness and damping as a function of pump speed. Alarms and trips can be set at any required rotor offset or bearing load to provide the operator with warnings or to trip the drive unit as necessary. Figure 6 illustrates a typical bearing transfer function, showing a statically stiff bearing, with a dynamic stiffness over the operating speed range designed to meet the rotor dynamics requirements of the unit. The stiffness is then rolled off above the operating range to avoid excitation of higher modes in the rotor or stator. Controller redundancy can be provided with the control loops switching to backup units upon sensing a failure. Also, backup power supplies should be provided, either through alternate sources or a battery system. Typically the power required is only one or two kW or less. A catcher (also known as auxiliary, backup, or touchdown) bearing (indicated in Figure 1) is required to protect the rotor stator interface during maintenance and in the event of 2.280 A CHAPTER 2 B FIGURE 5 Controllers for magnetic bearings, containing rectifier and amplifiers: a) digital controller; b) analog controller FIGURE 6 Bearing transfer function loss of power or a severe transient beyond the force capability of the bearing. Typically, the catcher bearing is designed for 5 to 20 lifetime drops from full speed, and will be a readily replaceable rolling element or sleeve bearing. The pump of Figures 1 and 2 has rolling element catcher bearings. The radial clearances G1 (see Figure 7) between the magnetic 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS 2.281 FIGURE 7 Clearance arrangement: Seal ring clearance G3 is greater than catcher-bearing clearance G2, but is less than magnetic-bearing clearance G1. bearing stator and rotor are of the order of 20 to 40 mil (0.5 to 1 mm), and those in the catcher bearings (G2) are about half of that value. Reasons for Using Magnetic Bearings There are several reasons to use magnetic bearings in pumps. While any one of these reasons may not be sufficient justification on its own, together they can provide a strong justification. Reliability is a key incentive. The components of a magnetic bearing are essentially the same components as are found in an electric motor: laminations and coils. Because no wear is involved due to the lack of contact, these components will generally last the life of the equipment involved. Thus maintenance of a magnetic bearing system is transferred from mechanical components inside the pump to the external controller, which has plugin card replacement maintenance. Pump reliability is therefore improved, whereas repair times and costs are reduced. Reduced power consumption is a second advantage, with the elimination of all losses associated with fluid film bearings and oil pumping equipment. This is replaced by the smaller power requirements of the bearing controller. Further, if the lifting force is supplied by a permanent magnet, supplemented by an active control circuit, this power requirement can be even smaller. The ability to submerge the bearing in the pump fluid is a major advantage that allows the outboard mechanical seal to be eliminated, thereby eliminating maintenance and replacement of this seal8. [This was not done for the pumps of Table 1 (and Figures 1 and 2), because in both cases magnetic bearings were retrofitted to existing machines.] More indirect savings are also possible in two other areas. Rotor dynamics can be controlled through the ability to adjust stiffness and damping as a function of pump speed, allowing higher imbalance without the need for shutdown. The diagnostic output inherent in the information provided in the controller can be fed into the overall plant operating system and the short-term and long-term health of the pump and the system can be monitored. This is done by inferring seal wear, transient hydraulic loads, and so on. The actual figures for the savings possible due to the previous advantages are very pump- and system-specific, and general numbers are not very useful. Reference 9 has developed methodology for considering the economic effect of the types of advantages given. Main Types of Magnetic Bearings and Their Selection There are two main types of magnetic bearings: passive and active. Passive bearings rely only on permanent magnets in repulsion and provide low stiffness, low damping, and no ability to control either of these parameters. Passive bearings are not applicable to pumps for this reason. 2.282 CHAPTER 2 FIGURE 8 Heteropolar and homopolar bearings Active bearings using the feedback system previously described are essential for pump applications. Within the active bearing systems are the options of heteropolar and homopolar and of electromagnetic and permanent magnet bias. The principles of the heteropolar and homopolar approaches are shown in the Figure 8. The main difference between the two types is that in the heteropolar design, the bias and control flux flow in the same magnetic circuit radially through the rotor, whereas in the homopolar design the bias flux flows axially along the rotor and only the control flux flows radially through the rotor. The homopolar design has two options for providing the bias flux for the bearing system10. One is to use an electromagnetic effect, and the other is to use a permanent magnet. The permanent magnet generation in the homopolar configuration results in a more linear relationship between force and distance. In a simple magnetic circuit, the attraction force of a magnet on a ferromagnetic target decreases as the square of distance (the target cuts 14 as many flux lines at twice the distance). With the permanent magnet in a homopolar circuit, the effect of the air gap is therefore reduced. DESIGN CONSIDERATIONS ____________________________________________ Design Loads The specification of a magnetic bearing requires a different approach to that required for a conventional bearing system. This section identifies the key areas where these differences occur. The rotor can move considerably within the magnetic bearing and catcher bearing clearances, typically up to 10 mil (0.25 mm), before any contact is involved. Thus the clearances G3 in the internal ring-seal system (see Figure 7) become the key controlling parameter in setting the bearing clearance design limits and in the degree of motion permitted during transients. Thus, very early in the design process, the magnetic bearing clearances, catcher bearing clearances, and sealing-ring clearances must be optimized with due consideration for manufacturing tolerances and assembly concerns. Synchronous filters or open-loop control methods can be used to handle imbalance loads so that the degree of imbalance acceptable (based on allowable bearing loads and shaft motion within the catcher bearing clearances) can often be significantly greater than in conventional bearings. 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS 2.283 There is an area where careful analysis is needed for each new pump configuration. A conventional bearing, if overloaded, will accept the load with a higher wear rate, but a magnetic bearing has a sharp cut-off at the point where the flux saturation level is reached. At that point, any additional load will be transferred to the catcher bearing. Thus it is important to include all loads in the design specification, with the appropriate margins. All applications to date have generally shown a) that there are loads which were mistakenly considered to be insignificant, or b) that the values of the loads were underestimated due to lack of knowledge. Magnetic bearings also must be carefully designed to handle transient loads, some of which may occur only once in a lifetime. Two approaches can be taken. One is to overdesign the bearing to handle the load without contacting the catcher bearing; the other is to allow momentary contact with the catcher bearing. Examples of these types of loads are hydraulic loads, seal touchdown loads, system loads such as water hammer, valve operation, pump switchover, pump to driver alignment loads, and seismic loading. Rotordynamics Considerations Magnetic bearings have the capability to control the rotor dynamics of a pump very effectively. If the pump is running below its first flexible mode, this is usually straightforward. If the pump has to traverse a flexible mode, the position of the bearings and the position sensor must be such that the modes can be recognized by the position sensor, and the bearings can exert a positive restoring force to control the mode. That is, the position sensors and bearings must not be positioned at or close to a node and certainly not positioned on opposite sides of a mode. Thus, rotordynamics considerations should be taken care of very early in the configuration of the system. The two typical rigid body modes are shown in Figure 9. The location of the bearing and position sensor is not usually an issue for these modes, but the flexible modes require that the position sensor and bearing be positioned in such a way that the rotor deflection can be measured and the bearing can exert the necessary restoring force and damping to control the mode. Magnetic bearing controllers are programmed with a transfer function designed specifically for the pump. This transfer function, or control algorithm, provides the necessary stiffness and damping at all operating speeds to control the rotor as previously described. The key requirements of the transfer function are that it • Provides correct damping and stiffness to handle the rotor rigid body modes • Provides sufficient force with the appropriate control bandwidth to handle the rotor flexible modes below the maximum operating speed FIGURE 9 Flexible mode design considerations 2.284 CHAPTER 2 • Does not excite any rotor modes above the maximum operating speed • Does not excite any stator vibration modes at any frequency • Takes into account the stiffness and damping contribution from wear rings and interstage annular seals Auxiliary or Catcher Bearings However reliable magnetic bearings become, a landing surface for maintenance will be required. Further, designing a bearing that will take all transient loads without any possibility of overload will usually result in over-sized and costly bearings. Hence catcher, bearings are expected to be required. The design limits for catcher bearings are usually established by calculating the forces that will result from a drop at operating speed. For applications running at significant speeds, a non-linear analysis is required to determine the motion and loading on the bearing during such a drop. The key design considerations are impact loads, heat generation during the rundown, and the response to imbalance if, when running on the catcher bearings, any critical speeds have to be traversed. Rolling element bearings have typically been used as catcher bearings; however, sleeve bearings and bushings have been used in several applications, and are better suited to a submerged application. MAGNETIC BEARING SIZING___________________________________________ The fundamental magnetic bearing sizing problem is to define the pole area at the bearing air gap that is necessary to achieve the desired force capacity without saturating the selected pole materials. Given the pole area, the minimum stator outside diameter and maximum rotor inside diameter can be determined using simple algorithms to ensure that no other part of the magnetic circuit saturates. The stator geometry must also include sufficient volume for the control and bias coils. Approximate size and load capability Table 2 contains rough approximations for the dimensions of magnetic bearings. (Refer to Figure 1.) These are based on the experience with the pumps of Table 1. Also included are the unit load capabilities, which are given a) for radial bearings in terms of radial load Fr divided by the projected area DL of the active polar area of the bearing at the air gap, and b) for axial thrust bearings in terms of the axial load Fz divided by the active pole area of the runner disk between the inner diameter Di and the outer diameter Do. These approximations provide the designer and user with an idea of the design configuration of a magnetic-bearing-equipped pump. The TABLE 2 Approximate magnetic bearing sizing relationships Sizing Parameter Heteropolar Homopolar Radial Bearings a) Dimensions, in multiples of the shaft diameter Air gap diameter, D Stator outer diameter Overall axial length Active axial length of poles at air gap, L b) Unit load capability, Fr/DL, lb/in2 (MPa) 2 4 2 0.9 40 (0.28)* 1.5 3 1.5 0.8 60 (0.41) Axial Thrust Bearings a) Diameter ratio Di/Do of active pole area b) Unit load capability, Fz / [p(Do2⫺Di2)], lb/in2 (MPa) 0.5 50 (0.34)* 0.5 70 (0.48) *Higher with special lamination material 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS 2.285 low unit loads of magnetic bearings result in more space being needed for them in comparison to conventional bearings. Theory of One-Dimensional Sizing To perform more accurate, in-depth sizing, the theory of both heteropolar and homopolar magnetic bearings is applied. Magnetic bearing sizing and geometry programs normally use simple one-dimensional magnetic circuit theory to obtain initial sizing and perform design iterations. This initial sizing is then followed up with design analysis using 2D or 3D magnetic FEA analysis to verify the design. The basis of the classic one-dimensional sizing for a magnetic bearing is discussed next, first for the heteropolar bearing and then for the homopolar bearing. HETEROPOLAR BEARING a. Magnetic circuit—The basic magnetic circuit equation, derived from Ampere’s Loop Law, is MMF ⫽ £ R (1) where MMF  magnemotive force ⌽ ⫽ magnetic flux R ⫽ path reluctance A sketch of one quadrant (one electromagnet) of a heteropolar bearing is shown in Figure 10. Assuming the air gap area and path areas are equal, Eq. 1 becomes lstat lrot g g ⫹ 2Ni ⫽ BA a m A ⫹ m A ⫹ m m b A m m o o o r, stat o r, rotA FIGURE 10 Simple electromagnetic circuit for one quadrant of a heteropolar magnetic bearing (2) 2.286 Where N i B A g mo mr l CHAPTER 2  number of turns per pole  coil current  magnetic flux density  pole area  air gap  permeability of free space  relative permeability  iron path length If the iron permeability is high relative to the air gap (mr, rot, mr, stat 77 mo), the iron reluctance terms are insignificant and the following equation can be obtained for the flux: B⫽ Nimo g (3) b. Force calculation—The basic force equation for an air gap is Fgap ⫽ B2A 2mo (4) This equation assumes negligible leakage and fringing and that the flux density is uniform in the air gap. The combined vector force along the center of the bearing for the two air gaps of the top magnet is Ftop ⫽ 2Fgapcos a (5) B2A mo (6) Substituting from Eq. 4 gives Ftop ⫽ cos a If the saturation flux density, Bsat, of the iron material is used for B, Eq. 6 defines the load capacity as a function of pole area for a heteropolar magnetic bearing. c. Linearization of the force/current characteristic—Substituting from Eq. 3 gives: Ftop ⫽ cos a i2 i2 A Nimo 2 ⫽ cos a AN2mo 2 ⫽ k1 2 mo a g b g g (7) k1 ⫽ cos a AN2mo Thus, the force in a given magnet is proportional to the square of the current, a result that makes the bearing more difficult to control. Additionally, a single electromagnet can only apply a force in one direction (an attractive force). For these two reasons, opposed electromagnets are used together with a bias current (or flux) in each coil. The current relationship is i ⫽ ibias ⫹ icon to increase the force, i  ibias ⫺ icon to decrease the force. The rotor may also be off-center in the air gap, described by 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS FIGURE 11 2.287 Force versus current for radial bearing of multistage pump g ⫽ g0 ⫺ y top air gap g ⫽ g0 ⫹ y bottom air gap Applying Eq. 7 to both top and bottom electromagnets yields Fy ⫽ Ftop ⫺ Fbot ⫽ k1 c a ibias ⫹ icon 2 ibias ⫺ icon 2 b ⫺ a b d g0 ⫺ y top g0 ⫹ y bot (8) With the rotor centered (y  0), this can be reduced to Fy ⫽ 4 k1 g02 ibiasicon (9) Thus the net bearing force is proportional to the control current. Figures 11 and 12 are examples of measured force versus current for radial and axial thrust bearings respectively. These measurements were made on the magnetic bearing of the multistage pump in Table 1 (refer to Figures 1 and 2). d. Force constant and negative stiffness—Eq. 8 can be also be linearized for small motion about the center (y 66 g0) by differentiating with respect to icon and y, the two quantities in Eq. 8 that can change in normal operation of the bearing. The result is Fy ⫽ ⫺kf i con ⫺ kn y (10) 2.288 CHAPTER 2 FIGURE 12 Force versus current for axial thrust bearing of multistage pump where kf ⫽ kn ⫽ ⫺4k1ibias g20 ⫺4k1i2bias g30 ⫽ force constant or current stiffness (11) ⫽ negative stiffness or position stiffness (12) The position stiffness is the passive stiffness of the bearing with a bias field but with no control current. The position stiffness is always negative, indicating that if the rotor is displaced from center, it will be pulled further from its equilibrium position if no control current is applied. The force constant defines the relationship of the control force to current (lb/amp or N/amp) with the bearing centered. It is also negative, indicating that applying a control current pulls the rotor from its centered position. Many practitioners use positive values for the force constant and the position stiffness as a matter of convenience. In this case, the minus signs in Eq. 10 become plus signs. The control current is determined by the measured displacement, y, and the characteristics of an adjustable sensor/compensator/amplifier transfer function: Gcon 1s2 ⫽ icon y (13) Substituting Eq. 11 into Eq. 10 gives Fy ⫽ ⫺kF Gcon 1s2y ⫺ kn y (14) 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS FIGURE 13 2.289 Control magnetic circuit for one axis of a homopolar magnetic bearing The characteristics of the control transfer function Gcon(s) are determined in order to stabilize the rotor/bearing system. In the homopolar bearing, a bias flux is used for linearization just as in the heteropolar bearing; however, the bias flux and control flux follow different paths. Additionally, the bias flux can be generated by either a permanent magnet (most common) or an electromagnetic coil. The use of a permanent magnet for bias reduces power consumption and makes the bearing more linear at large position offsets. HOMOPOLAR BEARING a. Magnetic control circuit—The control circuit for one half of the homopolar bearing, shown in Figure 13, is across the top air gap, through the rotor, out the bottom air gap, and around the back iron. When Eq. 1 is applied to this circuit, the result is similar to Eq. 2: 2Nicon  BconA a lstat lrot g g ⫹ ⫹ ⫹ b moA moA momr,statA momr,rotA (15) Again, if the iron permeability is high relative to the air gap, this can be reduced to the following: Bcon  Niconmo g (16) b. Magnetic bias circuit—The magnetic circuit equation for the bias circuit, shown in Figure 14, is Br lm lm g g  Bbias A ⫹ ⫹ momr, mag a mo A mo A momr, mag Am b Where Br  residual induction of the magnet lm  axial length of the magnet mr, mag  relative permeability of the magnet (⬵1.0–1.05) Am  magnet cross-sectional area per quadrant With the magnet permeability assumed to be 1.0, this can be reduced to (17) 2.290 CHAPTER 2 FIGURE 14 Bias magnetic circuit for one quadrant of a homopolar magnetic bearing Bbias ⫽ Br 2g A ⫹ lm Am (18) c. Force calculation—The basic force equation for an air gap is given by Eq. 4. In the homopolar bearing, the pole face covers a broader arc of the rotor than in the heteropolar bearing; therefore, the air gap force must be integrated over the surface to obtain the desired vector force, Ftop. Both the front and back control stacks must also be included. The result is Ftop ⫽ 2CgFgap Where: Cg ⫽ 2sin1ua>2 2 ua (19) This result assumes negligible leakage and fringing and that the flux density is uniform in the air gap. Substituting from Eq. 4 into Eq. 19 gives Ftop ⫽ Cg B2A mo (20) If the saturation flux density, Bsat, of the iron material is used for B, Eq. 20 defines the load capacity as a function of pole area for a magnetic bearing. d. Linearization of the force/current characteristic—The air gap flux density in the homopolar bearing is the superposition of the bias and control flux. The control coils and permanent magnet polarity are arranged such that when the control flux adds to the bias in the top air gaps, the control subtracts from the bias in the bottom air gaps. Thus the net vertical force is Fy  Ftop ⫺ Fbot ⫽ Cg A mo 3 1Bbias ⫹ Bcon 2 2top ⫺ 1Bbias ⫺ Bcon 2 2bot 4 (21) This can be reduced in a similar manner as before to produce Fy  4Cg A mo Bbias Bcon (22) 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS 2.291 Substituting from Eq. 16 Fy ⫽ 4Cg AN g Bbias icon (23) Thus the control force is proportional to the control current as desired. e. Force constant and negative stiffness—The expression given in Eq. 10 for the heteropolar bearing also applies to the homopolar bearing: Fy ⫽ ⫺kf icon ⫺ kn y (24) The definition of the force constant is kf  ⫺4Cg ANBbias g0 ⫽ force constant or current stiffness (25) The expression for negative stiffness is not easily reducible to analytical form due to the complexity of the bias circuit. However, the existence of the permanent magnet in the bias flux path as a fixed and large reluctance improves the linearity of this bearing for offcenter operation. Eqs. 13 and 14 for the heteropolar bearing apply to the homopolar bearing as well. INSTALLATION AND TUNING ___________________________________________ Mechanical Installation Magnetic bearings have the advantage that they can be set, by adjusting offsets in the controller, to center the rotor on the magnetic bearing stator, the catcher bearing, the seals, or any other mechanical reference in the pump. The rotor can even be offset vertically to cancel the effect of gravity, thus reducing static power requirements to near zero. Careful consideration is needed during design to decide which is the best approach. The position sensors in the bearings can also be used to measure the clearances between the rotor and any physical stops such as a sealing ring, without disassembly. Tuning The rotor dynamics analysis performed during the early design stage will be used to determine the initial controller compensation, which will have a transfer function matched to the pump requirements. During initial testing, this transfer function will have to be adjusted to match the as-built dynamics of the rotor and support structure. Normally the rotor will be accurately modeled and little change will be needed. If there are shrink fits or bolted joints, there may be some stiffness variation from the theoretical model. This may require on-site controller compensation adjustment. However, the stator is often a complex structure, and adjustments may be needed to avoid the excitation of stator modes. DIAGNOSTICS AND USER INTERFACE __________________________________ Diagnostic Capabilities The controller, in order to function, must analyze a continuous stream of information on the shaft location in each of the five control axes, two for each radial bearing and one for the thrust bearing. This information can be accessed for external diagnostic use, as can the corresponding information on bearing current, from which can be inferred the bearing load. This diagnostic information can be a very useful source of information on the health of the pump, its mechanical components, and on the system it is operating in. 2.292 CHAPTER 2 Interface Requirements The magnetic bearing system controller can also be interfaced with the plant control system with the following type of logic: • • • • No drive unit start without levitation No delevitation at speed Rotor offset and bearing load alarms Rotor offset and bearing load driver trips, with possible time delay RELIABILITY AND MAINTENANCE ______________________________________ Bearing Cartridges As explained earlier, the reliability of the bearing stator and rotor components should be such as to provide lifetime service. Controller The main life limiting component in the controller is likely to be the amplifier. In a redundant system, online replacement is possible without loss of levitation. In a non-redundant system, a preventative maintenance approach should be used for this component. OPERATING EXPERIENCE _____________________________________________ Multistage Boiler Feed Pump The multistage pump of Table 1 (refer to Figure 1) was installed as one of three otherwise identical pumps (two in parallel, one standby) in an electric utility generating plant (refer to Figure 2). The objective was to show that magnetic bearings would work in a typical field application of a pump of significant power level. This project took the first step of replacing the conventional bearings in this 610 hp (0.46 MW) eight stage centrifugal pump, which were outboard of the pump itself, and replacing them with heteropolar active magnetic bearings without any major design changes6. This was seen as the first of two steps, the second being a project where the bearings would be submerged in the operating fluid, allowing one seal system to be replaced8. As with many magnetic bearing projects, the main lesson learned was that the transient bearing loads could not always be predicted ahead of time, and that the magnetic bearings gave very precise and important feedback of this information. Table 3 contains the design and field data. In this case, the 2500 lb (11 kN) transient load occurred when the plant underwent a suction pressure transient in which the available NPSH became so low that the first stage of this horizontally-opposed staging configuration (refer to Figure 1) apparently lost pressure rise completely6. The axial thrust of this stage was accordingly lost, destroying the intended axial hydrodynamic thrust balance of the pump. (See Sections 2.1 and 2.2.1.) Nevertheless, the conservative design of the thrust bearing enabled it to accommodate this load. Single-Stage Process Pump The 800 hp (0.6 MW) single-stage double-suction process pump of size 8 ⫻ 26 (200mm ⫻ 660mm) of Table 1 was retrofitted with homopolar bearings7. Closed-loop testing was conducted in the pump manufacturer’s facility (see Figure 15). As indicated in Table 4, the results showed that a substantial operating margin exists for TABLE 3 Operating experience with multistage pump Expected Load, lb (kN) Bearing Radial Axial Steady Transient 1,280 (1.2) 1,000 (4.4) 1,560 (2.5) 2,000 (8.9) Design Load, lb (kN) 1,800 (3.6) 4,000 (17.8) Actual Load, lb (kN) Steady Transient 1,280 (1.2) 1,100 (4.9) 1,580 (2.6) 2,500 (11.1) 2.2.6 CENTRIFUGAL PUMP MAGNETIC BEARINGS FIGURE 15 (Reference 7) 2.293 Single-stage double-suction 10 ⫻ 26-size process pump [800 hp (0.6 MW)] with magnetic bearings. TABLE 4 Design and experimental loads for the single-stage pump Bearing Expected Load, lb, (kN) Design Load, lb (kN) Actual Load, lb (kN) Radial Axial 953 (4.2) 1,000 (4.4) 1,430 (6.4) 4,000 (17.8) 1,450 (6.4) 2,100 (9.3) the axial thrust bearing. Greater design capacity would provide the same margin with respect to radial loads. The margins evident in Table 3 for the multistage pump may appear excessive, but until more operating knowledge about such pumps is acquired, it would appear that this degree of conservatism in designing magnetic bearings for pumps is merited. COSTS _____________________________________________________________ Backed by the vision of complete magnetic suspension and the attendant benefits, magnetic bearing technology has been proven and demonstrated in pumping machinery. However, the major deterrent to further application of this technology is cost. Retrofitting the magnetic bearings to the multistage pump described above cost at least twice the price of the pump itself, and this included the analog controller, which accounted for about a third of the retrofit cost. Digital controllers, a later development, are one third the size and cost of analog controllers; this has significantly reduced the overall cost. Much of this overall cost is in the engineering of the magnetic bearings, which includes matching this system to the rotordynamic characteristics of the pump. Use of the same system in quantity production would reduce the cost by up to 80 percent. REFERENCES _______________________________________________________ 1. Marks Standard Handbook for Mechanical Engineers. 9th ed. E. A. Avallone and T. Baumeister, eds., McGraw-Hill, 1987, pp. 3—56. 2. McCloskey, T., and Jones, G. “Electric Utility Applications for Active Magnetic Bearings.” Proceedings of MAG ‘92, Magnetic Bearings, Magnetic Drives and Dry Gas Seals Conference and Exhibition, University of Virginia, July 1992, pp. 3—18. 2.294 CHAPTER 2 3. Cooper, P., McGinnis, G., Janik, G., Jones, G., and Shultz, R. “Application of Magnetic Bearings in a Multistage Boiler Feed Pump.” Proceedings of the Second International Symposium on Magnetic Bearings, July 1990. 4. McGinnis, G., Cooper, P., Janik, G., Jones, G., and Shultz, R. “A Boiler Feed Pump Employing Active Magnetic Bearings.” Presented at International Joint Power Generation Conference, Boston, MA, ASME, October 1990, ASME. 5. Cooper, P., and Jones, G. “Operating Experience, Including Transient Response, of a Magnetic-Bearing-Equipped Boiler Feed Pump.” Proceedings of MAG ‘92, Magnetic Bearings, Magnetic Drives and Dry Gas Seals Conference and Exhibition, University of Virginia, July 1992, pp. 19—28. 6. Jones, G., and Penfield, S. Jr. “Magnetic Bearing Boiler Feed Pump Demonstration”, Research Report EP89-39, Empire State Electric Energy Research Corporation, October 1992. 7. Brown, E., Thorp, J. M., Hawkins, L., and Sloteman, D. “Development and Application of a Homopolar, Permanent-Magnet-Bias Magnetic Bearing System for an API 610 Pump.” Saudi Aramco Journal of Technology, Spring 1997, pp. 2—13. 8. Hanson, L., and Imlach, J. “Development of a Magnetic Bearing API Process Pump with a Canned Motor.” Proceedings of the Ninth International Pump Users Symposium, Texas A&M University, March 1992, pp. 3—8. 9. “Guidelines for the Use of Magnetic Bearings in Turbomachinery.” Prepared by Technology Insights, San Diego, California and published by the Electric Power Research Institute, February 1996. 10. Meeks, C. R., DiRusso, E., and Brown, G. V. “Development of a Compact, Lightweight Magnetic Bearing.” Proceedings of the 26th Joint Propulsion Conference. AIAA/SAE/ ASME/ASEE, Orlando, FL, 1990. SECTION 2.2.7 SEALLESS PUMPS FREDERIC W. BUSE STEPHEN A. JASKIEWICZ Sealless pumps are developed to eliminate the liquid leakage to the atmosphere that occurs from pumps that employ packing or mechanical seals. This leakage is usually toxic or dangerous to the environment. Sometimes the leakage is valuable. Eighty percent of the applications are for pressures below 200 lb/in2 (13.8 bar) and below 250°F (120°C). Sealless pumps are divided into two categories: magnetic drive pumps and canned motor pumps. The two categories compete against themselves in certain applications, but for the most part they each have their own market niche into which they are applied. Both have encapsulated inner driven mechanisms. On a magnetic drive pump (see Figure 5 in Section 2.2.7.1), the impeller is mounted to an inner magnet carrier. The inner and outer magnetic carriers are sealed by what is called a shell, which contains pump internal pressure. On a canned motor pump (see Figure 1 in Section 2.2.7.2), the impeller is mounted directly to the motor rotor. The atmospheric sealing element between the motor stator and rotor is called a liner or “can.” Both magnetic drive and canned motor pumps use product lubricated bearings of compatible design and materials. These bearings are usually cooled and lubricated by the pump liquid. The following sections describe both magnetic drive and canned motor pumps. 2.296 SECTION 2.2.7.1 MAGNETIC DRIVE PUMPS FREDERIC W. BUSE The principle of a magnetic drive pump is the elimination of the seal or packing in an overhung pump (Figure 1) and cutting the shaft in two at the seal location (Figure 2a). On the inner or wet end half of the shaft (Figure 2b), an inner magnet assembly (a) is placed on a shaft (b) supported by product lubricated bearings (c). On the other cut portion of the shaft, an outer magnet assembly (d) is placed on the power shaft (e). Between the inner and outer magnet assemblies, a static seal (f) isolates the shell (g) or diaphragm and the pumped liquid from the atmosphere. The magnetic flux from the outer magnet assembly drives the inner magnet assembly and impeller(s). The outer assembly is mounted directly to either its own bearing assembly in a frame housing (Figure 3) or motor shaft (Figure 4). Figure 5 shows a cross-section of a magnetic drive frame mounted pump. Magnetics A magnetic circuit usually consists of two sets of permanent magnets and inner and outer conducting rings (Figure 6). The conducting rings can be cast iron, ductile iron, or a 400 series stainless steel. Magnet Materials The first permanent magnets, developed in the 1940s, were made of aluminum-nickel-cobalt (AlNiCo) and used in small chemical pumps. Development of rare earth magnets in the 1980s made it possible to have a small power package that could drive larger pumps. The two types of rare earth materials commonly used are neodymium iron boron (NdFeB) and samarium cobalt (SmCo). The SmCo is four times stronger than AlNiCo. NdFeB, at 70°F (21°C), is 20% stronger than SmCo. The advantage of SmCo is the maximum service temperature of 550°F (288°C),almost twice that of NdFeB, which is 300°F (149°C). Figure 7 shows the strength versus temperature characteristics of the two materials. The cost of SmCo, however, is about twice that of NdFeB. 2.297 2.298 CHAPTER TWO FIGURE 1 Pump with conventional bearing housing and mechanical seal (Flowserve Corporation) The magnets have a maximum temperature at which they lose all magnetism in an irreversible process; the magnets do not regain magnetism as they cool. This temperature is the Curie temperature and is shown for each magnet material in Table 1. TABLE 1 Operating and Curie temperatures for magnet materials Type of Magnet Material Operating Temperature (°F/°C) Curie Temperature (°F/°C) AlNiCo 660/305 800/425 NdFeB 250—300/120—137 (Depending on grade) 600/310 SmCo 500—660/260—350 (Depending on grade) 1300/700 Conduction Ring The amount of transmittal torque depends on the overall gap (Figure 8) between the poles of the magnets and the thickness of the conducting ring (Figure 9). If the thickness of the conducting ring is too small, it will become saturated with flux and the torque capability will be reduced. 2.2.7.1 MAGNETIC DRIVE PUMPS 2.299 A. B. FIGURE 2 Comparison of typical mechanical seal arrangement and sealless pump drive configuration: a) arrangement for typical mechanical through casing cover; b) sealless drive configuration FIGURE 3 Frame-mounted magnet drive arrangement Gap The “overall gap” (Figure 10) is made up of the air gap, containment shell thickness, liquid gap, and encapsulation. Gap dimensions are based on the pressure requirement for the shell, the number of magnets (single or dual), and the material of the shell (metallic or nonmetallic) in the gap. 2.300 CHAPTER TWO FIGURE 4 Close-coupled magnet drive arrangement FIGURE 5 Cross-section of a frame-mounted magnet-driven pump (Flowserve Corporation) 2.2.7.1 MAGNETIC DRIVE PUMPS FIGURE 6 Magnetic circuit FIGURE 7 Magnet strength versus temperature FIGURE 8 Magnet configuration 2.301 2.302 CHAPTER TWO FIGURE 9 FIGURE 10A and B Rows of magnets Encapsulation of magnets using nonmetallic composition and metallic clad 2.303 2.2.7.1 MAGNETIC DRIVE PUMPS EXAMPLE Dimensions shown as in (mm) Mean diameter (on radius dimensions) Air gap minimum Containment shell thickness: Hastelloy C Polymer Liquid gap Encapsulation Overall gap (basis Hastelloy C) Overall gap (basis polymer) 4.6 (117) .030 (0.762) .040 (1.016) .120 (3.048) .035 (0.889) .030 (0.762) .135 (3.429) .215 (5.461) 6.0 (152.4) .045 (1.143) .060 (1.524) .150 (3.810) .035 (0.889) .030 (0.762) .170 (4.318) .260 (6.604) In the previous example, the same overall gap was maintained for polymer and Hastelloy C shells. This allows for interchangeability of the magnetic assemblies independent of shell material. Transmittal Torque The torque transmitted by the magnets depends on the following: • • • • • Flux density of the magnets, Bg Operating temperature of the magnets (which will change Bg) Length of the magnet poles, L Number of magnets per ring, M Mean ratios between the ID of the outer assembly magnets and OD of the inner assembly magnets, r • Overall gap between the ID and OD of the magnets in the assembly, g •  constant, K, which changes as a function of the specific design Torque is determined from the following relationship, with appropriate units: T K  B2g  L  M  r g For a given design type and configuration, the torque varies inversely as the square of the overall gap. Depending on costs and specific design construction, an assembly ring of magnets can have either one continuous length of magnets of a series of 1, 2, 3 or more rows of individual magnets. TORQUE CAPABILITY_________________________________________________ The ultimate torque is the static “breakaway torque.” To determine this, the magnet carriers are assembled so the inner carrier is locked in position. Then a torque is applied by bar and weights or by a torque wrench to the outer carrier. The torque value at which the two carrier assemblies break loose from each other radially—or “decouple”—is called the “breakaway torque.” The designer has to account for the driver start-up acceleration time, start-up torque, and an appropriate safety factor to apply to the breakaway torque to determine the allowable applied torque. Table 2 gives examples of the torque capability of assemblies composed of NdFeB blocks of magnet 0.75 in wide  0.38 in high  1.125 in long (19 mm  9.65 mm  28.6 mm) with a 0.180 in (4.6 mm) thick conducting ring. Basic Dimensioning of Magnet Blocks The magnet blocks can be made too long or too short, resulting in handling, magnet molding, or flux density problems. The ratio of the dimensions for magnet blocks for successful designs is as follows: 2.304 CHAPTER TWO TABLE 2 Torque capability of magnet assemblies Mean Diameter in (mm) 4.3 (109) 6.0 (152) Magnets/Row Overall Gap in (mm) Breakaway Torque ft-lb (N • m) 12 18 0.240 (6.10) 0.260 (6.60) 30 (40) 60 (80) TABLE 3 Coefficients of thermal expansion (in/in/°F  106)/(cm/cm/°C  106) Magnet material NdFeB SmCo Parallel to Axis Perpendicular to Axis 5 (9) 20 (36) 9 (16.2) 16 (28.8) • Block width is two to three times the thickness. • Block length is three to five times the thickness. • Doubling the block thickness will increase the strength by approximately 20%. For reference: 1.0 in3 (16.387 cm3) of NdFeB per assembly at a mean diameter of 4.0 in (101.6 mm) with a 0.25 in (6.35 mm) overall gap produces approximately 7 ft-lb (9.5 N • m) of torque. A 3% reduction in flux density is equal to a 6% reduction in torque. Characteristics of magnet material: Density: 0.273 lb/in3 (7.56 g/cm3) Tensile: 12  103 lb/in2 (844 kg/cm2) Compression: 110  103 lb/in2 (7.7  103 kg/cm2) Flex stress: 36  103 lb/in2 (2.53  103 kg/cm2) Coefficients of thermal expansion are shown in Table 3. Radial and Axial Magnet Forces The inner and outer carriers have to be restrained radially by bearings from contacting each other. In the example of “Torque Capability,” a single row of 18 magnets with a 6 in (152 mm) mean diameter, the radial force with magnets concentric is 40 lb (18 kgf). When the magnets are offset by .005 in (0.127 mm), the radial force is 55 lb (25 kgf) (Figure 9). When the magnets are against each other (no gap), the radial force is 80 lb (36 kgf). In the previous example, it takes 60 lb (27 kgf) in an axial direction to separate a single row of concentric magnets and 180 lb (82 kgf) for three rows. It is strongly advisable that provisions be made for personnel to address these loads during assembly and disassembly of the carriers. Encapsulation of Inner Carrier Magnetics Encapsulation can be accomplished with either metallic or polymer materials. The encapsulation of the inner magnet and conducting ring is probably the most expensive and extensive process in a magnetic drive sealless pump. After encapsulation, the carrier should be nondestructively tested to confirm 100% effectiveness. The pros and cons of polymer encapsulation (Figure 10a) are as follows: 1. Limited to 250—300°F (120—150°C) 2. SmCo magnets are used because of the high exposure temperature when applying polymer over the magnets. 2.2.7.1 MAGNETIC DRIVE PUMPS 2.305 3. The overall gap is increased because of the required thickness of the polymer. 4. Magnets must be restrained mechanically on the conducting ring by adhesives or high-strength polymers around the magnets. 5. Polymers like PFA/PTFE or PEEK are more corrosive-resistant than most metals. 6. Production polymer construction is much less expensive than metallic construction. 7. Polymer tooling is expensive. The pros and cons of metallic encapsulation (Figure 10b) are as follows: 1. Temperature rating can be 500°F (260°C). 2. Thickness of the encapsulation material over the magnets can be 0.030 in (0.76 mm). 3. Welding of the components can be conventional, electron beam, or laser. However, care must be taken with conventional welding to prevent the arc from jumping toward the magnet flux. 4. If castings are used for the inner carrier, porosity can be a problem. 5. Gassing of polymers from welding heat coming out of the seams can be a problem. 6. Adhesives are not required to keep the magnets in place at 3600 rpm with metallic encapsulation; the outer shield performs this function. Encapsulation of Outer Carrier Magnets The magnets for the outer carriers do not have to be encapsulated (Figure 11). However, material like NdFeB has an infinity for water absorption that results in rusting and swelling. The magnets can then break loose and move in position relative to one another. It is highly recommended that they be encapsulated with an epoxy or metal sheathing for atmospheric protection and handling. Construction The outer carrier can be a casting or fabrication. The carrier is attached to the power end shaft in the bearing housing (Figure 3) or directly to a motor shaft, which is then called close coupled construction. When attached to the bearing housing shaft, there is very little axial or radial load applied to the bearings. This lightly loaded condition can result in internal skidding of the rolling element bearings within their races, resulting in premature bearing failure. Therefore, it is best to preload the bearings with a spring to prevent skidding. This can be accomplished outside of the bearing by a spring-loading feature (Figure 11). Containment Shell The containment shell shape and thickness depends on working pressure, material, and temperature. The shell thickness is usually uniform. To keep the FIGURE 11 Inner and outer carrier 2.306 CHAPTER TWO FIGURE 12 Shell ratio L/r for pressure TABLE 4 Pressure capabilities of shells with various end shapes for the same thickness and material Shape L/r Allowable pressure—lb/in2 (kPa) Ratio of allowable pressure to that for a flat plate Flat 533 9 (62) 1.0 Flat 267 23 (158) Elliptical 28 113 (779) 12.6 2.55 Elliptical 8 209 (1440) 23.2 Spherical 1 815 (5620) 90.5 length of the shell to a minimum, a square-ended shell can be used. However, unless the end of the shell is made extra thick, it would have a relatively low pressure capability. Therefore, an end shape which is elliptical or spherical is used to obtain higher pressure capability (Figure 12). Examples of how allowable shell pressure varies with the shape of the end plate are shown in Table 4. The classes of materials used for containment shells (Figure 11) include metals, polymers, and ceramics. The characteristics of containment shells of various materials are shown in Table 5. The main advantage of polymer or ceramic shells is that there are no eddy current losses. Therefore, cooling of the magnets is not required. Eddy Currents Metallic or metallic-lined shells will produce eddy currents. Depending on the thickness of the shell, the eddy current losses (PL) can amount to as much as 20% of the total power. K  T  L  N2 Bg 2D3 M PL  R where K  a constant, depending on the design T  thickness of the shell L  length of magnets (times the core of magnets) N  speed, rpm Bg  flux density of the magnets D  mean diameter M  number of sets of magnets R  electrical resistivity, microhms per cm3 (electrical resistivity for various shell materials is given in Table 6) 2.307 2.2.7.1 MAGNETIC DRIVE PUMPS TABLE 5 Characteristics of containment shells of various materials Material Metal Reinforced polymer Ceramic Metal with PTFE Construction Welded or hydroformed Injection molded Set and fired Spray coated Temperature limit—°F (°C) 300 to  750 (150 to 400) 40 to 250 (4 to 120) 32 to 2000 (0 to 1100) 32 to 350 (0 to 175) Thickness— in (mm) 0.030—0.040 (0.76—1.0) 0.170 (4.3) 0.250—0.380 (6.35—9.6) 0.050 (1.27) Heat conductivity* Hastelloy C  71 AISI 316  7.5 3 6 Higher than metal alone Creep None None None Some Thermal shock 1000 (537) 375 (190) 500 (260) — Eddy currents 316 is twice Hastelloy C None None Same as metal *Approximate thermal conductivity in Btu/hr/ft2/°F. Multiply by 0.488 to get calories/hr/cm2/°C TABLE 6 Electrical resistivity of various shell materials Shell Material Electrical Resistivity (R), microhms per cm3 Hastelloy C AISI 316L Inconel 625 Nimonic 90 Titanium K-Monel Alloy 20 130 74 129 115 53 58 75 TABLE 7 Effect of material selection on heat build-up Material of shell AISI 316 AISI 316 Hastelloy C Hastelloy C Thickness-in (mm) Time-minutes kW °F (°C) °F/min (°C/min) 0.43 (10.9) 0.43 (10.9) 0.43 (10.9) 0.43 (10.9) Start 5.0 0.5 5.0 0.40 0.30 0.24 0.21 80 (27) 500 (260) 170 (77) 370 (188) — 100 (38) 340 (171) 74 (23) The effect of material selection on heat build-up is shown in Table 7. The tabulation shows the temperature rise of the air space inside a shell with only the outer carrier spinning around a metallic shell (no inner carrier in place) at 3550 rpm. The carrier has 12 magnets, 1.25 in (3.175 cm) long, with a mean diameter of 4.25 in (11.43 cm). This illustrates the difference between AISI 316 and Hastelloy C shells at 3550 rpm. At 1750 rpm, the power loss for AISI 316 would be one-fourth that at 3550 rpm, which may not result in excess power consumption. There are also axially laminated metal shells available that substantially reduce eddy current (I2R) losses. These laminations work on the principle that when a shell length is 2.308 CHAPTER TWO FIGURE 13 Dual containment cut in half, the eddy current losses are reduced to one-fourth their original value. If the one-half shell length is cut in half again, the resulting pieces have eddy current losses onesixteenth that of the original shell. A shell design made of segments sealed together to make a full-length shell will thus reduce total eddy current losses. Dual Containment As a precaution against leakage due to a breech in the primary containment shell, a dual or double-containment arrangement can be used. This double layering of shells (Figure 13) usually consists of a combination of a nonmetallic and metallic shell. The pressure rating of the secondary shell is equal to that of the primary shell. It is designed to operate at least 48 to 120 hours after a breach in the primary shell occurs. A pressure monitor is inserted in the flange of the secondary shell to detect pressure buildup from primary shell leakage. Bearings The internal shaft system of the pump is supported by one or more bearings. Some designs use a rotating shaft; others use a mandrel on which the bearings rotate. The bearing, which consists of a journal and bushing, is made of various materials, depending on the loads and pumpage (used for product lubrication). The bearing loads are from the weight of the components and hydraulic forces from the impeller and inner carrier. The impeller forces are both radial and axial. Most magnetic drive sealless pumps are single-stage volute pumps that have the same radial bearing loads as comparable conventionally sealed pumps. The main difference with the sealless pump is that there is almost no overhang from the impeller to the first bearing. The load on the bearing, therefore, is almost equal to that of the impeller. In a conventionally sealed pump, the load on the radial bearing is almost twice that of the impeller. This can be seen by comparing Figures 1 and 5. The axial load will depend on whether the impeller is enclosed or semi-open (Figure 14). Enclosed impellers usually have horizontal front rings and may also have a back ring or pump out vanes (POV). Semi-open impellers have no rings, but they usually employ scallops in the shroud to reduce the effective pressure area (Figure 15). Axial thrust force is further reduced by employing pump-out vanes (POV) or pump-out slots (POS) on the back shroud of the impeller to reduce the amount of pressure on the impeller back shroud. The enclosed impeller can have less radial and axial load than a semi-open impeller. This is usually accomplished by employing back rings (14b). To ensure positive pumping in the lubrication flow path in magnetic drive pumps, POV-POS on an enclosed or semiopen impeller are recommended. Liquids pumped in chemical or petroleum plants may have low viscosity, low specific gravity, or low specific heat. These characteristics can result in boundary lubrication rather than hydrodynamic lubrication of the product lubricated bearings. Therefore, bearings are selected using PV (pressure-velocity) values. Some limiting PV values for various bearing material combinations are shown in the following section on bearing materials. 2.2.7.1 MAGNETIC DRIVE PUMPS 2.309 FIGURE 14A through D Various enclosed and semi-open impeller configurations FIGURE 15A and B Semi-open impellers with a full back shroud and with a partially scalloped back shroud Bearing Materials Table 8 shows PV values for various bearing journal and thrust face materials. When loads exceed a PV value of 150,000, the bearing should have an alignment compensator built in to correct for inaccuracies in perpendicularity, concentricity, and parallelism of parts. When designing the bearing, thermal expansion and heat conductivity properties must be considered. Some designs have a temperature range of 300°F to 750°F (150 to 400°C). Table 9 lists bearing material characteristics to facilitate selection of the proper material for the intended application. Particles For the softer bearing materials, the maximum particle size should be no more than 10% of the diametral clearance for a bearing with no grooves and no more than 20% 2.310 CHAPTER TWO TABLE 8 PV values for various bearing journal and thrust face materials Journal/Thrust Face PV ( 103)* Bushing Materials Carbon Graphite vs. AISI 316 Carbon Graphite vs. Chrome Oxide Hardened Coating Silicon Carbide vs. Hardened Coating Silicon Carbide vs. Carbon Graphite Silicon Carbide vs. Silicon Carbide PEEK—Carbon-filled vs. Hardened Coating Polyimide-carbon filled vs. Hard Coating 150 250 250 300 500 150 150 P  Net load/projected area (minus area for slots), psi V  Velocity at the shaft diameter or mean diameter of the thrust face in ft/min *Note: Multiplying the PV values in the table by 2.1 will give values equal to P in kPa and V in m/min. TABLE 9 Bearing material characteristics Material AISI 316 Hastelloy C Carbon graphite Silicon carbide— self-sintered Silicon carbide with carbon PEEK—Carbon-filled Polyimide—Carbon-filled Aluminum oxide Chrome oxide Thermal Expansion (note 1) Heat Conductivity (note 2) Hardness (note 3) 9.6 6.7 2.6 2.2 7.5 71 5 85 160 BHN 170 BHN 95 Vickers 2400 Vickers 2.0 75 2400 Vickers 8.0 5.0 6 100—150 BHN 1800 Vickers 1800 Vickers Note 1: in/in/°F  106 (multiplied by 1.411 to get cm/cm/°C) Note 2: BTU/hr/ft2/°F (multiply by 0.488 to get calories/hr/cm2/°C) Note 3: RC  25  Vickers; BHN  10  RC for a bearing with grooves. Polymers and graphites are much softer than silicon carbides or hard coatings, and they are not recommended for liquids with hard particles. Silicon carbide versus silicon carbide will grind up most particles. For particle concentrations of 50 parts per million (ppm) or less, particle hardness is generally not a factor in bearing performance. Strainers of 100 mesh are recommended to reduce the amount and size of particles going through the flow path. These strainers are installed for internal or external injection. Remember: Particles that will pass through a 100 mesh screen may be as large as 0.006 in (0.15 mm), whereas nominal diametral bearing clearances of 0.002 in (0.05 mm) are common. Note also that the bearing material combination of silicon carbide against silicon carbide or against carbon is electrically conductive. Running Dry Most sealless pump failures occur because the pump system is not monitored and the pump is allowed to run dry. When this happens, the bearings will run dry. When the bearings run dry, hard, brittle bearing materials such as self-sintered silicon 2.2.7.1 MAGNETIC DRIVE PUMPS 2.311 carbide will fail within minutes. Graphite silicon carbide may run dry for as many as 10 to 20 minutes with no damage. The polymers, which are poor heat conductors, expand inwardly and seize against the mating surface. The graphites will also move inwardly, but they tend to wear rather than seize, which results in excessive clearances when the pump is stopped and cooled. One optional design uses Teflon strips that expand inwardly when the unit runs dry so the shaft runs on the Teflon rather than on the silicon carbide, thus helping to avoid bearing failure. Another option is to employ an external circulating tank, with no external running parts, that will provide the bearings with external lubricating liquid should dry operation occur. This type of external tank system has allowed pumps to operate for two hours or more without incurring damage to the bearings. Flow Path The amount and direction of flow for cooling of the magnets and lubrication of the bearings is critical to the operation of a sealless pump (Figure 16). It is preferable for the liquid to lubricate the bearings before being heated by the magnets. This reduces the possibility of vaporization of the liquid occurring at the thrust-bearing faces. The amount of liquid circulated through the system is usually between 1 and 8 gpm (4 and 30 l/min). It is usually channeled to the front and back bearings, the thrust bearing face, across the magnets, and to the impeller hub. Some manufacturers have computer programs that calculate the flow, pressure, and temperature of the cooling/lubricating flow at various critical locations along the flow path. These programs can also account for the effects of the liquid specific gravity, specific heat, and viscosity. The local pressure and temperature is used to determine the vapor pressure at that point to ensure that the liquid is not flashing. The programs can also calculate axial thrust and determine if the lubrication at the thrust bearing face is hydrodynamic or boundary. This is done at various flow rates, impeller diameters, and pump speeds. The temperature rise of the liquid as it travels through the cooling-lubricating flow path depends to a great extent on the liquid’s characteristics, such as specific gravity, specific heat, vapor pressure, and viscosity. With a nonmetallic shell on ambient water service, a typical temperature rise might be 1 to 2°F (0.5 to 1°C). For a metallic shell, with its much higher eddy current losses, the temperature rise could be as much as 8 to 12°F (4 to 7°C). FIGURE 16 Magnetic sealless pump components for the internal flow system 2.312 CHAPTER TWO FIGURE 17 Typical performance for a 1.5  1.0  6 pump at 3550 rpm, comparing characteristics for sealless versus mechanical seal construction Performance The head capacity curve of a typical two-pole speed sealless pump matches that of a conventionally sealed pump; however, the overall efficiency is lower. With a nonmetallic shell, the efficiency may be only about two points less at the best efficiency point flow rate, but as much as six points less at one-half the best efficiency point flow rate. When a metallic shell is used, the efficiency at best efficiency point flow rate may be as much as 8 to 12 points lower than a comparable pump with a mechanical seal (Figure 17). APPLICATION ADVANTAGES OF SEALLESS PUMPS: • • • • • • • • No leakage to the environment No loss of valuable liquids Lower noise levels High suction pressure does not affect the axial thrust Can handle liquids from 0 to 4 toxicity rating Because of no leakage, there is much less chance of a fire Easier to obtain construction permits and permits for continued operation Less external piping required APPLICATIONS THAT SHOULD BE REVIEWED BEFORE SEALLESS PUMPS ARE APPLIED: • • • • • • • Dirty liquids High temperature Liquids that solidify Viscous liquids above 200 centipoise Oversize drivers that can cause decoupling during acceleration Cavitation of liquid in the impeller eye that can result in excess thrust Excessive entrained gas 2.2.7.1 MAGNETIC DRIVE PUMPS 2.313 REFERENCES AND FURTHER READING _________________________________ 1. American National Standard for Sealless Centrifugal Pumps, ANSI/HI 5.15.6-2000, Hydraulic Institute, Parsippany, NJ www.pumps.org. 2. Hydraulic Institute ANSI/HI 2000 Edition Pumps Standards, Hydraulic Institute, Parsippany, NJ www.pumps.org. 3. Sealless Pumps for Petroleum, Heavy Duty Chemical, and Gas Industry Services, API Standard 685, 2000, The American Petroleum Institute, 1220 L Street, Northwest, Washington, D.C. www.api.org. 4. Specification for Sealless Horizontal End Suction Centrifugal Pumps for Chemical Process, ANSI/ASME B73.3M-1997, The American Society of Mechanical Engineers, 345 East 7th Street, New York, NY www.ASME.org. 5. Eierman, R. “A User’s View of Sealless Pumps—Their Economics, Reliability, and the Environment.” Proceedings of the Seventh International Pump Users Symposium. Texas A&M University, College Station, TX, March 1990, pp. 127—133. 6. Hernandez, T. “A User’s Engineering Review of Sealless Pump Design Limitations and Features.” Proceedings of the Eighth International Pump Users Symposium. Texas A&M University, College Station, TX, March 1991, pp. 129—145. 7. Littlefield, D. “Sealless Centrifugal Pumps.” Proceedings of the Eleventh International Pump Users Symposium. Texas A&M University, College Station, TX, March 1994, pp. 115—119. 8. Guinzburg, A., and Buse, F. “Computer Simulation of the Flowpath in Magnetic Sealless Pumps.” Proceedings of the Fifteenth International Pump Users Symposium. Texas A&M University, College Station, TX, March 1998, pp. 1—9. SECTION 2.2.7.2 CANNED MOTOR PUMPS STEPHEN A. JASKIEWICZ A canned motor pump (CMP) is a combination of a centrifugal pump and a squirrel cage induction motor built together into a single hermetically sealed unit (see Figure 1). The pump impeller (A) is normally of the closed type and is mounted on one end of the rotor shaft that extends from the motor section into the pump casing. The rotor (B) is submerged in the fluid being pumped and is therefore “canned” to isolate the motor parts from contact with the fluid. The stator (C) is also “canned” to isolate it from the fluid being pumped. Bearings (D) are submerged in system fluid and are, therefore, continually lubricated. The canned motor pump has only one moving part, a combined rotor-impeller assembly that is driven by the magnetic field of an induction motor. A portion of the pumped fluid is allowed to recirculate through the rotor cavity to cool the motor and lubricate the bearings. A self-cleaning filter can be provided, on pumps having external circulation, to filter the recirculation fluid before it enters the bearing section of the motor. The stator windings and rotor armature are protected from contact with the recirculating fluid by a corrosion resistant, non-magnetic, alloy liner (E) that completely seals or “cans” the stator winding. Modifications to the recirculation flow system are available to allow canned motor pumps to be used in any application including fluids up to 1000°F (538°C), volatile liquids, and liquids with solids. BASIC DESIGN ______________________________________________________ Stator Assembly The stator assembly of canned motor pumps (see Figure 2) consists of a set of one or three-phase windings (A). Stator laminations (B) are constructed of lowsilicon grade steel. Laminations and windings are mounted inside the cylindrical stator band (C). End bells (D and E), welded to the stator band, close off the ends of the stator 2.315 2.316 CHAPTER TWO FIGURE 1 Typical canned motor pump assembly. The stator liner (F) is supported on the outside diameter by the steel lamination of the motor. Back-up sleeves (G) are provided to strengthen those areas of the stator liner not supported by the stator laminations. The stator liner is, in effect, a cylindrical “can” placed in the stator bore and welded to the rear end bell and front end bell shroud to hermetically seal off the windings from contact with the liquid being pumped. Terminal leads (H) from the windings are brought out through a pressure tight lead connector (I) mounted on the stator band and terminated in a standard connection box. Motors are either designed and manufactured specifically for use in canned motor pumps or components of conventional motor are modified. A variety of motor insulation types is available ranging from temperature limits of 266°F (130°C) to above 482°F (250°C). Because the pump and motor are one unit, the complete assembly must be tested and approved for explosion-proof applications. Explosion-proof pumps are rated as either Class 1, Group D, Division 1 or Class 1, Group C & D, Division 1 locations. It is not uncommon to operate canned motor pumps with variable speed drive controllers. Rotor Assembly The rotor assembly is a squirrel cage induction rotor constructed and machined for use in canned motor pumps (see Figure 3). It consists of a machined corrosion resistant shaft (A), laminated core (B) with copper or aluminum bars and end rings, corrosion resistant end covers (C), and a corrosion resistant can (D). Various methods are used to attach the impeller to the motor shaft (E). The rotor end covers are welded to the shaft and to the rotor can that surrounds the outside of the rotor, thus hermetically sealing off the rotor core from contact with the liquid being pumped. Some manufactures offer replaceable shaft sleeves (F) and axial thrust surfaces (G) for longer service life and ease of maintenance. Bearings Only two bearings are required for canned motor pumps. These bearings are normally cooled and lubricated by the pumped fluid; therefore, they must be compatible 2.2.7.2 CANNED MOTOR PUMPS FIGURE 2 2.317 Stator assembly FIGURE 3 Rotor assembly with the process fluid. A multitude of materials is available such as various grades of carbon graphite, silicon carbide, aluminum oxide, and many polymers. Bearing selection is dependent on the compatibility with the process fluid, amount of solids present, and pumping temperature. A hydrodynamic bearing is the most common type of bearing used. The bearings can be either stationary or rotating with the rotor assembly. In either case, the fluid passes between the bearing and shaft journal, resulting on the rotating assembly running on a thin film of liquid, not the journal and bearing. Most bearings have helical grooves in the inside diameter to increase the flow of process fluid through the journal area, thereby decreasing the temperature of the bearings. Internal Clearances The determination of the overall gap between the stator windings and the rotor armature is paramount in the design and operation of the pump. The wider the distance between the iron of the motor winding and the rotor armature, the less efficient the motor becomes. The material of construction of the stator liner and rotor sleeve 2.318 CHAPTER TWO also effects motor efficiency. Stainless steels and Hastelloy are the most common materials used for stator liners and rotor sleeves. Although stainless steel is less expensive, Hastelloy C has higher corrosion resistance, is a stronger material, and offers lower electrical losses. Motor efficiency in canned motor pumps is not only important for energy cost considerations, but also for the amount of heat input to the recirculation fluid. The stator liner (a wetted, pressure boundary component) ranges in thickness between 0.010 to 0.040 in (0.254 to 1.016 mm). For high-pressure applications, the liner remains at the same thickness, but the outside diameter is supported by the motor laminations and by back up sleeves located on both sides of the motor. Canned motor pumps have been designed to withstand working pressures up to 5,000 lb/in2 (345 bar) with 0.015 in (0.381 mm) stator liners and heavy walled back-up sleeves. The rotor armature (a wetted component) is also protected from the process fluid by a sleeve and two end covers. The thickness of the rotor sleeve ranges from 0.010 to 0.25 in (0.254 to 6.35 mm). The radial running clearance between the rotating motor armature and the stationary stator liner is usually about 0.020 in (0.508 mm). The total diametral clearance can range from 0.040 to 0.075 in (1.016 to 1.905 mm), or higher, depending on the manufacturer’s design. Secondary Containment Canned motor pumps offer a level of safety and process fluid containment unavailable with any other type of pump. Positive, secondary containment of the process fluid is a built-in feature with canned motor pumps when the motor lead wires are housed in a pressure retaining lead seal. In case of a failure of the primary containment shell (stator liner), the outer stator band becomes a secondary containment vessel, preventing the process fluid from entering the environment. The outer stator band is far removed from the rotation element, making it impossible for the rotating element to make contact. When secondary containment is required, the stator band assembly should be designed and tested to the same pressure and temperature rating as the pump. PRINCIPLE OF OPERATION ____________________________________________ Flow Path Most canned pumps, when pumping relatively clean fluids, will channel a small portion of the process fluid through the motor section. This fluid cools and lubricates the bearings and removes heat generated by the induction motor. The circulation path can be either external or internal to the pump. With external circulation (see Figure 4), the recirculation fluid is piped outside of the pump, through a filter, and then into the motor section of the unit. The filter assembly (see Figure 5) is self-cleaning and located in the discharge flange of the pump. Pumps having internal circulation have the recirculation contained within the pump. Filtering the recirculation liquid is not available with internal circulation. In either external or internal circulation, the flow path is from the high pressure area of the pump (pump discharge or pump chamber at the tip of the impeller) returning to the low pressure area (near the hub or eye of the impeller). The amount of liquid recirculated through the motor section ranges from 2 to 16 gpm (7.5 to 60 l/m). Many recirculation flow path modifications are available to allow a canned motor pump to pump any type of fluid. When pumping volatile fluids, the motor section can be pressurized by an auxiliary impeller located on the rotor. The recirculation fluid, which normally returns to the eye of the impeller, is channeled to the pressurized section of the liquid end, increasing the pressure in the motor section. This design allows a volatile fluid to remain liquid even with a temperature increase caused by motor heat (see Figure 6). Another method to handle fluids near their boiling point is to reverse the recirculation flow path. Instead of returning the heated liquid back to the eye of the impeller, the recirculation liquid is removed from the pump and returned to the suction vessel. High temperature and slurry applications can be handled by canned motor pumps by isolating the bearings from the pumped fluid. The recirculating fluid in the motor section 2.2.7.2 CANNED MOTOR PUMPS FIGURE 4 FIGURE 5 2.319 External circulation flow path Self-cleaning filter assembly is used for lubrication of the bearings and cooling of the motor. This fluid remains in the motor section and is circulated through the rotor cavity by an auxiliary impeller, which is an integral part of the rotor assembly. The fluid in the motor section is forced across the rotor and through the bearings, after which it flows through a heat exchanger. The heat exchanger is cooled by water or a suitable heat transfer fluid (see Figure 7). The motor section can be backflushed with a clean, cooled liquid when handling slurry. In Hydraulic Institute Standard ANSI/HI 5-1-5.6, Figure 5.9 provides an excellent description of various recirculation flow plans for sealless pumps. Thrust Balance Because most canned motor pumps use hydrodynamic bearings, the rotating assembly is allowed to float axially. This movement is known as “end play” and is defined as the movement of the rotor, in the axial direction, between the forward and 2.320 CHAPTER TWO FIGURE 6 Pressurized circulation FIGURE 7 Isolated motor section 2.2.7.2 CANNED MOTOR PUMPS 2.321 FIGURE 8 Automatic hydraulic thrust balance rear contact points (normally the thrust bearings). Many canned motor pumps incorporate a principle of hydraulic thrust balance to position the rotation assembly between the forward and rear mechanical contact points. Based on hydraulic principles, automatic thrust balance is accomplished by the pressure of the pumped fluid. Pressure chambers are designed either within the pump chamber using the movement of the impeller or rear of the pump controlling the rate of return of the process fluid. Figure 8 illustrates hydraulic thrust balance within the pump casing. Balance chambers are located on the front and rear of the impeller. When a change in axial load changes the position of the impeller, either forward or rear, there is an equalizing change of hydraulic pressure in the balance chambers, which immediately returns the rotation assembly into a balance position. Pressure-Temperature Profile The single most important consideration in the application and successful operation of canned motor pumps is bearing environment control. The process fluid cools and lubricates the bearings and removes the heat generated by the motor. The bearings must remain in a liquid state to provide adequate bearing lubrication. A number of variables must be considered when considering the state of the recirculation fluid. These variables include vapor pressure, specific heat, specific gravity, viscosity, pump efficiency, motor efficiency, motor load, recirculation flow rate, and the recirculation flow system. A vapor pressure curve versus temperature of the process fluid is also necessary. Figure 9 illustrates heat balance equations necessary to determine the temperature rise of the fluid within the motor section. Figures 10a, b, and c show pressure and temperature profiles for a specific fluid based on the heat balance equations within a canned motor pump modified for pressurized circulation. If the flow rate varies, a new profile should be calculated for each design condition. A pressure-temperature profile should be calculated for any application where there is doubt concerning the condition of the recirculation fluid. Installation The installation of a canned motor pump can be much less costly than a conventionally sealed unit because canned motor pumps do not require special baseplates or mounting pads. In fact, many canned motor pumps are stilt-mounted or bolted directly into the system piping. There is only one shaft in a canned motor pump; therefore, alignment of the pump to the driver is not necessary and pump orientation is not critical. The unit can be mounted either horizontally or vertically, with the pump casing 2.322 CHAPTER TWO FIGURE 9 Heat balance calculations up or down. The sleeved bearings are located in the bearing housing. Alignment of the bearings is accomplished by a register fit between the bearing housings and the stator assembly. Diagnostics Some canned motor pump manufacturers have developed diagnostic systems that monitor the condition of the internal wear surfaces of the pump. Because the 2.2.7.2 CANNED MOTOR PUMPS FIGURE 10A FIGURE 10B 2.323 Pressurized circulation—points of reference Pressurized circulation pressure profile rotating components of the pump are not visible, even the direction of rotation can not be easily determined. Advanced diagnostic systems indicate both axial and radial wear continuously. Continuous wear indication allows the pump user to trend the wear of the pump and schedule standard wear parts replacement long before a major failure occurs. Remote outputs are also available on these diagnostic systems. These outputs include either a digital or analog signal, or a relay designed to shut down the pump—or signal an alarm—when a certain amount of bearing wear has occurred. 2.324 CHAPTER TWO FIGURE 10C Pressurized circulation temperature rise The diagnostic systems available today are perhaps the most significant advancement in the technology of sealless pumps. Figure 11, rotor position output data, illustrates the type of information available when using digital output linked to a data collection system. 2.2.7.2 CANNED MOTOR PUMPS FIGURE 11 2.325 Rotor position output data REFERENCES AND FURTHER READING _________________________________ 1. American National Standard for Sealless Centrifugal Pumps, ANSI/HI 5.1-5.6-2000, Hydraulic Institute, Parsippany, NJ www.pumps.org. 2. Hydraulic Institute ANSI/HI 2000 Edition Pump Standards, Hydraulic Institute, Parsippany, NJ www.pumps.org. 3. Sealless Pumps for Petroleum, Heavy Duty Chemical, and Gas Industry Services, API Standard 685, 2000, American Petroleum Institute, 1220 L Street, Northwest, Washington, D.C. www.api.org. 2.326 CHAPTER TWO 4. Eierman, R. “A User’s View of Sealless Pumps—Their Economics, Reliability, and the Environment.” Proceedings of the Seventh International Pump Users Symposium. Texas A&M University, College Station Texas, March 1990, pp. 127—133. 5. Hernandez, T. “A User’s Engineering Review of Sealless Pump Design Limitations and Features.” Proceedings of the Eighth International Pump Users Symposium. Texas A&M University, College Station, TX, March 1991, pp. 129—145. 6. Littlefield, D. “Sealless Centrifugal Pumps.” Proceedings of the Eleventh International Pump Users Symposium. Texas A&M University, College Station, TX, March 1994, pp. 115—119. SECTION 2.3 CENTRIFUGAL PUMP PERFORMANCE 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS C. P. KITTREDGE PAUL COOPER DEFINITIONS ________________________________________________________ Nomenclature Many of the quantities involved in this subsection are also dealt with in Section 2.1. Therefore, a single nomenclature that applies to both sections appears at the beginning of Section 2.1. Differences in notation exist for some of these quantities as a result of the coexistence of different traditions and pump cultures, so the nomenclature shows the equivalence in each case. An example is the use in this subsection of “c” and “w” to denote absolute and relative velocity respectively, whereas the NASA system of capital letters V and W is employed in Section 2.1. Units The units used in this subsection are as defined in the nomenclature unless specifically noted in the text. In particular, the primary units for this subsection are those of the U.S. Customary System (USCS). A distinction in USCS usage in this subsection is that the pound force (lbf) is represented simply as “lb”. In keeping with the commentary on SI units in the front matter of this handbook, conversions to SI units are given throughout this subsection, or the actual equivalent SI values are given in parentheses. However, the number appearing in parentheses after the USGS value of specific speed ns is the equivalent value of the universal specific speed s. Note that the value of specific speed corresponding to the best efficiency point (BEP) operating conditions of the pump is the value of interest and is often used to identify the impeller geometry involved. Volume Flow Rate Abbreviated to “flow rate” and known traditionally as “pump capacity” Q, this is the volume of liquid per unit time delivered by the pump. In USCS units, Q is expressed in U.S. gallons per minute or USgpm, for which the abbreviation “gpm” is 2.327 2.328 CHAPTER TWO FIGURE 1A through C Elevation datum for defining pump head (Hydraulic Institute ANSI/HI 2000 Edition Pump Standards, Reference 27) used. (1 US gallon  231 in3.) For very large pumps, the units ft3/sec are used. The consistent SI units m3/s are implied when an SI value of Q is—unless the numerically convenient liters per second (l/s) are specifically called out. Datum for Pump Head As defined in Eq. 3 and Figure 1 of Section 2.1, the total head has components of pressure, velocity, and elevation Z (or Ze). Because pump head H (more precisely H) is the difference of the total heads evaluated at the discharge flange d and the suction flange s respectively, the elevation of the datum from which Z is measured cancels out. However, for purposes of identification, computing NPSH, and so on, the standard datum as shown in Figure 1 is used. The standard datum for horizontal-shaft pumps is a horizontal plane through the centerline of the shaft (Figure la). For vertical-shaft pumps, the datum is a horizontal plane through the entrance eye of the first-stage impeller (Figure lb) if single suction or through the centerline of the first stage impeller (Figure 1c) if double suction. Because pump head is the difference between the discharge and suction heads, it is not necessary that the standard datum be used, and any convenient datum may be selected for computing the pump head. 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.329 Power In USCS, the pump output is customarily given as liquid horsepower (lhp) or as water horsepower if water is the liquid pumped. It is given by lhp  QH1sp. gr.2 3960 (1) where Q is in gallons per minute, H is in feet, and sp. gr. is specific gravity. If Q is in cubic feet per second, the equation becomes lhp  QH1sp. gr.2 8.82 (2) In SI, the power P in watts (W) is given by P  9797QH 1sp. gr.2 (3) where Q is in cubic meters per second and H is in meters. When Q is in liters per second and H is in meters P  9.797QH 1sp. gr.2 (4) Efficiency The pump efficiency h is the liquid horsepower divided by the power input to the pump shaft. The latter usually is called the brake horsepower (bhp). The efficiency may be expressed as a decimal or multiplied by 100 and expressed as percent. In this subsection, the efficiency will always be the decimal value unless otherwise noted. Some pump driver-units are so constructed that the actual power input to the pump is difficult or impossible to obtain. Typical of these is the “canned” pump for volatile or dangerous liquids. In such case, only an overall efficiency can be obtained. If the driver is an electric motor, this is called the wire-to-liquid efficiency or, when water is the liquid pumped, the wire-to-water efficiency. CHARACTERISTIC CURVES____________________________________________ Pump with Non-Viscous Flow and Zero Slip The basic shapes of centrifugal pump performance characteristics arising from various geometries can be ascertained and compared without the necessity of evaluating the slip. (Illustrated in Figure 15 of Section 2.1, the slip phenomenon is explained in the related discussion1.) For this purpose, one employs the artifice of non-viscous flow through an impeller with an infinite number of blades having infinitesimal thickness and that therefore produce neither structural (geometric) nor flow (boundary layer) blockage. The result is the ideal head for no slip or blockage He as given by Eq. 5 (cf. Eq. 15b of Section 2.1): He  u1cu1 u2cu2  g g (5) The velocity components in Eq. 5 can be seen in the velocity diagrams of Figure 2, slip being neglected. Generally the inlet swirl term is small, and Eq. 5 can be approximated by He ⬵ u2cu2 g (6) As shown in Figure 2b, the absolute velocity vector c may be resolved into the meridional, or radial, velocity cm and the peripheral velocity cu. From the geometry of the figure cu2  u2  which, substituted into Eq. 6, gives cm2 tan b2 (7) 2.330 CHAPTER TWO FIGURE 2A and B Velocity diagrams for radial-flow impellers, neglecting slip and blockage FIGURE 3 Head-versus-flow rate characteristics for non-viscous, zero-slip impeller flow He  u2cm2 u22  g g tan b2 (7) Neglecting leakage flow, the meridional velocity cm must be proportional to the capacity Q. With the additional assumption of constant impeller speed, Eq. 8 becomes He  k1  k2Q (8) in which k1 and k2 are constants, with the value of k2 dependent on the value of the vane angle b2. Figure 3 shows the HevsQ characteristics for the three possible conditions on the vane angle at exit b2. The second right-hand term in Eq. 5 may be treated in like manner to the foregoing and included in Eqs. 8 and 9. The effect on Figure 3 would be to change 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.331 the value of He  u22>g at Q  0 and the slopes of the lines, but all head-flow rate characteristics would remain straight lines. Viscous Flow with Slip The real flow situation involves friction losses in an impeller with a finite number of relatively widely spaced blades. Thus, slip occurs, reducing the exit flow angle bf,2 below that of the blade b2, (or, more precisely, bb,2,) which in turn reduces cu2 (cf. Figure 15 of Section 2.1). Therefore, the ideal head Hi drops below He. Moreover, losses and recirculation occur to cause additional deviation of pump head H from He. While CFD flow analysis can be employed to predict H with fair accuracy2, lesser means, such as one-dimensional analysis, require experienced correlation and calibration skills to make such predictions. Therefore, many engineers commonly depend on testing and empirical modification of test results on the exact or similar geometry to make the final determination of the performance characteristics of a pump. This effort involves constantspeed plots of data as shown in the example of Figure 4. Pumps are designed to operate at the point of best efficiency. The head, power, and flow rate at best efficiency, often called the normal values, are indicated in this subsection by Hn, Pn, and Qn respectively. Sometimes a pump may be operated continuously at a flow rate slightly above or below Qn. In such case, the actual operating point is called the rated or guarantee point if the manufacturer specified this capacity in the guarantee. It is unusual to operate a pump continuously at a flow rate at which the efficiency is much below the maximum value. Apart from the unfavorable economics, the pump may be severely damaged by continued off-design operation, as described later. Backward-Curved Blades, B2 6 90° Figure 4 shows the characteristics of a double-suc- tion pump with backward-curved blades, b2  23°. The impeller discharged into a single volute casing, and the specific speed was ns ⬇ 2200 (s ⬇ 0.8) at best efficiency. At shutoff (Q  0), Eq. 8 predicts He to be 374 ft (114 m), whereas the pump actually developed about 210 ft (64 m), and this head remained nearly constant for the range 0 6 Q 6 1000 gpm (63 l/s). For Q 7 1000 gpm (63 l/s), the head decreased with increasing capacity but not in the linear fashion predicted by Eq. 9. At best efficiency, where Qn  3200 gpm (202 l/s), Eq. 8 predicts He  281 ft (86 m), whereas the pump actually developed Hn  164 ft (50 m). FLOW RATE FIGURE 4 Typical pump characteristics, backward-curved blades (ft ⫻ 0.3048 = m; hp ⫻ 745.7 = W; gpm ⫻ 0.06309 = l/s) 2.332 CHAPTER TWO Radial Blades, B2 ⴝ 90° Large numbers of radial-blade pumps are used in many applications, from cellar drainers, cooling-water pumps for internal combustion engines, and other applications where low first cost is more important than high efficiency to highly engineered pumps designed for very high heads. The impellers are rarely more than 6 in (15 cm) in diameter, but the speed range may be from a few hundred to 30,000 rpm or more. The casings usually are concentric with the impellers and have one or more discharge nozzles that act as diffusers. The impellers usually are open, with three or more flat blades. The clearance between blades and casing is relatively large for easy assembly. Such pumps exhibit a flat head-capacity curve from shutoff to approximately 75% of best efficiency flow rate, and beyond this flow the head-flow curve is steep. The pumps develop a higher head, up to 8000 ft (2400 m) per stage, than pumps with backward-curved blades, but the efficiency of the former usually is lower. (See also Subsection 2.2.1.) Figure 5 shows the characteristics of a pump as reported by Rupp.3 The impeller was fully shrouded, D  5.25 in (13.3 cm), and fitted with 30 blades of varying length. The best efficiency, h  55%, was unusually high for the specific speed ns  475 (0.174), as may be seen from Figure 6. The head-flow curve showed a rising (unstable) characteristic for 0 6 Q 6 25 gpm (1.6 l/s) and a steep characteristic for Q 7 25 gpm (1.6 l/s). Figure 7 shows the characteristics of a pump as reported by Barske.4 The impeller was open, D ⬇ 3 in (7.6 cm), and fitted with six radial tapered blades. The effective b2 may have been slightly greater than 90° due to the taper. At 30,000 rpm, the best efficiency was more than 35% at ns  355 (0.130), which is much higher than for a conventional pump of this specific speed and capacity (Figure 6). The head-flow curve showed a nearly flat characteristic over most of the usable range, as predicted by Eq. 8, but the head was always lower than He. The smooth concentric casing was fitted with a single diffusing discharge nozzle. When two or more nozzles were used, the head-flow curves showed irregularities at low flow rates and became steep at high flow rates. Manson5 has reported performance characteristics for jet engine fuel pumps having straight radial blades in enclosed impellers. The head curves showed unstable characteristics at low flow rates and steep characteristics at higher flow rates. The best efficiency reported was 54.7% for an impeller diameter of 3.300 in (8.382 cm) and speed n  28,650 rpm. Forward-Curved Blades, B2 7 90° Pumps with forward-curved blades have been proposed,6 but the research necessary to achieve an efficient design appears never to have been carried out. Tests have been made of conventional, backward-curved-blade, doublesuction pumps with the impellers mounted in the reversed position but with rotation correct for the volute casing. As tested, these pumps therefore had forward-curved blades. FLOW RATE FIGURE 5 Pump characteristics, radial blades (ft ⫻ 0.3048 = m; hp ⫻ 745.7 = W; gpm ⫻ 0.06309 = l/s) (Reference 3) 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.333 FIGURE 6 Pump efficiency versus specific speed and size (gpm ⫻ 0.06309 = l/s) (Flowserve Corporation) (s = ns/2733) FLOW RATE FLOW RATE FIGURE 7 Pump characteristics, radial blades (lb/in2 ⫻ 6.894 = kPa; gpm ⫻ 0.06309 = l/s) (Reference 4) Table 1 shows the pertinent results for six different pumps. Both flow rate and efficiency were drastically reduced, and there was only a modest increase in head for five of the six pumps. The sixth pump showed a 38% increase in head over that obtained with the impeller correctly mounted. Published estimates7,8 of the head-flow curves to be expected from reversed impellers predict an unstable characteristic at the low end of the flow rate range and a steep characteristic at the high end of the range. PERFORMANCE EFFECTS_____________________________________________ Classification of Curve Shapes A useful method for comparing characteristics of pumps of different specific speeds is to normalize on a selected operating condition, usually best efficiency. Thus, 2.334 CHAPTER TWO TABLE 1 Effects of reversed mounting of impeller Number of stages Specific speed per stage ns (s) Percent of normal shutoff head 2 2 1 1 1 1 828 (0.303) 1024 (0.375) 1240 (0.454) 1430 (0.523) 2570 (0.940) 2740 (1.003) 86 82 75 82 74.5 77.5 Percent of normal values at best efficiency Head Flow Rate Power Efficiency 111 112 105 106 117 138 65 88 38.5 69.7 62 61.5 104 145 68.5 138 138 180 71 68 59 53.5 52.5 47 Source: Flowserve Corporation q Q Qn h H Hn p P Pn (10) where the subscript n designates values for the best efficiency point. Figures 8, 9, and 10 show approximate performance curves normalized on the conditions of best efficiency and for a wide range of specific speeds as defined in Table 2. These curves are applicable to pumps of any size because absolute magnitudes have been eliminated. In Figure 8, curves 1 and 2 exhibit a rising head or unstable characteristic where the head increases with increasing flow rate over the lower part of the flow rate range. This may cause instability at heads greater than the shutoff value, particularly if two or more pumps are operated in parallel. Curve 3 exhibits an almost constant head at low flow rates and is often called a flat characteristic. Curves 4 to 7 are typical of a steep or stable head characteristic, in which the head always decreases with increasing flow rate. Although the shape of the head-flow curve is primarily a function of the specific speed, the designer has some control through selection of the vane angle b2 number of impeller vanes nb, and capacity coefficient f  cm2/u2, as described in Section 2.1 (see also Figure 2). For pumps having a singlesuction specific speed approximately 5000 (1.83) and higher, the power is at its maximum at shutoff and decreases with increasing flow rate. This may require an increase in the power rating of the driving motor over that required for operation at normal capacity. Efficiency The efficiency h is the product of three component efficiencies (defined in Section 2.1): h  hmhvhh (11) The mechanical efficiency hm accounts for the bearing, stuffing box, and all disk-friction losses including those in the wearing rings and balancing disks or drums if present. The volumetric efficiency hv accounts for leakage through the wearing rings, internal labyrinths, balancing devices, and glands. The hydraulic efficiency hh accounts for liquid friction losses in all through-flow passages, including the suction elbow or nozzle, impeller, diffusion vanes, volute casing, and the crossover passages of multistage pumps. Figure 11 shows an estimate of the losses from various sources in double-suction single-stage pumps having at least 12-in (30-cm) discharge pipe diameter. Minimum losses and hence maximum efficiencies are seen to be in the vicinity of ns ⬇ 2500 (0.91), which agrees with Figure 6. Effects of Pump Speed Increasing the impeller speed increases the efficiency of centrifugal pumps. Figure 7 shows a gain of about 15% for an increase in speed from 15,000 to 30,000 rpm. The increases are less dramatic at lower speeds. For example, Ippen9 reported about 1% increase in the efficiency of a small pump, D  8 in (20.3 cm) and hs  1992 (0.73), at best efficiency, for an increase in speed from 1240 to 1880 rpm. Within limits, the cost of the pump and driver usually decreases with increasing speed. Abrasion 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS FLOW RATE FIGURE 8 Head curves for several specific speeds, as defined in Table 2 (Reference 12) FIGURE 9 Power curves for several specific speeds, as defined in Table 2 (Reference 12) FLOW RATE FLOW RATE FIGURE 10 Efficiency curves for several specific speeds, as defined in Table 2 (Reference 12). 2.335 2.336 CHAPTER TWO TABLE 2 Characteristic curves as a function of specific speed (Figures 8, 9, and 10) Curve Number on USCS Specific Figures 8 , 9 , and 10 Speed ns 1 2 3 4 5 6 7 900 1500 2200 3000 4000 5700 9200 Metric (SI) Specific Speed nq Universal Specific Speed s Impeller Suction Configuration 17 29 43 58 77 110 178 0.33 0.55 0.80 1.10 1.46 2.09 3.37 Double Double Double Double Double Single Single UNIVERSAL SPECIFIC SPEED ΩS 0 FIGURE 11 Power balance for double-suction pumps at best efficiency (Reference 12) and wear increase with increasing speed, particularly if the liquid contains solid particles in suspension. The danger of cavitation damage usually increases with increasing speed unless certain suction requirements can be met, as described later. Effects of Specific Speed Figues 6 and 11 show that maximum efficiency is obtained in the range 2000 (0.73) 6 ns 6 3000 (1.10), but this is not the only criterion. Pumps for high heads and small flow rates occupy the range 500 (0.18) 6 ns 6 1000 (0.37). At the other extreme, pumps for very low heads and large flow rates may have ns  15,000 (5.49) or higher. For given head and flow rate, the pump having the highest specific speed that will meet the requirements probably will be the smallest and least expensive. However, Figure 9 of Section 2.1 shows that it will run at the highest speed and be subject to maximum wear and cavitation damage, as previously mentioned. Effects of Clearance Details of wearing-ring construction are given in Subsection 2.2.1. Schematic outlines of two designs of rings are shown in Figure 12. The L-shaped construction shown in Figure 12a is very widely used with the close clearance between the cylindrical portions of the rings. Leakage losses increase and pump performance WEARING-RING CLEARANCE 2.337 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS FIGURE 12A and B Typical wearing rings TABLE 3 Effects of increased wearing-ring clearance on centrifugal pump performance Specific speed ns (s) Design head, ft (m) Ring clearance, % of normal value 2100 (0.77) ... ... ... 3500 (1.28) 4300 (1.57) 4800 (1.76) 63 (19.2) ... ... ... 65 (19.8) 41(12.5) 26 (7.9) 178 356 688 1375 354 7270 5220 Percent of values at shutoff with normal ring clearance Percent of values at best efficiency with normal ring clearance Q H P h H0 P0 100 100 100 100 100 62 96 98.3 97.5 96.0 94.3 90.0 65.5 78.8 98.9 99.0 98.9 97.4 99.1 81.7 89.2 99.4 98.5 97.1 96.8 90.8 49.8 84.8 97.0 93.6 91.2 88.8 85.0 44.3 78.2 100 98.2 94.8 92.5 96.2 106 83.3 Source: Flowserve Corporation falls off as the rings wear. Table 3 shows some of the effects of increasing the clearance of rings similar to Figure 12a. The labyrinth construction shown in Figure 12b has been used to increase the leakage path without increasing the axial length of the rings. If the pressure differential across these rings is high enough, the pump shaft may take on lateral vibrations with relatively large amplitude and long period, which can cause serious damage. One remedy for these vibrations is to increase clearance 2 in Figure 12b relative to clearance 1, at the expense of an increased leakage flow. High-pressure breakdown through plain rings may cause vibration,10 but this is not usually a serious problem. It is considered good practice to replace or repair wearing rings when the nominal clearance has doubled. The presence of abrasive solids in the liquid pumped may be expected to increase wearing-ring clearances rapidly. Many impellers are made without an outer shroud and rely on close running clearances between the vane tips and the casing to hold leakage across the vane tips to a minimum. Although this construction usually is not used with pumps having specific speeds less than about 6000 (2.20), Wood et al.11 have reported good results with semiopen impellers at 1800 (0.66)  ns  4100 (1.50). It appears that both head and efficiency increase with decreasing tip clearance and are quite sensitive to rather small changes in clearance. Reducing the tip clearance from about 0.060 in (1.5 mm) to about 0.010 in (0.25 mm) may increase the efficiency by as much as 10%. Abrasive solids in the liquid pumped probably will increase tip clearances rapidly. VANE-TIP CLEARANCE 2.338 CHAPTER TWO MODIFICATIONS TO IMPELLER AND CASING _____________________________ Diameter Reduction To reduce cost, pump casings usually are designed to accommodate several different impellers. Also, a variety of operating requirements can be met by changing the outside diameter of a given radial impeller. Eq. 6 shows that the head should be proportional to (nD)2 provided that the exit velocity triangles (Figure 2b) remain similar before and after cutting, with w2 always parallel to itself as u2 is reduced. This can be achieved if the impeller meridional exit area Am,2 is the same before and after cutting—and if the flow angle bf,2 (Figure 15, Section 2.1) also stays the same. [bf,2 would stay the same if the blade angle b2 ( bb,2) does, the difference being due to slip velocity Vs that should also scale down with D.] For radial discharge impellers, area Am,2 equals pDb2 (minus blade and boundary layer blockage) and requires that b2 increase as D decreases. This is typical of many impellers—as is constancy of b2 over the cutting range—and, together with Eqs. 27–33 of Section 2.1, leads to the so-called “affinity laws” for predicting performance: Q1 n1D1  n2D2 Q2 (12a) n21D21 H1  2 2 H2 n2D2 (12b) P1 n31D31  3 3 P2 n2D2 (12c) which apply only to a given impeller with altered D and constant efficiency but not to a geometrically similar series of impellers. The assumptions on which Eqs. 12 were based are rarely if ever fulfilled in practice, so exact predictions by the equations should not be expected. A common example is the low-ns radial discharge impeller with parallel radial hub and shroud profiles over most of the path from inlet to exit (Figure 6). Here, Am,2 decreases with cutting, and H falls more than would be predicted by Eq. 12b. [This type of impeller is often found in multistage pumps, particularly those in which the designer, driven by cost reduction goals, has minimized a) the axial length occupied by each stage and b) the number of stages, thereby pushing down the ns of the individual stage to the point that a tolerable sacrifice in efficiency results.] Impellers of low specific speed may be cut successfully provided the following items are kept in mind: RADIAL DISCHARGE IMPELLERS 1. The angle b2 may change as D is reduced, but this usually can be corrected by filing the blade tips. (See the discussion on blade-tip filing that follows.) 2. Tapered blade tips will be thickened by cutting and should be filed to restore the original shape. (See the discussion below on blade-tip filing.) 3. Bearing and stuffing box friction remain constant, but disk friction should decrease with decreasing D. 4. The length of flow path in the pump casing is increased by decreasing D. 5. Because cm1 is smaller at the reduced capacity, the inlet triangles no longer remain similar before and after cutting, and local flow separation may take place near the blade entrance tips. 6. The second right-hand term in Eq. 5 was neglected in arriving at Eqs. 12, but it may represent a significant decrease in head as D is reduced. 7. Some blade overlap should be maintained after cutting. Usually the initial blade overlap decreases with increasing specific speed, so the higher the specific speed, the less the allowable diameter reduction. 8. Diameter reductions greater than from 10 to 20% of the original full diameter of the impeller are rarely made. 2.339 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS Most of the losses are approximately proportional to Q2, and hence to D2 by Eq. 12. Because the power output decreases approximately as D3, it is reasonable to expect the maximum efficiency to decrease as the wheel is cut, and this often is the case. By Eq. 12 and the ns-definition (Eq. 38a of Section 2.1), the product nsD should remain constant so the specific speed at best efficiency increases as the wheel diameter is reduced (Table 4). The characteristics of the pump shown in Figure 13 may be used to illustrate reduction 5 of diameter at constant speed. Starting with the best efficiency point and D  1616 in (41.4 cm), let it be required to reduce the head from H  224.4 to H¿  192.9 ft (68.4 m to 58.8 m) and to determine the wheel diameter, capacity, and power for the new conditions. Because the speed is constant, Eqs. 12 may be written H  kHQ2 and P  kpQ3 (13) 5 l616 in (41.4 where kH and kp may be obtained from the known operating conditions at D  cm). Plot a few points for assumed capacities and draw the curve segments as shown by the solid lines in Figures 13b and 13c. Then, from Eqs. 12 D¿  D2H¿>HQ¿  Q2H¿>H P¿  P1H¿>H2 3>2  D¿  D1Q¿>Q2H¿  H1Q¿>Q2  2 P¿  P1Q¿>Q2 3 (14a) (14b) 1518 from which D  in (38.4 cm), Q  3709 gpm (234 l/s), and P¿  215.5 hp (160.7 kW). In Figure 13, the initial conditions were at points A and the computed conditions after cutting at points B. The test curve for D  1518 in (38.4 cm) shows the best efficiency point a at a lower flow rate than predicted by Eqs. 14, but the head curve satisfies the predicted values very closely. The power prediction was not quite as good. Table 4 and Figure 13 give actual and predicted performance for three impeller diameters. The error in predicting the best efficiency point was computed by (predicted value minus test value) (100)/(test value). As the wheel diameter was reduced, the best efficiency point moved to a lower flow rate than predicted by Eq. 14 and the specific speed increased, showing that the conditions for Eqs. 12 to hold were not maintained. Wheel cutting should be done in two or more steps with a test after each cut to avoid too large a reduction in diameter. Figure 14 shows an approximate correction, given by Stepanoff,12 that may be applied to the ratio D¿/D as computed by Eqs. 12 or 14. The accuracy of the correction decreases with increasing specific speed. Figure 15 shows a correction proposed by Rütschi13 on the basis of extensive tests on low-specific-speed pumps. The corrected diameter reduction D is the diameter reduction D  D¿ given by Eqs. 14 and multiplied by k from Figure 15. The shaded area in Figure 15 indicates the range of scatter of the test points operating at or near maximum efficiency. Near shutoff the values of k were smaller and at maximum flow rate the values of k were larger than shown in Figure 15. Table 5 shows the results of applying Figures 14 and 15 to the pump of the preceding example. There is no independent control of Q and H in impeller cutting, although Q may be increased somewhat by underfiling the blade tips as described later. The flow rate and power will automatically adjust to the values at which the pump head satisfies the system head-flow curve. Diameter reduction of mixed-flow impellers is usually done by cutting a maximum at the outside diameter D0 and little or nothing at the inside diameter Di, as shown in Figure 16. Stepanoff14 recommends that the calculations be based on the average diameter Dav  (Di  D0)/2 or estimated from the blade-length ratio FK/EK or GK/EK in Figure 16d. Figure 16 shows a portion of the characteristics of a mixed-flow impeller on which two cuts were made as in Figure 16b. The calculations were made by Eqs. 14 using the mean diameter MIXED-FLOW IMPELLERS Dm  21D2o  D2i 2>2 instead of the outside diameter in each case. The predictions and test results are shown in Figure 16 and Table 6. It is clear that the actual change in the characteristics far exceeded TABLE 4 Predicted characteristics at different impeller diameters on a radial-flow pump Test values Predicted from 5 D  1616 in (41.4 cm) Predicted from D  1518 in (88.4 cm) 2.340 D, in (cm) Q, gpm (l/s) H, ft (m) P, hp (kW) ns (s) nsD (sD) Q¿, gpm (l/s) H¿, ft (m) P¿, hp (kW) Q¿, gpm (l/s) H¿ ft (m) P¿, hp (kW) 5 1616 (41.4) 4000 (252) 224.4 (68.4) 270.4 (201.6) 1953 (0.7146) 31,860 (29.54) ... ... ... 1518 (38.4) 3600 (227.1) 195.4 (59.6) 217.0 (161.8) 2055 (0.7519) 31,080 (28.87) 163.6 (49.9) 167.2 (124.7) 2214 (0.8101) 31,000 (28.84) 215.5 (160.7) 0.69% errora 170.9 (127.4) 2.33% errora 272.2 (203.0) 0.68% errora ... 3200 (201.9) 192.9 (58.8) 1.28% errora 165.3 (50.4) 1.03% errora 227.3 (69.3) 1.29% errora ... 14 (35.6) 3709 (234) 3.02% errora 3433 (217) 7.28% errora 3888 (245) 2.93% errora ... 3332 (210.2) 4.13% errora 167.4 (51.0) 2.33% errora 172.1 (128.3) 2.93% errora Error in predicting best efficiency point. a 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.341 (0.715) (0.752) (0.810) FLOW RATE Q, GPM FIGURE 13A through C Diameter reduction of radial-flow impeller (gpm ⫻ 0.06309 = l/s). D is measured in inches (centimeters). the predicted values. Except for the use of the mean diameters, the procedure was essentially the same as that described for Figure 13, and all points and curves are similarly labeled. The corrections given in Figure 14 would have made very little difference in the computed diameter reductions, and those of Figure 15 were not applicable to impellers having specific speeds greater than ns  2000 (0.73). In this case, the product nsDm did not remain constant and the maximum efficiency increased as Dm was reduced. Although the 2.342 CHAPTER TWO FIGURE 14 FIGURE 15 Corrections for calculated impeller diameter reductions (Reference 12) Corrections for calculated impeller diameter reductions (Reference 13). (s  ns/2733) TABLE 5 Impeller diameter corrections D before cutting D¿ predicted by Eqs. (14) D¿ corrected by Figure 14 D¿ corrected by Figure 15 in in in in 16.3125 15.125 15.25a 15.60c 16.3125 14.000 14.26b 14.93d a D¿/D  15.125/16.3125  0.927; by Figure 14, corrected D¿/D  0.935. Corrected D¿  (0.935)(16.3125)  15.25 in. b D¿/D  14.000/16.3125  0858; by Figure 14, corrected D¿/D  0.874. Corrected D¿  (0.874)(16.3125)  14.26 in. c D  D¿  16.3125—15.125  1.1875; by Figure 15, K ⬵ 0.6 at ns  1,953. Corrected D  D¿  (0.6)(1.1875)  0.7125 and corrected D¿  16.3125—0.7125  15.60 in. d D  D¿  16.3125—14.000  2.3125; by Figure 15, K ⬵ 0.6 at ns  1,953. Corrected D  D¿ (0.6)(2.3125)  1.3875 and corrected D¿  16.3125—1.3875  14.93 in. changes in diameter were small, the area of blade removed was rather large for each cut. The second cut eliminated most of the blade overlap. The characteristics of mixed-flow impellers can be changed by cutting, but very small cuts may produce a significant effect. The impellers of propeller pumps are not usually subject to diameter reduction. Shaping Blade Tips If the discharge tips of the impeller blades are thick, performance usually can be improved by filing over a sufficient length of blade to produce a long, gradual taper. Chamfering, or rounding, the discharge tips may increase the losses and should never be done. Reducing the impeller diameter frequently increases the tip thickness. 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.343 FLOW RATE Q, GPM FIGURE 16A through C Diameter reduction of mixed-flow impeller (ft ⫻ 0.3048 = m; in ⫻ 2.540 = cm; gpm ⫻ 0.06309 = l/s; hp ⫻ 0.7457 = kW) OVERFILING This is shown at B in Figure 17, and the unfiled blade is shown at A. Usually there is little or no increase in the blade spacing d before and dF after filing, so the discharge area is practically unchanged. Experience indicates that any change in the angle b2 due to overfiling usually produces a negligible change in performance. UNDERFILING This is shown at C in Figure 17. If properly done, underfiling will increase the blade spacing from d to df and hence the discharge area, which lowers the average meridional velocity cm2, at any given flow rate Q. The angle b2 usually is increased slightly. Figure 18A and Eq. 6 show that the head and consequently the power increase at the same flow rate. The maximum efficiency usually is improved and may be moved to a higher flow rate. At the same head, Figure 18B shows that both cm2 and the flow rate will increase. The change both in the area and in Cm2 may increase the flow rate by as much as 10%. Table 7 shows the results of tests before and after underfiling the impeller blades of nine different pumps. In general, they confirm the foregoing predictions based on changes in area and in the velocity triangles. TABLE 6 Predicted characteristics at different impeller diameters on a mixed-flow pump Predicted from D  16.00 (40.64 cm), Dm  13.17 (33.45 cm) Test values 2.344 D, in (cm) Dm, in (cm) Q, gpm (l/s) H, ft (m) P, hp (kW) ns nsDm 16 (40.64) 13.17 (33.45) 5800 (365.9) 18.6 (5.67) 32.5 (24.2) 7385 (2.702) 1517 32 (39.45) 12.89 (32.74) 5100 (321.8) 17.1 (5.21) 25.7 (19.2) 1 1516 (38.26) 12.62 (32.05) 4600 (290.2) 16.8 (5.12) 22.6 (17.1) Error in predicting best efficiency point. a Q¿, gpm (l/s) H¿, ft (m) P¿, hp (kW) 97,300 (90.39) ... ... ... 7385 (2.702) 95,200 (88.47) 7100 (2.598) 89,600 (83.26) 5680 (358.4) 10.2% errora 5560 (350.8) 17.3% errora 17.8 (5.43) 4.50% errora 17.1 (5.21) 1.75% errora 30.4 (22.7) 15.5% errora 28.6 (21.6) 21.0% errora Predicted from D  15.53 (39.45 cm), Dm  12.89 (32.74 cm) Q¿, gpm (l/s) H¿, ft (m) 5210 17.9 (329) (5.46) 11.3% 3.91% error error 5000 (315.5) 8.00% errora 16.4 (5.00) 2.44% errora P¿, bp (kW) 27.4 (20.4) 18.6% error 24.1 (18.2) 6.22% errora 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS FIGURE 17 FIGURE 18 2.345 Underfiling and overfiling of blade tips Discharge velocity triangles for underfiled blades (neglecting slip) If the inlet blade tips are blunt, as shown at D in Figure 17, the cavitation characteristics may be improved by sharpening them, as shown at E. In this case overfiling increases the effective flow area, which reduces cm1 for a given flow rate. If more area is needed, it may be advantageous to cut back part of the blade and sharpen the leading edge. Overfiling tends to increase b1, which is incompatible with a decrease in cm1 (Figure 19). The increase in b1 increases the angle of attack of the liquid approaching the blade. In Figures 2 and 19, w1 is tangent to the centerline of the blade at entrance and w0 INLET BLADE TIPS 2.346 CHAPTER TWO TABLE 7 Changes in performance when impeller blades are underfiled Specific speed ns s No. of stages Impeller diameter D2, in (cm) 862 (0.3154) 1 945 (0.3458) 2 1000 (0.3659) 4 1080 (0.3952) 2 1525 (0.5580) 2 1112 838 918 913 16 1012 1950 (0.7135) 1 3300a (1.2075) 1 Change in blade spacing dF/d HF  H H %b Changes at best efficiency point after filing, % Hoc H Q P h (29.21) 1.13 4 1.5 4 0 9 (21.27) 1.05 5.5 0 3 4 0.5 7.5d (23.18) 1.055 5.5 0 5.5 0 3 2.5 (24.92) 1.08 6.8 3 (26.67) 1.035 3.2 3 0.5 4.5 3.5 1.2 22 (55.88) 1.02 1.5 1 1.5 0 1.5 0 12 (30.48) ... 7.8 2.5 7.8 0 8.5 0.6 ... 6.5 0.5 ... 5 3450a (1.2623) 1 3012 4300 (1.5734) 1 4141 (77.47) (104.8) 10 2.5 10 0 5 6.5 0 7.8 2.2 5 0.5 0 4.5e Double suction. HF is head after filing. c Shutoff head. d Due in part to changes in pump casing. e Due in part to rounding of inlet blade tips. Source: Flowserve Corporation a b FIGURE 19 Effect of overfiling at inlet on the velocity triangle just inside the blades is the velocity of the approaching liquid as seen by an observer moving with the blade. The angle d between w0 and w1 is the angle of attack, which increases to dF after overfiling. Although the increase in d tends to increase the opportunity for the liquid to separate from the low-pressure face of the blade, this disadvantage usually is outweighed by the improvement that results when cm1 is reduced.* 1 *The distinction between a true velocity diagram formed just upstream of the impeller blades and the triangles shown just inside the blades in Figure 19 should be noted. w1 in this figure is shown aligned with the blade, having undergone an adjustment in direction via incidence from that of the approaching w0. Only by plotting the velocity diagram just upstream of this point can one determine the true value of the circumferential component cu1 of the absolute velocity being delivered to the impeller by the upstream piping, vanes, and so on. It is this delivered value of cu1 that must be used in computing input or ideal head of the impeller (Eq. 15b, Section 2.1 or, neglecting slip, Eq. 6 of this subsection). Thus w0 is properly the true w1 vector and should be understood as such (see Section 2.1). 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.347 CASING TONGUE The casing tongue or cutwater forms part of the throat of the discharge nozzle of many volute casings (Figure 20a, Section 2.1). Frequently the throat area is small enough to act as a throttle and reduce the maximum flow rate otherwise obtainable from the impeller. Cutting back the tongue increases the throat area and increases the maximum flow rate. The head-versus-flow rate characteristic is then said to carry out farther. Shortening the discharge nozzle may increase the diffusion losses a little and result in a slightly lower efficiency. CAVITATION _________________________________________________________ The formation and subsequent collapse of vapor-filled cavities in a liquid due to dynamic action are called cavitation. The cavities may be bubbles, vapor-filled pockets, or a combination of both. The local pressure must be at or below the vapor pressure of the liquid for cavitation to begin, and the cavities must encounter a region of pressure higher than the vapor pressure in order to collapse. Dissolved gases often are liberated shortly before vaporization begins. This may be an indication of impending cavitation, but true cavitation requires vaporization of the liquid. Boiling accomplished by the addition of heat or the reduction of static pressure without dynamic action of the liquid is arbitrarily excluded from the definition of cavitation. When a liquid flows over a surface having convex curvature, the pressure near the surface is lowered and the flow tends to separate from the surface. Separation and cavitation are completely different phenomena. Without cavitation, a separated region contains turbulent eddying liquid at pressures higher than the vapor pressure. When the pressure is low enough, the separated region may contain a vapor pocket that fills from the downstream end,15 collapses, and forms again many times each second. This causes noise and, if severe enough, vibration. Vapor-filled bubbles usually are present and collapse very rapidly in any region where the pressure is above the vapor pressure. Knapp16 found the life cycle of a bubble to be on the order of 0.003 s. Bubbles that collapse on a solid boundary may cause severe mechanical damage. Shutler and Mesler17 photographed bubbles that distorted into toroidal-shaped rings during collapse and produced ring-shaped indentations in a soft metal boundary. The bubbles rebounded following the initial collapse and caused pitting of the boundary. Pressures on the order of 104 atm have been estimated18 during collapse of a bubble. All known materials can be damaged by exposure to bubble collapse for a sufficiently long time. This is properly called cavitation erosion, or pitting. Figure 20 shows extensive damage to the suction side of pump impeller vanes after about three months’ operation with cavitation. At two locations, the pitting has penetrated deeply into the 83-in (9.5-mm) thickness of stainless steel. The unfavorable inlet flow conditions, believed to have been the cause of the cavitation, were at least partly due to elbows in the approach piping. Modifications in the approach piping and the pump inlet passages reduced the cavitation enough to extend impeller life to several years.19 It has been postulated that high temperatures and chemical action may be present at bubble collapse, but any damaging effects due to them appear to be secondary to the mechanical action. It seems possible that erosion by foreign materials in the liquid and cavitation pitting may augment each other. Controlled experiments20 with water indicated that the damage to metal depends on the liquid temperature and was a maximum at about 100 to 120°F (38 to 49°C). Cavitation pitting, as measured by weight of the boundary material removed per unit time, frequently increases with time. Cast iron and steel boundaries are particularly vulnerable. Controlled experiments have shown that cavitation pitting in metals such as aluminum, steel, and stainless steel depends strongly on the velocity of the fluid in the undisturbed flow past the boundary. On the basis of tests of short duration, Knapp15 reported that damage to annealed aluminum increased approximately as the sixth power of the velocity of the undisturbed flow past the surface. Hammitt21 found a more complicated relationship between velocity and damage, as shown in Figure 21. It seems clear that after cavitation begins, it will increase rapidly with increasing velocities. Frequently the rate of pitting accelerates with elapsed time. Comprehensive discussions of cavitation and erosion together with additional references are given by Preece.22 2.348 CHAPTER TWO FIGURE 20 FIGURE 21 Impeller damaged by cavitation (Demag Delaval, Reference 19) Cavitation damage exponent versus test time for several materials in water (Reference 21) Centrifugal pumps begin to cavitate when the suction head is insufficient to maintain pressures above the vapor pressure throughout the flow passages. The most sensitive areas usually are the low-pressure sides of the impeller vanes near the inlet edge and the front shroud, where the curvature is greatest. Axial-flow and high-specific-speed impellers without front shrouds are especially sensitive to cavitation on the low-pressure 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.349 sides of the vane tips and in the close tip-clearance spaces. Sensitive areas in the pump casing include the low-pressure side of the tongue and the low-pressure sides of diffusion vanes near the inlet edges. As the suction head is reduced, all existing areas of cavitation tend to increase and additional areas may develop. Apart from the noise and vibration, cavitation damage may render an impeller useless in as little as a few weeks of continuous operation. In multistage pumps cavitation usually is limited to the first stage, but Kovats23 has pointed out that second and higher stages may cavitate if the flow is reduced by lowering the suction head (submergence control). Cavitation tends to lower the axial thrust of an impeller. This could impair the balancing of multistage pumps with opposed impellers. A reduction in suction pressure may cause the flow past a balancing drum or disk to cavitate where the liquid discharges from the narrow clearance space. This may produce vibration and damage resulting from contact between fixed and running surfaces. Net Positive Suction Head (NPSH) Several criteria exist for establishing the NPSH of a pump. These are connected with a) inception of cavitation, b) loss of hydraulic performance, and c) protecting the pump against cavitation erosion or damage. These are defined and found as follows: a. Inception NPSH. Cavitation can be completely prevented so long as the static pressure within the pump is everywhere greater than the vapor pressure of the liquid. This can be achieved if the NPSH (also called hsv), defined in Section 2.1 as the total head of the pumped liquid at the pump inlet datum or suction flange above that vapor pressure, is sufficiently large. The value of NPSH that achieves this is called NPSHi, namely, the “inception NPSH.” b. Performance NPSH. In a typical NPSH-test at constant speed n and flow rate Q, a substantial reduction of NPSH below NPSHi is usually necessary to reach the value that produces an identifiable drop in performance—usually 3% of pump stage total pressure rise or pump head. This value is called the “required NPSH,” NPSHR or NPSH3%. Table 1 in Section 2.1 provides empirical correlations for NPSH3% for common pumps and inducers. Other criteria for NPSH3% will be given further on. The domains of cavitation within a pump are widespread at this condition, and experience shows that a head drop in the neighborhood of 3% must occur in order to obtain a repeatable value of NPSHR on test12. Lower percentages have been demanded by users, but they invariably give rise to a large scatter in the measured NPSHR for even small variations in any of the test variables. It used to be thought throughout the pump community that there are no cavities or bubbles present at zero percent head drop due to cavitation activity within a pump. In recent decades, however, it has been proven through many observations that not only is the head drop equal to zero at inception (where NPSH  NPSHi), it remains so for an enormous range of lesser NPSH-values over which extensive bubble activity is observed. In fact, NPSHi is commonly from two to five times the magnitude of the NPSHR that is associated with any noticeable drop in pump head. At the 3% (or lesser percentage) head drop condition, actual observations of the cavitating flow show the cavities to extend all the way from the leading edge of each blade to the throat formed by the leading edge of the next blade24. In the face of these learnings, it is evident that to demand a test for the NPSH required for a head drop of much less than 3% usually causes only needless misunderstandings and expenditure of time and money. c. Damage-Limiting NPSH or Life NPSH. If it is desired to substantially reduce or eliminate cavitation activity within the pump, an acceptance test should be conducted wherein the cavitating flow can be observed visually25. This is indeed a serious issue if the pump energy level is high enough (and therefore the inlet pressure) for the collapsing cavities to do damage (as in Figure 20; see also Section 2.1). In such a case, “NPSH3%” has virtually no meaning; rather, the truly required NPSH or “NPSHR” is that larger value which satisfactorily limits the extent of this visually observed cavitation within the pump or pump model26. Therefore, the available NPSH at the installation must be at least equal to the required NPSH if the above consequences are to be avoided—be they significant loss of 2.350 CHAPTER TWO FIGURE 22 Definition sketch for computing NPSH pump head (that is, loss of hydraulic performance) or damage due to collapsing cavities. Increasing the available NPSH above the required value provides a margin of safety against unpredictable variations of conditions, including transient behavior. Figure 22 and the following symbols will be used to compute the available NPSH: pa  absolute pressure in atmosphere surrounding gage, Figure 22 ps  gage pressure indicated by gage or manometer connected to pump suction at section s-s; may be positive or negative pt  absolute pressure on free surface of liquid in closed tank connected to pump suction pvp  vapor pressure of liquid being pumped corresponding to the temperature at section s-s (if liquid is a mixture of hydrocarbons, pvp must be measured by the bubble point method) hf  lost head due to friction in suction line between tank and section s-s V  average velocity at section s-s Z, Zps  vertical distances defined by Figure 22; may be positive or negative g  specific weight of liquid at pumping temperature It is satisfactory to choose the datum for small pumps as shown in Figures 1 and 22, but with large pumps, the datum should be raised to the elevation where cavitation is most likely to start. For example, the datum for a large horizontal-shaft propeller pump should be taken at the highest elevation of the impeller-blade tips. The available NPSH is given by hsv  pa  pvp g  ps V2  Zps  g 2g (15) or hsv  pt  pvp g  Z  hf (16) 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.351 Consistent units must be chosen so that each term in Eqs. 15 and 16 represents feet (or meters) of the liquid pumped. Equation 15 is useful for evaluating the results of tests. Equation 16 is useful for estimating available NPSH during the design phase of an installation. In Eq. 15, the first term represents the height of a liquid barometer, hb, containing the liquid being pumped and the sum of the remaining terms represents the suction head hs; that is, the total head H (Eq. 3, Section 2.1) evaluated at the suction flange S in Figure 22. Therefore (17) hsv  hb  hs Usually, a positive value of hs is called a suction head and a negative value of hs is called a suction lift. Many pumps draw cold water from reservoirs exposed to atmospheric pressure, so the suction lift is limited if there is to be a reasonable value of hsv ( NPSH) available at the pump suction flange or port. In normal practice, rather small values of suction lift or suction head are encountered in many installations, so one can depend on the value of hsv being a substantial portion of hb —say about 60% or 20 ft (6m) of water column. Recognizing that the suction specific speed capability S of most pumps falls within a rather small range has led to charts of recommended operating speeds such as those in Section 9.10 or the speed recommendations of Figure 23 for condensate pumps for varying amounts of NPSH available. In fact, S  8500 (ss  3.11) is recommended by the Hydraulic Institute27. S ( Nss) and ss are defined in Eqs. 41 and 42 of Section 2.1. The relationship of suction-specific speed to other suction parameters is treated further on. FLOW RATE Q, GPM FLOW RATE Q, GPM Cavitation Tests In addition to the constant-Q tests for NPSHR that were described in the foregoing review of the various NPSH-criteria, one can establish the curve of performance NPSH versus flow rate by varying the NPSH available and seeing how much flow FIGURE 23 Capacity (flow rate) and speed limitations for condensate pumps with shaft through eye of impeller (ft ⫻ 0.3048 ⫺ m; gpm ⫻ 0.06309 = l/s) (adapted from Hydraulic Institute ANSI/HI 2000 Edition Pump Standards, Reference 27) 2.352 CHAPTER TWO FIGURE 24 Test of a 1.5-in (3.81-cm) single-stage pump at 3470 rpm, on water, 70ºF (2lºC) (ft ⫻ 0.3048 = m; gpm ⫻ 0.06309 = l/s) (Reference 12) rate Q the pump can handle. This is illustrated in Figure 24, wherein a different head curve is found for each value of NPSH. For each such curve, as Q is increased, there is a point at which the head departs from the next one of higher NPSH. This, then, is the Q for which the given NPSH is that required to maintain head, essentially NPSH3%. At higher flow rate, performance is lost, each head curve possessing a “cutoff” flow rate. References 27 and 28 may be consulted for details of test procedures. Thoma Cavitation Parameter S All the terms in Eqs. 15 to 17 may be made dimensionless by dividing each by the pump head H. The resulting parameter s hsv H (18) has proven to be useful especially for high-specific speed pumps and turbines. The loading on the blades of a variable-pitch propeller, and therefore the minimum suction-side static pressure, is directly connected to the head developed. As this head increases with increasing blade setting angle, the loading increases, the minimum pressure drops, and the NPSHR goes up. Hence an identifiable limiting value of s is found to exist for such a machine. s-limits are presented for a range of specific speeds in Figure 25. Performance-NPSH data are sometimes given in terms of s; for example, Rütschi13 presents the effect of pump hydraulic efficiency on NPSH in these terms, a consistently higher value of s being required for pumps with lower efficiency (Figure 26). Suction-Specific Speed For pumps in which the eye diameter of the impeller is smaller than that at exit, the pressures on the suction-sides of the blades that are associated with most of the head addition are in a region of higher pressure than in the vicinity of the eye. Thus Wislicenus29 was able to decouple the eye from outer diameter of the pump (where most of the head is generated in a radial-flow machine) and show that only the eye geometry determines performance-NPSH and not the magnitude of the head created by the impeller. He demonstrated that for most centrifugal pumps (except the propellers just described), there is a small range of suction specific speed—from which the performance-NPSH can be determined. As in defined in Section 2.1, this parameter is defined as follows: S N2Q 1hsv 2 3>4 (19) Note that Q  half the discharge of a double-suction impeller when computing S. Equations 18 and 19 and the definition of ns may be combined to yield 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS FIGURE 25 FIGURE 26 2.353 Cavitation limits of centrifugal and propeller pumps (Flowserve Corporation) Cavitation parameter versus specific speed for different efficiencies (s = ns/2733) (Reference 13) 2.354 CHAPTER TWO FIGURE 27 Commercial pumps applicable to zero suction head. Suction specific speed S = 8000 (ss = 2.93). For centrifugal impellers, ns = 1,000-5,000 (s = 0.4-1.8); mixed-flow, 5,000-9,000 (1.8-3.3); propeller, 9,000-13,000 (1.3-4.8). s a ns 4>3 b S ns  S1s2 3>4 (20) (21) Figure 25 shows lines of constant S. For most pumps, 7500 6 S 6 11,000 (2.7 6 SS 6 4.0), S  8500 (3.11) being recommended by the Hydraulic Institute ANSI/HI 2000 Pump Standards (Reference 27). Higher values may apply to special designs or service conditions, such as an inducer ahead of the first-stage impeller. For a given specific speed, the lower the value of S, the safer the pump against cavitation. Experience with large European pumped storage installations has shown that cavitation effects began at S ⬇ 6000 (2.2), and this value is recommended for large pumps. Figure 27 shows a summary of data and formulas that may be useful with commercial pumps. German practice differs considerably from that in the United States in computing suction specific speed. Pfleiderer1 defined a hub correction k as k1 a dh 2 b Do (22) where dh  the hub diameter and Do  the diameter of the suction nozzle, in any consistent units. The suction specific speed SG is defined as SG  1n>1002 2Q khsv 3>2 (23) where n is measured in rpm, Q in cubic meters per second per impeller inlet, and hsv in meters of liquid pumped. It follows that S  51642SGk (24) NPSH for Liquids Other Than Cold Water Field experience, together with carefully controlled laboratory experiments, has indicated that pumps handling hot water or cer- 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS FIGURE 28 2.355 Cavitation tests with different liquids at constant speed and constant flow rate (Reference 34) tain liquid hydrocarbons may be operated safely with less NPSH than would normally be required for cold water. This may lower the cost of an installation appreciably, particularly in the case of refinery pumps. A theory for this has been given by Stepanoff and others.14,30–33 Figure 28 shows the results of cavitation tests on two liquids for constant capacity and constant pump speed. No head loss due to cavitation is present at point C or at hsv hsvc. With cold deaerated water, lowering hsv, slightly below hsvc produces limited cavitation and a decrease in pump head H to point C1, but hsvw usually is negligible. With hot water (T 100°F  37.8°C) or with many liquid hydrocarbons, a much larger decrease in hsv will be required to produce the same drop in head H that was shown by the cold water test. The NPSH reduction, or NPSH adjustment, is Hsv⬇ hsvl. In practice, H has been limited to h  0.03H, for which there is a negligible sacrifice in performance. The pumps usually are made of stainless steel or other cavitation-resistant materials, and the lower NPSH results in lower collapse pressure of the vapor bubbles, reducing the damage potential. Chart for NPSH Reductions A composite chart of NPSH reductions for deaerated hot water and certain gas-free liquid hydrocarbons is shown in Figure 29. The curves of vapor pressure versus temperature and the curves of constant NPSH reduction were based on laboratory tests with the liquids shown and should be used subject to the following limitations: 1. No NPSH reduction should exceed 50% of the NPSH required by the pump for cold water or 10 ft (3.0 m), whichever is smaller. 2. NPSH may have to be increased above the normal cold-water value to avoid unsatisfactory operation when (a) entrained air or other noncondensable gas is present in the liquid or (b) dissolved air or other noncondensable gas is present in the liquid and the absolute suction pressure is low enough to permit release of the gas from solution. 3. The vapor pressure of hydrocarbon mixtures vary significantly with temperature and so should be determined at pumping temperature (see Reference 27). 4. If the suction system may be susceptible to transient changes in absolute pressure or temperature, a suitable margin of safety in NPSH should be provided. This is particularly important with hot water and may exceed the reduction that would otherwise apply with steady-state conditions. 5. Although experience has indicated the reliability of Figure 29 for hot water and the liquid hydrocarbons shown, its use with other liquids is not recommended unless it is clearly understood that the results must be accepted on an experimental basis. 2.356 CHAPTER TWO FIGURE 29 NPSH reductions for pumps handling liquid hydrocarbons and hot water. This chart has been constructed from test data obtained by using the liquids shown. For applicability to other liquids, refer to the text (ft ⫻ 0.3048 = m; lb/in2 ⫻ 6.895 = kPa). (Hydraulic Institute ANSI/HI 2000 Edition Pump Standards, Reference 27) USE OF FIGURE 29 Given a fluid having a vapor pressure of 30 lb/in2 (210 kPa) abs at 100°F (37.8°C). Follow the arrows on the key shown on the chart and obtain an NPSH reduction of about 2.3 ft (0.70 m). Since this does not correspond to one of the liquids for which vapor pressure curves are shown on the chart, the use of this NPSH reduction should be considered a tentative value only. Given a pump requiring 16-ft (4.9-m) cold-water NPSH at the operating capacity, the pump is to handle propane at 55°F (12.8°C). Figure 29 shows the vapor pressure to be about 105 lb/in2 (733 kPa) abs and the NPSH reduction to be about 9.5 ft (2.9 m). Since this exceeds 8 ft (2.44 m), which is half the cold-water NPSH, the recommended NPSH for the pump handling propane is half the cold-water NPSH, or 8 ft (2.44 m). If the temperature of the propane in the previous example is reduced to 14°F (⫺10°C), Figure 29 shows the vapor pressure to be 50 lb/in5 (349 kpa) abs and the NPSH reduction to be about 5.7 ft (1.74 m), which is less than half the coldwater NPSH. The NPSH required for pumping propane at 14°F (⫺10°C) is then 16 ⫺ 5.7  10.3 ⬇ 10 ft (4.87  1.74  3.13 ⬇ 3 m). Reduction of Cavitation Damage After the pump has been built and installed*, there is little that can be done to reduce cavitation damage. As previously mentioned, sharpening the leading edges of the blades by filing may be beneficial. Stepanoff12 has suggested cutting back part of the blades in the impeller eye together with sharpening the tips, for low-specific-speed pumps, as a means of reducing the inlet velocity c1 and thus lowering s. Although a small amount of prerotation or prewhirl in the direction of impeller rota*Sometimes it is possible to lower the pump, and this should be considered before other alterations are made. 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.357 tion may be desirable34, excessive amounts should be avoided. This may require straightening vanes ahead of the impeller and rearranging the suction piping to avoid changes in direction or other obstructions. The cavitation damage to the impeller shown in Figure 20 was believed to have been at least partly due to bad flow conditions produced by two 90° elbows in the suction piping. The planes of the elbows were at 90° to each other, and this arrangement should be avoided. Straightening vanes in the impeller inlet may increase the NPSH requirement at all flow rates. Three or four radial ribs equally spaced around the inlet and extending inward about one-quarter of the inlet diameter are effective against excessive prerotation and may require less NPSH than full-length vanes. This is very important with axial-flow pumps, which are apt to have unfavorable cavitation characteristics at partial flow rates. Operation near the best efficiency point usually minimizes cavitation. The admission of a small amount of air into the pump suction tends to reduce cavitation noise.7 This rarely is done, however, because it is difficult to inject the right amount of air under varying head and flow rate conditions and frequently there are objections to mixing air with the liquid pumped. If a new impeller is required because of cavitation, the design should take into account the most recent advances described in the literature. Gongwer35 has suggested (1) the use of ample fillets where the vanes join the shrouds, (2) sharpened leading edges of vanes, (3) reduction of b, in the immediate vicinity of the shrouds, and (4) raking the leading edges of the vanes forward out of the eye. Increasing the number of vanes for propeller pumps lowers s for a given submergence. A change in the impeller material may be very beneficial, as described below. Resistance of Materials to Cavitation Damage Table 8 shows the relative resistance of several metals to cavitation pitting produced by magnetostriction vibration. It will be seen that cast iron, the most commonly used material for impellers, has relatively little pitting resistance relative to bronze and stainless steel, which are readily cast and finished. Damage due to cavitation erosion is commonly assessed in terms of the depth of penetration. The life of an impeller is generally considered to be the time required for cavitation erosion to reach a depth of 75% of the blade thickness at any point36. The life of any material in years can be expressed as the product of the mean depth of penetration rate TABLE 8 Cavitation erosion resistance of metals Alloy Magnetostriction weight loss after 2 h, mg Rolled stellite Welded aluminum bronze Cast aluminum bronze Welded stainless steel (2 layers, 17 Cr-7 Ni) Hot rolled stainless steel (26 Cr-13 Ni) Tempered rolled stainless steel (12 Cr) Cast stainless steel (18 Cr-8 Ni) Cast stainless steel (12 Cr) Cast manganese bronze Welded mild steel Plate steel Cast steel Aluminum Brass Cast iron a 0.6 3.2 5.8 6.0 8.0 9.0 13.0 20.0 80.0 97.0 98.0 105.0 124.0 156.0 224.0 a Despite the high resistance of this material to cavitation damage, it is not suitable for ordinary use because of its comparatively high cost and the difficulty encountered in machining and grinding. Source: Reference 69. 2.358 CHAPTER TWO (MDPR, in mm per year) times the thickness of the material in mm. Values of MDPR have been deduced37 from the weight loss of test samples of known diameter in a magnetostriction test (Table 8)38. Unfortunately, MDPR-values obtained in such tests are not the same as those of actual pumps, as the mode of cavitation varies with pump operating conditions and is generally different from the laboratory results. Nevertheless, laboratory results for MDPR have been used to rank the ability of various materials to resist cavitation erosion in pumps. See Section 5.1 for material selection guidelines. Cavitation-resistant coatings, either metallic or nonmetallic, have found some niche applications. Elastomeric coatings are resilient and resist cavitation through a different erosion mechanism than that of metal. As such, they can be very effective69. At least two considerations are involved in the use of coatings for resisting cavitation damage: 1) even if a contemplated coating demonstrates a reduced damage or erosion rate, this reduction must be enough to justify the cost of establishing a satisfactory bond between the coating and the base metal; 2) erosion of both the coating and the base material must be considered in determining the life according to the above 75%-depth criterion. Therefore, the life in this case would be equal to the MDPR of the coating times the coating thickness plus the MDPR of the base material times the allowable erosion depth of that material. Inducers It is sometimes difficult or impossible to provide the required NPSH for an otherwise acceptable pump. Besides normal industrial situations that might produce a very low available NPSH, the need to keep the weight down in aircraft and rocket liquidpropellant pumps has led to high rotative speeds, which, for typical values of NPSH, produce extremely high suction specific speeds. The performance-NPSH required by the impeller under these circumstances can be provided by a small, axial-flow booster pump, called an inducer, placed ahead of the first-stage impeller39. Inducers are designed to operate with very low NPSH and to provide enough head to satisfy the NPSH required by the impeller. In fact, long stable cavities are established on the suction sides of the long, lightly-loaded blades of an inducer, which enable it to operate at about twice the suction specific speed of a conventional impeller40,41. At lower than normal flow rates, however, inducers readily produce swirling, destabilizing backflow at the inlet, which can cause excessive pump vibration in high-head pumps. These instabilities can be overcome by various passive design features, such as that described in Reference 39. The inducers described in Reference 42 (Figure 30) have “constant lead” helical blades. They contribute not more than 5% of the total pump head. Although the efficiency of the inducer alone is low, the reduction in overall pump efficiency is not significant. Because this type of inducer causes prerotation, a careful match between inducer and suction impeller is required. In vertical multistage pumps, where a long shaft can be better supported, a vaned diffuser may be inserted between the inducer and the first-stage impeller. Such an arrangement is very beneficial for operation at reduced flow rate. Reference 42 shows that a suitable inducer-impeller combination can operate at about 50% of the NPSH required for the impeller alone at flow rates not exceeding the normal value. The NPSH requirement increases rapidly for flow rates above normal. Unless a variable-lead inducer is used32, 39, operation in this range should be avoided. Entrained Air Air or other gases may enter the impeller inlet from several sources. The immediate effect usually will be a drop in pump pressure rise, flow rate, and power. This will be followed by loss of prime if more gas is present than the impeller can handle. A typical limit for commercial industrial pumps is an inlet gas-to-liquid volume fraction (GVF) of 0.03, although specialty pumps such as those used in aircraft (Section 9.19) can handle higher GVF. See also Reference 9, Section 2.1. Air may be released from solution or enter through leaks in the suction piping. Stuffing box air leakage may be prevented by lantern rings supplied with liquid from the pump discharge. If the pump takes water from a sump with a free surface, a vortex may form from the free surface to the impeller inlet. The remedy may be the introduction of one or more baffle plates or even major changes in the sump. For information on proper sump design and the prevention of airentraining vortices, see Sections 10.1 and 10.2, pp. 457 and 460 of Reference 7, and Reference 43. It is sometimes permissible to inject a small amount of air into the pump 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS FIGURE 30 2.359 Pump fitted with inducer (Reference 42) suction to reduce the noise and damage from cavitation caused by inadequate NPSH or recirculation in the impeller (see Subsection 2.3.2). STARTING CENTRIFUGAL PUMPS ______________________________________ Priming Centrifugal pumps usually are completely filled with the liquid to be pumped before starting. When so filled with liquid, the pump is said to be primed. Pumps have been developed to start with air in the casing and then be primed.44 This procedure is unusual with low-specific-speed pumps but is sometimes done with propeller pumps12. In many installations, the pump is at a lower elevation than the supply and remains primed at all times. This is customary for pumps of high specific speed and all pumps requiring a positive suction head to avoid cavitation. Pumps operated with a suction lift may be primed in any of several ways. A relatively inexpensive method is to install a special type of check valve, called a foot valve, on the inlet end of the suction pipe and prime the pump by filling the system with liquid from any available source. Foot valves cause undesirable frictional loss and may leak enough to require priming before each starting of the pump. A better method is to close a valve in the discharge line and prime by evacuating air from the highest point of the pump casing. Many types of vacuum pumps are available for this service. A priming chamber is a tank that holds enough liquid to keep the pump submerged until pumping action can be initiated. Self-priming pumps usually incorporate some form of priming chamber in the pump casing. Section 2.4 and Reference 7 may be consulted for further details. Torque Characteristics of Drivers Centrifugal pumps of all specific speeds usually have such low starting torques (turning moments) that an analysis of the starting phase of operation seldom is required. Steam and gas turbines have high starting torques, so no special starting procedures are necessary when they are used to drive pumps. If a pump is directly connected to an internal combustion engine, the starting motor of the engine should be made adequate to start both driver and pump. If the starter does not have enough torque to handle both units, a clutch must be provided to uncouple the pump until the driver is started. 2.360 CHAPTER TWO Electric motors are the most commonly used drivers for centrifugal pumps. Direct current motors and alternating current induction motors usually have ample starting torque for all pump installations, provided the power supply is adequate. Many types of reduced voltage starters are available7 to limit the inrush current to safe values for a given power supply. Synchronous motors are often used with large pumps because of their favorable power-factor properties. They are started as induction motors and run as such up to about 95% of synchronous speed. At this point, dc field excitation is applied and the maximum torque the motor can then develop is called the poll-in torque, which must be enough to accelerate the motor and connected inertia load to synchronous speed in about 0.2 s if synchronous operation is to be achieved. Centrifugal pumps usually require maximum torque at the normal operating point, and this should be considered in selecting a driver, particularly a synchronous motor, to be sure that the available pull-in torque will bring the unit to synchronous speed. Torque Requirements of Pumps The torque, or turning moment, for a pump may be estimated from the power curve in USCS units by M 5252P n M 9549P n (25) and in SI units by where M  pump torque, lb ⴢ ft (N ⴢ m) P  power, hp (kW) n  speed, rpm Equation 25 makes no allowance for accelerating the rotating elements or the liquid in the pump. If a 10% allowance for accelerating torque is included, the constant should be correspondingly increased. The time t required to change the pump speed by an amount n  n2  n1 is given by ¢t  I¢n k1Mm  M2 (26) where t  time, s I  moment of inertia (flywheel effect) of all rotating elements of driver, pump, and liquid, lb ⴢ ft (kg ⴢ m2) n  change in speed, rpm k  307 in USCS (9.549 in SI) Mm  driver torque, lb ⴢ ft (N ⴢ m) M  pump torque, lb ⴢ ft (N ⴢ m) (Eq. 25) The inertia I of the driver and pump usually can be obtained from the manufacturers of the equipment. The largest permissible n for accurate calculation will depend on how rapidly Mm and M vary with speed. The quantity Mm  M should be nearly constant over the interval n if an accurate estimate of t is to be obtained. Torque-speed characteristics of electric motors may be obtained from the manufacturers. Horizontal-shaft pumps fitted with plain bearings and packed glands require a breakaway torque of about 15% of Mn, the torque at the normal operating point, to overcome the static friction. This may be reduced to about 10% of Mn if the pump is fitted with antifriction bearings. The breakaway torque may be assumed to decrease linearly with speed to nearly zero when the speed reaches 15 to 20% of normal. Construction of torque-speed curves requires a knowledge of the pump characteristics at normal speed as well as details of the entire pumping system. Some typical examples taken from Reference 7 are given 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.361 below. The following forms of the affinity laws (Eqs. 12) are useful in constructing the various performance curves: Q2  Q1 n2 n1 (27) H2  H1 a n2 2 Q2 2 b  H1 a b n1 Q1 (28) P2  P1 a n2 3 b n1 (29) M2  M1 a n2 2 b n1 (30) where Q  flow rate n  speed H  head P  power M  torque in any consistent units of measure. After speeds n1 and n2 are chosen, the subscripts 1 and 2 refer to corresponding points on the characteristic curves for these speeds. Low-Specific-Speed Pumps Figure 31 shows the constant-speed characteristics of a pump having ns ⬇ 1740 (0.64) at best efficiency. This pump usually would be started with a valve in the discharge line closed. During the starting phase, the pump operates at shutoff with P1  25.8 hp (19.2 kW) and n1  1770 rpm. Then, by Eq. 25, M1  76.6 lb ⴢ ft (103.9 N ⴢ m). These values may be used in Eq. 30 to evaluate starting torques M2 at as many speeds n2 as desired and plotted in Figure 32 as curve BCD. Section AB of the starting torque curve is an estimate of the breakaway torque. If the discharge valve is now opened, the speed remains nearly constant but the torque increases as the capacity and power increase. If the normal operating point is Q  1400 gpm (88.3 l/s) and Pn  53.2 hp (39.7 kW), the motor torque will be Mn  158 lb ⴢ ft (214 N ⴢ m) by Eq. 25. The vertical line DE in Figure 32 shows the change in torque produced by opening the discharge valve. Instead of starting the pump with the discharge valve closed, let the pump be started with a check valve in the discharge line held closed by a static head of 100 ft (30.5 m). The frictional head in the system may be represented by kQ2. The value of k may be estimated from the geometry of the system or from a frictional-loss measurement at any convenient flowrate Q, preferably near the normal capacity Qn. In this example, Qn  1400 gpm (88.33 l/s) and k  14.4/106 (0.00109). The curve labeled system head 1 in Figure 31 was computed from H  100  (14.4/106)Q2 ft (H  30.48  0.00109Q2 m) and intersects the head curve at H  128 ft (39.01 m) and Qn  1400 gpm (88.33 l/s). The normal shutoff head is H1  153 ft (46.6 m) at n1  1770 rpm. By Eq. 28, the pump will develop a shutoff head H2  100 ft (30.5 m) at n2  1430 rpm. By Eq. 30, the torque at 1430 rpm will be 50 lb ⴢ ft (68 N ⴢ m), corresponding to point C in Figure 32. The portion ABC of the starting torque curve has already been constructed. Trial-and-error methods must be used to obtain the portion ABC of the starting torque curve. The auxiliary curves in Figure 33 are useful in constructing the CE portion of the starting torque curve. Select a value of n2 intermediate between 1427 and 1770 rpm, say n2  1600 rpm. In Figure 31, read values of Q1, H1, and P1 for speed n1  1770 rpm. By Eqs. 27 and 28, determine values of Q2 and H2 and plot as shown in Figure 31 until an intersection with the system-head 1 curve is obtained that provides Q1600 corresponding to n  1600 rpm. By Eq. 29, determine value of P2 and plot as shown in Figure 31 until an intersection is obtained with the Q1600 line which provides P1600 corresponding to n  1600 rpm. Eq. 34 is now used to obtain M1600, which is one point on the desired starting torque curve. The process is repeated for various speeds n2 until the curve CE in Figure 32 can be drawn. The 2.362 CHAPTER TWO FLOW RATE Q, GPM FIGURE 31 Characteristics of a 6-by-8 double-section pump at 1770 rpm (ft ⫻ 0.3048 = m; gpm ⫻ 0.06309 = l/s; hp ⫻ 0.7457 = kW) ns = 1,740 (s = 0.64) (Reference 7) FIGURE 32 Torque characteristics of 6-by-8 pump shown in Figure 31 (lb • ft ⫻ 1.356 = N • m) (Reference 7) complete starting torque curve for this example is ABCE in Figure 31, with steady-state operation at point E. Assume that the pump of the preceding examples is installed in a system having zero static head but a long pipeline with friction head given by H  (65.4/106)Q2 ft (H  0.00500Q2 m), as shown in Figure 31 by the curve labeled system head 2. The valve in the discharge line is assumed to open instantaneously when power is first applied to the pump. The procedure described to construct curve CE of the preceding example must now be used together with the system-head 2 curve of Figure 31 to obtain the curve BE of Figure 32. The complete starting torque curve for this example is ABE in Figure 32, with steady-state operation at point E. The inertia of the fluid in the system was neglected in solving the previous examples. Some of the power must be used to accelerate the liquid, and this may be appreciable in the case of a long pipeline. Low-specific speed pumps, which are used with long pipelines, have 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.363 FLOW RATE Q, GPM FIGURE 33 Analysis of 6-by-8 pump shown in Figure 41 (ft ⫻ 0.3048 = m; gpm ⫻ 0.06309 = l/s; lb • ft ⫻ 1.356 = N • m; hp ⫻ 0.7457 = kW) (Reference 7) FLOW RATE Q, GPM FIGURE 34 Characteristics of a 16-in (40.6cm) volute pump with mixed-flow impeller with flat power characteristic (ft ⫻ 0.3048 = m; gpm ⫻ 0.06309 = l/s; hp ⫻ 0.7457 = kW). ns = 4570 (s = 1.672) (Reference 7) rising power curves with minimum power at shutoff and maximum power at normal flow rate. Experience has shown that the starting torque-speed curves computed by neglecting the inertia of the liquid are conservative, so inertia effects need not be included. The inertia effect of the liquid does slow the starting operation. If the time required to reach any event, such as a particular speed or flow rate, is required, the inertia of the liquid should be considered, but including it greatly increases the difficulty of computation. References 45 through 48 give general methods for handling problems involving liquid transients. High-specific-speed pumps have falling power flow rate curves with maximum power at shutoff and minimum power at normal flow rate. Neglecting the inertia of the liquid probably will result in too low a value for the computed starting torque for such pumps. If liquid inertia is to be included, consult References 44 to 47 and Section 8.1. Figure 34 shows constant-speed characteristics for a medium-specific-speed pump, ns  4570 (1.672) at best efficiency. The shutoff power MEDIUM- AND HIGH-SPECIFIC-SPEED PUMPS 2.364 CHAPTER TWO FIGURE 35 Torque characteristics of pump shown in Figure 34 (Reference 7) FLOW RATE Q, GPM FIGURE 36 Characteristics of a 30-in (76.2-cm) discharge propeller pump at 700 rpm (ft ⫻ 0.1048 = m; gpm ⫻ 0.06309 = l/s; hp ⫻ 0.7457= kW). ns = 12,000 (s = 4.391) (Reference 7) is the same as the power at best efficiency, and the starting torque-speed curve is but little affected by the method of starting, as shown by Figure 35. Figure 36 shows the constant-speed characteristics of a high-specific-speed propeller pump, ns ⬇ 12,000 (4.391) at best efficiency. Figure 37 shows the starting torque-speed curve when the pump is started against a static head of 14 ft (4.3 m) and a friction head of 1 ft (0.3 m) at Qn  12,500 gpm (789 l/s). The system was assumed full of water with a closed check valve at the outlet end of the short discharge pipe. The methods of computation for Figures 35 and 37 were the same as for Figure 32. Sometimes propeller pumps are started with the pump submerged but with the discharge column filled with air. In such a case, the torque-time characteristic for the driver must be known and a step-by-step calculation carried out. If the discharge column is a siphon initially filled with air, the starting torque may exceed the normal running torque during some short period of the starting operation. If the pump is driven by a synchronous 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS FIGURE 37 2.365 Torque characteristics of pump shown in Figure 36 (Reference 7) motor, it is particularly important to investigate the starting torque in the range of 90 to 100% of normal speed to make sure that the pull-in torque of the motor is not exceeded. For additional information regarding starting high-specific-speed pumps discharging through long and large diameter systems, see Section 8.1. Miscellaneous Requirements Pumps handling hot liquids should be warmed up to operating temperature before being started unless they have been especially designed for quick starting. Failure to do this may cause serious damage to wearing rings, seals, and any hydraulic balancing device that may be present. A careful check of the installation should be made before starting new pumps, pumps that have had a major overhaul, or pumps that have been standing idle for a long time. It is very important to follow the manufacturer’s instructions when starting boiler-feed pumps. If these are unavailable, Reference 7 may be consulted. Ascertain that the shaft is not frozen, that the direction of rotation is correct, preferably with the coupling disengaged, and that bearing lubrication and gland cooling water meet normal requirements. Failure to do this may result in damage to the pump or driver. A pump may run backwards at runaway speed if the discharge valve fails to close following shutdown. Any attempt to start the pump from this condition will put a prolonged overload on the motor. Figure 38 shows one example of the torque-speed transient for a pump, ns  1700 (0.622), started from a runaway reversed speed while normal pump head was maintained between the section and discharge flanges. In most practical cases, water hammer effects would make this transient even more unfavorable than Figure 38 indicates. The duration of such a transient will always be much longer than the normal starting time, and so protective devices would probably disconnect the motor from the power supply before normal operation could be achieved. Consult Section 8.1 for additional information on this subject. REGULATION OF FLOW RATE__________________________________________ Flow rate variation ordinarily is accomplished by a change in pump head, speed, or both simultaneously. The flow rate and power input of pumps with specific speeds up to about 4000 (1.464) double suction increase with decreasing head, so the drivers of such pumps may be overloaded if the head falls below a safe minimum value. Increasing the head of high-specific-speed pumps decreases the flow rate but increases the power input. The drivers of these pumps should either be able to meet possible load increases or be equipped with suitable overload protection. Flow rate regulation by the various methods given below may be manual or automatic (see also References 1, 7, 12, 34 and 49). 2.366 CHAPTER TWO FIGURE 38 Torque characteristics of a double-suction pump, ns L 1700 (0.64), from reversed runaway speed to normal forward speed (Reference 7) FLOW RATE Q, GPM FIGURE 39 Power requirements of two double-suction pumps in series operated at constant head and variable flow rate. Total Hn = 382 ft (116 m) for both pumps at 1800 rpm (gpm ⫻ 0.06309 = l/s hp ⫻ 0.7457 = kW) (Reference 50). Curve AA: constant speed with discharge throttling. Curve BB: synchronous motor with variable-speed hydraulic coupling on each pump. Curve CC: variable-speed wound-rotor induction motor. Curve DD: dc motor with rectifier and shunt field control. Curve EE: synchronous motor with variable-speed constant-efficiency mechanical speed reducer Discharge Throttling This is the cheapest and most common method of flow rate modulation for low- and medium- specific-speed pumps. Usually its use is restricted to such pumps. Partial closure of any type of valve in the discharge line will increase the system head so the system-head curve will intersect the pump head curve at a smaller flow rate, as shown in Figure 40. Discharge throttling moves the operating point to one of lower efficiency, and power is lost at the throttle valve. This may be important in large installations, where more costly methods of modulation may be economically attractive. Throttling to the point of cutoff may cause excessive heating of the liquid in the pump. This may require a bypass to maintain the necessary minimum flow or use of different method of modulation. This is particularly important with pumps handling hot water or volatile liquids, as previously mentioned. Refer to Section 8.2 for information regarding the sizing of a pump bypass. Suction Throttling If sufficient NPSH is available, some power can be saved by throttling in the suction line. Jet engine fuel pumps frequently are suction throttled5 because discharge throttling may cause overheating and vaporization of the liquid. At very low flow rate, the impellers of these pumps are only partly filled with liquid, so the power input and temperature rise are about one-third the values for impellers running full with discharge throttling. The capacity of condensate pumps frequently is submergence-controlled,7 which 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.367 is equivalent to suction throttling. Special design reduces cavitation damage of these pumps to a negligible amount, the energy level (Section 2.1) being quite low. Bypass Regulation All or part of the pump flow may be diverted from the discharge line to the pump suction or other suitable point through a bypass line. The bypass may contain one or more metering orifices and suitable control valves. Metered bypasses are commonly used with boiler-feed pumps for reduced-flow operation, mainly to prevent overheating. There is a considerable power saving if excess capacity of propeller pumps is bypassed instead of using discharge throttling. Speed Regulation This can be used to minimize power requirements and eliminate overheating during flow rate modulation. Steam turbines and internal combustion engines are readily adaptable to speed regulation at small extra cost. A wide variety of variable-speed mechanical, magnetic, and hydraulic drives are available, as well as both ac and dc variable-speed motors. Usually variable-speed motors are so expensive that they can be justified only by an economic study of a particular case. Figure 39 shows a study by Richardson50 of power requirements with various drivers wherein substantial economies in power may be obtained from variable-speed drives. Regulation by Adjustable Vanes Adjustable guide vanes ahead of the impeller have been investigated and found effective with a pump of specific speed ns  5700 (2.086). The vanes produced a positive prewhirl that reduced the head, flow rate, and efficiency. Relatively little regulation was obtained from the vanes with pumps having ns  3920 (1.204) and 1060 (0.39). Adjustable outlet diffusion vanes have been used with good success on several large European storage pumps for hydroelectric developments. Propeller pumps with adjustable-pitch blades have been investigated with good success. Wide flow rate variation was obtained at constant head and with relatively little loss in efficiency. These methods are so complicated and expensive that they have very limited application in practice. Reference 34 may be consulted for further discussion and bibliography. Air Admission Admitting air into the pump suction has been demonstrated as a means of flow-rate regulation, with some savings in power over discharge throttling. Usually air in the pumped liquid is undesirable, and there is always the danger that too much air will cause the pump to lose its prime. The method has rarely been used in practice but might be applicable to isolated cases. PARALLEL AND SERIES OPERATION ___________________________________ Two or more pumps may be arranged for parallel or series operation to meet a wide range of requirements in the most economical manner. If the pumps are close together, that is, in the same station, the analysis given below should be adequate to secure satisfactory operation. If the pumps are widely separated, as in the case of two or more pumps at widely spaced intervals along a pipeline, serious pressure transients may be generated by improper starting or stopping procedures. The analysis of such cases may be quite complicated, and References 46 to 48 should be consulted for methods of solution. Parallel Operation Parallel operation of two or more pumps is a common method of meeting variable-flow-rate requirements. By starting only those pumps needed to meet the demand, operation near maximum efficiency can usually be obtained. The head-flow characteristics of the pumps need not be identical, but pumps with unstable characteristics may give trouble unless operation only on the steep portion of the characteristic can be assured. Care should be taken to see that no one pump, when combined with pumps of different characteristics, is forced to operate at flows less than the minimum required to prevent recirculation. See the discussion that follows on operation at other than normal flow rate. Multiple pumps in a station provide spares for emergency service and for the downtime needed for maintenance and repair. 2.368 CHAPTER TWO FLOW RATE Q, GPM FIGURE 40 Head-flow curves of pumps operating in parallel The possibility of driving two pumps from a single motor should always be considered, as it usually is possible to drive the smaller pumps at about 40% higher speed than a single pump of twice the capacity. The saving in cost of the higher-speed motor may largely offset the increased cost of two pumps and give additional flexibility of operation. One of the first steps in planning for multiple-pump operation is to draw the systemhead curve, as shown in Figure 40. The system head consists of the static head Hs and the sum Hf of the pipe-friction head and the head lost in the valves and fittings (see Sections 8.1 and 8.2). The head curves of the various pumps are plotted on the same diagram, and their intersections with the system-head curve show possible operating points. Combined pump head curves are drawn by adding the flow rates of the various combinations of pumps for as many values of the head as necessary. The intersection of any combined HQ curve with the system-head curve is an operating point. Figure 40 shows two pump head curves and the combined curve. Points 1, 2, and 3 are possible operating conditions. Additional operating points may be obtained by changing the speed of the pumps or by increasing the system-head loss by throttling. Any number of pumps in parallel may be included on a single diagram, although separate diagrams for different combinations of pumps may be preferable. The overall efficiency h of pumps in parallel is given by h where H sp. gr. k 兺Q 兺P H1sp. gr.2 k ⫻ 兺Q 兺P (31)  head, ft (m)  specific gravity of the liquid  3960 USCS (0.1021 SI)  sum of the pump flow rates, gpm (l/s)  total power supplied to all pumps, hp (W) Series Operation Pumps are frequently operated in series to supply heads greater than those of the individual pumps. The planning procedure is similar to the case of pumps in parallel. The system-head curve and the individual head-flow curves for the pumps are plotted as shown in Figure 41. The pump heads are added as shown to obtain the combined pump head curve. In this example, Pump 2 operating alone will deliver no liquid because its shutoff head is less than the system static head. There are two possible operating points, 1 and 2, as shown by the appropriate intersections with the system-head curve. As with parallel operation, other operating points 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.369 FLOW RATE, Q FIGURE 41 Head-flow curves of pumps operating in series could be obtained by throttling or by changing the pump speeds. The overall efficiency of pumps in series is given by h Q1sp. gr.2 k ⫻ 兺H 兺P (32) wherein the symbols are the same as for parallel operation. It is important to note that the stuffing box pressure of the second pump is increased by the discharge pressure of the first pump. This may require a special packing box for the second pump with leakoff to the suction of the first pump. The higher suction pressure may increase both the first cost and the maintenance costs of the second pump. OPERATION AT OTHER THAN THE NORMAL FLOW RATE___________________ Centrifugal pumps usually are designed to operate near the point of best efficiency, but many applications require operation over a wide range of flow rates, including shutoff, for extended periods of time. Pumps for such service are available but may require special design and construction at higher cost. Noise, vibration, and cavitation may be encountered at low flow rates. Large radial shaft forces at shutoff as well as lack of through flow to provide cooling may cause damage or breakage to such parts as shafts, bearings, seals, glands, and wearing rings of pumps not intended for such service. Some of the phenomena associated with operation at other than normal flow rate are described below. Recirculation There is a small flow from impeller discharge to suction through the wearing rings and any hydraulic balancing device present. This takes place at all flow rates, but does not usually contribute to raising the liquid temperature very much unless operation is near shutoff. When the flow rate has been reduced by throttling (or as a result of an increase in system head), a secondary flow called recirculation begins. Recirculation is a flow reversal due to separation at the suction and at the discharge tips of the impeller vanes. All impellers 2.370 FIGURE 42 CHAPTER TWO Suction recirculation FIGURE 43 Discharge recirculation have a critical flow rate at which recirculation occurs. The flow rates at which suction and discharge recirculation begin can be controlled to some extent by design, but recirculation cannot be eliminated (see Figure 6 in Section 2.1). Suction recirculation is the reversal of flow at the impeller eye. A portion of the flow is directed out of the eye at the eye diameter, as shown in Figure 42 , and travels upstream with a rotational velocity approaching the peripheral velocity of the diameter. A rotating annulus of liquid is produced upstream from the impeller inlet, and through the core of this annulus passes an axial flow corresponding to the output flow rate of the pump. In pumps equipped with long, straight suction nozzles but no suction elbow, this rotating fluid has been detected over considerable distances upstream from the impeller eye. Suction pressures measured at wall taps where this phenomenon is present are always higher than the true average static pressure across the measuring section. This means that the pump head as determined from wall taps is less than it would be if true average static pressures were measured. The high shear rate between the rotating annulus and the axial flow through the core produces vortices that form and collapse, producing noise and cavitation in the suction of the pump. Discharge recirculation is the reversal of flow at the discharge tips of the impeller blades, as shown in Figure 43. The high shear rate between the inward and outward relative velocities produces vortices that cavitate and can attack the pressure side of the blades. This phenomenon, which tends to occur at a lower flow rate than the highest Q for suction recirculation, also involves stalled flow from the diffuser vanes or volute tongue(s). Separated reversed flow recirculates and emerges from these vane systems back into the impeller with negative swirl (that is, swirl opposite to the direction of rotation). The impeller must expend significant power to redirect the portion of this fluid (that reenters it) out again—with positive swirl. As discussed in Section 2.1, the portion of this backflow from the diffuser or volute that enters the spaces outside the impeller shrouds and adjacent to the casing walls has the potential to reverse the axial thrust of the impeller, and this reversal can fluctuate if the backflow is unsteady (as separated, recirculating flow normally is) and not always feeding the same side of the impeller. The flow rate Qsr below which suction recirculation occurs is directly related to the design suction-specific speed S of the pump. The higher the suction-specific speed, the closer will be the beginning of recirculation to the flow rate at best efficiency. Figure 44 shows the relation between the suction specific speed and suction recirculation for pumps 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.371 FIGURE 44 Influence of the design value of suction specific speed S on the flow rate QSR below which suction recirculation occurs. 500 6 ns 6 2500; (0.18 6 s 6 0.91) for single-suction or one side of a double-suction impeller. The ordinate is QSR/QBEP in percent. (To obtain ss, divide S by 2733.) FIGURE 45 Influence of the design value of suction specific speed S on the flow rate QSR below which suction recirculation occurs. 2500 6 ns 6 10,000; (0.91 6 s 6 3.66) for single-suction or one side of a double-suction impeller. The ordinate is QSR/QBEP in percent. (To obtain ss, divide S by 2733.) up to 2500 (0.915) specific speed, and Figure 45 shows the same relation for pumps up to 10,000 (3.659) specific speed. Despite the existence of suction and discharge recirculation, the mechanical response of the pump will not be serious unless the energy level is high. In other words, most pumps can indeed be operated at Q 6 QSR. The minimum flow rate or simply “minimum flow” Qmin is quantified in Section 2.1. As energy level is increased, Qmin approaches QSR in the limit. Examples of the difference between Qmin and QSR are as follows: For water pumps rated at 2500 gpm (158 l/s) and 150 ft (45.7 m) total head or less, the minimum operating flows can 2.372 CHAPTER TWO be as low as 50% of the suction recirculation values shown for continuous operation and as low as 25% for intermittent operation. For hydrocarbons, the minimum operating flows can be as low as 60% of the suction recirculation values shown for continuous operation and as low as 25% for intermittent operation.51, 52, 53 Temperature Rise Under steady-state conditions, friction and the work of compression increase the temperature of the liquid as it flows from suction to discharge. A further temperature increase may arise from liquid returned to the pump suction through wearing rings, a balancing device, or a minimum-flow bypass line that protects the pump when operating at or near shutoff. Assuming that all heat generated remains in the liquid, the temperature rise is ¢T  gH ⫻ 11 ⫺ h2 goCphJ ⫹ ¢Tc (33) where g/go  1 lbf/lbm; but when using SI units, g/go is replaced by 9.80665 m/s2 ( g in the SI system). Tc is due to the compression of the liquid and is not a consequence of loss or dissipation as is the term involving the pump efficiency h (see Section 2.1). As shown in Reference 1 of Section 2.1, Tc is 3°F per 1000 psi (0.24°C per MPa) of pump pressure rise for hydrocarbon fuels. For boiler feedwater at 350°F (177°C), Tc  1.6°F per 1000 psi (0.129°C per MPa), but it is much smaller for cold water. By consulting tables of properties for the liquid phase of the fluid being pumped and assuming the compression process between the actual inlet and discharge pressures to be isentropic, Tc can be determined. This is important if Eq. 33 is used to evaluate overall pump efficiency from temperature rise measurements. T and Tc are often of the same order of magnitude at BEP, and serious errors have been made by excluding Tc from the efficiency computation. At very low, off-BEP flow rates, T will be high in comparison to Tc; so, the latter can be safely ignored in temperature rise calculations at such low-efficiency conditions. In practice, determination of efficiency from T-measurements is accomplished by the direct thermodynamic method54, rather than by the Tc-method. Both approaches are based on the definition of pump efficiency as the ratio of an isentropic rise of total enthalpy ( gH) to the actual rise of total enthalpy (Eq. 1 of Section 2.1), allowances being made for the usually small external power losses that do not appear in the pumped fluid (such as bearing drag) and the similarly small effects of heat transferred between pump and surroundings. In the direct thermodynamic method, the enthalpy rise h is found from the chain rule, ¢h  兰 dh  兰 3 1 0h>0p2 Tdp  1 0h> 0T2 pdT4  a¢p  CpJ¢T the coefficients a and CpJ being average values of the two partial derivatives as found from tables of thermodynamic properties of the fluid. Values of these partial derivatives are conveniently tabulated for water in Reference 54. General service pumps handling cold liquids may be able to stand a temperature rise as great as 100°F (56°C). Most modern boiler-feed pumps may safely withstand a temperature rise of 50°F (28°C). The NPSH required to avoid cavitation or to prevent flashing of hot liquid returned to the pump suction may be the controlling factor. Minimum flow may be dictated by other factors, such as recirculation and unbalanced radial and axial forces on the impeller. Axial forces can be the controlling factor with single-stage double-suction pumps. It is especially important to protect even small pumps handling hot liquids from operation at shutoff. This is usually done by providing a bypass line fitted with a metering orifice to maintain the minimum required flow through the pump. In the case of boiler-feed pumps, the bypass flow usually is returned to one of the feed-water the water heaters. Unless especially designed for cold starting, pumps handling hot liquids should be warmed up gradually before being put into operation. Radial Thrust Ideally, the circumferential pressure distribution at the impeller exit is uniform at the design condition (as explained in Section 2.1); however, it becomes non- 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.373 uniform at off-BEP flow rates. An exception is that concentric collecting configurations will produce non-uniform pressure distributions at the BEP. Any non-uniformity leads to a radial force on the pump shaft called the radial thrust or radial reaction. The radial thrust Fr in pounds (newtons) is Fr  kKr 1sp. gr.2HD2b2 (34) where k  0.433 USGS (9790 SI) Kr  experimentally determined coefficient sp. gr.  specific gravity of the liquid pumped (equal to unity for cold water) H  pump head, ft (m) D2  outside diameter of impeller, in (m) b2  breadth of impeller at discharge, including shrouds, in (m) Values of K, determined by Agostinelli et al.55 for single-volute pumps are given in Figure 46 as functions of specific speed and flow rate. The magnitude and direction of Fr on the pump shaft may be estimated from Figure 47 , but Eq. 34 probably will be more accurate for determining the magnitude of the force. The radial thrust usually is minimum near Q  Qn, the flow rate at best efficiency, but rarely goes completely to zero. Near shutoff, Fr usually is maximum and may be a considerable force on the shaft in high-head pumps. The radial thrust can be made much smaller throughout the entire flow-rate range by using a double volute (twin volute) or a concentric casing. These designs should be considered, particularly if the pump must operate at small flow rates. Figures 48 to 50 compare radial forces generated by three types of casings: a standard volute, a double volute, and a modified concentric casing. The latter casing was concentric with the impeller for 270° from the tongue and then enlarged in the manner of a single volute to form the discharge nozzle. The magnitude and direction of Fr on the pump shaft for the modified concentric casing may be estimated from Figure 51. The direction of Fr on the pump shaft with a double volute was somewhat random but in the general vicinity of the casing tongue. Radial forces on pumps fitted with diffuser vanes usually are rather small, although they may be significant near shutoff due to stall in some of the passages and not in others. EXAMPLE Consider a single-stage centrifugal pump, ns  2000 (0.732) at best efficiency, handling cold water, sp. gr.  1.0. Estimate the radial thrust on the impeller at half the normal flow rate when fitted with (a) a single volute, (b) a modified concentric casing, and (c) a double volute. Impeller dimensions are D2  15.125 in (38.4 cm) and b2  2.5 in (6.35 cm). The shutoff head is H  252 ft (76.8 m), and the head at half capacity is H  244 ft (74.4 m). FIGURE 46 Kr as a function of specific speed and flow rate for single-volute pumps (to obtain s, divide by 2733.) (Reference 55) 2.374 CHAPTER TWO FIGURE 47 Polar plot showing direction of resultant radial forces for single-volute pumps at various flow rates (“capacities”) and specific speeds. To obtain s, divide by 2733. (Reference 55) FLOW RATE, % OF NORMAL FIGURE 48 Comparison of the effect of three casing designs on radial forces for ns = 1165 (0.426) (Reference 55) FIGURE 49 Comparison of the effect of three casing designs on radial forces for ns = 2120 (0.776) (Reference 55) FLOW RATE, % OF NORMAL 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.375 FLOW RATE, % OF NORMAL FIGURE 50 Comparison of effect of three casing designs on radial forces for ns = 3500 (1.281) (Reference 55) FIGURE 51 Polar plot showing direction of resultant radial forces for modified concentric casings at various flow rates and specific speeds; namely, 1,165 (0.426), 2,120 (0.776), and 3500 (1.281). The casings were concentric for 270º from the tongue. (Reference 55) Solution (a) Kr  0.2 from Figure 46, by Eq. 34. Fr  10.433210.22 11.02 12442 115.125212.52  799 lb 13554 N2 Estimating between the curves for nS,  2370 (0.867) and 1735 (0.635) in Figure 47, the direction of Fr on the shaft should be about 65° to 70° from the casing tongue in the direction of rotation. (b) Use Figure 49, ns  2120 (0.776), which is nearest to ns  2000 (0.730), to find the radial force for a modified concentric casing at half flow, which is about 33% 2.376 CHAPTER TWO of shutoff value for a single-volute casing. From Figure 46, for a single-volute casing at shutoff, Kr  0.34 and Fr  10.433210.34211.02 12522 115.1252 12.52  1403 lb 16241 N2 Then, for a modified concentric casing at half flow, Fr  10.332114032  463 lb 12059 N2 From Figure 51, the direction of Fr should be about 75° to 80° from the casing tongue in the direction of rotation. (c) From Figure 49, the radial force for a double-volute casing is about 8% of the shutoff value for the single-volute casing, and so Fr,  10.082114032  112 lb 1499 N2 According to Agostinelli et al.,55 the direction of the radial thrust in double-volute casings was found to be generally toward the casing tongue. Stepanoff12 has found this direction to follow approximately that in single-volute casings (see also Biheller56). Axial Thrust See Section 2.1 and 2.2.1. ABNORMAL OPERATION ______________________________________________ Complete Pump Characteristics Many types of abnormal operation involve reversed pump rotation, reversed flow direction, or both, and special tests are required to cover these modes of operation. Several methods of organizing the data have been proposed, and each has certain advantages. The Thoma diagrams shown in Figures 52 and 5345 are easily understood and are truly complete characteristics diagrams because all possible modes of operation are covered (see also Reference 57). Figure 54 shows schematic cross-sections of the two pumps tested by Swanson58 for which characteristics are given in Figure 53. The Karman circle diagram (Reference 59) attempted to show the complete characteristics as a four-quadrant contour plot of surfaces representing head and torque with speed and flow rate as base coordinates. Because the head and torque tend to infinity in two zones of operation, another diagram would be required to show the complete pump characteristics. The data presented in such a diagram are, nevertheless, adequate for almost all requirements. One example of a circle diagram is given in Figures 55 and 56. Other examples may be found in References 12, 58, and 59. Frequently, tests with negative head and torque have been omitted so that only half of the usual circle diagram could be shown. This has been called a three-quadrant plot, but the information necessary to predict an event, such as possible water-column separation following a power failure, is lacking. Power Failure Transient A sudden power failure that leaves a pump and driver running free may cause serious damage to the system. Except for rare cases where a flywheel is provided, the pump and driver usually have a rather small moment of inertia, and so the pump will slow down rapidly. Unless the pipeline is very short, the inertia of the liquid will maintain a strong forward flow while the decelerating pump acts as a throttle valve. The pressure in the discharge line falls rapidly and, under some circumstances, may go below atmospheric pressure, both at the pump discharge and at any points of high elevation along the pipeline. The minimum pressure head which occurs during this phase of the motion is called the downsurge, and it may be low enough to cause vaporization followed by complete separation of the liquid column. Pipelines have collapsed under the external atmospheric pressure during separation. When the liquid columns rejoin, following separation, the shock pressures may be sufficient to rupture the pipe or the pump 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.377 FIGURE 52 Complete pump characteristics. Specific speeds: Voith pump ns = 1,935 (0.708); Delaval pump  ns = 1,500 (0.549). (Reference 45) 2.378 CHAPTER TWO FIGURE 53 Complete pump characteristics. Specific speeds: Radial flow  ns = 1800 (0.659); Peerless 10MH mixed flow  ns = 7,550 (2.763); Peerless 10PL axial flow  ns = 13,500 (4.940). (Reference 45) 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.379 FIGURE 54 Schematic cross sections of high-specific-speed pumps. ns = 13,500 (4.940) for axial flow pump; ns = 7,550 (2.763) (Reference 45) casing. Closing a valve in the discharge line will only worsen the situation, and so valves having programmed operation should be closed very little, if at all, before reverse flow begins. Reversed flow may be controlled by valves or by arranging to have the discharge pipe empty while air is admitted at or near the outlet. If reversed flow is not checked, it will bring the pump to rest and then accelerate it with reversed rotation. Eventually the pump will run as a turbine at the runaway speed corresponding to the available static head diminished by the frictional losses in the system. However, while reversed flow is being established, the reversed speed may reach a value considerably in excess of the steadystate runaway speed. Maximum reversed speed appears to increase with increasing efficiency and increasing specific speed of the pump. Calculations indicate maximum reversed speeds more than 150% of normal speed for ns  1935 (0.708) and h  84.1%45. This should be considered in selecting a driver, particularly if it is a large electric motor. There will be a pressure increase, called the upsurge, in the discharge pipe during reversed flow. The maximum upsurge usually occurs a short time before maximum reversed speed is reached and may cause a pressure as much as 60% or more above normal at the pump discharge. A further discussion of power-failure transients is given in Section 8.3. The data of Figures 52 and 53 have been presented in a form suitable for general application to pumps having approximately the same specific speeds as those tested. The symbols are h  H/Hn, q  Q/Qn, m  M/Mn, and n  n/nn, wherein H, Q, M, and n represent instantaneous values of head, flow rate, torque, and speed respectively and the subscript n refers to the values at best efficiency for normal constant-speed pump operation. Any consistent system of units may be used. According to the affinity laws (Eqs. 12), q is proportional to n, and h and m are proportional to n2. Thus the affinity laws are incorporated in the scales of the diagrams. Figures 52 and 53 are divided into sections for convenience in reading data from the curves. The curves of sections 1 and 3 extend to infinity as q/n increases without limit in either the positive or negative direction. This difficulty is eliminated by sections 2 and 4, where the curves are plotted against n/q, which is zero when q/n becomes infinite. Usually any case of transient operation would begin at or near the point q/n  n/q  1, which appears in both sections 1 and 2 of Figures 52 and 53. The detailed analysis of transient behavior is beyond the scope of this treatise. An analytical solution by the rigid column method, in which the liquid is assumed to be a rigid body, is given in Reference 45. Friction is easily included, and the results are satisfactory for many cases. The same reference includes a semigraphical solution to allow for elastic waves in the liquid, but friction must be neglected. Graphical solutions including both elastic waves and friction are discussed in References 45 to 48. Computer solutions of a variety of transient problems are discussed in Reference 47. These offer considerable flexibility in the analysis once the necessary programs have been prepared. ANALYSIS OF TRANSIENT OPERATION 2.380 CHAPTER TWO FIGURE 55 Circle diagram of pump characteristics. Specific speed ns = 4200 (1.537). (Courtesy Combustion Engineering) Some extreme conditions of abnormal operation can be estimated at points where the curves of Figures 52 and 53 cross the zero axes and are listed in Table 9. The data for Columns 2 to 4 of Table 9 were read from Sections 2 of Figures 52 and 53, and the data for Columns 6 and 7 were read from Sections 3. Column 8 was computed from Column 7 by assuming h  1. Let the pump having ns  1500 (0.549) deliver cold water with normal head Hn  100 ft (30.48 m), and let the center of the discharge flange be 4 ft (1.2 m) above the free surface in the supply sump. The discharge pressure head following power failure may be estimated by assuming the inertia of the rotating elements to be negligible relative to the inertia of the liquid in the pipeline. Then q  1 and, from Column 2 of Table 9, the downsurge pressure head is (—0.22)(100)  4  —26 ft (—7.9 m), which is not low enough to cause separation of the water column. 2.381 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS FIGURE 56 Explanatory diagram for Figure 55 TABLE 9 Abnormal operating conditions of several pumps Downsurge Specific speed ns (s) 1 Free-running m/q2  0 2 h/q n/q Runaway turbine Locked-rotor n/q  0 h/q 2 m/q m/n2  0 2 q/n h1 2 h/n n 2 3 4 5 6 7 8 1500 (0.549)a 0.22 0.36 0.55 0.53 0.46 0.77 1.14 1800 (0.659) 0.24 0.31 0.60 0.44 0.56 0.75 1.16 1935 (0.708) 0.36 0.32 0.94 0.60 0.41 0.66 1.23 7550 (2.763) 0.23 0.56 1.57 1.38 0.99 0.38 1.62 13,500 (4.940) 0.12 0.67 0.96 0.60 1.08 0.33 1.73 a Double-suction. Actually the downsurge would be less than this because of the effects of inertia and friction, which have been neglected. If this pump were stopped suddenly by a shaft seizure or by an obstruction fouling the impeller, Column 4 of Table 9 shows the downsurge to be (—0.55)(100)  4  —59 feet (—18 m), which would cause water column separation and, probably, subsequent water hammer. If, following power failure, the pump were allowed to operate as a no-load turbine under the full normal pump head, Column 8 of Table 9 shows the runaway speed would be 1.14 times the normal pump speed. The steady-state runaway speed usually would be less than this because the effective head would be decreased by friction, but higher speeds would be reached during the transient preceding steady-state operation. Column 8 shows that runaway speeds increase with increasing specific speed. 2.382 CHAPTER TWO FIGURE 57A through C Pump assembly and rotation (Reference 7) Incorrect Rotation Correct rotation of the driver should be verified before it is coupled to the pump (Figure 57). Sections 3 of Figures 52 and 53 show that reversed rotation might produce some positive head and flow rate with pumps of low specific speed, but at very low efficiency. It is unlikely that positive head would be produced by reversed rotation of a high-specific-speed pump. Reversed Impeller Some double-suction impellers can be mounted reversed on the shaft. If the impeller is accidentally reversed, as at B in Figure 57, the flow rate and efficiency probably will be much reduced and the power consumption increased. Care should be taken to prevent this, as the error might go undetected in some cases until the driver was damaged by overload. Table 1 shows performance data for six pumps with reversed impellers. At least one of these would overload the driver excessively. Further discussion of abnormal operating conditions may be found in References 7, 12, and 34. Vibration Vibration caused by flow through wearing rings and by cavitation has been discussed in the foregoing and some remedies indicated. Vibration due to unbalance is not usually serious in horizontal units but may be of major importance in long vertical units, where the discharge column is supported at only one or two points. The structural vibrations may be quite complicated and involve both natural frequencies and higher harmonics. Vibration problems in vertical units should be anticipated during the design stage. If vibration is encountered in existing units, the following steps may help to reduce it: (1) dynamically balance all rotating elements of both pump and motor; (2) increase the rigidity of the main support and of the connection between the motor and the discharge column: (3) change the stiffness of the discharge column to raise or lower natural frequencies as required. A portable vibration analyzer may be helpful in this undertaking. Kovats60 has discussed the analysis of this problem in some detail. Structural vibrations can occur in most pump types. Typical sources are a) bearing housings—due to the commonly encountered cantilever construction, b) couplings, c) rotor instabilities stemming from excessive ring clearances and consequent loss of Lomakin stiffening of long-shaft multistage pumps, and d) hydraulic unbalance—due to dimensional variations in flow passages and clearances61. PREDICTION OF EFFICIENCY FROM MODEL TESTS _______________________ Many pumps used in pumped storage power plants and water supply projects are so large and expensive that extensive use is made of small models to determine the best design. It is often necessary to estimate the efficiency of a prototype pump, as a part of the guaran- 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.383 tee, from the performance of a geometrically similar model. A model and prototype are said to operate under dynamically similar conditions when Dn> 2H  D¿n¿> 2H¿, where D  impeller diameter n  pump speed H  pump head in any consistent units of measure. Primed quantities refer to the model and unprimed quantities to the prototype. Dynamic similarity is a prerequisite to model-prototype testing so that losses that are proportional to the squares of fluid velocities, called kinetic losses, will scale directly with size and not change the efficiency. Surface frictional losses are boundary-layer phenomena which depend on Reynolds number Re  D2H>n, where n  kinematic viscosity of the liquid pumped. (See also Section 2.1, Eqs. 36—40.) Reynolds numbers increase with increasing size, and, within limits, surface frictional-loss coefficients decrease with increasing Reynolds number. This leads to a gain in efficiency with increasing size. Computational difficulties have forced an empirical approach to the problem. Details of the development of a number of formulas are given in Reference 62. Moody-Staufer Formula In 1925, L. F. Moody and F. Staufer independently developed a formula that was later modified by Pantell to the form a 1  hh h¿ h D¿ n b a b b a hh 1  h¿ h D (35) where hh  hydraulic efficiency, discussed previously, and D  impeller diameter. Primed quantities refer to the model, and n is a constant to be determined by tests. The model must be tested with the same liquid that will be used in the prototype; that is, cold water in most practical cases. The original formula contained a correction for head, which is negligible if H¿ 0.8H, and this requirement is now virtually mandatory in commercial practice. The meager information available indicates 0.2 n 0.1 approximately, with the higher value currently favored. Improvements in construction and testing techniques very likely will move n toward lower value in the future. The Moody-Staufer formula has been widely used since first publication. In practice, both h¿h and hh are usually replaced by the overall efficiencies h¿ and h, respectively, because of the difficulty in determining proper values for the mechanical and volumetric efficiencies (see Eq. 11). Rütschi Formulas The general form of several empirical formulas due to K. Rütschi62 and others was originally given as hh  f h¿ f¿ h (36) where hh and h¿h are the hydraulic efficiencies of the prototype and model, respectively, and f and f ¿ are values of an empirical f function for both the prototype and the model. The f function was obtained from tests of six single-stage pumps. nS 2000 (0.732) and is shown in Figure 58 based on the eye diameters Do of the pumps in millimeters. Thus the f function depends on actual size in addition to scale ratio. The extrapolated portion of the curve, shown dashed in Figure 58, checked well with values for a model and large prototype, shown by E¿ and E, respectively. One of several formulas that have been proposed to fit the curve in Figure 58 is, in SI units, f1 3.15 D1.6 o (37) 2.384 FIGURE 58 and 62) CHAPTER TWO The f function for the Rütschi formula (to obtain Do in inches, multiply by 0.03937) (References 13 FIGURE 59 Chart for the solution of the Rütschi formula (to obtain Do in meters, multiply by 0.3048) (Reference 62) where the eye diameter Do is in centimeters or, in USCS units, f1 0.0133 D1.6 o (38) where Do is in feet. Figure 59 gives a graphical solution of Eq. 38. The Society of German Engineers (VDI) has adopted a slightly modified version of the Rütschi formula as standard. Rütschi later recommended, in discussion of Reference 62, that the internal effi- 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.385 ciency hi  h/hm be used instead of the hydraulic efficiency hh in Eq. 45. The mechanical efficiency hm will probably be very high for both model and prototype for most cases of interest, so good results should be obtained if the overall efficiency is used in Eq. 36. Model-prototype geometric similarity should include surface finish and wearing ring or tip clearances, but this may be difficult or impossible to achieve. Anderson (see Section 2.1: Figure 10 and Reference 6) proposed a method that includes a correction for dissimilarity in surface finish. OPERATION OF PUMPS AS TURBINES __________________________________ Centrifugal pumps may be used as hydraulic turbines in some cases where low first cost is paramount. Because the pump has no speed-regulating mechanism, considerable speed variation must be expected unless the head and load remain very nearly constant. Some speed control could be obtained by throttling the discharge automatically, but this would increase the cost, and the power lost in the throttle valve would lower the overall efficiency. Pump Selection After the head, speed, and power output of the turbine have been specified, it is necessary to select a pump that, when used as a turbine, will satisfy the requirements. Assuming that performance curves for a series of pumps are available*, a typical set of such curves should be normalized using the head, power, and flow rate of the best efficiency point as normal values. These curves will correspond to the right part of Section 1 of either Figure 52 or 53. In normalizing the power P, let p  P/Pn and the curve of p/n3 will be identical with the curve of m/n2 in Figure 52 or 53. The normalized curves may be compared with the curves in Sections 1 of Figures 52 and 53 to determine which curves best represent the characteristics of the proposed pump. When a choice has been made, the approximate turbine performance can be obtained from the corresponding figure of Figures 60 to 63. Assume that the turbine specifications are HT  20 ft, PT  12.75 hp and n  580 rpm, and that the characteristics of the DeLaval L1O/8 pump are representative of a series of pumps from which a selection can be made. The turbine discharge QT in gallons per minute after substituting the above values is EXAMPLE QT  3960PT 2520  hT HThT (39) FIGURE 60 Dimensionless characteristic curves for DeLaval L 10/8 pump, constant-discharge turbine operation (Reference 63) *It is assumed that turbine mode characteristics of the proposed pump are not available when the initial selection is made. 2.386 CHAPTER TWO FIGURE 61 Dimensionless characteristic curves for Voith pump, constant-discharge turbine operation (Reference 63) FIGURE 62 Dimensionless characteristics curves for Peerless 10MH pump, constant-discharge turbine operation (Reference 63) FIGURE 63 Dimensionless characteristic curves for Peerless 10PL pump, constant-discharge turbine operation (Reference 63) where hT is turbine efficiency, shown in Figure 60. Only the normal values Qn, Hn Pn, and so on, are common to the curves of both Figures 52 and 60, so these alone can be used in selecting the required pump. Values of n/q, h/q2, and hT are read from the curves of Figure 60 and corresponding values of Q computed by Eq. 39. Values of Qn and Hn are then given by (40) Qn  Q1n>q2 Hn  H1n>q2 2 h>q2  201n>q2 2 h>q2 (41) 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.387 FLOW RATE, GPM FIGURE 64 Head-flow curves of the pump selection (ft ⫻ 0.3048 = m; gpm ⫻ 0.06300 = l/s) (Reference 63) For example, in Figure 60 at n/q  0.700, read h/q2  0.595 and hT  0.805. By Eq. 39, QT  2520/0.805  3130 gpm; by Eq. 40, Qn  (3130)(0.700)  2190 gpm; and by Eq. 41, Hn  (20)(0.700)2/0.9595  16.5 ft. In a similar manner, the locus of the best efficiency points for an infinite number of pumps, each having the same characteristics as shown in Figures 52 and 60, is obtained, and each pump would satisfy the turbine requirements. This locus of best efficiency points is plotted as curve A in Figure 64. The head curve for a DeLaval L 16/14 pump having a 17-in-diameter impeller tested at 720 rpm is shown as curve B in Figure 64. The best efficiency point was found to be at Qn  3500 gpm and Hn  37.2 ft. The locus of the best efficiency points for this pump for different speeds and impeller diameters is given by Eqs. 12 as Hn  37.2 a Qn 2 3.04 2 Qn b  3500 106 (42) and is shown by curve C in Figure 64. Curve C intersects curve A at two points, showing that the L 16/14 pump satisfies the turbine requirements. Only the intersection at the higher turbine efficiency is of interest. At this point, Qn  2490 gpm and Hn  18.8 ft. Because the turbine speed was specified to be 580 rpm, the required impeller diameter is given by Eqs. 12 as D 1172 12490>35002 580>720  15 in or by D 17118.8>37.2 580>720  15 in The computed head curve for the 15-in-diameter impeller at 580 rpm is shown as curve D in Figure 64. The turbine discharge is 3200 gpm from Eq. 39 with hT  0.787. The optimum solution would be to have Curve C tangent to Curve A at the point corresponding to maximum turbine efficiency, in this case hT  0.812. Because this 2.388 CHAPTER TWO would require a smaller pump, Curve E in Figure 64 shows a head curve for a K 14/12 pump, which was the next smaller pump in the series. Curve F, the locus of the best efficiency points, does not intersect Curve A, showing that the smaller pump will not satisfy the turbine requirements. It is important to note that the turbine head, 20 ft, specified for this example was assumed to be the net head from the inlet to outlet flange of the pump when installed and operated as a turbine. The procedure outlined above should lead to the selection of a pump large enough to provide the required power. However, it probably will be necessary to apply the affinity laws over such wide ranges of the variables that the usual degree of accuracy should not be expected. Considerable care should be exercised if it becomes necessary to interpolate between the curves of Figures 60 to 63. The computed performance will very likely differ from the results of subsequent tests. The curves of Figure 60 may be converted to show the constant-head characteristics of the L 16/14 pump when installed and operated as a turbine. Details of the method of computation are given in Reference 63, and the computed characteristics are shown in Figure 65. A comprehensive study by Acres American64,65 led to a computer program to aid in selecting a pump to meet specific requirements when operating in the turbine mode. Known pump characteristics are entered in the program according to a specified format. The computer compares them with stored characteristics of pumps for which turbine mode characteristics are known and provides estimated turbine mode characteristics for the proposed pump. Vols. I and III of Reference 64 describe the method of computation and give complete instructions for using the program. Optimized Hydraulic Turbines A logical outgrowth of applying pumps as hydraulic turbines is the optimization of these machines as turbines. Several points of efficiency improvement have been demonstrated over that of pumps running in reverse as turbines.66 The fluid accelerates through both the stator or nozzles and the turbine wheel or runner, whereas it decelerates through the same elements when they are acting as impeller and diffuser or volute in the pumping mode. If the fluid never has to decelerate FIGURE 65 Computed constant-head turbine characteristics for DeLaval L 16/14 pump (gpm ⫻ 0.06309 = l/s; hp ⫻ 0.7457 = kW) (Reference 63) 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.389 because the turbine will never be used as a pump, the nozzles and runner can be designed to produce more aggressive and efficient acceleration in shorter distances than if the same machine is simply a pump being used as a turbine. If it is a radial-inflow runner, the resulting optimized turbine wheel employs a radial blade (with a blade angle b  90 deg. from the tangential direction) at the outer diameter (inlet). This outer diameter is about 75% of the diameter of the impeller of the pump-as-turbine, which further improves the efficiency by reason of the reduction of the disk friction drag of the runner. Design and application of both approaches is explained and compared in Reference 67. Many of the applications of these hydraulic turbines are for power recovery in the pressure let-down processes that occur in petroleum refining. Another variable is introduced in such processes; namely, the evolution of large volumes of dissolved gas as the pressure decreases through successive stages or portions of the turbine. This phenomenon affects the performance of hydraulic turbines for such applications because somewhat more power is produced when gas evolves from the liquid than when a single-phase liquid flows through the turbine at the same pressure drop and mass flow rate.68 VORTEX PUMPS _____________________________________________________ A typical vortex pump is shown in Figure 66.* The ability of this type of pump to handle relatively large amounts of suspended solids as well as entrained air or gas more than offsets the relatively low efficiency. Table 10 lists performance data for four typical vortex pumps and four radial-flow centrifugal pumps of nearly the same head and flow rate. Figure 67 shows the head characteristics of a typical vortex pump with impellers of different diameters together with curves of constant efficiency and constant NPSH. Power curves for the same impellers are shown in Figure 68. Curves for a conventional radial-flow pump have been added for comparison in Figures 67 and 68. Note that the head of the vortex pump does not decrease as rapidly with an increasing flow rate as does the head of the conventional pump. The power requirement of the vortex pump increases almost linearly with an increasing flow rate, whereas the power required by a conventional pump of about the same specific speed reaches a maximum and then decreases with the increasing flow rate. Thus if the motor of the vortex pump has been selected to match the power required at the normal flow rate for best efficiency, it will be overloaded if the pump operates much beyond that point. FIGURE 66 *See also Section 9.2. Vortex pump (courtesy Fybroc Division, METPRO) TABLE 10 Characteristics of typical vortex pumps and comparable radial-flow centrifugal pumps at 1750 rpm Impeller Flow rate, gpm (l/s) 2.390 Vortex 100 Radial 108 Vortex 200 Radial 225 Vortex 850 Radial 900 Vortex 1050 Radial 1250 Vortex average Radial average (6.3) (6.8) (12.6) (14.2) (53.6) (56.8) (66.2) (78.9) Total head, ft (m) 24 26 61 60 108 102 150 154 (7.3) (7.9) (18.6) (18.3) (32.9) (31.1) (45.7) (46.9) Specific speed, ns (s) 1614 1580 1134 1218 1523 1636 1323 1415 1399 1462 (0.591) (0.578) (0.415) (0.446) (0.557) (0.599) (0.484) (0.518) (0.512) (0.535) Suction specific speed, S (ss) Impeller diameter, in (mm) Sphere diameter,a in (mm) 7700 6430 7400 6450 9820 8690 7088 578 578 838 838 1112 1112 13 13 2 (2.82) (2.35) (2.71) (2.36) (3.59) (3.18) (2.59) Diameter of the largest sphere that will pass through the pump Subscript n designates values at the best efficiency point a b (149) (149) (213) (213) (292) (292) (330) (330) 5 8 2 3 4 4 7 8 334 1 132 (51) (16) (51) (19) (102) (22) (95) (26) Efficiencyb at Q  Qn, % Shutoff power,b %Pn Powerb at 1.5Qn, %Pn Shutoff head,b %Hn Headb at 1.5Qn, %Hn 39.5 59 45 61 59 76 56 83 50 70 37.5 50 45 50 49 47 48 52 45 50 131 108 129 130 134 113 137 123 133 119 120 144 123 125 120 135 115 116 120 130 71 50 82 65 83 44 87 45 81 51 NPSHRb at Qn, ft (m) 3 4 5 6.5 9 11 16 11 (0.9) (1.2) (1.5) (2.0) (2.7) (3.4) (4.9) (3.4) NPSHRb at 1.5Qn, ft (m) 8 7.5 8 14 15 21 20 - (2.4) (2.3) (2.4) (3.4) (4.6) (6.1) (6.1) (-) 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.391 FLOW RATE, GPM FIGURE 67 Head characteristics of a typical vortex pump. Curves show approximate characteristics when pumping clear water (in ⫻ 2.54 = cm) (Flowserve Corporation). FLOW RATE, GPM FIGURE 68 Power characteristics of a typical vortex pump. Curves show approximate characteristics when pumping clear water (in ⫻ 2.54 = cm) (Flowserve Corporation). REFERENCES _______________________________________________________ 1. Pfleiderer, C. Die Kreiselpumpen für Flüssigkeiten und Gase. 5te Auflage, SpringerVerlag, 1961. 2. Gülich, J., Favre, J. N., and Denus, K. “An Assessment of Pump Impeller Performance Predictions by 3D-Navier Stokes Calculations.” Third International Symposium on Pumping Machinery (S239), ASME Fluids Engineering Division Summer Meeting, Paper No. FEDSM97-3341, June 1997. 2.392 CHAPTER TWO 3. Rupp, W. E. High Efficiency Low Specific Speed Centrifugal Pump. U.S. Patent No. 3,205,828, September 14, 1965. 4. Barske, U. M. “Development of Some Unconventional Centrifugal Pumps.” Proc. Inst. Mech. Eng., London, 174(11):437, 1960. 5. Manson, W. W. “Experience with Inlet Throttled Centrifugal Pumps, Gas Turbine Pumps.” Cavitation in Fluid Machinery, Symposium Publication, ASME, 1972, pp. 21—27. 6. Wislicenus, C. F. “Critical Considerations on Cavitation Limits of Centrifugal and Axial-Flow Pumps.” Trans ASME, 78:1707, 1956. 7. Karassik, I. J., and Carter, R. Centrifugal Pumps: Selection, Operation and Maintenance. McGraw-Hill, New York, 1960. 8. Holland, F. A., and Chapman, F. S. Pumping of Liquids. Reinhold, New York, 1966. 9. Ippen, A. T. “The Influence of Viscosity on Centrifugal Pump Performance.” Trans. ASME, 68(8):823, 1946. 10. Black, H. F., and Jensen, D. N. “Effects of High-Pressure Ring Seals on Pump Rotor Vibrations.” ASME Paper No. 71-WA/FF-38, 1971. 11. Wood, C. M., Welna, H., and Lamers, R. P. “Tip-Clearance Effects in Centrifugal Pumps.” Trans. ASME, J. Basic Eng., Series D, 89:932, 1965. 12. Stepanoff, A. J. Centrifugal and Axial Flow Pumps, 2nd ed., Krieger Publishing, Malabar, FL, 1957. 13. Rütschi, K. “Untersuchungen an Spiralgehäusepumpen veraschiedener Schnelläufigkeit,” Schweiz. Arch Angew, Wiss. Tech. 17(2):33, 1951. 14. Stepanoff, A. J. Pumps and Blowers: Two Phase Flow. Krieger Publishing, Malabar, FL, 1965. 15. Knapp, R. T. “Recent Investigations of the Mechanics of Cavitation and Cavitation Damage.” Trans. ASME 77:1045, 1955. 16. Knapp, R. T. “Cavitation Mechanics and Its Relation to the Design of Hydraulic Equipment.” James Clayton, Lecture, Proc. Inst. Mech. Eng., London, Sec. A, 166:150, 1952. 17. Shutler, N. D., and Mesler, R. B. “A Photographic Study of the Dynamics and Damage Capabilities of Bubbles Collapsing Near Solid Boundaries.” Trans. ASME, J. Basic Eng., Series D, 87:511, 1965. 18. Hickling, R., and Plesset, M. S. “The Collapse of a Spherical Cavity in a Compressible Liquid.” Division of Engineering and Applied Sciences, Report No. 85-24, California Institute of Technology, March 1963. 19. Pilarczyk, K., and Rusak, V. “Application of Air Model Testing in the Study of Inlet Flow in Pumps.” Cavitation in Fluid Machinery, ASME, 1965, p. 91. 20. Plesset, M. S. “Temperature Effects in Cavitation Damage.” Trans. ASME, J. Basic Eng., Series D, 94:559, 1972. 21. Hammitt, F. C. “Observations on Cavitation Damage in a Flowing System.” Trans. ASME, J. Basic Eng., Series D, 85:347 (1963). 22. Preece, C. M., ed. Treatise on Materials Science and Technology. Vol. 16, Erosion, Academic Press, New York, 1979. 23. Kovats, A. Design and Performance of Centrifugal and Axial Flow Pumps and Compressors. Macmillan, New York, 1964. 24. Palgrave, R., and Cooper, P. “Visual Studies of Cavitation in Pumping Machinery.” Proceedings of the Third International Pump Symposium, Texas A&M University, 1986, pp. 61—68. 25. Cooper, P., Sloteman, D. P., Graf, E., and Vlaming, D. J. “Elimination of CavitationRelated Instabilities and Damage in High-Energy Pump Impellers.” Proceedings of the Eighth International Pump Users Symposium, Texas A&M University, 1991, pp. 3—19. 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.393 26. Vlaming, D. J. “Optimum Impeller Inlet Geometry for Minimum NPSH Requirements for Centrifugal Pumps.” Pumping Machinery—1989, ASME, July 1989, pp. 25—29. 27. Hydraulic Institute ANSI/HI 2000 Edition Pump Standards, Hydraulic Institute, Parsippany, NJ www.pumps.org. 28. “Centrifugal Pumps.” PTC 8.2-1965, American Society of Mechanical Engineers, New York, 1965. 29. Wislicenus, C. F., Watson, R. M., and Karassik, I. J. “Cavitation Characteristics of Centrifugal Pumps Described by Similarity Considerations.” Trans. ASME 61:17, 1939; 62:155, 1940. 30. Stahl, H. A., and Stepanoff, A. J. “Thermodynamic Aspects of Cavitation in Centrifugal Pumps.” Trans. ASME 78:1691, 1956. 31. Salemann, V. “Cavitation and NPSH Requirements of Various Liquids.” Trans. ASME, J. Basic Eng., Series D, 81:167, 1959. 32. Stepanoff, A. J. “Cavitation Properties of Liquids.” Trans. ASME, J. Eng. Power, Series A, 86:195, 1964. 33. Cooper, P. “Analysis of Single- and Two-Phase Flows in Turbopump Inducers.” Transactions of the ASME, Series A, Vol. 89, 1967, pp. 577—588. 34. Lazarkiewicz, S., and Troskolan¿ ski, A. T. Impeller Pumps, Pergamon Press, New York, 1965. 35. Gongwer, C. A. “A Theory of Cavitation Flow in Centrifugal-Pump Impellers.” Trans. ASME, 63:29, 1941. 36. Guelich, J. F. Guidelines for Prevention of Cavitation in Centrifugal Feedpumps. EPRI CS-6398, 1989. 37. ASTM Standard G-32. Cavitation Erosion Using Vibratory Apparatus, American Society for Testing Materials, 1992. 38. Cooper, P., and Antunes, F. F. “Cavitation Damage in Boiler Feed Pumps.” Symposium Proceedings: Power Plant Feed Pumps—The State of the Art, EPRI CS-3158, July 1983, pp. 2—24 to 2—49. 39. Cooper, P., Sloteman, D. P., and Dussourd, J. L. “Stabilization of the Off-Design Behavior of Centrifugal Pumps and Inducers.” Proceedings of the Second European Congress on Fluid Machinery for the Oil, Petrochemical and Related Industries, I Mech E Conference Publications, 1984-2, Paper No. C41/84, 1984, pp. 13—20. 40. Stripling, L. B., and Acosta, A. J. “Cavitation in Turbopumps—Part 1.” Transactions of the ASME, Series D, Vol. 84, 1962, pp. 326—338. 41. Stripling, L. B. “Cavitation in Turbopumps—Part 2.” Transactions of the ASME, Series D, Vol. 84, 1962, pp. 339—350. 42. Grohmann, M. “Extend Pump Application with Inducers.” Hydrocarbon Processing, Dec. 1979, p. 121. 43. Doolin, J. H. “Centrifugal Pumps and Entrained Air Problems.” Pump World, 4(3), 1978. 44. Mechanical Engineering. 93(6):89, 1971. 45. Kittredge, C. P. “Hydraulic Transients in Centrifugal Pump Systems.” Trans. ASME, 78(6):1807, 1956. 46. Parmakian, J. Waterhammer Analysis. Prentice-Hall, Englewood Cliffs, NJ, 1955. 47. Streeter, V. L., and Wylie, E. B. Hydraulic Transients. McGraw-Hill, New York, 1967. 48. Bergeron, L. Waterhammer in Hydraulics and Wave Surges in Electricity. Wiley, New York, 1961. 49. Addison, H. Centrifugal and Other Rotodynamic Pumps. 3rd ed. Chapman & Hall, London, 1966. 2.394 CHAPTER TWO 50. Richardson, C. A. “Economics of Electric Power Pumping.” Allis-Chalmers Elec. Rev. 9:20, 1944. 51. Fraser, W. H. “Flow Recirculation in Centrifugal Pumps.” Tenth Turbomachinery Symposium. Texas A&M University, College Station, TX, 1981, p. 95. 52. Fraser, W. H. “Recirculation in Centrifugal Pumps.” Materials of Construction of Fluid Machinery and Their Relationship to Design and Performance. ASME, November 1981, pp. 65—86. 53. Gopalakrishnan, S. “A New Method for Computing Minimum Flow.” Proceedings of the Fifth International Pump Users Symposium. Texas A&M University, 1988, pp. 41—47. 54. “Code of Practice for Pump Efficiency Testing by the Direct Thermodynamic Method.” Report 695/27, The Pump Centre, AEA Technology plc, Birchwood Science Park, Warrington WA3 6AT, UK, June 1995. 55. Agostinelli, A., Nobles, D., and Mockridge, C. R. “An Experimental Investigation of Radial Thrust in Centrifugal Pumps.” Trans. ASME, J. Eng. Power, Series A, 82:120, 1960. 56. Biheller, H. J. “Radial Force on the Impeller of Centrifugal Pumps with Volute, Semivolute, and Fully Concentric Casings.” Trans. ASME, J. Eng. Power, Series A, 87:319, 1965. 57. Donsky, B. “Complete Pump Characteristics and the Effects of Specific Speeds on Hydraulic Transients.” Trans. ASME, J. Basic Eng., Series D, 83:685, 1961. 58. Swanson, W. M. “Complete Characteristic Circle Diagrams for Turbomachinery.” Trans. ASME 75:819, 1953. 59. Knapp, R. T. “Complete Characteristics of Centrifugal Pumps and Their Use in the Prediction of Transient Behavior.” Trans. ASME 59:683, 1937; 60:676, 1938. 60. Kovats, A. “Vibration of Vertical Pumps.” Trans. ASME, J. Eng. Power, Series A, 84:195, 1962. 61. Bolleter, U., Leibundgut, E., Sturchler, R., and McCloskey, T. “Hydraulic Interaction and Excitation Forces of High Head Pump Impellers.” Pumping Machinery—1989, FED-Vol. 81, ASME, 1989, pp. 187—193. 62. Kittredge, C. P. “Estimating the Efficiency of Prototype Pumps from Model Tests.” Trans. ASME, J. Eng. Power, Series A, 90:129, 301, 1968. 63. Kittredge, C. P. “Centrifugal Pumps Used as Hydraulic Turbines.” Trans. ASME, J. Eng. Power, Series A, 83:74, 1961. 64. Acres American Inc. for U.S. Department of Energy, Idaho National Engineering Laboratory, Small Hydro Plant Development Program. Vols. I, II, and III, subcontract No. K-1574, Oct. 1980. Available from National Technical Information Service, U.S. Department of Commerce, Springfield, VA 22161. 65. Lawrence, J. D., and Pereira, L. “Innovative Equipment for Small-Scale Hydro Developments: Waterpower '81.” An International Conference on Hydropower, Proceedings. Vol. II, Washington, DC, June 22-24, 1981, pp. 1622—1639. 66. Cooper, P., and Nelik, L. “Performance of Multi-Stage Radial-Inflow Hydraulic Power Recovery Turbines.” Paper presented at ASME Winter Annual Meeting, New Orleans, December 1984. 67. Gopalakrishnan, S. “Power Recovery Turbines for the Process Industry.” Proceedings of the Third International Pump Symposium, Texas A&M University, 1986, pp. 3—11. 68. Hamkins, C. P., Jeske, H. O., Apfelbacher, R. and Schuster, O. “Pumps as Energy Recovery Turbines With Two-Phase Flow.” Pumping Machinery—1989, FED-Vol. 81, ASME, 1989, pp. 73—81. 69. Kallas, D. H., and Lichtman, J. Z. “Cavitation Erosion.” In Environmental Effects on Polymeric Materials, D. V. Rosato and R. T. Schwartz, eds., Wiley-Interscience, New York, 1968, pp. 223-280. 2.3.1 CENTRIFUGAL PUMPS: GENERAL PERFORMANCE CHARACTERISTICS 2.395 FURTHER READING __________________________________________________ Anderson, H. H. Centrifugal Pumps, Trade and Technical Press, Surrey, England, 1980. Brennen, C. E. Hydrodynamics of Pumps, Concepts ETI, Inc., and Oxford University Press, 1994. Carter, R. “How Much Torque Is Needed to Start Centrifugal Pumps?” Power 94(1):88, 1950. Church, A. H. Centrifugal Pumps and Blowers. Krieger Publishing, Malabar, FL, 1944. Heald, C. C. Cameron Hydraulic Data, 18th ed., 1998. Flowserve Corporation, Irving, TX 75039. Japikse, D., Marscher, W. D., and Furst, R. B. Centrifugal Pump Design and Performance. Concepts ETI, Inc., Wilder, VT, 1997. Moody, L. F., and Zowaki, T. “Hydraulic Machinery.” 3rd ed., sec. 26 of Handbook of Applied Hydraulics. Davis and Sorenson, eds., McGraw-Hill, New York, 1969. Spanohake, W. Centrifugal Pumps, Turbines, and Propellers. Technology Press, MIT, Cambridge, MA, 1934. Wislicenus, C. F. Fluid Mechanics of Turbomachinery. Dover, New York, 1965. 2.3.2 CENTRIFUGAL PUMP HYDRAULIC PERFORMANCE AND DIAGNOSTICS WARREN H. FRASER Any successful mathematical model of the mechanics of head generation in centrifugal pumps should do more than just make accurate predictions of pump performance; it should also be capable of identifying the cause of operational difficulties. Unlike mechanical malfunctions, which can be detected, analyzed, and corrected, many of the problems caused by hydraulic forces cannot be corrected in the mechanical sense—they are the unavoidable side effects of head generation in a rotating pressure field. For example, a thrust bearing may fail from lack of lubrication, from misalignment, or because an underrated bearing has been used. These are mechanical failures that can be corrected. A thrust bearing may also fail from a complex pattern of dynamic loading that reflects pressure pulsations of high intensity and a broad spectrum of frequencies during operation at reduced flow. This is an example of a failure from hydraulic causes and can be corrected only by resorting to an oversized bearing or modifying the operation of the pumping system to avoid low-flow operation. Whenever a consistent correlation can be made between the known dynamics of head generation and operational problems, it is possible to devise a strategy to improve operation and reduce mechanical failures. The most significant problems caused by hydraulic dynamic forces can be listed as follows. CAVITATION _________________________________________________________ Cause and Effect Cavitation is the formation of vapor bubbles in any flow that is subjected to an ambient pressure equal to or less than the vapor pressure of the liquid being pumped. Cavitation damage is the loss of material produced by the collapse of the vapor bubbles against the surfaces of the impeller or casing. Formation of these bubbles cannot occur if the net positive suction head supplied, or NPSHA, exceeds the NPSH required for 2.397 2.398 CHAPTER 2 cavitation inception. This NPSH inception value is usually significantly higher than the usually mentioned NPSH required, or NPSHR, which is based on a certain amount of cavitation being present to create a prescribed deviation in pump performance. Diagnosis from Pump Operation Pump operation in the presence of sufficient cavitation activity will reduce both the total head and the output capacity. A steady crackling noise in and around the pump suction indicates cavitation. A random crackling noise with high-intensity knocks indicates suction recirculation but does not indicate a degradation of performance if NPSHA is greater than NPSHR. The random crackling is the unsteady occurrence of cavitation that generally accompanies suction recirculation, which, in turn, is an unsteady phenomenon. Diagnosis from the Visual Examination of Surface Damage Cavitation damage from inadequate NPSH (that is, NPSHA 6 NPSHR) occurs on the low-pressure or the visible surface of the impeller inlet vane. Instrumentation A suction gage or manometer in the pump suction can be used to determine whether the NPSH available is equal to or greater than the NPSH required from the manufacturer’s rating curve (NPSHR). Corrective Procedures If additional NPSH cannot be supplied, the capacity of the pump should be reduced until the required NPSH is equal to or less than the available NPSH. If this is not possible, impeller improvement may be necessary. This may be accomplished on some impeller designs by reworking the geometry or impeller surface finish to reduce losses, improve flow characteristics, or increase the flow inlet area (thus lowering impeller inlet velocity). If this is not possible, consideration should be given to replacing the impeller (first stage or suction impeller on multistage pumps) with one of improved suction performance. The pump manufacturer should be consulted to determine the most satisfactory course of action. SUCTION AND DISCHARGE RECIRCULATION ____________________________ Cause and Effect Recirculation occurs at reduced flows and is the reversal of a portion of the flow back through the impeller. Recirculation at the inlet of the impeller is known as suction recirculation. Recirculation at the outlet of the impeller is discharge recirculation. Suction and discharge recirculation can be very damaging to pump operation and should be avoided for the continuous operation of pumps of significant energy level or pressure rise per stage. Diagnosis from Pump Operation Suction recirculation will produce the previously mentioned loud crackling noise in and around the suction of the pump. Recirculation noise is of greater intensity than the noise from low-NPSH cavitation and is a random knocking sound. Discharge recirculation will produce the same characteristic sound as suction recirculation except that the highest intensity is in the discharge volute or diffuser. Diagnosis from Visual Examination Suction and discharge recirculation produce cavitation damage to the pressure side of the impeller vanes. Viewed from the suction of the impeller, the pressure side would be the invisible, or underside, of the vane. Figure 1 shows how a mirror can be used to examine the pressure side of the inlet vane for cavitation damage from suction recirculation. This is unlike cavitation damage from inadequate NPSH that occurs on the low pressure surface of the inlet vanes. Damage to the pressure side of the vane from discharge recirculation is shown in Figure 2. Guide vanes in the suction may show cavitation damage from impingement of the backflow from the impeller eye during suction recirculation. Similarly, the casing tongue or diffuser vanes may show cavitation damage on the impeller side from operations in discharge recirculation. 2.3.2 CENTRIFUGAL PUMP HYDRALIC PERFORMANCE AND DIAGNOSTICS 2.399 FIGURE 1 Examining the pressure side of the inlet vanes for suction recirculation damage FIGURE 2 Damage to the pressure side of the vane from discharge recirculation Instrumentation The presence of suction or discharge recirculation can be determined by monitoring the pressure pulsations in the suction and in the discharge areas of the casing. Piezoelectric transducers installed as close to the impeller as possible in the suction and in the discharge of the pump can be used to detect pressure pulsations. The data may be analyzed with a spectrum analyzer coupled to an XY plotter to produce a record of the 2.400 CHAPTER 2 FIGURE 3 Pressure pulsations versus capacity pressure pulsations versus the frequency for selected flows. Figure 3 shows a typical plot of pressure pulsations versus capacity. As can be seen, a sudden increase in the magnitude of pressure pulsations indicates the onset of recirculation. The onset of suction recirculation can be determined by an impact head tube (or pitot tube) installed at the impeller eye, as shown in Figure 4. With the tube directed into the eye, the reading in the normal pumping range is the suction head minus the velocity head at the eye. At the point of suction recirculation, however, the flow reversal from the eye impinges on the head tube with a rapid rise in the gage reading. Corrective Procedures Every impeller design has specific recirculation characteristics. These characteristics are inherent in the design and cannot be changed without modifying the design. An analysis of the symptoms associated with recirculation should consider the following as possible corrective procedures: 1. Increase the output capacity of the pump. 2. Install a bypass between the discharge and the suction of the pump. 3. Substitute an improved material for the impeller that is more resistant to cavitation damage. 4. Modify the impeller design. 2.3.2 CENTRIFUGAL PUMP HYDRALIC PERFORMANCE AND DIAGNOSTICS 2.401 FIGURE 4A and B Installation of the impact head tube to detect suction recirculation during (a) normal flow and (b) recirculation flow AXIAL THRUST ______________________________________________________ Cause and Effect Axial thrust is the thrust imposed in the direction of the shaft. It may occur in either the inboard or the outboard direction and is usually composed of a dynamic cyclic component superimposed on a steady-state load in either direction. The dynamic cyclic component increases in the recirculation zone and may impose excessive stresses in the shaft, which could ultimately result in failure from metal fatigue. The static component may impose an excessive load in the thrust bearing, causing unacceptable bearing temperatures. The majority of thrust-bearing failures are caused by fatigue failure of the bearing components from dynamic cyclic axial loads. Diagnosis from Pump Operation High axial loads usually produce high thrust-bearing temperatures and short thrust-bearing life. Diagnosis from Visual Examination of Damage Static thrust in excess of the bearing rating will cause cracking of the balls or rollers and of the race in rolling element bearings and metal scorings of the shoes in tilting-pad bearings. Bearing failure from dynamic loading in excess of the bearing rating will cause fatigue failures of the balls or rollers and race in rolling element bearings. It is important to differentiate clearly between static load failure and fatigue failure. This can be done by examination of a cross section of the failure zone under the microscope. Fatigue failure from dynamic loading will show a hammering effect caused by points of impact. Fatigue failure from excessive static loading will show metal fatigue without the hammering effect of impact loading. BEARING DAMAGE Shaft failure at the outboard, or unloaded, end of the shaft in multistage or double-suction pumps may be a fatigue failure in tension resulting from the high cyclic stresses induced in the shaft when the pump is operated in the discharge recirculation zone. Axial cyclic stresses can be reduced by increasing the pump output or, if this is not possible, by installing a recirculation line to bypass sufficient flow to move the pump total flow rate beyond the point where damaging discharge recirculation occurs. The pump manufacturer can advise the recommended minimum continuous flow for a specific pump design. SHAFT FAILURE 2.402 CHAPTER 2 FIGURE 5 Load cell to monitor axial load Instrumentation Displacement-type pickups should be used to determine the axial movement of the shaft relative to the bearing housing. The deflection of the thrust bearing housing can be determined by seismic instruments. Axial loading of the tilting-shoe type of thrust bearing can be monitored by a load cell permanently installed in the leveling plate. A typical installation is shown in Figure 5. Corrective Procedures To determine the most effective procedure to correct axial thrust problems, it is necessary to determine whether the loads are static, dynamic, or a combination of both. If it is a static failure, the thrust can usually be reduced by restoration of the internal clearances. Most shaft and bearing failures from axial thrust, however, are fatigue failures. If the failure is a fatigue failure, the loading can usually be decreased by increasing the capacity of the pump. If this is not possible, shaft failures can be reduced by substituting a shaft material of higher endurance limit. Rolling element bearing failures can be addressed in large, between bearings pumps by substituting a tilting-shoe type of thrust bearing. The high cyclic axial forces are better absorbed in the oil film of the tilting-shoe bearing than in the rolling element bearing. RADIAL THRUST _____________________________________________________ Cause and Effect Radial thrust is the thrust imposed on the pump rotor and directed toward the center of rotation of the shaft. The forces are usually composed of a dynamic cyclic component superimposed on a steady-state load. The dynamic cyclic component increases rapidly at low-flow operation when the pump is operating in the recirculation zone. The static load also increases with low- and high-flow operation, with the minimum value at or near the maximum efficiency capacity. 2.3.2 CENTRIFUGAL PUMP HYDRALIC PERFORMANCE AND DIAGNOSTICS 2.403 Diagnosis from Pump Operation High radial thrust is difficult to determine from pump operation. Persistent packing or mechanical seal problems may indicate excessive shaft deflection from radial loads. As in the case of high axial loads, high radial loads may produce high bearing temperatures with reduced life. Diagnosis from Visual Examination of Damage Static radial loads in excess of the bearing rating will cause cracking of the balls or rollers and the races in rolling element bearings. In the case of sleeve bearings, the bearing metal will be worn in one direction only and the journal will be worn uniformly. If the opposite is true (that is, the bearing is worn uniformly and the journal excessively in one direction), the cause of the failure is most likely unbalance or a bent shaft and not excessive bearing loads. BEARING DAMAGE Shaft failures from excessive radial loads usually occur at the midpoint of the shaft span in double-suction or multistage pumps. In the case of end-suction pumps, shaft failures usually occur at the shoulder of the shaft, where the impeller hub joins the shaft sleeve, or at the location of the highest stress concentration, if elsewhere. SHAFT FAILURES Instrumentation It is difficult to devise instrumentation to determine excessive radial loading of the shaft and bearings. Temperature rise of the bearings may or may not be symptomatic of excessive radial loading. High bearing temperatures may occur from misalignment, inadequate lubrication, or excessive axial loading of the thrust bearing. These causes should be eliminated before concluding that the radial loads are excessive. Corrective Procedures Most bearing and shaft failures caused by excessive radial loads occur when the pump operates at low flow rates. Radial loads can be reduced by operating the pump at higher capacities or by installing a bypass from the pump discharge back to the pump suction or suction source. For pumps handling water, the life of the shaft may be extended by substituting a martensitic stainless (13% chrome) steel shaft for carbon steel. If there are signs of corrosion as well as fatigue failure, an austenitic stainless steel shaft may also be considered. Physical properties should be evaluated carefully, as the endurance limit of the 300 series steels may be lower than that of chrome steels in fresh water. For liquids other than water, the endurance limit of the shaft material in the liquid being pumped may be a significant determining factor in the life of the shaft in the presence of high dynamic loading. PRESSURE PULSATIONS______________________________________________ Cause and Effect Pressure pulsations are present in both the suction and the discharge of any centrifugal pump. The magnitude and frequencies of the pulsations depend upon the design of the pump, the head produced by the pump, the response of the suction and discharge piping, and the point of operation of the pump on its characteristic curve. The observed frequencies in the discharge may be the running frequency, the vane passing frequency, or multiples of each. In addition, random frequencies with pressure pulsations higher than either the rotating or the vane passing frequencies have been observed. The cause of these random frequency pulsations is sometimes difficult to determine. System resonance, acoustic behavior, eddies from valves and poor upstream piping, and so on, are sometimes involved. However, such random pressure pulsations should not be dismissed as spurious or irrelevant data in any analysis of symptomatic operational problems. The observed frequencies in the pump suction are much lower than in the discharge. Typical frequencies are in the order of 5 to 25 cycles/s, and they do not appear to bear any direct relation to the rotational speed of the pump or the vane passing frequency. Diagnosis from Pump Operation In most pumping installations of 435 lb/in2 (3MPa) [that is, 1000 ft (305 m) of head in water] or less of head per stage, there is little outward 2.404 CHAPTER 2 manifestation of pressure pulsations during normal pumping operation. Other than for specialized applications, such as white water pumps for paper machines (where the discharge pressure pulsations may affect the quality of the paper) or quiet pumps in marine service, there are few external symptoms of internal pressure pulsations. For high-head pumps, however, suction and discharge pressure pulsations may cause instability of pump controls, vibration of suction and discharge piping, and high levels of pump noise. Diagnosis from Visual Examination of Damage In the case of high-head pumps, any failure of internal pressure-containing members should be investigated with consideration given to the possibility that the failures are fatigue failures from internal pressure pulsations. Examination of the fracture will determine whether the failure is a fatigue failure or not. Fatigue failures may have one or more origins. Characteristic markings, known as striations, are often present on the fracture surface. Metallurgical examination of the fracture surface will also disclose striations on a microscopic scale. These markings represent growth of the crack front under cyclic stress. If it is a fatigue failure, the cause can usually be traced to high cyclic stress induced in the pressure-containing member from high-frequency pressure pulsations. Instrumentation Pressure pulsations are usually measured with piezoelectric pressure transducers and recorded as peak-to-peak pressure pulsations over a broad frequency band. Recorded on tape or strip charts, a spectral analysis may be performed for any operating condition. Corrective Procedures A spectral analysis of the pressure pulsations at the suction and at the discharge of the pump is necessary before a strategy for corrective procedures can be developed. After the spectral analysis is available, problems associated with pressure pulsations can usually be reduced by implementing the procedures shown in Table 1. TABLE 1 Corrective procedures for various problems Problem Corrective procedure 1. Vibration of suction or discharge piping a. Search for responsive resonant frequencies in the piping or supports. If any part of the system responds to the frequency of the pressure pulsations, alter the system to shift the resonant frequencies. b. If possible, increase the output of the pump by changing the mode of operation or by installing a bypass from the discharge to the suction of the pump. c. If the piping responds to the vane passing frequency of the pump, the impellers can be replaced with a unit containing either one fewer or one more vane. a. If possible, increase the output of the pump by changing the mode of operation or by installing a bypass from the discharge to the suction of the pump. b. Install acoustical filters to reduce the magnitude of the pressure pulsations. a. If possible, increase the output of the pump by changing the mode of operation or by installing a bypass from the discharge to the suction of the pump. b. Redesign the failed components to reduce the induced cyclic stresses to below the endurance limit of the material. c. If the spectral analysis shows that the maximum pressure pulsations correspond to the vane passing frequency of the impeller, the impeller can be replaced by one having either one fewer or one more vane of the same design. 2. Instability of pump controls 3. Fatigue failure of internal pressurecontaining components of the pump from pressure pulsations 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE, INSTRUMENTATION, AND DIAGNOSTICS CHARLES JACKSON JAMES H. INGRAM MECHANICAL PERFORMANCE _________________________________________ The mechanical performance of a pump would imply only the rotating mechanical masses, with no consideration given to hydraulic (process) effects. The rotating masses (impellers, sleeves, nuts, coupling, bearings, seals, and so on) can be examined as pure mechanics. A person concerned with mechanical performance should be intimately familiar with pump design, construction, and maintenance to be successful. In discussing the mechanical performance of centrifugal pumps, two examples will be used. The first will be a horizontal, 500-hp (373-kW), single-stage (overhung impeller) American Petroleum Institute (API) process pump. The second will be a six-stage, horizontal, 1000-hp (746-kW), multistage boiler-feed pump. Normally, the rotor dynamics will involve (a) a review of the shaft stiffness of the bearings and structure, (b) a mass model of the rotor, and (c) a critical speed analysis with mode shapes of the rotor or shaft. SINGLE STAGE PUMP ________________________________________________ An 8  6  13 pump is operating on water at 3550 rpm with a design flow of 2500 gpm (567 m3/h) at 600 ft (183 m) total head, 1.0 sp. gr., requiring approximately 500 hp (373 kW). The pump operated extremely rough, and the bearings and bearing housing failed. The impeller weighs 61.4 lb (27.9 kg). If the impeller is fitted on the shaft with an eccentricity of 0.002 in (0.051 mm), a calculated centrifugal force Fe of 44 lb (196 N) would cause a deflection of 0.0026 in (0.066 mm) from y  wl3>3EI 2.405 2.406 where w l E I CHAPTER TWO  weight (force) of impeller, lb (N)  length of overhang, in (m)  modules of elasticity, lb/in2 (N/m2 or Pa)  shaft section moment of inertia, in4 (mm4) This pump has a 5212 line bearing and tandem mounted (DB) 7311DB angular contact thrust bearings (40° contact angle). An extremely loose fit of the radial bearing in the bearing housing could cause the outer race to spin, which could cause a vibration equal to twice the rotation frequency. Interference fitting could lead to radial bearings’ accepting thrust (for which many are not designed) from thermal expansion of the shaft or from the thrust bearing. Frequencies Generated The following data and definitions are needed to compute the frequencies generated by defective bearings1: rpm rps FTF BPFI BPFO BSF Bd Nb Pd Ø  revolutions per minute  revolutions per second  fundamental train frequency, Hz  ball passing frequency of inner race, Hz  ball passing frequency of outer race, Hz  ball spin frequency, Hz  ball or roller diameter, in (mm)  number of balls or rollers  pitch diameter, in (mm)  contact angle The formulas are rps  FTF  rpm 60 rps Bd a1 ⫺ cos 0b 2 Pd BPF1  a Nb Bd cos 0b rpsb a 1 ⫹ 2 Pd BPFO  a Nb Bd cos 0b rpsb a 1 ⫺ 2 Pd BSF  a Pd Bd 2 rpsb c 1 ⫺ a b cos2 0 d 2Bd Pd The pitch diameter is the diameter measured across the bearing from ball or roller center to ball or roller center. The contact angle is measured from a line perpendicular to the shaft to the point at which the balls or rollers contact the race. The contact angle of a deep groove ball bearing is zero. It is necessary to distinguish between the ball frequency and the impeller vane passing frequency, which is 17,750 cpm (5 vanes  3550 rpm  1 casing cutwater) for this example. The mode shape of the pump shaft is conical (pivotal) in the first mode of a cantilevered shaft mount. The stiffness map of the rotor looks like that shown in Figure 1. This pump has two design faults, as can be seen from the stiffness map. The first is that an excessively large coupling is used. This heavy overhung mass at the coupling forces the first shaft resonance to be very near the pump operating speed. Normally, the shaft in this type of pump is considered to be “rigid;” that is, operating safely below the first undamped shaft resonance. In this case, the pump is affected by two negatively additive errors. The 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.407 FIGURE 1 Undamped critical speed map of single-stage overhung pump, comparing normal and heavy coupling weights (lb/in  175 = N/m) TABLE 1 Logic of spring equivalent stiffness Ke heavy mass coupling effect is compounded by a weak baseplate that is not properly grouted, leaving a void under the pump supports. In rotor (shaft) supports, two spring supports in series reciprocally add, similar to electric resistors in parallel (Table 1). The effective stiffness is Ke  1 1>Kb ⫹ 1>Ks If the bearing stiffness Kb, is 2.5  106 lb/in (4.4  106 N/m) and the support stiffness Ks, is low; that is, 7.0  105 lb/in (1.2  108 N/m), the effective stiffness is 5.47  105 lb/in (9.57  107 N/m), which moves the first mode resonance from 4300 cpm to 3550 cpm, which is the running speed of the pump. The mode shapes of the rotor are shown in Figures 3 and 4. An animated display is used to better show the rotor gyrations in synchronous whirl. The first modes are shown with a 2.408 CHAPTER TWO FIGURE 2 Rotor cross section of single-stage overhung pump with normal coupling weight. Rotor weight = 110.4 lb (50kg); rotor length = 30.3 in (77.0 cm); number of stations = 22; number of bearings = 2 FIGURE 3 First, second, and third resonate animated mode shapes of single-stage overhung pump with light coupling of 15 lb (6.8 kg) and rigid foundation. Rotor weight = 110.4 lb (50 kg); rotor length = 30.3 in (77.0 cm); number of stations = 22; number of bearings = 2. (a) Mode 1: frequency = 4717 cpm; (b) mode 2: frequency = 53,482 cpm; (c) mode 3: frequency = 67,522 cpm 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.409 FIGURE 4 Synchronous critical speed summary with first three mode shapes of a single-stage overhung pump with light coupling of 15 lb (6.8 kg) and rigid or flexible foundation. Rotor weight = 110.4 lb (50 kg); rotor length = 30.3 in (77.0 cm); number of stations = 22; number of bearings = 2. (a) Rigid support, modes 1, 2, and 3 respectively: 4717, 58,482, and 67,522 rpm; (b) flexible support, modes 1, 2, and 3 respectively: 4085, 21,251, and 40,036 rpm. (lb  0.454 = kg; lb/in  175 = N/m) lighter, normal, and correct coupling (15 lb; 6.8 kg). Figure 2 shows the mathematical model of this pump with the impeller at station 2, the radial bearing at station 11 at 2.5  106 lb/in (4.4  108 N/m) stiffness, the outboard thrust/radial bearing at station 18 at 3.5  106 lb/in (6.1  108 N/m) and the coupling at station 21. Figure 3a, obtained from a finite element computer analysis of the mathematical model in Figure 1, shows the first mode with a rotor weight of 110.4 lb (50 kg), a rotor length of 30.3 in (77.0 cm), and a first mode undamped resonance (critical) of 4717 cpm. This is a pivotal mode, with 100% of the normalized motion at the impeller end. Any motion at the antifriction bearings is greatly restrained. 2.410 CHAPTER TWO Figure 3b shows the second resonance mode, at 53,482 cpm, which is not to be encountered. The coupling motion is now the greatest motion. Figure 3c shows the third mode, at 67,522 cpm. Figure 4a shows an overlay of all three modes with a summary of the criticals, the modal mass, and the relative strain energy (91) in the shaft at station 10 (impeller side of radial bearing). A lesser strain energy is at the radial bearing (station 11). Figure 4b summarizes what happens to critical speed modes if either a more flexible bearing or a soft structure is provided intentionally or unintentionally. Also note that the criticals are lowered significantly and the strain energy is transferred more from the shaft into the bearings; that is, strain values under the U-shaft column are less than under U-bearing column. The first critical is 4085 cpm at a pump speed of 3550 rpm (⫹15%). A 15% margin of separation may be close enough to excite (cause a rise in vibration) the rotor if the resonance response envelope is too wide. However, this is unlikely on antifriction bearings (spiky/narrow response), but possible on sleeve bearings (low/broad response). Figure 5a is a summary which shows the response of a rigid support and an excessively heavy (62 lb  28 kg) coupling, which is as heavy as the impeller. Note that the first mode is again only slightly above the operating speed; that is, 4279 cpm compared with 3550 (⫹21%). The bearing stiffness is assumed to be the controlling stiffness. Many assume that the structure or base stiffness is one order above the bearing stiffness (Ks  l0Kb). This assumption that the bearing stiffness is the controlling stiffness variable is often a very poor assumption. The larger the pump size, the more this is true. That is why an 8  6  13 pump was used as an example. Further, the second mode, at 15,865 cpm, is in an area where the blade passing frequency (5  3550  17,750 cpm) can easily excite this mode, given little variation in support stiffness. Figure 5b is a summary sheet that best illustrates the problem: • The baseplate was improperly installed and grouted. • The elastomeric coupling designed for low-duty, low-speed, and torsional damping was too heavy; that is, too much overhung weight. Note that the first critical is in sympathy with the pump operating speed, which becomes intolerable with the operating time limited to one to two days, due to bearing failures. The stiffness on antifriction bearings was determined from a program written by M. E. Leader of Monsanto, using values projected by an article written by F. F. Garguilo, DuPont.2 The correction consisted of converting the 62-lb (28-kg) coupling to a 15-lb (6.8-kg) series dry flex disk-type coupling and stiffening the support by flushing the baseplate cavity with a degreasing fluid and pressure injection of epoxy to fill the baseplate voids. It should be remembered that the blade passing frequencies will normally be the strongest exciting force. On this pump, the frequency is five times running speed (five vanes times each cutwater). Because there are two cutwaters, there can also be a frequency at 10 times running speed. The 5 frequency is shown on Figure 1. Also, this 5 frequency excitation could excite the second mode because the second mode critical could fall anywhere between the solid and dashed lines, depending on baseplate stiffness. The instruments used in diagnosing this problem were force-effective seismic sensors (velocity or piezoelectric accelerometers). They are preferred for pumps, particularly those with antifriction bearings. MULTISTAGE PUMP EXAMPLE _________________________________________ To show the mechanical rotor variations, a six-stage boiler-feed pump with a design capacity of 1250 gpm (284 m3/h), 2200 ft (670 m) total head, and driven by a 1000-hp (746-kW), two-pole motor has been selected. This pump utilizes interstage bushings as support bearings to the rotor. The contribution of these bushings as bearings will probably be less than might be assumed. 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.411 FIGURE 5 Synchronous critical speed summary with first three mode shapes of a single-stage overhung pump with heavy coupling of 62 lb (28 kg) and rigid foundation. Rotor weight = 157.4 lb (71.4 kg); rotor length = 30.3 in (77.0 cm); number of stations = 22; number of bearings = 2. (a) Rigid support, modes 1, 2, and 3 respectively: 4279, 15,865, and 64,970 cpm; (b) flexible support, modes 1, 2, and 3 respectively; 3580, 9339, and 35,565 cpm. (lb  0.454 = kg; lb/in  175 = N/m) Pressure or seal leakage control bushings contribute rotor support if they are long. The hot feedwater has very low viscosity and little damping. The bearing stiffness will be relative to the eccentricity ratio of the shaft in the bushings. An eccentricity ratio of unity (maximum) implies that the shaft is rubbing directly on its bushing. The impeller weight is increased by the water trapped in each impeller. Many pump manufacturers improperly list pump undamped critical speeds from dry pump data or calculations. The bushings, labyrinths, and wear rings all contribute to the actual critical speed. Also, bearing housing resonances are more common than expected. 2.412 CHAPTER TWO FIGURE 6 Undamped critical speed map of multistage boiler-feed pump with plain journal bearings (lb/in  175 = N/m) FIGURE 7 Rotor cross-section of multistage high-pressure boiler-feed pump with plain journal bearings. Rotor weight = 377.7 lb (171.3 kg), rotor length = 84.6 in (215 cm); number of stations = 47, number of bearings or bushings = 4 (in  2.54 = cm) Figures 6 to 9 illustrate it in the same fashion as the previous example the design audit of a steam-turbine-driven, 4  8  1021, six-stage, boiler-feed pump using hydrodynamic radial and thrust bearings. No problems were experienced with this pump. A more complete listing of data from analysis is shown with an added breakdown of the rotor model and the output of the first, second, and third resonant modes. The first critical (rotor resonance) is now a cylindrical mode and not a conical mode, as previously seen for the overhung impeller of the single-stage process pump. ALIGNMENT OF PUMPS AND DRIVERS __________________________________ Outside of serious unbalance of pump components, there is no single contributor of poor mechanical performance more significant than poor alignment. Incorrect alignment between a pump and its driver can cause • Extreme heat in couplings • Extreme wear in gear couplings and fatigue in dry element couplings 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.413 FIGURE 8 First, second, and third resonate animated mode shapes of multistage high-pressure boiler-feed pump with plain journal bearings. Rotor weight = 377.7 lb (171.3 kg); rotor length = 84.6 in (215 cm); number of stations = 47; number of bearings or bushings = 4. (a) Mode 1: frequency = 2614 cpm; (b) mode 2: frequency = 5223 cpm; (c) modes: frequency = 8134 rpm. (in  2.54 = cm) • Cracked shafts and totally failed shafts, with failure due to reverse bending fatigue transverse to the shaft axis initiating at the change of section between the large end of the coupling hub taper and the shaft • Preload on bearings (evident by an elliptical and flattened orbit resembling a deflated beach ball); pure asymmetry of vertical and horizontal vibration can be misleading because the bearing spring constants could vary greatly in the kyy (vertical) and the kxx (horizontal) axis. • Bearing failures plus thrust transmission through the coupling, which can be totally locked (axial vibration checks across the coupling; that is, at each adjacent machine, will generally confirm this condition) Significant changes in the cold nonrunning alignment of a pump and driver can take place if the temperature rise in each machine is different and if the piping imposes forces on the pump. 2.414 CHAPTER TWO FIGURE 9 First three critical speed mode shapes of multistage high-speed boiler-feed pump superimposed (in  2.54 = cm) Therefore, alignment under actual operating conditions must be predicted or, if unknown, confirmed by instrumentation. In either case, an allowance must be made in the initial cold alignment to compensate for changes in alignment from cold idle to hot running. There are several techniques for measuring cold and hot alignment. The cold alignment is generally measured by either face and rim (Figure 10) or reverse dial indicator (Figure 11) methods. The face and rim method has a sensitivity advantage when the diameter of a coupling exceeds the indicator span of reverse indicator bracket tooling. This is rare, as the pump will generally have a spacer coupling and the reach of the reverse indicators can be increased by clamping onto the shaft behind each coupling half. The face and rim method would also have an advantage if either the driver or the gear could not be rotated, as it seems unlikely that the pump could not be rotated. In order to compensate for the measuring surface’s not being circular or smooth, both shafts should be rotated together when using this method. Disadvantages of Face and Rim Method 1. Diameters of the rim must be true (circular) and smooth and the face reading surface must be flat and smooth, unless both shafts are rotated together. 2. The driver and pump cannot float axially while a reading is being taken or an error will be introduced into the face (angular) reading. (A fixed axial stop will assist in reducing errors.) Reverse Dial Procedure for Measuring Alignment (Hot or Cold) Several procedures have been suggested by various people to estimate or actually measure alignment while a pump is running at operating temperature. Some techniques are 1. Shutdown after temperatures have stabilized for “hot check” by dial indicators 2. Optical measurements cold to hot (A. J. Campbell, Compressor Engineering Corp., Houston) 3. Dodd bars (DynAlign) technique (B. Dodd, Chevron3) 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.415 Check for Angular Misalignment Dial indicator measures maximum longitudinal variation in hub spacing through 360° rotation. Check for Parallel Misalignment Dial indicator measures displacement of one shaft center line from the other. 1. Attach dial indicator to hub, as with a hose clamp; rotate 360° to locate point of minimum reading on dial; and then rotate body or face of indicator so the zero reading lines up with pointer. 4. Reset pointer to zero and repeat operations 1 and 2 when either driven unit or driver is moved during aligning trials. 2. Rotate both half couplings together 360°. Watch indicator for misalignment reading. 3. Driver and driven units will be lined up when dial indicator reading comes within maximum allowable variation for that coupling style. Refer to specific installation instruction sheet for the coupling being installed. Note: If both shafts cannot be rotated together, connect dial indicator to the shaft that is rotated. FIGURE 10 5. Check for parallel misalignment as shown. Move or shim units so parallel misalignment is brought within the maximum allowable variations for the coupling style. 6. Rotate couplings several revolutions to make sure no “end-wise creep” in connected shafts is measured. 7. Tighten all lockouts or capscrews. 8. Recheck and tighten all locknuts or capscrews after several hours of operation. Face and rim dial indicator method (Courtesy Rexnord) FIGURE 11 Model used for training machinist in the use of reverse dial indicator technique for alignment of machine shafts. Misalignment can be measured as parallel or angular offset. 2.416 CHAPTER TWO 4. Acculign bench mark gauges (I. Essinger, Shell Oil) 5. Water-cooled probe stands (C. Jackson, Monsanto) 6. Instrumented coupling (for example, Indikon) Refer to Reference 4 for more specific information on the above techniques. After measurements have been made and actual misalignment is determined for the hot running condition of the pump, the next step is to calculate the alignment correction required between machines to bring them into alignment during operation. Trial-anderror methods should be discouraged. Plotting actual shaft positions to scale allows one to graphically measure the required cold alignment corrections. Other methods—for example, those that can best be carried out with the aid of a programmable calculator—quickly and accurately calculate vertical, horizontal. inboard or outboard changes in machine positions to accomplish the correct cold alignment. See Reference 11 for a calculator program for both reverse dial and face and rim methods. Offered here are a graphical procedure and an example for obtaining the desired alignment between a pump and driver for a case where the thermal growth has been estimated and a calculated cold alignment is desired to achieve the final alignment during operation. The example is illustrated in Figure 12. Consider a steam-turbine-driven boiler-feed pump that has the following heat rise predictions by the manufacturers: steam end 0.002 in (0.051 mm), exhaust end 0.012 in (0.305 mm), pump inboard support 0.006 in FIGURE 12 Alignment example of a turbine-driven boiler-feed pump with heat-rise data from the manufacturers plotted to calculate the desired cold alignment 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.417 Actual alignment readings FIGURE 13 Desired dial indicator readings corrected for indicator bar sag FIGURE 14 FIGURE 15 Dial readings corrected to position zero at top FIGURE 16 Actual alignment readings corrected for indicator bar sag FIGURE 17 Plot of absolute vertical shaft positions. It can be seen from the graph that a 0.019-in (0.483-mm) shim is required to raise the inboard end of the pump and a 0.035-in (0.889-mm) shim is required to raise the outboard end of the pump to compensate properly for thermal growth (0.152 mm), and outboard bearing 0.004 in (0.102 mm). The horizontal length is laid on a graph with 1 div.  1 in (25.4 mm). The vertical movement plots are laid out with 1 div.  0.001 in (1 mil; 0.0254 mm). Based on 0.001-in (0.0254-mm) sag (see Figure 19), the field readings needed to meet the above absolute requirements are shown in Figure 13. The actual cold alignment is checked, and the reverse dial readings are as recorded in Figure 14. There are 0.125-in (3.175-mm) shims under all support feet. It is decided to move the pump rather than the turbine, and therefore the correct cold position of the pump must be calculated. To correct the turbine dial readings to position zero at the top, simply add ⫺12 to all four turbine readings (Figure 15). To correct for sag (see Figure 19 for explanation), subtract 1 from the left and right readings and 2 from the bottom reading (Figure 16). Finally, plot the absolute shaft positions on graph paper, leaving the turbine “in place,” so to speak, thereby determining two points across the 16-in (40.64-cm) indicator span to define where the pump shaft lies with respect to the turbine (Figure 17). To plot the horizontal corrections, reduce the final horizontal readings only to zero on the least numerical reading, by adding 4 to the turbine readings and ⫺5 to the pump reading, as shown in Figure 18. Dial indicator readings can be in metric units and the scale in centimeters rather than inches. A scale of between 500:1 and 1000:1 is suggested. A 1000:1 scale is in use here (horizontal scale equals 1000 times vertical scale). 2.418 CHAPTER TWO FIGURE 18 Required horizontal corrections As shown in Figure 12, the driver and pump are purposely misaligned so actual operating temperatures will put the two shafts within acceptable limits. The acceptable limits for pump final alignment are 0.001 in/in (0.025 mm/mm) of coupling flex plane separation. If a spacer coupling is 5 in (13 cm) between flex planes on the flexible coupling, the shafts must be within 0.005 in (0.127 mm) vertical or horizontal offset. Pure angular misalignment in one plane is not desired, as it reduces the tolerance by 2:1 for gear couplings (dry coupling would have severe fatigue in one flex plane). The above limits should be reduced by one-half. INSTALLATION SUGGESTIONS AND USE OF DIAL INDICATORS _____________ 1. Nonferrous shim packs should be installed under all feet of the pump and driver, particularly when installing a new pump. The amount should be 0.125 to 0.250 in (3.175 to 6.35 mm) in no more than three pieces to start; for example, one 0.125-in (3.175 mm) and two 0.0625-in (1.59-mm) full shims of stainless steel. 2. Motors have four feet generally, and any “soft foot” should be compensated first. A soft foot is one that is shorter than the other two or three feet, a condition that puts a twist or strain in the equipment. Simply place a dial indicator stem vertically against the motor foot and release the hold-down bolts sequentially around the unit, recording and retightening at each step. If a 0.002-in (0.050-mm) spring-up occurs on three feet, for example, and 0.006 in (0.152 mm) occurs on the fourth foot, add 0.004 in (0.10 mm) of shim to the fourth foot, eliminating the soft foot. 3. Provide low-sag tooling to reach over the coupling (coupling left in place) for reverse indicator alignment. A 0.001- to 0.0015-in (0.025- to 0.038-mm) sag is easy to accomplish on indicator reach bars. 4. Let the indicator indicate on its own bracket or bracket pin, thus preventing any poor surface condition of shaft or coupling from contributing to poor measurements. 5. Support the dial indicator weight on the motor or pump shaft so it does not contribute to “reach bar” sag. 6. Do not overlook the fact that many times one can clamp to the shaft behind each coupling hub and obtain more span and therefore better accuracy. 7. Record all data looking the same way down the unit; that is, top east, bottom, west or top, north bottom, south or top, right bottom, left. It is suggested that the driver-pump always be viewed from the driver end. 8. Turn the shafts in the direction they normally turn and approach the 90° points in a precise manner (do not back up and introduce backlash errors). Turning in the normal direction is good training because, on gear units, it reduces helix angle lift errors. 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.419 FIGURE 19A through D Illustrative procedure to determine the amount of sag in an indicator bar (bracket) (Courtesy Reference 3) 9. If the motor can be turned down from the end opposite the end from which the measurements are taken, do so. Regardless, always release the strap wrench or spanner bar before recording each 14-point reading. 10. Obtain center zero dial indicators or revolution counter indicators or carefully note all indicator movements with a mirror to assure, for example, that 0.090 in was not really ⫺0.010 in. The algebraic sum of horizontal and vertical readings should be near equal. INSTRUMENTS FOR VIBRATION ANALYSIS_______________________________ One fact about end-suction and between-bearing pumps is that external visual evidence of mechanical problems is very limited. Only three gauges for mechanical trouble exist: temperature, vibration, and sound. 2.420 CHAPTER TWO FIGURE 20 Complex vibration signal resolves into sine wave spectrum It is normal for a machine to vibrate at some level; such vibrations are caused by manufacturing defects, design limits of the pump, casting irregularities, less than optimum application, and a maintenance/installation problem. When the velocity vibration level starts to increase 0.1 in/s (2.5 mm/s) zero to peak (0-P) above the “as new installed level,” the vibration should be analyzed to determine the possible sources of the mechanical and/or hydraulic problem. Several mechanical and/or hydraulic problems may be producing, for instance, the 1 running speed frequency vibration. The key in using vibration to define the mechanical and/or hydraulic problems is to determine the frequency at which the vibration occurs. Vibration amplitude is also an important factor because it indicates the severity of the vibration. Field vibration data are normally a complex vibration waveform. By using a tunable analyzer, the complex vibration signal, as shown in Figure 20, can be filtered or tuned into its basic frequency components; that is, all complex signals are summations of the harmonics and subharmonics 1, 0.5, 6, 30, and so on. By comparing these filtered components of the complexed vibration signal with an analysis chart and some common-sense experience, probable causes of the vibration can be listed. The first step toward resolving the vibration problem is to convert the mechanical movement to an equivalent electrical signal so it can be filtered and measured. Often used analyzer systems are 1. A turbine ac-battery-powered analyzer with strobe light, providing amplitude, frequency, and phase. A plotter accessory can also be attached for copies of the data. 2. A small battery-powered, internally driven, tunable analyzer with a built-in plotter using an accelerometer or velocity sensor. 3. A spectrum analyzer, ac powered, that receives the signal directly from a vibration transducer or the recorded signal from a battery-powered four-channel FM/AM cassette tape recorder. For startup or where the problem is tougher, one can add 1. Eight-channel FM tape recorder 2. Four-channel oscilloscope with blanking and time display 3. Tracking filter displaying revolutions per minute, amplitude, and phase, capable of tracking runup/rundown data 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.421 FIGURE 21 Limitations on machinery vibration analysis systems and transducers (mils  0.0254 = mm; in/s  25.4 = mm/s) (Reference 10) Provisions should be made for the use of all types of sensors, as there are advantages in each. As more complex problems continue to appear, tunable analyzers with a sensor are not just a requirement but a necessity in any maintenance reliability program. The choice of a displacement sensor (eddy current probe), velocity or seismic sensor, or an accelerometer depends on the frequency range to be analyzed and the type of pumping equipment. There is no one vibration sensor for all jobs. Of the three types of vibration measurements, acceleration and displacement are dependent on frequency and velocity is independent of frequency. Most engineers and technicians select a measurement that is independent of frequency for a datum to judge the general health of new and used pumps. With the exception of low-speed pumps and motors, 1750 rpm or less, unfiltered velocity and filtered velocity are used for most basic data. Figure 21 shows the frequency relationships (log) versus output (log) of three different measurement sensors with reference to a constant velocity of 0.3 in/s (7.6 mm/s). The figure gives an overview of present sensor limits and shows that each sensor is like a window through which portions of the frequency spectrum may be observed. The figure also shows that the accelerometer is the choice sensor at high frequency because it measures the square of the frequency. The advantage of displacement at low frequencies is due to its high output; the disadvantage of displacement at high frequencies is that the output signal will disappear into the background noise of most measuring systems. One should not confuse the measurement parameters (displacement, velocity, and acceleration) with the sensors (eddy current probes for displacement, velocity sensors, and accelerometers). The basic relationship of these measurement parameters with commonly used units are shown on a simple sine wave in Figure 22. Although the velocity sensor is not necessarily the best all-around type of sensor, it does have the advantage of high self-generating output (up to 1000 ft [300 m] of cable), can be mounted in any position, and is influenced only slightly (less than 5%) by transverse sensitivity (side forces). The disadvantages are that the output signal below 600 cpm is significantly nonlinear but can be corrected, the accuracy is limited at ⫾8% to 1000 Hz, and 2.422 CHAPTER TWO FIGURE 22 Basic relationship of measured parameters with a simple sinusoidal vibration the sensor will most likely have problems in one to two years when mounted in field applications where vibration is high, especially vane passing frequencies. The piezoelectric accelerometer is a very light and compact sensor that measures vibration using a mass mounted on a piezoelectric crystal. Its output is low and requires a charge amplifier in the lead even with very short leads. The accelerometer is small and can be mounted virtually anywhere; it has a 1 to 3% influence factor from transverse side forces. A good rule of thumb on the usable frequency range is one-fifth to one-third of the resonant frequency. The disadvantages are that the sensor is sensitive to mounting torque, although stud mount is the best method to mount accelerometers. A lot of data are produced, of which some may be the data from an excited accelerometer resonance or cable noise. An impedance matching device can be built into the accelerometer for use at temperatures below 250°F (120°C), and cable noise can be greatly reduced with the voltage and charge sensitivity greatly improved; for example, 100 mV/g and 50 to 100 pC/g (where g  number of accelerations of gravity). For higher temperatures, the accelerometer will need a separate charge amplifier and may need heat insulation, such as MICA wafers (refer to API 678, dated 1981). Most so-called ultrasonic analyzers use the accelerometer as a structural microphone. Many have chosen a carrier frequency in the megahertz range to improve the signal-tonoise ratio and make characteristic high-frequency patterns. Most of these systems are still in the development stage. Techniques for Taking Data The second most important part of a vibration analysis program is the type of data taken and the techniques used to take the data. The purpose of taking vibration data on a pump is to either perform an analysis because someone noticed a noise or increased vibration level, or as a part of a periodic preventative maintenance program. It has been proven from experience that the velocity measurement is the best method for determining acceptable levels of centrifugal pump vibration. This is not to say that displacement and acceleration are not measures of vibration severity; they are, but it is necessary to know the frequency of the vibration. Displacement is preferred by a few for frequencies less than 6000 cpm. Accelerometers matched with analyzers can be purchased with signal integration that will give reliable readings in velocity in the 3000- to 60,000-cpm range. For readings above 60,000 cpm, the sensor would generally be an accelerometer reading in g’s, peak or integrated to read velocity zero to peak or 0-P. Vibration amplitude is an important parameter because it indicates the approximate severity of the dynamic stress levels in the pump. Experience has shown that the shaft 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.423 bearings and seal will probably fail in a pump with a velocity reading of 0.5 in/s (13 mm/s) 0-P. Also, catastrophic failures will probably occur when a pump is at 1 in/s (2.5 mm/s) 0-P. Pumps with velocity readings of 0.05 to 0.15 in/s (1.3 to 3.8 mm/s) 0-P, will perform well mechanically. Vibration readings are taken in the horizontal and vertical planes on the bearing housing of horizontal-shaft pumps. To have a worthwhile maintenance reliability program with pumps, vibration readings must be recorded regularly (that is, monthly). This can range from a trend plotting of unfiltered vibration to a full vibration analysis using a real-time analyzer to generate the frequency spectrum. A standard method used by many companies consists of taping pump vibrations with a battery-powered cassette recorder using a velocity sensor. Readings can then be processed through a real-time analyzer and recorded on an XYY¿ plotter. The best application of this method is during startup and repair evaluation. As an alternate method, a spectrum analyzer/plotter that produces a spectrum on a 4 in  6 in (l0 mm  15 mm) card with a frequency plot versus amplitude can be used. This procedure has in some installations detected and corrected 95% of the mechanical problems before failure. Experience has shown that had unfiltered displacement readings been taken, only 60 to 70% of the mechanical problems would have been observed. During these recordings, emphasis should be placed on the change in vibration levels, which is a better indication of a mechanical problem than absolute vibration. One of the best pieces of data available for the pump’s equipment file is a vibration record taken during the manufacturer’s test or during water batching or commissioning. It is advisable to request a witness performance test on key or critical pumps. The purpose of this test is to assure mechanical reliability along with performance. The manufacturer should be asked about the availability and type of vibration analysis equipment and sensors. Regardless of the instruments used, the vibration data sheet for the tested pump should have a sketch of where all vibration points were taken. The manufacturer should also supply a complete mechanical description of the number of impeller vanes, number of casing volute cutwaters or diffuser vanes, type of coupling, length of coupling spacer, and so on. There are several different methods for taking periodic vibration data on pumps: 1. Using a handheld battery-powered velocity probe/readout, a machinist or operator logs unfiltered readings taken at one or two points on the bearing housing. When the reading reaches 0.3 to 0.5 in/s (8 to 13 mm/s) 0-P, the pump is pulled for maintenance. Readings are usually taken every two weeks. 2. Vibration points in the vertical, horizontal, and axial directions are recorded on a tabulated chart in unfiltered and filtered velocity at the various peak amplitudes, using a battery-powered tunable analyzer with a velocity sensor. 3. Vibrations in the vertical, horizontal, and axial direction are taken at each bearing, using a velocity sensor. The signal is recorded on a tape recorder, preferably a batterypowered FM/AM cassette. These data are then processed through a real-time analyzer. A spectrum hard copy is made on an XYY¿ plotter of velocity versus frequency. 4. Key vibration points are fed directly from a velocity sensor or an accelerometer/charge amplifier through a long extension cable to a safe area, where a real-time analyzer processes the signal into a velocity versus frequency spectrum or a g’s (acceleration) versus frequency spectrum. Hard copies for records are made on a XYY¿ plotter. This method requires two technicians with radios. The most accurate are methods 3 and 4. The most costly to run in workerhours per point is method 2. The least accurate is, of course, method 1, but it is a popular screening technique. Use of Vibration Sensors The use of a handheld velocity sensor with an aluminum extension rod or a light-duty vise grip with the probe mounted on the top of the grip has produced some high and misleading vibration readings because of extension resonances. For instance, the vise grip should not be used because of a 5000-cpm resonance. A 9-in (23 cm) long by 38-in (0.95-cm) diameter extension to the velocity pickup should not be used above 16,000 cpm. The approximate axial natural frequency in cycles per minute for a rod extension from the probe, in tension and compression, can be expressed as 2.424 CHAPTER TWO in USCS units in SI units where W L A E AE fn  188 B WL AE fn  946 B WL  pickup weight (force), lb (N)  length of rod, in (m)  cross-sectional area of rod, in2 (m2)  modulus of elasticity of rod,° lb/in2 (kPa) Use of attachments above these listed frequencies will produce a higher amplitude. The best and simplest method of holding a velocity probe to a pump is a two-bar magnetic holder on the end of a velocity probe. Proper cleaning and some paint removal are generally necessary for good attachment. Periodic wiping of the magnetic bar to remove iron filings is also necessary. After mounting the probe, give it a light twist and a rocking motion; if it twists easily or rocks, change locations or reclean the surface. This location should be marked and future readings taken on the same spot; otherwise, the trend plots will vary. If there is a concern about an extension or magnetic holder resonance, test this by holding the sensor, without the extension or magnetic holder, directly to a reasonably flat spot and noting any differences. Holding the sensor on a flat spot is generally safe up to 60,000 cpm. When measuring vibration on an electric motor, there is always the possibility of false readings at 60 and/or 120 Hz due to electrical induction by the motor. This can be checked by two methods: 1. Hold the sensor by its cord and move it toward the motor, noting any increase in amplitude. 2. Using a two-channel oscilloscope, trigger the filtered signal against line voltage. Inphase signals mean the vibration is electrically induced. For field use, one usually does not have to contend with temperatures above 250°F (120°C) direct to the sensor; thus, accelerometers with built-in impedance matching devices can be used and 100 mV/g voltage sensitivities can be obtained and transmitted 300 ft (90 m) if necessary. If the frequency range is low, and it is for pumps, a charge sensitivity in the order of 100 pC/g can also be obtained. Techniques for Taking Preliminary Vibration Readings Some key points to remember before you start your analysis of the vibration problem: 1. Do not reach a decision on what the problem is before you record and analyze the data. By deciding too quickly what the problem is, you will most likely neglect other important factors. 2. Before you take the data, take time to review maintenance logs, talk with the area mechanic and operator, and make notes on the following: a. Are there any unusual sounds (cavitation, bearings, and so on)? b. Is there any movement in the discharge pressure gage? c. What is the direction of rotation? d. Are the flush and cooling lines lined up properly? e. Is there any movement in the coupling shim pack? f. Are there any foundation cracks? g. Are pipe supports functioning properly? h. Has a suction screen been installed? °For aluminumm, E  10.3  106 lb/in2 (71  106 kPa). For steel, E  30  106 lb/in2 (207  106 kPa) 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.425 What is the magnitude of the liquid velocity in the suction line? Is the automatic oiler level adjusted correctly? Where is the pump being operated? What are the flow, suction, and discharge pressure? Is the pump’s minimum flow bypass system in service? Has the process changed? What is the suction valve stem orientation? Have there been any color changes in the paint? Are there any loose parts, including the coupling guard? Determine the color and feel of the oil (if possible). What is the bearing housing temperature? How is the coupling guard attached (attachment to the bearing housing is poor practice)? You will be surprised how much this information aids in an analysis. Example: you record a high 1 in the radial plane, low values of 2 and 5, and several high-frequency components at about 0.15 in/s (4 mm/s) 0-P. The 1 could be a bent shaft, loose coupling, plugged impeller, bad coupling unbalance, upper and lower case halves misaligned, and a whole list of running frequencies symptoms. During your review of maintenance logs, you noted that the impeller had been replaced because there had been distillation column tray part damage. The next questions you should ask are, “Did maintenance reinstall the suction screen?” (if not, another tray part may have lodged in the impeller) and, “Was the impeller rebalanced after it was trimmed from maximum diameter?” 3. Do not try to interpret partial vibration readings for someone looking over your shoulder before you have even taken all the readings. Sit down in a quiet place with your notes on installation and maintenance, a symptoms list, and a severity chart and then make the analysis. Analysis is not a simple task, but with some experience you will build confidence and it will become second nature. i. j. k. l. m. n. o. p. q. r. s. t. VIBRATION DIAGNOSTICS _____________________________________________ Analysis Symptoms The vibration severity chart and vibration identification chart are guides, but experience, a set of procedures, and study of the literature will make diagnostics easier. To be effective, one must be thoroughly familiar with the machinery’s internal construction, installation, and basic control system. Both mechanical and hydraulic mechanisms can produce symptoms of vibration. The vibration analysis symptoms, or vibration severity criteria, chart has taken on many forms since the Rathbone chart of 1939. Perhaps the most widely used symptoms chart in the turbo-machinery field today is the original paper published by Sohre.5 A condensation and revision of the original paper is shown in Table 2. Although this chart includes some symptoms that will never appear in pumps, it is one of the better references for vibration analysis. The way the table gives percentages of cases showing the symptoms for the causes listed is unique. As one learns to use the chart and modifies it with experience, a good diagnostic tool will be developed. A good guide for unfiltered bearing cap velocity limits on field installed pumps is given in Table 3. A guide for shop testing new and rebuilt pumps is given in Reference 6. Comments on Table 2 In the following comments, the numbers correspond to “Cause of vibration” in Table 2. 1. Long, high-speed rotors often require field balancing at full speed to make adjustments for rotor deflection and final support conditions. Corrections can be made at balancing rings or at coupling bolts. TABLE 2 Vibration analysis symptoms Numbers indicate percent of cases showing previous symptoms, for causes listed in vertical column at left. Source: The Practical Vibration Primer by Charles Jackson. Copyright (c) 1979 by Gulf Publishing Company, Houston, Texas. Used with permission. All rights reserved. TABLE 2 Continued. 2.427 TABLE 2 Continued. 2.428 TABLE 2 Continued. 2.429 2.430 CHAPTER TWO TABLE 3 Bearing cap data-velocity unfiltered Smooth 0.1 in/s (p) and less Acceptable Marginal Planned shutdown for repairs 0.1—0.2 in/s (p) 0.2—0.3 in/s (p) 0.3—0.5 in/s (p) Immediate shutdown 0.5 in/s (p) Note: For gearing, add 0.1 in/s to all values. p  peak mm/s  25.4  in/s 2. Bent rotors can sometimes be straightened by the “hot-spot” procedure, but this should be regarded as a temporary solution because bow will come back in time. Several rotor failures have resulted from this practice. If blades or disks have failed, check for corrosion fatigue, stress corrosion, resonance, off-design operation. 3. Straighten bow slowly, running on turning gear or at low speed. If rubbing occurs, trip unit immediately and keep the rotor turning 90° using a shaft wrench every 5 minutes until the rub clears; resume slow run. This may take 12 to 24 h. 4. Often requires complete rework or new case, but sometimes a mild distortion corrects itself with time (requires periodic internal and external realignment). Usually caused by excessive piping forces or thermal shock. 5. Usually caused by poor mat under the foundation or thermal stress (hot spots) or unequal shrinkage. May require extensive and costly repairs. 6. Slight rub may clear, but trip the unit immediately if a high-speed rub gets worse. Turn by hand until clear. 7. Unless thrust bearing has failed, this is caused by rapid changes of load and temperature. Machine should be opened and inspected. 8. Usually caused by excessive pipe strain or inadequate mounting and foundation, but is sometimes caused by local heat from pipes or the sun’s heating the base and foundation. 9. Most trouble is caused by poor pipe supports (should use spring hangers), improperly used expansion joints, and poor pipe line up at casing connections. Foundation setting can also cause severe strain. 10. Bearings may become distorted from heat. Make a hot check, if possible, observing contact. 11. Watch for brown discoloration, which often precedes recurring failures. This indicates very high local oil film temperatures. Check rotor for vibration. Check bearing design and hot clearances. Check condition of oil, especially viscosity. 12. Check clearances and roundness of journal, as well as contact and tight bearing fit in the case. Watch out for vibration transmission from other sources and check the frequency. May require antiwhirl bearings or tilting-shoe bearings. Check especially for resonances at whirl frequency (or multiples) in foundation and piping. 13. This can excite resonances and criticals and combinations thereof at two times running frequency. Usually difficult to field balance because, when horizontal vibration improves, vertical vibration gets worse and vice versa. It may be necessary to increase horizontal bearing support stiffness (or mass) if the problem is severe. 14. Usually the result of slugging the machine with fluid, solids built up on rotor, or offdesign operation (especially surging). 15. The frequency at rotor support critical is characteristic. Disks and sleeves may have lost their interference fit by rapid temperature changes. Parts usually are not loose at standstill. 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.431 16. It is often confused with oil whirl because the characteristics are essentially the same. Before suspecting any whirl, make sure everything in the bearing assembly is absolutely tight with an interference fit. 17. This should always be checked. 18. It usually involves shading pedestals and casing feet. Check for friction, proper clearance, and piping strains. 19. To obtain frequencies, tape a microphone to the gear case and record noise on magnetic tape. 20. Loose coupling sleeves are notorious troublemakers, especially in conjunction with long, heavy spacers. Check tooth fit by placing indicators on top, then lifting by hand or a jack and noting looseness (should not be more than 1—2 mils [0.025 to 0.05 mm] at standstill, at most). Use hollow coupling spacers. Make sure coupling hubs have at least 1 mil/in (1 mm/m) interference fit on shaft. Loose hubs have caused many shaft failures and serious vibration problems. 21. Try field balancing; more viscous oil (colder); larger, longer bearings with minimum clearance and tight fit; stiffen bearing supports and other structures between bearing and ground. This is basically a design problem. It may require additional stabilizing bearings or a solid coupling. It is difficult to correct in the field. With high-speed machines, adding mass at the bearing case helps considerably. 22. These are criticals of the spacer-teeth-overhang subsystem. Often encountered with long spacers. Make sure of tight-fitting teeth with a slight interference at standstill and make the spacer as light and stiff as possible (tubular). Consider using a solid or membrane coupling if the problem is severe. Check coupling balance. 23. Overhang criticals can be exceedingly troublesome. Long overhangs shift the nodal point of the rotor deflection line (free-free mode) toward the bearing, robbing the bearing of its damping capability. This can make critical speeds so rough it is impossible to pass through these speeds. Shorten the overhang or put in an outboard bearing for stabilization. 24. Casing resonance is also called case drumming. It can be very persistent but is sometimes harmless. The danger is that parts may come loose and fall into the machine. Also, rotor/casing interaction may be involved. Diaphragm drumming is serious because it can cause catastrophic failure of the diaphragm. 25. Local drumming is usually harmless, but major resonances, resulting in vibration of the entire case as a unit, are potentially dangerous because of possible rubs and component failures, as well as possible excitation of other vibrations. 26. Similar problems exist as in 24 and 25 with the added complications of settling, cracking, warping, and misalignment. This cause may also produce piping troubles and possible case warpage. Foundation resonance is serious and greatly reduces unit reliability. 27. Pressure pulsations can excite other vibrations with possible serious consequences. Eliminate such vibrations using restraints, flexible pipe supports, sway braces, shock absorbers, and so on, plus isolation of the foundation from piping, building, basement, and operating floor. 28. It occurs mostly at two times line frequency (7200 cpm), coming from motor and generator fields. Turn the fields off to verify the source. It is usually harmless, but if the foundation or other components (rotor critical or torsional) are resonant, the vibrations may be severe. There is a risk of catastrophic failure if there is a short circuit or other upsets. 29. This can excite serious vibrations or cause bearing failures. Isolate the piping and foundation and use shock absorbers and sway braces. 30. Valve vibration is rare but sometimes very violent. Such vibrations are aerodynamically excited. Change the valve shape to reduce turbulence and increase rigidity in the valve gear. Make sure the valve cannot spin. 31. The vibration is exactly one-half, one-quarter, one-eighth of the exciting frequency. It can be excited only in nonlinear systems; therefore, look for such things as looseness 2.432 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. CHAPTER TWO and aerodynamic or hydrodynamic excitations. It may involve rotor “shuttling.” If so, check the seal system, thrust clearances, couplings, and rotor-stator clearance effects. The vibrations are at two, three, and four times exciting frequency. The treatment is the same as for direct resonance: change the frequency and add damping. If the cause is intermittent, look into temperature variations. Usually the rotor must be rebuilt, but first try to increase stator damping, add larger bearings (tilting-shoe), increase stator mass and stiffness, and improve the foundation. This problem is usually caused by maloperation, such as quick temperature changes and fluid slugging. Use membrane-type coupling. This is basically a design problem, but is often aggravated by poor balancing and a poor foundation. Try to field-balance the rotor at operating speed, lower oil temperature, and use larger and tighter bearings. Add mass or change stiffness to shift the resonant frequency. Add damping. Reduce excitation and improve system isolation. Reducing mass or stiffness can leave the amplitude the same even if resonant frequency shifts because of stronger amplification. Check “mobility.” Stiffen the foundation or bearing structure. Add mass at the bearing, increase critical speed, or use tilting-shoe bearings (which is the best solution). First, check for loose fit of bearings in bearing case. Same comments as 36 with additional resonance of rotor, stator, foundation, piping, or external excitation; find the resonant members and the sources of excitation. Tiltingshoe bearings are the best. Check for loose bearings. Sometimes you can bear the squeal of a bearing or seal, but frequency is usually ultrasonic—very destructive. Check for rotor vanes hitting the stator, especially if clearances are smaller than the oil film thickness plus rotor deflection while passing through the critical speed. Usually accompanied by rocking motions and beating within clearances. It is serious especially in the bearing assembly. Frequencies are often below running frequency. Make sure everything is absolutely tight, with some interference. Line-in-line fits are usually not sufficient to positively prevent this type of problem. This problem is very destructive and difficult to find. The symptoms are gear noise, wear on the hack side of teeth, strong electrical noise or vibration, loose coupling bolts, and fretting corrosion under the coupling bolts. There is wear on both sides of coupling teeth and possibly torsional-fatigue cracks in keyway ends. The best solution is to install properly tuned torsional vibration dampers. It is similar to 40, but encountered only during startup and shutdown because of very strong torsional pulsations. It occurs in reciprocating machinery and synchronous motors. Check for torsional cracks. Impeller Unbalance Impeller unbalance appears as a 1 running speed frequency vibration approximately 90% of the time and may be mechanical or hydraulic in origin. Impeller mechanical unbalance is a frequent cause of mechanical seal and bearing failures. Many mechanics will never think of checking impeller balance until heavily pitted areas appear. Because of the nonhomogeneous nature of most castings, corrosion is usually more aggressive in one area of the impeller. The degree of etching or surface pitting is a judgmental indicator of balance change. Impeller balancing should be part of the shop repair procedure for impellers over 10 in (25 cm) at 3600 rpm. It is good practice, when balancing an impeller, to keep the impeller bore to balance mandrel fit no greater than 0.001 in (0.0254 mm) loose. Installing the impeller with the keyway up on the balance mandrel and pump shaft will help eliminate some of the unbalance due to shaft centerline shift. A 38-lb (17.2 kg), 15 12-in (39.4-cm) diameter impeller operating at 3600 rpm is balanced on a machine good to 25  10⫺6 in (635  10⫺6 mm) using an expanding manEXAMPLE 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.433 drel. The impeller is then installed on its shaft, which has a loose fit of 0.0035 in (0.0889 mm). The forces created by this shift in the center of mass is calculated as follows: in USCS units Unbalance  eccentricity of impeller 1in2  impeller wt. force 1oz2  0.0035  38  16  1.064 oz # in 2 Unbalance force  1.77 a  1.77 a rpm 2 b  unbalance 1oz # in2 1000 3600 2 b  1.064  24.4 lb 1000 in SI units Unbalance  eccentricity of impeller 1mm2  impeller wt, mass 1g2  0.0889  17.2  1000  765 g # mm 2 Unbalance force  0.01094 a  0.01094 a rpm 2 b  unbalance 1g # mm2 1000 3600 2 b  765  108.5 N 1000 The example also points out that the unbalance force generated by loose fit impellers with keyways mounted in one plane could be quite high. This force could be minimized by staggering keyways or randomly orienting the impellers on the balance mandrel. Shifting of shrink-fitted, well-balanced impellers on multistage and highspeed double-suction pumps after a period of operation can result in unbalance. The shifting of the impeller is due to the relaxation of residual stresses that built up as the impeller cooled and contracted around the shaft. Shaft vibration and flexing tend to relieve the residual stress and cause the impeller to cock or bow the shaft from the original balance centerline. Standards should be referred to for balancing pumps and their drivers. When balancing, consideration must be given to the need for balancing at rated speed in order to properly evaluate the importance of shaft deflection due to modal components of unbalance. See References 12 and 13. Hydraulic Unbalance Uneven flow distribution entering the impeller can cause a 1 running speed frequency type of vibration. The unbalance occurs because the flow is not equal in all vane passages. An example of this is a double-suction impeller with a short, straight run to the pump and an elbow in the horizontal plane. Flow from the elbow does not have time to straighten and therefore enters both sides of the impeller unequally. A similar condition results if suction is taken from a tee off the main header. Unequal and unsteady flow into the impeller may cause axial thrust and high axial vibrations. Thus, it is good design practice to install elbows vertically in double-suction pumps. In double-suction pumps, the nonsymmetrical positioning of the impeller or the offset of the upper case half of the lower case half will cause a 1 unbalance due to nonsymmetrical flow. Recirculation forces and pulsation recirculation within a pump (which can occur when flow is less than design) may manifest themselves in the form of a noise and/or vibration with random frequencies, along with pressure pulsation that may be seen on a pressure gage. Recirculation may also appear in the piping system as vibration and noise. Increased NPSHA has helped in a few cases, especially if the recirculation is mainly on the suction side of the impeller. After a pump has a recirculation problem in a given system and the 2.434 CHAPTER TWO flow cannot be increased using a bypass, little can be done to the pump itself unless the system characteristics allow an impeller change. Antifriction Bearings Vibrations generated by ball bearings cover a wide range of high frequencies that are not necessarily a multiple of the shaft running speed. The frequency readings obtained during analysis are somewhat unsteady because of the resolution of the filters in a hand-tuned analyzer. The amplitude reading may also be somewhat unsteady. Experience has shown that hand-tuned field analyzers tend to show the last stage of the bearing failure. Monitoring of stress waves or shock pulses (impact energy) on the pump bearing housing will show failure trends that will generally precede an increase in the detectable level of mechanical vibration. This method of failure detection is called acoustic high-frequency monitoring or incipient-failure detection (1FD). A comparison between conventional methods and the acoustic high-frequency method is shown in Figure 23. Accurate analysis of pump bearings and other machinery can also be made using a velocity sensor good to 1500 Hz or an accelerometer. Data should be recorded and processed through a real-time analyzer with at least 256-line resolution capability and a band selectable analysis option. Analysis of antifriction bearings using a real-time analyzer and equations for calculating frequencies generated by defective bearings can be found in Reference 7. Accurate and extended analysis of pump bearing vibration is generally not needed. A majority of bearing problems are currently identified by an acoustic noise during operation. What is needed by maintenance personnel is a quick and reliable method to monitor bearings and determine when a bearing is failing and the rate of bearing deterioration. At present, there are some expensive instruments that can be purchased, but none of them meet maintenance requirements. Currently, the best method for reducing bearings analysis is prevention of the failure. Most antifriction pump bearings fail for one of several reasons: (1) water gets into the oil, (2) automatic oiler is not adjusted properly (this cause is most often overlooked and will continue to produce a short life to failure cycle), (3) product gets into the oil, (4) acidic vapors condense and break down the oil, (5) mounting techniques or fits are improper, (6) new hearing is defective. The solution to high-humidity problems and problems with acidic units is the use of an oil mist lubrication system. If this cannot be economically installed, an aggressive preventative maintenance program on a monthly to every-othermonth basis is required. FIGURE 23 Relative signal strength versus days to failure for acoustic IFD and conventional vibration monitoring methods 2.3.3 CENTRIFUGAL PUMP MECHANICAL PERFORMANCE 2.435 Baseplates With the change from low rotating speeds and cast iron baseplates to the less rigid fabricated steel baseplates and higher rotating speeds, higher operating temperatures, and larger impellers have greatly increased the probability of baseplate vibrations, distortion, and a decreased stiffness for rotor dynamics. Reference 6 has increased coverage for piping loads and the option of a heavy-duty baseplate over the proposed standard baseplate. The reference specifies a standard baseplate and pedestal support that are twice as rigid as specified in the fifth edition of the reference standards. Baseplate vibration problems can be resolved at the design stage or during construction. Engineering specifications should call for leveling screws, grout filling holes [4 in (10 cm) minimum, 6 in (15 cm) preferred for each bulkhead section], venting holes [421 in (11 cm) holes to each bulkhead section], and corrosion protection. A check of the outline dimension approval drawing should also be made for proper grout hole placement (bulkhead and cross bracing must be shown on the drawing). Construction specifications should call for proper baseplate preparation before grouting.8 API pumps should have epoxy grout bonding the baseplate to the concrete foundation. After proper cure, the baseplate should be tapped for voids, especially between and under the pump centerline supports. EXAMPLE A high vertical vibration occurred on the coupling end bearing at vane passing frequency on a multistage volute pump. The vertical vibration was 2.5 times the horizontal; thus, a check of the bearing pedestal and baseplate was in order. The hammer test on the pan of the baseplate under the bearing pedestal showed complete lack of grout. Vertical vibration on the pan was 1.8 times the horizontal. Regrouting eliminated the pan vibration and reduced the vertical vibration to one-half of the horizontal. Lack of proper stiffness from the baseplate can and will lower some pump critical speeds into the operating range. When a complete analysis is being done on a problem pump, take several pedestal readings (top, middle, and bottom on the side and end) and several readings on the pan of the baseplate. The middle, bottom, and pan readings should show good attenuation. REFERENCES _______________________________________________________ 1. IRD Mechanalysis. “Methods of Vibration Analysis.” Technical paper no. 104-1975, IRD, Cleveland. 2. Garguilo, E. P. Jr. “A Simple Way To Estimate Bearing Stiffness.” Machine Design, July 24, 1980, p. 107. 3. Dodd, V. R. Total Alignment. Petroleum Publishing, Tulsa, 1975. 4. “Alignment Tutorium.” Proc. Tax. A&M Turbomachinery Sym., December 1980. 5. Sobre, J. S. “Operating Problems with High Speed Turbomachinery: Causes and Corrections,” ASME Petroleum Mechanical Engineering Conference, Dallas, Sep. 1968. 6. American Petroleum Institute. Centrifugal Pumps for General Refinery Services. API Standard 610, 6th ed. Washington, D.C., 1981. 7. Taylor, J. I. “An Update of Determination of Anti-Friction Bearing Condition by Spectral Analysis.” Vibration Institute, Machinery Vibrations Monitoring Analysis Seminar. New Orleans, April 1981. 8. Murray, M. G. Jr. “Better Pump Grouting.” Hydrocarbon Processing. February 1974. 9. Jackson, C. A Practical Vibration Primer. Gulf Publishing, Houston, 1979. 10. Mitchell, J. S. An Introduction to Machinery Analysis and Monitoring. 1st ed., PennWell Books, Tulsa, 1981. 11. Messina, J. P., and S. P. D’Alessio. “Align Pump Drives Faster by Well-Planned Procedure.” Power, March 1983, p. 107. 2.436 CHAPTER TWO 12. American National Standard. Procedures for Balancing Flexible Rotors, ANSI S2.42, New York, 1982. 13. American National Standard. Balance Quality of Rotating Rigid Bodies, ANSI S2.19, New York, 1975. FURTHER READING __________________________________________________ Bloch, H. P. “Improve Safety and Reliability of Pumps and Drivers.” Hydrocarbon Processing, May 1977, p. 213. Bussemaker, E. J. “Design Aspects of Baseplates for Oil and Petrochemical Industry Pumps.” IMechE Paper C45/81, Netherlands, 1981, p. 135 (English ed.). Jackson, C. “Alignment of Pumps.” Centrifugal Pump Engineering Seminar, ASME South Texas Section Professional Development, Sec. 9, Houston, 1979. Jackson, C. “Alignment of Rotating Equipment.” NPRA, MC-74-7, Houston, 1974. Sprinker, E. K., and F. M. Patterson. “Experimental Investigation of Critical Submergence for Vortexing in a Vertical Cylindrical Tank.” ASME Paper 69-FE-49, June 1969. Von Nimitz, W. W. “Dynamic Design Criteria for Reciprocating Compressor and Pump Installations.” Presented at the 27th Annual Petroleum Mechanical Engineering Conference, New Orleans, September 19, 1972. Wachel, J. C., and C. L. Bates. “Escape Piping Vibrations While Designing.” Hydrocarbon Processing, October 1976, p. 152. 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS HORACE J. MAXWELL DAVID A. KALIX In designing centrifugal pumps, engineers strive to develop specific internal geometry that will produce head and flow with low energy loss. Each pump is designed for a specific head versus flowrate for the given impeller speed. The head and flowrate on this family of “characteristic” curves where the energy loss is minimum is known as the Best Efficiency Point (BEP). In application, pumps spend a significant portion, if not all of their life, operating at conditions other than BEP. This is normal and is due to a combination of system design conditions (for example, static head, piping and valve impedances, and so on), available pump designs (that is, capacities and heads) and actual plant operating requirements. Low flow related problems occur when pumps continuously operate in or repetitively cycle into flow regions that are significantly below the BEP (for example, 50% of BEP). In these low-flow regions, pump and system component performance and longevity can be adversely affected. Low flow problems are known to be worst for large high-energy pumps (for example, boiler feedwater), for pumps handling hot liquids, for pumps that handle liquids that have solid particles, and for pumps for low net positive suction head (NPSH) service. A welldesigned minimum flow control system can establish an environment that will substantially improve pump and system performance. The following chapter will briefly review the topics that should be considered when designing a minimum flow system for a centrifugal pump. FACTORS AFFECTING LOW FLOW RATE PUMP OPERATION________________ Collectively, the following broadly classed factors have been recognized as the major contributors to low flow pump problems. 2.437 2.438 CHAPTER 2 Thermal Factors The increase in temperature of the liquid within the pump is directly related to the pump’s efficiency. The energy that is available to heat the flowing liquid and the pump casing is basically the difference between the power input to the pump (brake horsepower) and the useful work done by the pump (liquid horsepower). At low flow conditions, centrifugal pumps are very inefficient and a significant amount of input energy is lost and heats the liquid and the pump assembly. Refer to Subsection 2.3.1 and Chapter 12 for more discussion on thermal effects. Hydraulic Instabilities When the pump is operating significantly below the BEP, flow streamlines (that is, patterns) within the pump change considerably from the rated design streamlines. Fluid eddies are most likely to develop at the inlet and discharge of the impeller resulting in flashing, cavitation, and shock waves that often produce vibration and serious component erosion. This phenomenon is classically known as internal recirculation. It can occur at the pump inlet (suction) and discharge. Refer to Subsection 2.3.1 and 2.3.2 for a more in-depth discussion on this topic. Mechanical Loads As the flowrate through the pump decreases, steady state loads increase and superimposed dynamic cyclic loads appear radially and axially on the impeller and shaft. The dynamic cyclic component increases significantly when recirculation within the pump occurs. Bearing damage, shaft and impeller breakage, and rubbing wear on casing, impeller and wear rings can occur. See Subsections 2.3.2 and 2.3.3 for discussion of this subject. Axial-flow and mixed-flow pumps with high specific speed produce comparatively higher head and take comparatively more power at low flow. A bypass system may be necessary not only to reduce component loading and stress but also to prevent motor overload. See Subsection 2.3.1 and Section 8.1 for discussion. Abrasive Fluids Liquids containing a large amount of abrasive particles, such as sand or ash, must flow continuously through the pump. If flow decreases, the particles can circulate inside the pump passages and quickly erode the impeller casing, wear rings, and shaft. ESTABLISHING MINIMUM PUMP FLOW REQUIREMENTS ___________________ The bypass system designer must know the minimum pump flow specified by the pump manufacturer in order to properly design a bypass system. The four previously discussed topics should be evaluated in detail by the pump manufacturer to establish the minimum flowrate specification. Minimum flowrate specifications are generally established through a combination of analytical and experimental techniques coupled with field performance data. Thermal Considerations The maximum allowable temperature rise of the pump is primarily based on two points: the permissible pump casing and shaft thermal growth and the flash point temperature of the pumpage. Pump manufacturers use analytical, laboratory and field data to validate their thermal analysis to ensure that pumps do not seize within the allowable temperature operating ranges. Refer to Subsection 2.3.1 for temperature rise calculations. Applications involving extremely high or low temperature fluids may require more indepth analysis to determine if individual component thermal growth is the limiting factor in determining minimum flowrate. Additionally, certain chemicals, which polymerize or solidify at particular temperatures, may establish the minimum flowrate specification. Chapter 12 and Subsection 2.3.1 provide more detail on this subject. Hydraulic Considerations The minimum bypass flow requirement for most pumps is based on minimum continuous stable flowrate, a hydraulic criterion, rather than a temperature rise. Pump internal recirculation will occur at both the impeller inlet and 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS 2.439 impeller outlet as flowrates are reduced. Internal recirculation will occur at flowrates well above those that cause temperature concerns. Refer to Subsection 2.3.1 for a detailed evaluation of this topic. Mechanical Considerations It is necessary to know how head, radial thrust, axial thrust and power vary with capacity before deciding on minimum allowable flow. Bearing capacity, motor rating, and stresses in drive and driven components are important influences. Abrasive Wear Considerations Relatively high bypass flowrates may be required to protect the pump against abrasives in the liquid. Heavy wear can occur at flows below 85% of the best efficiency point. The designer must establish the minimum pump flow specification using the pump manufacturer’s recommendation and his experience with comparable pumps and liquid/solid mixtures. MINIMUM FLOW CONTROL SYSTEM DESIGN FACTORS ____________________ Pump Size Capacity, power, specific speed, and suction specific speed are all factors that must be examined when designing a bypass system. These factors have a direct impact on the cost of building and operating the bypass system. Use of a continuous bypass system will require an even larger pump and driver to supply both the process and bypass flow requirements simultaneously. Discharge Pressure High discharge pressures result in high head loss in bypass valves, components, and lines. Liquids that can flash and cavitate demand special precautions to minimize damage in valving, orifices, and piping. Available Heat Sink Bypass flow must be reintroduced into the system far enough upstream to prevent progressive temperature buildup or flow disturbance in the pump suction. This may mean a simple discharge back to an uninsulated inlet line or discharge to a receiving tank or cooler with enough area and enough inflow of cool liquid to handle the thermal load. Bypass flow can discharge into a deaerator storage tank, a condenser, a flash tank, or a cooling pond. Elevation, distance, and pressure inside the receiving tank are also factors, as is the fact that the interior must be available for inspection and for repair of spargers, spray or distribution pipes, orifices, and backpressure regulators. Pump Design A pump’s design and materials of construction often affect the minimum allowable percentage of flow. With thermal effects, pumps vary in the length of time that they will tolerate shut off or low flow. This is important in designing the bypass system valves, instrumentation, and controls. Hydraulic effects at low flow are most apparent in high-energy pumps. The pump manufacturer should state the continuous minimum hydraulically acceptable flow for a given pump—and how it was determined. With axial-flow pumps, the shape of the pump head curve may be a factor in selecting the required bypass flow percentage. If feasible, a witness shop test of the pump should be specified to demonstrate and verify minimum flow recommendations. Liquid Pumped Liquids that flash and cavitate generally required a high bypass flow percentage. Examples are liquids near the boiling point or at high pressure. Abrasives in the liquid may require more bypass than would be needed for thermal reasons alone. Energy Costs A high energy cost to operate the pump requires careful consideration of bypass system design. The evaluation should compare equipment installation costs, maintenance, and energy costs for various bypass configurations. Figure 1 shows the annual pumping costs for a continuous bypass type system based on bypass flow, pressure and energy rates. The example shows a pump with a discharge head of 500 ft (152 2.440 CHAPTER 2 FIGURE 1 Annual pumping cost estimate for continuous bypass systems (metric conversions: ft  0.3048 = m, gpm  0.277 = m3/h, hp  0.746 = W) m) and a bypass flow requirement of 400 gpm (91 m3/h). Based on an energy cost of 5.0 cents/KWH, the annual cost for continuous bypass is $24,000. An automatic bypass system will only open the bypass when process flow demand is low. The total design life energy costs of a continuous bypass system can easily exceed the hardware costs of an automatic bypass system. Noise Considerations Bypass systems can easily exceed OSHA requirements for occupied spaces if not designed properly. High pressure drops and high fluid velocity increase noise. Multi-stage pressure reduction, heavy wall pipe, insulation, and silencers will all combine to reduce noise to acceptable levels. See Section 8.4 for a further discussion of this topic. Process System Design and Operational Expectations Bypass system design will depend on the plant design life and the expected process operational requirements. For example, a swing-loaded electric power plant will have far different pump operating requirements than a low-pressure emergency fire water system. The bypass system designer must evaluate installation, maintenance, and operating costs for the life of the process system with the expected utilization. If the actual operating conditions differ significantly from the design, the configuration of the bypass system should be reevaluated. For example, a process designed for normal operation with one pump at 75% of maximum capacity may put severe demands on the bypass system if the pump is operated continuously at 20% of maximum capacity. 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS 2.441 LOW FLOW PROTECTION SYSTEMS ____________________________________ Bypass Systems—Types/Design Considerations As the title implies, continuous bypass systems provide continuous flow whenever the pump is running, regardless of the process demand. Figure 2 illustrates a simple system, with a bypass line branching off the pump discharge upstream of the main line check valve and containing a fixed orifice dimensioned by analysis to provide minimum required pump flow. The bypass line discharges into a reservoir that is at a lower pressure than the pump discharge. The bypass line can also discharge directly into the pump supply line. However, the piping system design must ensure that the bypass liquid temperature does not increase to an unacceptable level. Additionally, vapor bubbles formed in the bypass by the pressure reduction process may be introduced into the pump. This will affect pump performance and longevity. Locating the bypass branch-off before the discharge check valve as shown keeps backflow from the process or from a parallel operating pump from going back to the receiving tank or back through the pump during a pump shutdown. Consideration should also be given to installing a check valve in the bypass line. The size of the bypass pipe depends on flow and piping configuration. If the pressure drop through the orifice results in flashing flow, the orifice should be located at the end of the bypass piping and should discharge directly into the larger receiving tank or larger downstream piping. The discharge flow should be directed so it does not impinge on the walls of the receiving tank but rather into the liquid to absorb the force of cavitation implosion. Valves and fittings located immediately downstream of the orifice may be damaged by cavitation. Full ported gate or ball valves are less susceptible to damage. Waterhammer and erosion are not problems in well-designed continuous bypass systems. The pump and prime mover must be sized to simultaneously supply both the bypass flow and the maximum process flow at the required pump discharge pressure. CONTINUOUS BYPASS SYSTEMS Automatic recirculation systems control the bypass flow in relation to the process flow. The sum of the process and bypass flowrates will always exceed the minimum flowrate specified to protect the pump. Two essential elements of these systems are a device to measure the process flow and a device to control the bypass flow. Compared with continuous bypass, these systems reduce energy consumption and pump horsepower requirements. The bypass flow can be regulated by either “Modulated” or “On-Off” control. When bypass flow is “Modulated,” the bypass flow is inversely proportional to the process flow. AUTOMATIC BYPASS SYSTEMS FIGURE 2 Typical continuous bypass system 2.442 CHAPTER 2 FIGURE 3 Comparison of On-Off versus Modulated automatic bypass flow control. The bypass is shut when process flow is greater than the control setpoint (that is, minimum flowrate required), opens more as the process flow is reduced, and is full open when the process flow is zero. With “On-Off” bypass control, the bypass is open fully when process flow is below the setpoint and is shut when the process flow is above the setpoint. Figure 3 illustrates these control methods and the effect on total pump flow. The total pump flow is the sum of the process and bypass flows; it is the flow that enters the pump suction. The “On-Off” control graph shows that there is a step change in the total flow when the bypass opens or closes. The “Modulated” control regulates the total flow smoothly when the process flow is below the setpoint. Either method will protect the pump from damage due to low flow. However, “On-Off” control may produce large pump discharge pressure variations depending on the shape of pump head curve. Rapid pump flow changes may produce damaging hydraulic or mechanical shocks. This will reduce pump life and increase maintenance. Modulated bypass control is required when the minimum flowrate specification exceeds approximately forty percent of the rated pump capacity. If “On-Off” bypass control is used above this point, the pump may exceed its capacity rating with resultant pressure pulsation and unsteady flow. A typical automatic bypass system, based on an instrumented flow control loop, is shown in Figure 4. This system requires a flow meter, a check valve, a bypass control valve, a valve controller and a pressure-reducing orifice. The flow meter can be placed at the pump inlet or discharge before the bypass tee (as opposed to it having to be placed after the bypass tee in an “On-Off” system). The controller compares the flow meter signal with the setpoint (minimum pump flow required) and operates the bypass control valve. Single stage control valves are commonly used in low pressure systems with an orifice for pressure reduction located downstream. However, these are not suitable for high pressure differentials or for the controlling of cavitation or flashing. Pressure reducing valves with multi-stage trim, good seat tightness, and low pressure recovery characteristics are widely used in this service. Refer to Chapter 7 for a detailed discussion of control valve characteristics and selection. CONTROL LOOP SYSTEMS 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS 2.443 FIGURE 4 Typical automatic bypass system based on flow control FIGURE 5 Typical automatic bypass system based on pump differential head The bypass pressure reducing orifice handles the high velocity and erosive forces of the bypass flow. However, the handling of cavitation and flashing by this fixed restriction is limited. The orifice is sized to operate properly at one bypass flow condition. At low flows, the orifice does not provide the correct pressure reduction as designed. The check valve is placed in the main line, as shown in Figure 4, to prevent reverse flow through the pump. The pump and motor bearings can be damaged if rotated in reverse. Pressure drop through the check valve must be considered in pump sizing in addition to other piping and valve pressure losses. A typical automatic bypass system based on pump differential head is shown in Figure 5. This system is identical to the flow controlled bypass system except that pump differential head is measured instead of flow. This system can only be used with continuously rising pump head characteristic curves. The system will have to be adjusted for each individual pump characteristic and later adjusted if this characteristic changes over the life of the pump. 2.444 CHAPTER 2 Depending on the bypass valve and control instrumentation selected, these systems can provide either “Modulating” or “On-Off” bypass control. Valves, instrumentation, and controls must be individually sized for the service conditions. All components must be integrated together to provide bypass flow control that meets the pump and process operational design criteria. The Automatic Recirculation Control (ARC ®) Valve This valve provides bypass flow control, pressure reduction, and reverse flow pump protection all within a single unit. This single valve combines the functions of the check valve, pressure reducing orifice, pipe tee, control instrumentation flow meter, and bypass control valve that are all required in the instrumented control loop bypass system. Valves can be designed to provide either “Modulated” or “On-Off” bypass control. The operation of all ARC valves is fundamentally the same. The flow sensing element (check valve disc) responds to the process flow demand and opens or closes the self-contained bypass control valve. Differences in ARC valve design center around bypass valve actuation method, repairability, serviceability, and adjustability. Figure 6 illustrates a typical ARC valve bypass system. Pilotless Trim Design The simplest of ARC valve designs operate the bypass flow control valve directly by the flow sensing element. The construction of a typical ARC for low pressure service is shown in Figure 7. The basic operation of this valve is shown in Figure 8. At zero main flow (“A”), the bypass provides full recirculation flow. As main flow increases (“B”), the bypass modulates proportionally closed. At full main flow (“C”), the bypass is shut. The pressure reduction from the pump discharge to the bypass is accomplished by the characterized orifices in the bypass element that is attached to the disc. However, the differential pressure from inlet to bypass imposes static and dynamic forces on the disc assembly. These unbalanced forces interfere with the normal disc motion and limit the maximum differential pressure for pilotless trim. There is an upper differential pressure limit at which this design will not operate satisfactorily and pilot operated bypass control must be used. Pilot Operated Trim Pilot operated bypass control valves utilize a combined hydraulicmechanical force to control the bypass flow. Only a relatively small mechanical force is necessary to hydraulically generate the large forces required to operate a bypass flow control valve in a high-pressure system. Figure 9 illustrates a pilot-operated bypass that utilizes a lever connected to the check valve disc to actuate the pilot. Figure 10 illustrates a pilot operated multistage bypass operated by an in-line pilot valve. Pilot-operated valves require more components and seals than unbalanced valves and are therefore more costly. FIGURE 6 Typical ARC valve bypass system 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS 2.445 FIGURE 7 Construction of a typical ARC valve for low-pressure service FIGURE 8A through C Basic ARC valve operation at three process flow conditions Serviceability ARC valves can be designed for in-line serviceability. Access is provided to the valves internal components for inspection or repair without disturbing the process or bypass piping. Figure 11 illustrates a valve designed to be serviced in-line. SAFETY SHUTDOWN SYSTEMS ________________________________________ These controls are normally provided as backup pump protection for the continuous bypass and automatic bypass control systems. They should only operate if the normal system fails due to mechanical or operator causes. In each case, the protection system trips the pump off. 2.446 CHAPTER 2 FIGURE 9 Operation of a typical pilot-operated ARC bypass valve actuated through a lever mechanism Head This requires only an accurate pressure switch, set for a discharge pressure corresponding to the total pump head when the pump is close to the shutoff capacity. For conservative results, calculate based on minimum suction pressure. To use this method, the pump characteristic must rise continuously to shutoff. Control circuitry may require delays to reject spurious pressure surges. Temperature A temperature sensor mounted on the pump casing or in the discharge and close to the pump can signal for a shutdown of all or some of the operating pumps. The trip can be on either a high temperature sensing or a high differential temperature sensing between inlet and discharge. This method can protect against large temperature rises (occurring at perhaps 5% or less of capacity) but is less sensitive to smaller rises at higher percentages of capacity where adverse hydraulic effects begin for large and highpower pumps. Low-Flow Systems These protection systems can be configured in various ways. A lowflow switch can trip the pump off at a predetermined setpoint. A flow meter can signal a controller to shut down the pump. If this protection circuit is installed, the pump start circuit must temporarily bypass this trip in order to start the pump. This is normally accomplished by a spring-loaded switch that bypasses the low-flow trip when held in the “Start” position, but activates the trip protection in the “Run” position. 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS FIGURE 10 Pilot operated multistage bypass trim actuated by an in-line pilot valve FIGURE 11 Typical in-line serviceable ARC valve 2.447 2.448 CHAPTER 2 OTHER BYPASS SYSTEM HARDWARE __________________________________ Orifices In high pressure applications, the system often does not provide adequate pressure in the bypass line to prevent cavitation or flashing. Either of these conditions is undesirable because it can cause damage to both valves and the pipe system or cause a reduction in flow below the minimum desired, jeopardizing the pump protection system. All pressurereducing valves will experience a velocity-induced recovery effect that will limit the amount of pressure drop a valve can take and cause a reduction in flow capacity. The requirement of backpressure is generic to all pressure-reducing applications. Pressure reduction even by multiple stage cascading can minimize the requirement; however, no valve design will redefine a fluid’s physical properties. This becomes especially important in modulating systems. A fixed orifice will not provide the proper backpressure at all flow levels. As the flow in the bypass line is reduced, the orifice becomes less effective. Proper system design should be used to optimize valve pressure reduction and consider all fluid dynamic effects downstream of any pressure-reducing device. When adequate backpressure is not available downstream of a pressure reducing valve, vapor bubbles will form in the zone just downstream of the valve last stage control surface. This zone is defined as the “Vena Contracta” and represents the point of highest fluid velocity and lowest pressure. The potential for 1) damage to downstream piping components and 2) flow reduction exists from this point. When line pressure remains below the fluid vapor pressure, any existing bubbles will remain and expand as piping friction further reduces line pressure. This can be defined as “flashing condition” and is characterized by a polished appearance on affected surfaces. When the line pressure drops below the fluid vapor pressure and then recovers, any entrapped vapor bubbles will collapse (implode). This is defined as a “cavitating condition” and is characterized by a cinder-like appearance on affected surfaces. The resolution of either condition is best addressed by eliminating vapor formation. This can be assured by the provision of adequate back pressure through the use of a fixed or variable orifice. Fixed Orifice Simple, easily replaced orifices that reduce the pressure are an effective way to reduce bypass head and provide adequate backpressure in bypass systems. Several stages may be necessary, however, to break down high-pressure drops without flashing. For calculations of flow through standard-shaped orifices, see Section 8.1 and 8.2. Coefficients of discharge for oddly shaped multistage orifices are difficult to calculate. However, manufacturers of these specialties can furnish curves of delivery as a function of pressure. Variable Orifice In modulating systems, a fixed orifice will not provide the proper backpressure over a wide flow range. A backpressure regulator (BPR) has a variable orifice with a spring-loaded plunger that is designed to open at a specified differential pressure. If flow and differential pressure increase, it opens further to maintain the differential pressure and backpressure constant. Figure 12 illustrates a typical BPR construction. Figure 13 shows a standard BPR installation. The BPR is normally located as close to the receiver vessel as possible so that the correct backpressure is maintained in the entire bypass line. Valves, Piping, and Fittings For cold water at low pressure, a simple power-actuated globe-type bypass valve is often adequate. In modulating bypass systems, the bypass valve must resist throttling damage, particularly if the water is hot. Staging the pressure drop in the valve is the most common way to reduce or eliminate flashing and cavitation damage to the valve trim or body. Figure 14 illustrates a typical multi-stage pressure reducing valve (PRV). Pressure is reduced in stages to ensure that the pressure never decreases below the fluid vapor pressure. This prevents cavitation and the resultant valve damage and noise. Figure 15 illustrates typical calculations for reducing pressure in sequential stages. Refer to Chapter 7 for detailed information regarding valve sizing and selection. 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS FIGURE 12 2.449 Construction of a typical backpressure regulator FIGURE 13 Backpressure regulator installation Pipe material like that used for the main discharge is adequate for nearly all of the bypass line. Near the orifice and control valve outlets, however, heavier wall or higher chrome content will lengthen life. Welded piping is common for high pressures. To prevent erosion, pipe fittings (especially elbows) should not immediately follow an orifice. Flow Meters Flow rate is the variable that must be measured for most automatic bypass multi-component control systems. The meter may have any type of primary element that will produce an accurate signal at the process flow for which the bypass must be controlled. A simple orifice meter or venturi tube is commonly used. The user must have the required straight upstream and downstream pipe lengths or use flow straighteners to obtain an accurate reading. The device must be properly sized to provide both accurate indication at relatively low process flows and satisfactory pressure drop at maximum process flow conditions. Flow meters can be located either upstream or downstream of the pump. Meters located upstream are at lower internal pressure but are larger in diameter with larger flanges. In addition, pressure drop at rated flow is important to avoid low pump NPSH. Meters located downstream are normally smaller but must be rated for the maximum pump discharge pressure. 2.450 CHAPTER 2 FIGURE 14 Construction of a typical multistage pressure-reducing valve Controls Controls range in complexity from a simple pressure switch that stops a pump drive motor to modulating systems that meter flow and maintain a selected minimum flow through the pump. The control system may have to provide a deadband near the opening point and be sufficiently stable to hold bypass flow nearly constant in spite of erratic flow during upsets and startup. For example, when the primary flow-measuring element is upstream of the bypass branch off, sufficient decease in flow will cause the bypass control to open in an “On-Off” bypass system. The main control valve, farther downstream, will then also open to maintain flow to the boiler or process. The primary element will sense the added flow and, in a simple bypass control system, close the bypass valve, initiating hunting in flow and valve action. To prevent this, the setpoint for bypass valve closing must be greater than twice the minimum bypass flow. If minimum flow is in the 30–50% range, this wastes energy during bypass. This control problem is avoided if a modulating system is installed. SPECIAL BYPASS SYSTEMS ___________________________________________ Pumps in Parallel It is possible to provide a single bypass valve for two or more pumps, as shown in Figure 16. The bypass line and valve must be able to handle simultaneously 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS FIGURE 15 FIGURE 16 2.451 Calculations for pressure reduction through a multistage valve Examples of a single bypass valve that provides minimum flow control for two parallel pumps the total bypass flow of all pumps. The pumps must operate together and they must have identical head-capacity curves. If they do not, the bypass flow will differ between pumps. The pump with the lower head may have insufficient bypass flow or could be completely shut off. In the system of Figure 16 (left), the main line check valve prevents backflow through an idle pump. If the branch off were installed as shown in Figure 16 (right), an additional check valve in each bypass connection would be necessary to stop the backflow. Therefore, this latter configuration is not recommended. An alternative arrangement that provides better operational flexibility and pump protection is illustrated in Figure 17. Each pump is individually protected by its own automatic recirculation control (ARC) valve. Pumps can be operated individually or in any combination and the pump head characteristic curves do not have to be identical. Pumps in Series A single bypass line and valve can protect two pumps in series (see Figure 18). The designer must provide for the larger of the two separately evaluated bypass flows and must take into account the heating effect of the upstream pump. The pumps cannot operate singly unless additional piping and controls are installed. 2.452 CHAPTER 2 FIGURE 17 FIGURE 18 Multiple parallel pumps individually protected by ARC valves Example of a single bypass line that protects two pumps in series VARIABLE SPEED PUMP BYPASS CONSIDERATIONS______________________ Large horsepower pumps are often powered by steam turbines or with motors connected through a variable speed coupling. The effect of decreased speed and discharge head is to reduce the minimum flow requirement for the pump. Therefore, if the minimum required bypass flow is calculated and sized with the pump at normal operating speed and pressure, the bypass flow will be adequate when the pump is operated at lower speeds. Warning If a constant backpressure device such as a backpressure regulator is used to control bypass pressure, precautions must be taken to ensure that the pump is never operated at pressures below the BPR setting. If the pump is operated below the BPR setting, the pump will have no bypass flow and may be damaged in a short time. In a typical boiler system, a constant speed booster pump, requiring less minimum flow than the boiler-feed pump, could be protected by the bypass flow for the variable speed boiler feed pump. The designer must determine if the booster pump bypass flow is sufficient when the high pressure feed pump is running at minimum discharge pressure. 2.3.4 CENTRIFUGAL PUMP MINIMUM FLOW CONTROL SYSTEMS 2.453 REFERENCES _______________________________________________________ 1. Dufour, J. W., and W. E. Nelson. Centrifugal Pump Source Book. McGraw-Hill, 1992. 2. Garay, P. N. Pump Application Desk Book. 2nd ed. Fairmount Press, 1992. 3. Anderson, H. H. Centrifugal Pumps and Allied Machinery. 4th ed. Elsevier Science Publishers Ltd., 1994. 4. Heald, C. C. “Cameron Hydraulic Data.” 18th ed. Ingersoll-Dresser Pump Company, now Flowserve Corporation, 1995. SECTION 2.4 CENTRIFUGAL PUMP PRIMING IGOR J. KARASSIK C. J. TULLO A centrifugal pump is primed when the passageways of the pump are filled with the liquid to be pumped. The liquid replaces the air, gas, or vapor in the passageways. This may be done manually or automatically. When a pump is first put into service, its passageways are filled with air. If the suction supply is above atmospheric pressure, this air will be trapped in the pump and compressed somewhat when the suction valve is opened. Priming is accomplished by venting the entrapped air out of the pump through a valve provided for this purpose. Unlike a positive displacement pump, a centrifugal pump that takes its suction from a supply located below the pump, which is under atmospheric pressure, cannot start and prime itself (unless designed to be self-priming, as described later in this section). At its rated capacity, a positive displacement pump will develop the necessary pressure to exhaust air from its chambers and from the suction piping. Centrifugal pumps can also pump air at their rated capacity, but only at a pressure equivalent to the rated head of the 1 pump. Because the specific weight of air is approximately 800 that of water, a centrifugal 1 pump can produce only 800 of its rated liquid pressure. For every 1 ft (1 m) water has to be raised to prime a pump, the pump must produce a discharge head of air of approximately 800 ft (m). It is therefore apparent that the head required for a conventional centrifugal pump to be self-priming and to lift a large column of liquid (and in some cases to discharge against an additional static liquid head) when pumping air is considerably greater than the rating of the pump. Centrifugal pumps that operate with a suction lift can be primed by providing (1) a foot valve in the suction line, (2) a single-chamber priming tank in the suction line or a two-chamber priming tank in the suction and discharge lines, (3) a priming inductor at the inlet of the suction line, or (4) some form of vacuumproducing device. 2.455 2.456 CHAPTER TWO FOOT VALVES _______________________________________________________ A foot valve is a form of check valve installed at the bottom, or foot, of a suction line. When the pump stops and the ports of the foot valve close, the liquid cannot drain back to the suction well if the valve seats tightly. Foot valves were very commonly used in early installations of centrifugal pumps. Except for certain applications, their use is now much less common. A foot valve does not always seat tightly, and the pump occasionally loses its prime. However, the rate of leakage is generally small, and it is possible to restore the pump to service by filling and starting it promptly. This tendency to malfunction is increased if the liquid contains small particles of foreign matter, such as sand, and foot valves should not be used for such service. Another disadvantage of foot valves is their unusually high frictional loss. The pump can be filled through a funnel attached to the priming connection or from an overhead tank or any other source of liquid. If a check valve is used on the pump and the discharge line remains full of liquid, a small bypass around the valve permits the liquid in the discharge line to be used for repriming the pump when the foot valve has leaked. Provision must be made for filling all the passageways and for venting out the air. PRIMING CHAMBERS _________________________________________________ The Single-Chamber Tank A single-chamber primer is a tank with a bottom outlet that is level with the pump suction nozzle and directly connected to it. An inlet at the top of the tank connects with the suction line (Figure 1). The size of the tank must be such that the volume contained between the top of the outlet and the bottom of the inlet is approximately three times the volume of the suction pipe. When the pump is shut down, the liquid in the suction line may leak out, but the liquid in the tank below the suction inlet cannot run back to the supply. When the pump is started, it will pump this entrapped liquid out of the priming chamber, creating a vacuum in the tank. The atmospheric pressure on the supply will force the liquid up the suction line into the priming chamber. FIGURE 1 Single-chamber priming tank 2.4 CENTRIFUGAL PUMP PRIMING 2.457 Each time the pump is stopped and restarted, a quantity of liquid in the priming tank must be removed to create the required vacuum. Because of possible back siphoning, the liquid volume in the tank is reduced. Unless the priming tank is refilled, it has limited use. Automatic refilling of the priming tank cannot occur if the pump has a discharge check valve unless there is sufficient backflow from the discharge system before the check valve seats. Because of their size, the use of single-chamber priming tanks is restricted to installations of relatively small pumps, usually to a 12-in (305-mm) suction line (approximately 2000 gpm [450 m3/h]). The Two-Chamber Tank This type of priming tank is an improvement over the singlechamber design. It consists of integral suction and discharge chambers connected by a vacuum breaker or equalizing line. The operation and automatic refilling feature of this device are shown in Figure 2. Commercial priming chambers are readily available with proper automatic vents and other features. PRIMING INDUCTORS_________________________________________________ If a separate source of liquid