Vous êtes sur la page 1sur 68

TECHNICAL AND RESEARCH BULLETIN NO.

2-29A

Measurement and Evaluation of Structural and Machinery Vibration in Ships


Panel HS-7 (Vibrations)
of the

Hull Structure Committee


and

Panel M-20 (Machinery Vibrations)


of the

Ships' Machinery Committee

Published by

The Society of Naval Architects and Marine Engineers


601 Pavonia Avenue, Jersey City, New Jersey 07306
www.sname.org

Copyright O 2004 by the Society of Naval Architects and Marine Engineers

T&R Bulletin No. 2-29A has been prepared for THE SOCIETY OF NAVAL ARCHITECTS AND MARINE ENGINEERS TECHNICAL AND RESEARCH PROGRAM

Reviewed and Approved by: (Joint with ANSI Working Group S2-11) PANEL HS-7 (Vibration) Richard J. Sonnenschein, Chairman, HS-7 Arthur F. Kilcullen, Chairman, S2-11 Gary P. Antonides Steven P. Antonides William Blake Richard K. Brown, Jr. Frederick J. Burke Arthur J. Cautilli Jiang-Ren Chang Yung-Kuang Chen Kevin F. Danahy Joe Gelsomino George D. Hill Allan K. Kukk Joel H. Leifer Mark T. McGown Edward F. Noonan Anthony R. Paladino SNAME PANEL M-20 (Machinery Vibration) Richard T. Woytowich, Chairman Gary P. Antonides Robert Gerlach Jaroslav V. Havel Moses W. Hirschkowitz Patrick 0. Prendergast Approved by: HULL STRUCTURE COMMITTEE Phillip G. Rynn, Chairman Frederick H. Ashcroft Bilal M. Ayyub Rameswar Bhattacharyya Harry Paul Cojeen Mark Debbink Allen H. Engle William J. Fallon Owen F. Hughes Roger G. Kline Chao H. Lin Naresh M. Maniar Thomas Miller SHIPS' MACHINERY COMMITTEE David R. Rodger, Chairman Robert S. Behr Karl E. Briers Roger K. Butturini Allen Chin Joseph H. Comer 111 Thomas F. Conroy Jr. James J. Corbett W. Mark Cummings Richard D. Delpizzo Earl W. Fenstermacher Joseph P. Fischer Richard W. Harkins John F. Hennings Bahadir lnozu Thomas P. Mackey William L. McCarthy Charles A. Narwicz Mark F. Nittel Charles H. Piersall Kevin D. Prince Alan L. Rowen E. Gregory Sanford Peter G. Schaedel William J. Sembler Kenneth Siegman Richard P. Thorsen Matthew F. Winkler Richard T. Woytowich Ivan Zgaljic Lewis E. Motter Stephen E. Sharpe Robert A. Sielski Richard J. Sonnenschein Robert J. vom Saal Christopher J. Wiernicki Steve Stroubakis Ivan Zgaljic J. Allen Parkes Paul C. Shang John J. Slager Steve Stroubakis Richard F. Taddeo Michael B. Wilson William A. Wood

It is understood and agreed that nothing expressed herein is intended or shall be construed to give any person, firm, or corporation any right, remedy, or claim against SNAME or any of its officers or members.

ABSTRACT This Bulletin supplements ANSl documents S2.19, S2.25, S2.26,S2.27, S2.28 and S2.29 pertaining to ship vibration. The ANSl documents contain guidelines for the measurement and evaluation of structural and machinery vibration on ships. They also specify data acquisition and processing procedures which will result in consistent and reliable data to compare with the guidelines. However, the ANSl documents do not provide a lot of detail about how to make the vibration measurements or do the analytical calculations that are required. This T&R Bulletin is designed to assist in understanding the rationale for many of the requirements, and help the reader in choosing the appropriate techniques for the required measurements and calculations. Where appropriate, it discusses standards issued by other organizations, and references texts, reports, etc. where more detail can be found, particularly for the analytical prediction of vibration magnitudes. This Bulletin, together with the underlying ANSI documents supercedes the existing SNAME T&R Bulletins 2-25 and 2-29 and SNAME Codes C1, C4, and C5.

ACKNOWLEDGEMENTS This document was a joint project of SNAME Panel HS-7, chaired by Rich Sonnenschein, and Panel M-20, chaired by Rich Woytowich. The panels took advantage of research and a number of reports that were done in connection with Navy and Coast Guard sponsored tasks to improve ship vibration standards in the 1990's. That work resulted in a number of ANSl standards, initiated by ANSl Working Group S2-11, chaired by Arthur Kilcullen. It is the function of this document to support, explain, and expound on the ANSl standards. There was one subject outside the expertise of most of the panel members, that of hull vibration calculations. To write that section, Panel HS-7 turned to Jerry Hill, Daniel Curtis, and Arthur Symmes of John J. McMullen Associates, Inc., Barton McPheeters of MSCSoftware Corporation, Kevin Arden of Northrop Grumman Newport News, and George Camp of Bath Iron Works.

Table of Contents PART I .GENERAL ...................................................................................................................... 1

I I

INTRODUCTION ................................................................................................................ 1 1 1.1.1 Objective and Scope .............................................................................................. 1 1.1.2 Background ............................................................................................................ 1.1.3 SNAME and ANSI Standards ................................................................................ 2 3 1.1.4 History .................................................................................................................. DEFINITIONS AND TERMINOLOGY ................................................................................. 4 TEST & ANALYSIS PROCEDURES (General) .................................................................. 6 1.3.1 Test Conditions .................................................................................................... 6 7 1.3.2 Test Procedures..................................................................................................... 1.3.3 Instrumentation ..................................................................................................... 8 1.3.4 Vibration Severity ................................................................................................. 8 1.3.5 Data Processing................................................................................................... 10 REFERENCES ................................................................................................................ 11

1.2 1.3

1.4

PART II .STRUCTURAL VIBRATION ......................................................................................... 13 2.1 2.2 INTRODUCTION .............................................................................................................. 13 HABITABILITY GUIDELINES ........................................................................................... 13 2.2.1 Background .......................................................................................................... 13 14 2.2.2 ANSI S2.25 Criteria .............................................................................................. 2.2.3 Data Processing................................................................................................... 15 EQUIPMENT SURVIVABILITY (Mechanical Suitability) GUIDELINES ............................16 2.3.1 General ................................................................................................................ 16 16 2.3.2 Shake Table Tests vs Shipboard Environment ................................................... 2.3.3 Major Propulsion System Problems vs Shipboard Environment .........................17 PROPELLER FORCES (Pressure & Bearing forces) ....................................................... 18 18 2.4.1 General ................................................................................................................ 18 2.4.2 Computer Programs ........................................................................................... 2.4.3 Preliminary Estimates .......................................................................................... 20 21 DIESEL ENGINE EXCITATION ..................................................................................... 21 2.5.1 Introduction ......................................................................................................... 2.5.2 Inertia Forces ....................................................................................................... 22 22 2.5.3 Inertia Moments (Couples) ................................................................................ 2.5.4 Firing Forces ........................................................................................................ 22 23 2.5.5 Summary of Likely Forces and Moments ............................................................ 24 2.5.6 Design Considerations ......................................................................................... 25 HULL VIBRATION CALCUATIONS ................................................................................. 2.6.1 Approaches .......................................................................................................... 25 2.6.2 General ................................................................................................................ 25 2.6.3 Empirical Formulas .............................................................................................. 26 2.6.4 Beam Models ....................................................................................................... 26 28 2.6.5 Plate Models ....................................................................................................... 2.6.6 Analysis ................................................................................................................ 29

2.3

2.4

2.5

.6

2.7

TESTING REQUIREMENTS ............................................................................................ 31 2.7.1 Transducer Locations ......................................................................................... 31 2.7.2 Test Report .......................................................................................................... 31 31 SHAKE TABLE TESTS ..................................................................................................... LOCAL VIBRATION .......................................................................................................... 33 REFERENCES ................................................................................................................ 34

2.8 2.9 2.10

PART Ill .MAIN PROPULSION & AUXILIARY MACHINERY VIBRATION ..................................37 3.1 37 INTRODUCTION ........................................................................................................... LONGITUDINAL VIBRATION ........................................................................................... 37 38 3.2.1 Historical Background and Perspectives ............................................................. 3.2.2 Conventional and Nonconventional Systems ...................................................... 38 3.2.3 Mathematical Model ............................................................................................. 39 3.2.3 Operational Factors ............................................................................................. 40 41 3.2.4 Measuring Alternating Thrust in the Main Thrust Bearing ................................... 42 3.2.5 Turbine and Diesel Thrust Bearings .................................................................... 43 3.2.6 Vibration of Propulsion System Components ..................................................... 43 3.2.7 Gear Tooth Stresses ............................................................................................ 3.3 TORSIONAL VIBRATION ................................................................................................. 44 3.3.1 Torsional Calculations .......................................................................................... 45 47 3.3.2 Torsional Vibration Criteria .................................................................................. 47 3.3.3 ANSI S2.27 Criteria for Torsional Vibration ...................................................... 50 3.3.4 Comparing Criteria to Measurements .............................................................. LATERAL VIBRATION OF PROPULSION MACHINERY ................................................ 51 3.4.1 Lateral Calculations ............................................................................................. 52 53 3.4.2 Lateral Measurements ......................................................................................... 53 3.4.3 Lateral Criteria ..................................................................................................... BALANCING ..................................................................................................................... 54 55 3.5.1 Balance Quality Grade ........................................................................................ 3.5.2 Balancing and Vibration Documentation ..............................................................55 55 3.5.3 Shipboard Measurements ................................................................................... 55 DIESEL ENGINES ............................................................................................................ 56 3.6.1 ANSI S2.27 Diesel Engine Criteria .................................................................... 56 3.6.2 Other Criteria ..................................................................................................... 57 GEAR CASE VIBRATION ................................................................................................. 3.7.1 Calculations and Measurements ......................................................................... 58 58 3.7.2 Criteria ................................................................................................................ AUXILIARY MACHINERY ................................................................................................. 58 59 3.8.1 ANSI S2.28 - Measurements on Bearing Housings ............................................ 3.8.2 ANSI S2.29 - Measurements on Shafts .............................................................. 59 REFERENCES ................................................................................................................ 60

3.4

3.5

3.6

3.7

3.8

3.9

PART I GENERAL
1.1
I.I .I
INTRODUCTION Objective and Scope

The primary objective of this Bulletin is to supplement ANSl guidelines for assessing the acceptability of a ship's performance with respect to structural (hull, superstructure, and local), equipment, and machinery vibration. It applies to seagoing ships and for inland ships of all lengths. It applies to turbine, electric, and diesel driven ships, with single or multiple shafts. This Bulletin also provides information on the measurement and data processing procedures which result in reliable data to compare with the guidelines, and gives guidance for calculations required for both machinery and structural vibration during the design of the ship. It also includes some background on how the ANSl standards were developed, and the rationale for them. ANSl guidelines for structural vibration (ships' hulls, superstructures, and locat structures) are based on human habitability and equipment survivability. If these guidelines are met there will also be assurance that there are no major shaft rate vibration problems associated with the propulsion system, or structural fatigue problems. Machinery vibration guidelines in the ANSl documents are given for the main propulsion machinery (torsional, lateral, and longitudinal), and for the internally excited vibration of auxiliary machinery. To indicate probable compliance with the machinery guidelines certain mathematical analyses are required, and these are discussed. This Bulletin is intended for use by naval architects, marine engineers, ship designers, shipbuilders, equipment manufacturers, and ship operators.
1.I .2 Background

Starting in the 1960's, the Technical Panels of the Society of Naval Architects and Marine Engineers (SNAME) were the driving forces in developing ship vibration standards in the United States. An indication of their general acceptance when they were written is the fact that many of the documents were adopted by I S 0 as International Standards. In recent years, however, the American National Standards Institute (ANSI) has included ship vibration as part of their purview and has issued a number of standards dealing with ship hull and machinery vibration. Much of this work was sponsored by the United States Navy. The ANSl documents, being more recent, are more compatible with new developments, and, if the SNAME documents were updated, they would essentially duplicate the ANSl documents, or would confuse shipbuilders and buyers if they contained different criteria. The members of the SNAME vibration panels have concluded that they can provide a more valuable service by issuing documents that explain how best to comply with the ANSl requirements with respect to calculations and measurements. This document, issued by Panels HS-7 (Vibrations) and M-20 (Machinery Vibrations), taken together with the underlying ANSl Standards, supersedes most of the existing SNAME documents on both hull and machinery vibration, including T&R Bulletins 2-25 and 2-29 and Codes C1, C4, and C5. Many different types of vibration are treated in this Bulletin, and each type is associated with its own set of problems and analysis methods. In recent years, the treatment of some types of vibration has changed significantly, while other types are treated essentially the same as years ago. For this reason, the depth of coverage of the different types varies considerably. Also, some new innovations in ship design, propulsion systems in particular, do not have a long history, so not as much is known about them, making the corresponding coverage sketchier. It is hoped that this Bulletin will be updated to account for new developments in shipboard vibration, and will eventually cover all types of vibration in greater depth.

Numerous industry standards addressing vibration of marine machinery and machinery spaces, in particular, currently exist. The prediction processes and vibration limits contained in these standards are sufficient in some cases, but in other cases lack sufficient definition. Reference is made to the existing industry standards where they may be of assistance. Specifically, some of the relevant industry standards are as follows: 1. 2. 3. 4. American Bureau of Shipping (ABS) Rules for Building Steel Vessels; American Society of Testing and Materials (ASTM) Standards; American National Standards Institute (ANSI) Standards; Bureau Veritas (BV) Building and Operation of Vibration-Free Propulsion Plants and Ships; 5. International Standards Organization (ISO) Standards; 6. U.S. Dept. of Labor, Occupational Safety and Health Administration (OSHA); 7. Society of Naval Architects and Marine Engineers (SNAME) T&R Bulletins

1.I .3

SNAME and ANSl Standards

The existing SNAME documents on ship vibration are as follows: T&R Code C-I, "Code for Shipboard Hull Vibration Measurements," 1975. T&R Code C-4, "Shipboard Local Structure and Machinery Vibration Measurements," 1976. T&R Code C-5, "Acceptable Vibration of Marine Steam and Gas Turbine Main and Auxiliary Machinery Plants," 1976. T&R Bulletin 2-25, "Ship Vibration and Noise Guidelines," 1980. T&R Bulletin 2-29, "Guide for the Analysis & Evaluation of Shipboard Hull Vibration Data," 1992. T&R Bulletin 3-42, "Guidelines for the Use of Vibration Monitoring for Preventative Maintenance," 1987. Note that three of the six were written in the 701s,two in the 80's, and only one in the 90's. Some of these documents deal only with measurement procedures and analysis methods, both of which, it was felt, had to be standardized to yield consistent results before the actual criteria could be developed. Criteria for machinery vibration magnitudes were established in Code C-5, but the first guidelines for hull vibration came from an I S 0 document (IS0 6954)(Ref. 1-1). Even after it was issued, the discussions continued about what is the appropriate measure of severity (broadband or narrowband, velocity or acceleration, etc.), and whether the criteria for hull vibration should be related to habitability, equipment survivability, or that which results from unacceptable machinery problems. The ANSI documents referred to are: ANSl S2.19-1989, "Mechanical Vibration - Balance quality requirements of rigid rotors - Part I: Determination of permissible residual unbalance." ANSl S2.25-2001, "Guide for the Measurement, Reporting, and Evaluation, of Hull and Superstructure Vibration in Ships. ANSl S2.26-2001, "Vibration Testing Requirements and Acceptance Criteria for Shipboard Equipment." ANSl S2.27-2002, "Guidelines for the Measurement & Evaluation of Ship Propulsion Machinery Vibration." ANSl S2.28-2003, "Guide for the Measurement and Evaluation of Vibration of Shipboard Machinery." ANSl S2.29-2003, "Guide for the Measurement and Evaluation of Vibration of Machine Shafts on Shipboard Machinery." Except for the balancing standard, these were started in the go's, but were approved in 2001, 2002, or should be approved in the near future. They all give quantitative guidelines for

shipboard vibration, and, where appropriate, they specify acceptable data acquisition and processing procedures. This Bulletin, which supersedes all of the SNAME documents listed above except T&R Bulletin 342 (on condition monitoring), is divided into three parts: (I) General (11) Structural Vibration (Habitability & Equipment Survivability) (Ill) Main Propulsion & Auxiliary Machinery Vibration Panel HS-7 (Vibrations) has the responsibility of maintaining Part II, Panel M-20 (Machinery Vibrations) has responsibility for Part Ill, and both panels are jointly responsible for Part I. It is recognized that vibration of hull and machinery (particularly when slow speed diesels are involved) should often be considered together, perhaps by using an integrated hull/machinery computer model. However, the exact procedure will differ from ship to ship, and there is not yet sufficient agreement within the industry to include such an integrated approach in the ANSI documents or these guidelines. Ship buyers and sellers are encouraged to agree on such an approach if the design of the ship warrants it.

I .I .4

History

In April 1959, a Task Group that became the HS-7 Panel (Vibrations), began evaluating instrumentation systems and methods of shipboard hull vibration measurement, for the purpose of establishing a recommended procedure for the collection of shipboard vibration data. Three shipboard studies were done (1959, 1960, and 1963). A shipboard vibration measurement system was developed for the U.S. Maritime Administration (MARAD) in 1965. The first "Code for Shipboard Hull Vibration Measurements" was published by SNAME as T&R Bulletin 2-10 in 1964, was revised in 1967 to include details on the new MARAD instrumentation system, and again in 1969 to include longitudinal vibration measurements on main machinery components. It was published as T&R Code C-I in 1974 by SNAME Panels HS-7 and M-20. Code C-I was used as the basis for the International Standards Organization 180-4867, 1984 (Ref. 1-2), which was expanded to include the treatment of large diesel drives. In like manner, T&R Code C-4, "Local Shipboard Structures and Machinery Vibration Measurements" was published in December 1976 and was used as the basis for 1S0-4868, "Code for the Measurement and Reporting of Local Vibration Data of Structures and Equipment," published in 1984 (Ref. 1-3). These events indicated a measure of international agreement at the time on the methods of evaluating shipboard vibration. In all cases, these codes called for the evaluation and reporting of each frequency component assessed separately, in terms of their maximum repetitive amplitudes (MRA). The MRA was the most used measure of the severity of vibration since about 1960, when the primary recording device was the oscillograph, and when manual analysis of data was performed as described by Manley (Ref. 1-4). In practice, this procedure proved to be time-consuming and subject to human error. Rather than using the manual method, a filter can be used to isolate the components of interest, and the MRA can be found from the filtered signal. A typical filtered signal modulates considerably, and the MRA is the maximum value of the envelope of this filtered signal. However, filters cannot isolate components properly if there are several components present in the frequency range of interest. These problems have led to the recent conclusion that MRAs of individual frequency components are not appropriate for use in criteria, and that the MRA is only useful when it is applied to a broadband or unfiltered signal. At the same time, it has become increasingly accepted that hull vibration should be evaluated based on all the frequency components together.

The use of automatic data analysis equipment also helped to change attitudes significantly. In 1972, Panel HS-7 started the process of updating data measurement and analysis procedures, the development of a ship vibration data bank and the publication of a guide for shipboard vibration control. The T&R "Vibration Data Sheets were published by SNAME starting in 1976. Many events took place before the "Ship Vibration Design Guide" (Ref. 1-5) was finally published in 1990. In 1982 SNAME recommended a test program to develop "crest factors" to bridge the gap between the MRAs and the rms results often obtained by the widely used spectral analyzers. Limited data indicated that the crest factor (the MRA divided by the rms value) for hull vibration is normally about 2.5, at least for blade frequency vibration. If tests could refine that number for various types of ships, a convenient method for estimating the MRAs would be to obtain the rms values of each component and simply multiply it by the appropriate crest factor. Further work on this project was deferred primarily because of cost considerations. Meanwhile, efforts to set reasonable hull vibration limits began in the IS0 in 1970. 180-6954 was published in December 1984 (Ref.1-I), with the same criteria that appeared in SNAME T&R Bulletin 2-25. It included a tentative crest factor of 2.5. T&R Bulletin 2-29 was authorized to clarify its usage. However, because of the continuing controversy over the application of the FFT analyzer in relation to 130-6954, the Bulletin was not issued until 1992. lS0-6954 had been well received and was in wide use, but was frequently misinterpreted. It satisfied the criteria for human exposure to whole body vibration, as defined in 1S0-263111 (Ref.1-6), and is applicable to each frequency component assessed separately. The attitudes toward MRAs changed also. Rather than finding MRAs by multiplying rms values by 2.5, it made more sense for the criteria to address rms magnitudes directly. This is particularly true for broadband magnitudes, which, as noted above, were becoming accepted as more appropriate for criteria. Work began on a new standard for hull vibration in 1995 when the U.S. Naval Sea Systems Command sponsored a task to recommend a commercial standard that could be used by the Navy instead of the existing military standard. Work on the Navy task was done in conjunction with SNAME Panel HS-7 (Vibration) which also wore an ANSl hat as Working Group S2-77. Over a period of several years, many improvements were made by this committee, and resulted in ANSl S2.25. For machinery vibration, the U. S. Coast Guard, which had been using a military standard, funded the development of a new propulsion system vibration standard which they issued in 1999. That formed the basis of ANSl S2.27, developed by Working Group S2-77.
1.2 DEFINITIONS AND TERMINOLOGY

990 (Ref. 1-7) apply. In addition, the following The definitions in ANSI S2.1-2000/ISO 2041 :I definitions and terminology apply. Added Mass. When a plate structure is in contact with a fluid, motion of the structure will force the fluid to move, causing a reaction force on the structure. For hull girder frequencies, putting additional mass on the structure can model this reaction. This is termed Added or Virtual Mass. Axial Vibration. Any forced or natural vibration wherein the various propulsion system mechanical elements move in the direction defined by the axis of the shafting. Balancing. A procedure by which the mass distribution of a rotor is checked and, if necessary, adjusted in order to ensure that the vibration of the journals and/or forces on the bearings at a frequency corresponding to operational speed are within specified limits.

Balance Quality Grade (G). A measure of the allowable unbalance expressed in terms of velocity (mmls). G = we, where w = angular rotational speed (radls), and e = eccentricity (mm). Blade Rate. Blade rate or propeller blade frequency is the product of rotational speed and the number of propeller blades. Broad-band Vibration. Vibration having its frequency components distributed over a broad frequency band, usually an octave or greater. Consistent Mass. A real structure usually has Distributed Mass rather than idealized Lumped Masses. If the mass is modeled as lumped, it will generate a diagonal mass matrix. A method that is appropriately consistent with the shape functions used to generate the stiffness matrices in the Finite Element method gives a banded mass matrix. This Consistent Mass matrix gives cross coupling terms that allows the motion of one grid point to produce an inertia reaction at an adjacent grid point. This can give more accurate results in some cases. Coordinate Axes. x longitudinal, + is forward y, transverse, + is to port z, vertical, + is upwards cp, angle of roll, right hand is +x 8, angle of pitch, right hand is +y ly, angle of yaw, right hand is +z Coupled Mass. See Consistent Mass. Coupling (mechanical). A mechanical coupling is a fastener connecting two shafts, the rotating elements of two pieces of equipment, or a shaft and a rotating element of a piece of equipment. Direct Frequency Response Analysis. A method that determines the response of a structure to dynamic loading at a particular frequency or a series of frequencies by solving the equations of motion at each of the loading frequencies. (Compare with Modal Frequency Response Analysis.) Distributed Mass. Mass that is modeled by describing the distribution of mass on a plate element or along a bar element. (Compare with Lumped Mass.) Engine Firing-rate Frequency. The product of rotational speed and the number of engine cylinders firing in each revolution. Environmental Vibration. Environmental vibration with respect to shipboard equipment is foundation or support vibration excited by sources external to the equipment, either by ship motion resulting from wave action (rolling, pitching, yawing, heaving, or surging), impacts, propeller or shaft forces, or by other ship machinery. Finite Element. The basic division of a finite element model. A finite element model represents a structure with an assemblage of connected units (e.g., plate and beam elements) that can individually be easily solved for forces, stresses and motion. The individual elements are combined using matrix algebra to solve the complete problem and to recover the forces, stresses and motions of the individual elements. Full Matrix. A matrix with no (or few) zeros. Fully Coupled Matrix. A mass, stiffness, or damping matrix with no (or few) zeros. This means that the local force or motion of any part of the structure will directly affect all other parts of the structure.

Grid Point. A location used in for defining the geometry of a finite element. Gyroscopic Moment. Gyroscopic moment is the turning moment which opposes any change in inclination of the rotational axis of a mass having a moment of inertia. Lateral Vibration. Lateral vibration of a shafting system is vibration normal to the shaft axis Local Vibration. The accentuated vibration that occurs in various components or at discrete locations within the vessel's structure. Longitudinal Vibration. Longitudinal vibration is synonymous with axial vibration. Lumped Mass. Mass that is modeled by putting discrete values of mass at a grid point; not a coupled mass. (Compare with Distributed Mass; see also Consistent Mass.) Machinery-excited Vibration. Vibration excited by forces or moments generated by machinery. Maximum Repetitive Amplitude (MRA). The maximum displacement, velocity, or acceleration of a vibratory signal that repeats over a given time period. In a modulating signal, the time period examined should include several crests. The fact that it repeats indicates that it can be used where fatigue or wear over a long period is a concern. Modal Frequency Response Analysis. A method that determines the response of a structure to dynamic loading at a particular frequency or a series of frequencies by first determining the natural frequencies and mode shapes of the structure for a large frequency range (perhaps five or ten times the range of loading frequencies). The natural frequencies and mode shapes are then used to solve the equations of motion for the desired loadings, which usually are not at any of the natural frequencies of the structure. (Compare with Direct Frequency Response Analysis.) Nonstructural Mass. The structural mass is the mass of the structure that is modeled as plate and beam elements (volume times density). Any additional mass of equipment, outfitting, or structure can be included by adding lumped or distributed masses to the model. These masses add no stiffness to the model, and are termed nonstructural mass. Propeller Blade Frequency. Propeller blade frequency or blade rate is the product of rotational speed and the number of propeller blades. Rayleigh Damping. Damping that is a linear combination of the stiffness and mass proportional damping of a structure. Under these conditions, classical normal modes, where everything vibrates in phase, will exist for the damped system. Shaft-rate Frequency. The frequency of shaft rotation. Structural Damping Coefficient. A measure of the hysteretic damping of a structure, equal in the energy sense to twice the Viscous Damping Ratio. Torsional Vibration. Torsional vibration of a rotating shafting system is an oscillatory rotational movement of the elements comprising that system. Vibratory Torque. Vibratory torque in a shafting system is a torque fluctuation which generates or results from torsional vibration. Virtual Mass. See Added Mass. Whirling Shaft Vibration. Whirling shaft vibration is a combination of shaft rotation and lateral vibration.

1.3 1.3.1

TEST AND ANALYSIS PROCEDURES (General) Test Conditions

The test conditions required for hull vibration measurements and for propulsion system measurements are listed in S2.25 and S2.27 respectively. There are two significant differences in the requirements for hull vibration and for propulsion system vibration, as follows: Sea state must be 3 (SS 3) or less for hull vibration. For propulsion systems: SS 1 or less for small craft, SS 2 for ships of less than 10,000 tons, SS 3 for ships greater than 10,000 tons. (S2.27 does not specify type of tonnage, but it is recommended that the gross tonnage be used to determine sea state.) Both cases require full immersion of the propeller, but for hull vibration, the stern plating in the vicinity of the propeller should be underwater. (For this requirement, judgement must be used to include the area of the plating subjected to significant pressure fluctuations from the propeller. It is suggested that any plating that comes within a propeller radius of the blade tips in any direction as the propeller rotates be included.) If both tests are being done concurrently, the lesser sea state applies, as required by S2.27, and the stern plating should be underwater, as per S2.25. Individual ship specifications can alter the test conditions for any particular ship.

1.3.2

Test Procedures

The test procedures are generally the same for both the structural and the machinery measurements. All transducers should be calibrated in the laboratory before being installed. After installation, each instrumentation channel of the vibration measurement system should be checked to ensure proper functioning. A field calibration (which normally excludes transducer calibrations, but includes conditioning and recording equipment) should be made before and after each major phase of the test. Transducer calibration can sometimes be done on site by placing a calibrated transducer on or immediately adjacent to the installed device, exciting the neighboring structure (with a shaker or by running nearby machinery) and comparing transducer outputs. In order to allow for limited diagnostics, if they become necessary, provision should be made for obtaining the phases between different locations. This can be done by making all measurements in a particular direction simultaneously, or by recording a shaft position marker as a reference.

Operating Conditions. The differences in the operating conditions for the hull and propulsion system during data recording are as follows:
a) Steady acceleration (or deceleration) run from half of rated shaft speed or less to the rated speed to determine critical speeds. The rate of change in speed should be 10% of the rated RPM per minute or less for the propulsion system, 5% for the hull. If the tests are concurrent, 5% should be used. b) Free-route runs at constant speeds from half of rated shaft speed or less to the rated speed. The words describing the increments differ somewhat, but, simply put, require 10 or more equally spaced increments, with additional runs at service speeds and near resonances. In addition, the following optional test conditions, described in Annex A of S2.27 may be required by the purchaser for the propulsion system tests, depending on the type of ship and its normal operating requirements.

(c) Hard turns to port and starboard at rated speed or speed(s) agreed upon by the purchaser and the builder. Recording should commence before the turn and continue until the maximum vibration has passed (when settled into the turn). The maximum vibration occurs at the beginning of the turn and is very dependent on how fast the rudder is applied. Some helmsmen are reluctant to apply the rudder quickly, so that data will be inconsistent from ship to ship. The vibration after the ship settles into the turn will give more consistent results. Note - turns would normally not be required for ships such as cargo ships and tankers, but would be required for ships such as combatant ships and patrol craft, where high speed turns are expected as a part of their normal operation. (d) Bollard tests (optional). While securely anchored, or tethered to a dock or another ship, perform tests (a) and (b) described above. This would be appropriate for tug and tow boats, etc. Although no guidelines are given for crashbacks (changing speed from full ahead to full astern as quickly as possible), this maneuver sometimes excites natural frequencies that may not otherwise be detected. If this test is done, it should be done with caution because crashbacks may result in structures or equipment being subjected to very large displacements and/or stresses. In multiple-shaft ships, all shafts should normally be run as closely as possible to the same speeds. However, in certain instances, it may be appropriate to run with a single shaft. With controllable pitch propellers, all runs should be made using the design pitch, or the range of pitches, normally used at each shaft speed. In some cases, it may be appropriate to make vibration measurements at different or additional speeds to those mentioned above, such as the maximum possible speed. In those cases, such changes should be detailed in the ship specifications. Often, data can be obtained during acceptance trials in which ships normally perform endurance runs for several hours. This may be an opportunity to use roving transducers at successive locations to define mode shapes or to acquire additional habitability or enviromental data which may have been limited due to the number of transducers or limited trial time.

1.3.3

Instrumentation

Acceleration and velocity transducers suitable for acquiring data between 1 and 100 Hz are generally acceptable for hull measurements, and between 1 and 1000 Hz for machinery measurements. The limits may be modified for particular ships depending on the expected excitation frequencies, and for special requirements such as machinery diagnostics. Virtually all modern vibration transducers generate signals proportional to acceleration, and it is necessary to integrate the signals with respect to time to obtain velocities. Integration at low frequencies can be a problem, and it may be necessary to calibrate at low frequencies. An alternative to integrating the signal in the time domain is to generate spectra from acceleration data and divide by the angular frequencies to obtain velocity spectra. Measurements should be recorded on an electronic system which produces time domain records in a form (magnetic tape, computer disk, etc.) which can be replayed for subsequent analysis. Either digital or analog records are acceptable.

1.3.4

Vibration Severity

Perhaps the most significant change in recent thinking, reflected in the new ANSI documents, is the use of broadband values as measures of vibration severity. Many of the guidelines in the superseded documents are based on the maximum repetitive amplitude (MRA) of individual

frequency components. There are problems in processing data in order to obtain MRAs, and often it was not clearly stated that the guidelines applied to MRAs. Test engineers using spectral analyzers often obtained the results in a format convenient for those machines. To understand one of the reasons for using broadband values, consider the characteristics of a typical ship vibration record. There could be a single dominant frequency as shown in Figure 1-1, normally blade frequency, or there could be several frequencies involved, perhaps including engine firing frequency or blade harmonics. In either case, there is likely to be significant modulation, due to ship motion, waves, etc. Frequency spectra of single modulating components have sidebands as shown. If the modulation frequency is constant, as in the upper trace, the sidebands will be distinct and will be spaced at the modulation frequency away from the carrier frequency, most often blade frequency. If the modulation frequency varies, as in the lower case, the sidebands will be spread over some range above and below the carrier frequency.

Time Histories

Frequency Spectra

Figure 1-1 - Modulating Signals and Corresponding Spectra. Where it is required to find the MRA or peak magnitude of a single frequency, it can be found from the time history. (The only difference between MRA and peak is that the MRA excludes the largest excursions, which may be a result of electrical problems, for example, or an impact or other transient phenomena.) However, if there are several frequencies present, they must be separated before finding the peaks. This is normally done with a filter, and if the components are spread out, they can be separated, and the peaks found from the filtered signals. In filtering, however, the pass band must include all of the associated sidebands, or the filter will reduce the modulation, and the peak value found will be too low. This is normally not a problem with turbine drives where the propeller blade excitation is relatively isolated. Unfortunately, with diesel drives, there are usually too many frequency components too close together to separate them with a filter and still include all of the appropriate sidebands. Consequently, the MRA (or peak) values of individual frequency components are often impossible to measure accurately, and are not appropriate for use in standards. Three useful criteria are left: (a) broadband rms, (b) narrowband rms, and (c) broadband peak. Even though narrowband data are essential in understanding the nature of a vibration, the broadband quantities are more appropriate for limiting criteria. If there are several significant frequency components involved, it would be the combined magnitude that would cause wear or fatigue, or that would be felt by humans. Rms values have the added advantage that they are affected less by waves, ship motion, etc. than peak values. Even though damage is often associated more with peak values, the magnitudes of vibration that will cause damage are not usually known, so there seems to be little advantage in measuring peaks.

In the case of alternating torque, alternating thrust, and torsional stress, however, the magnitudes that cause damage are usually known. When the alternating torque or thrust exceeds the mean, the resulting torque or thrust reversals could damage the gears or thrust bearing. The reversals would be associated with the peak unfiltered magnitudes, not the rms, and not just blade frequency. In evaluating the torsional stresses with respect to fatigue, it is the peak values that do the most damage. For these three quantities, the criteria in ANSl S2.27 are expressed in peak values, and apply to the unfiltered signal. In propulsion system vibration, peak values are sometimes used in the measurement of the relative radial motion between a shaft and an adjacent bearing. Adequate clearances must be maintained in the bearing, so the unfiltered peak-to-peak displacement is the quantity that best reflects vibration severity in this application. In recent years, with data being processed by spectrum analyzers, average amplitudes of individual frequency components were often considered instead of MRAs. Frequency spectra are valuable (and sometimes required by the new ANSl documents) in pointing out the sources of any excessive vibrations as well as ensuring the quality of the data and allowing for diagnostics if a problem is discovered. However, for the new standards, broadband rrns values have to be determined. Complicating the analysis of the habitability data somewhat is the fact that the human body is less sensitive to low frequencies, and therefore the frequency content of the vibration signal must be attenuated below 7 Hz. 1.3.5
Data Processing

In general there are four types of data presentation that are required by the ANSl documents, and the data processing procedures must be carefully executed to provide the results required at each of the measurement locations and directions: a) Broadband rms magnitudes for free-route runs, including speeds at hull girder, superstructure, and/or machinery resonances. After filtering the appropriate band, such as from 1 to 100 Hz for hull measurements, rrns magnitudes can be found from an rrns magnitude detector and averaged. Ship vibration usually has considerable modulation and irregularity, so a long enough sample has to be analyzed to get a good average. Alternatively, the rrns value of a signal within the specified frequency band can be obtained with most spectral analyzers. If the signals have low signal-to-noise ratios, the rrns magnitudes determined by these methods will be too large. In this case, it is possible to find the broadband rrns magnitude by obtaining the spectral amplitudes of all frequency components above the noise level (excluding spurious components such as electrical line frequencies) and finding the square root of the sum of the squares of those amplitudes. b) Broadband peak magnitudes for free-route runs or maneuvers as required. Peak values are useful when mechanical damage is associated with the largest forces, displacements, velocities, accelerations or strains. However, it must be kept in mind that peak values are very sensitive to the conditions present during sea trials, particularly the sea state. The simplest method of obtaining broadband peak values is to record the broadband time histories on a strip chart or display them on an analyzer, and manually measure the peaks or find them with a cursor. c) Rms magnitude frequency spectra, averaged over the length of each run, for the freeroute/acceleration runs, including at resonances. To obtain the required amplitude spectra, a Hanning window and a 114 Hz bandwidth are normally appropriate. Reports must contain frequency spectra for any free-route runs where the criteria are not met. d) Plots of rms amplitudes versus rpm of major orders (such as shaft rate, blade rate, firing rate, and other significant engine orders), as obtained from the spectra. The ANSl documents

require test reports to contain plots of significant orders of vibratory velocity, thrust, or torque versus rpm, even though there are no criteria associated with these. Low frequency vibration. Care must be taken with respect to the low frequency components in ship vibration records. The magnitudes associated with rigid body motions (roll, pitch, yaw) can be very large, but they are normally well below I Hz for large seagoing ships. If the instrumentation does not measure below 1 Hz, those signals will not overwhelm the signals of interest (those excited by the propeller or machinery). The next lowest frequencies are usually those associated with the fundamental hull modes (in the neighborhood of up to 3 Hz for large seagoing ships). Normally, the only excitation of interest in that range is propeller shaft rate. Unfortunately, during trials, the seaway often excites those modes as well. Ideally, sea trials should be conducted in calm water, but ships' schedules often do not allow that luxury. Sometimes, it will be necessary to judge whether these low frequency components are due to the seas or machinery, based on factors such as the constancy of the vibration, or the shaft speeds at which they occur. Any components due to the seas should not be included in the magnitudes compared to guidelines. In some cases, such as with smaller ships, it can be reasonably expected that there will be no shaft, machinery, or propeller excitations at very low frequencies, and that there may be ship motion near 1 Hz due to the seaway. In such cases, the owner and builder should agree on a appropriate lower limit greater than 1 Hz. On the other hand, there are some ships with hull fundamental frequencies below 1 Hz which could be excited by shaft rate. It may be appropriate for the buyer and seller to agree on a lower limit less than 1 Hz. However, note that, at least for habitability purposes, the weighting will minimize such low frequency components. Beating. Beating occurs when there are two excitations close in frequency so that their responses alternately add and subtract, such as two propeller shafts rotating at slightly different speeds. The resulting wave form appears similar to the modulation of a single frequency component which is shown in Figure 1-1 (top left). It also has "bulges" and "waists," but the spectrum of a beating vibration shows two distinct components rather than the single dominant component with sidebands such as occurs with modulation. The difference can also be detected from the waveform because the period of a modulating signal is constant, whereas the period of a beating signal is different in the "bulges" than in the "waists." If the period in the waist is shorter than in the bulge, the larger vibration is at a higher frequency, and the converse also applies. Evaluating data expressed in terms of acceleration and displacement. Much of the vibration data presented in years past are in terms of individual frequency components (e.g., blade rate) and may be expressed as acceleration or displacement instead of velocity. Data which contain two or more significant frequency components can be converted to broadband rms velocities that can be evaluated with the guidelines in this Bulletin. In such cases, where spectra consist of clear discrete components, conversion to rms velocity can be made by multiplying or dividing each frequency component expressed as an rms displacement or acceleration, respectively, by angular frequency, applying the weighting factor if appropriate (see Section 2.2 Habitability Guidelines), and then taking the square root of the sum of the squares of all the rms velocity components.
1.4

REFERENCES Overall Evaluation of Vibration in Merchant Ships," Dec. 15, 1984.

(1-1) International Standard I S 0 6954, "Mechanical Vibration and Shock - Guidelines for the

(1-2) International Standard I S 0 4867, "Code for the measurement and recording of shipboard vibration data," Dec. 15, 1984.

(1-3) lnternational Standard I S 0 4868, "Code for the Measurement and reporting of local vibration data of ship structures and equipment," Nov. 15,1984. (1-4) Manley, R.G., "Waveform Analysis," John Wiley and Sons, Inc., 1946. (1-5) "Ship Vibration Design Guide," Ship Structure Committee, SSC 350, 1990. (1-6) International Standard I S 0 263111, "Guide for the evaluation of human exposure to wholebody vibration - Part 1: General requirements." (1-7) ANSI S2.1-2000lISO 2041 :1990, Nationally Adopted International Standard "Vibration and Shock - Vocabulary."

PART II - STRUCTURAL VIBRATION


2.1 INTRODUCTION

In ANSl S2.25 (Ref. 2-I), guidelines are given for human habitability and equipment survivability. If the vibration is low enough to satisfy those guidelines, it would also provide an indication that there are no major vibration problems associated with the low speed components of the propulsion system such as unbalance, misalignment, or pitch variance between propeller blades (pitch unbalance). It is also likely that the hull stresses are not large enough to cause structural fatigue failure. When ANSl 52.25 was written, IS0 6954 (Ref. 2-2) was the most established hull vibration standard, but it only gave levels at which habitability complaints were (a) probable, or (b) unlikely. It used peak magnitudes of individual frequency components. The IS0 document has now been revised, but the U.S. delegation to IS0 still has some reservations about the new version. The guidelines in ANSl S2.25 can be used for the acceptance of new ships or for evaluating a ship's performance after an overhaul or at other times as deemed necessary. In some cases, such as for high performance craft, or for short term operation, criteria other than those contained in S2-25 may be more appropriate. Such cases should be identified by the manufacturer, and changes made to the criteria as agreed to by the manufacturer and the buyer.
2.2 HABITABILITY GUIDELINES 2.2.1

Background

Whereas the old version of IS0 6954 used two vibration levels for criteria, the new version uses three levels: A, B and C, intended generally for "passenger cabins, crew accomodations, and working areas" respectively. For ANSl S2.25, it was felt that a description of human reactions to various vibration magnitudes would also be useful for shipbuilders and buyers, allowing them determine what limits are appropriate for which spaces. Guidance on human reactions is given in ANSl S3.18 (Ref. 2-3), which is similar to I S 0 2631, "Mechanical vibration and shock - Evaluation of human exposure to whole-body vibration," which has two parts: "Part 1: General requirements" (Ref.2-4), and "Part 2: Continuous and shockinduced vibration in buildings (1 to 80 Hz)" (Ref. 2-5). It is anticipated that there will be a Part 4 "Vibration on board sea-going ships (1 to 80 Hz). In the meantime, Part 2 on buildings helped in the development of the new version (2000) of I S 0 6954, and, as a result, of ANSI S2.25. But it was an older version of 2631 that was most useful. It recommended three different vibration levels for the categories "comfort, work proficiency, and safety or health." There were limits for various times of exposure, ranging from one minute to 24 hours. In the newest version of 2631, there seems to be a realization that research has not adequately defined the relationship between the time of exposure and comfort or work proficiency. (However, there is some indication of how the time of exposure affects health.) While much has changed with the new version of 2631, the concepts of comfort, work proficiency and healthlsafety seem valid and are used in S2.25. Another basic difference between the old 6954 and the new one (and S2.25) is that the old criteria applied to individual frequency components, but the new criteria, including S2.25 apply to broadband rms vibration magnitudes. IS0 2631 has more restrictive guidelines, in general, for vertical than for horizontal vibration, but it does provide for "combined" criteria for cases where humans might be sitting, standing, or lying

down, which is certainly the case for ships' crews. The combined criterion is the same as the vertical criterion for most frequencies, but it is the same as the horizontal criterion from about 1 to 3 Hz where the horizontal is the most restrictive. The criteria, expressed as rms velocities or accelerations, are accompanied by "weighting curves" which account for the sensitivity of the human body to vibration. The vibration levels in 263112 apply to buildings, and for ships they are somewhat higher and based on past research and documents peculiar to ships. The weighting curves are felt to be applicable to ships as well as buildings, so the same weighting curves are used in I S 0 6954. They reflect equivalent sensitivities to vibration across the frequency range of interest. They attenuate the measured vibration at the parts of the frequency range at which the body is less sensitive before calculating the broadband rms vibration. This allows more vibration at those frequencies. A simplified approximation of those curves is used in ANSl S2.25, and reflects the fact that human perception of the vibration severity is the same for equal acceleration magnitudes at frequencies below 7 Hz, and also the same for equal velocities at frequencies above 7 Hz. Each vibration record (of one minute or more) must be expressed as a function of frequency (normally with a Fast Fourier Transform), and the amplitudes "weighted" with respect to frequency, that is, each component amplitudes must be multiplied by the weighting function for that frequency. The 7 Hz break point is a compromise between the horizontal break point, which occurs at about 2 Hz, and the vertical, which occurs at about 12 Hz. Note that on ships horizontal hull vibration is usually less severe than the vertical. While I S 0 6954 was often used for merchant ships in the past, the Navy used MIL-STD-1472D (Ref. 2-6) for habitability concerns, including vibration on ships. MIL-STD-1472D covers vibration of vehicles, mostly involving shorter time periods than crews normally encounter on ships. MILSTD-1472D is based on the parts of the 1978 version of 2631 that pertain to land, sea, and air vehicles, much of which has been superceded in the new version of 2631. In the Navy's recent program to switch to commercial standards, I S 0 6954 was identified as the most comparable commercial standard available dealing with hull vibration, and efforts were made to improve it, basing it on mechanical suitability as well as habitability. ANSl S2.25 is a result of those efforts.
2.2.2 ANSl S2.25 Criteria

In determining which spaces should be assigned to which categories, it must be recognized that some types of work, such as reading, writing, calculating, or precision work, are more difficult in a severe environment than physical labor and the more appropriate category may be "comfort, rather than "work proficiency." Normally unoccupied spaces do not require long term work proficiency, but should be "safe." Typical applications for the three habitability categories are given in the table below, along with the recommended broadband criteria in ANSl S2.25 in terms of both velocities and accelerations, so either one can be used. Types of Spaces Cabins, lounges, hospitals Some offices and shops Broadband rms velocity mmls (inls)
< 2.2 (0.09)

Category Comfort

Broadband rms acceleration mm/sA2 (inlsA2) < 98 (3.8) < 220 (8.9)

Work Workshops, galleys, proficiency machinery and control rooms Safety or Health Unoccupied spaces

< 5.0 (0.20)

Note - A range of values of 2 to 4 mmls is reasonable for the upper limit of comfort to take into account the environments of various ship types. For example, it is recommended that cruise ships use 2 mm/sec for cabins, while for tugboats, other work boats, or military vessels, 4 mmlsec may

be appropriate. Ship specifications should state this criterion for each ship. Using these recommended limits as a reference, the shipowner could use the individual ship specifications to require some spaces to have intermediate limits, or to clarify magnitudes for spaces not included in one of the categories above. Velocity weighting is in accordance with the first equation below, and acceleration weighting is given in the second equation below. Weighting function (velocity) = f / 7, 1 Hz < f < 7 Hz = I , f >or= 7 Hz Weighting function (acceleration) = 1, 1 Hz < f < 7 Hz = 7 / f. f >or= 7 Hz The velocity and acceleration guidelines, with their respective weightings amount to the same thing, and reflect the fact that vibration at low frequencies (1 to 7 Hz) are perceived as less severe with respect to habitability than higher frequencies. I S 0 8041 (Ref. 2-7) describes other weighting functions for evaluating human response to vibration that may be of interest. S2-25 differs from I S 0 6954 in the following respects: The categories are associated with human reactions, rather than just the type of space. It has a category for unoccupied spaces (safety) whereas 6954 has berthing and working categories only. It gives definite guidelines rather than a range of values reflecting the vibration "commonly experienced and accepted." It requires that the stern vibration be measured as a reference. It uses simplified weighting functions. It applies to a frequency range of 1 to 100 Hz, whereas 6954 applies to 1 to 80 Hz. It has criteria for "Mechanical Suitability" (Equipment survivability) as well as for habitability. (See Section 2.3.)

2.2.3

Data Processing

If the measured quantity is acceleration, the weighting curve for acceleration can be applied by using a custom built weighting filter on the acceleration signal before some type of rms level recorder. Note that the filter should cut off frequencies below 1 and above 100 Hz, as well as attenuate the signal above 7 Hz. (Other instrumentation characteristics may preclude the need for cutoffs at I and 100 Hz.) If the measured quantity is velocity, the procedures are the same, except the frequencies below 7 Hz are attenuated. The weighting can also be done with a spectral analyzer by obtaining the spectrum of the acceleration or velocity, multiplying the spectrum by the appropriate weighting curve, and integrating the area under the spectrum between 1 and 100 Hz. If the weighting functions are applied while processing the data rather than while recording, many of the same records can be used for evaluating both human habitability and mechanical suitability. Vertical, athwartship, and fore-aft vibrations should be evaluated separately and each should satisfy the guidelines, after applying the necessary weighting.

2.3 EQUIPMENT SURVIVABILITY (Mechanical Suitability) GUIDELINES 2 . 3 . 1


General

Equipment or machinery to be used on a ship must be designed for the ship's dynamic environment, which can be categorized as follows: Low frequency motions evolving from the motion of the ship in the seaway. The vibration of the hull structure on which the equipment is mounted. The vibration where the equipment or instruments are mounted, which may be on machinery, or on a structure, such as a mast, that might amplify the hull vibration level. A moving ship responds in a highly complex motion to the forces of waves. The periodic low frequency motions of a ship are rolling, pitching, and heaving. Sway, yaw, and surge are unrestrained and therefore not periodic. Roll and pitch are generally objectionable because of the large linear motions that they generate at distances remote from the center of gravity of the ship. They may also cause rotating machinery to generate gyroscopic moments. While these motions are not addressed by the recent ANSl documents, it is sometimes necessary to distinguish between vibration and ship motion while evaluating ship vibration measurements. 2.3.2
Shake Table Tests vs Shipboard Environment.

In ANSl S2.25 the recommended hull vibration level for equipment survivability is based on the shake table tests required by ANSl S2.26 (Ref. 2-8), extrapolated from the testing period to the expected life of the equipment. That extrapolation requires that the severity of the vibration environment on the ship be correspondingly less than that during the tests. 52.26 requires that the equipment be able to perform its normal functions during the shake table tests, and survive with no damage. The hull vibration criteria assume that major low-frequency propeller or shaft problems, such as unbalance or propeller pitch irregularities are absent, as discussed in the next section. For those areas where equipment is mounted, broadband (1-100 Hz) hull vibration should be below about 10.0 mmls rms, with no weighting curve. This is the same as the velocity requirement for "safety" as identified above, except for the weighting. The criteria for both habitability and mechanical suitability must be satisfied, and because most unoccupied spaces contain equipment, this results in a broadband velocity limit of 10.0 mmls in the 1 to 100 Hz range, with no weighting curve. For occupied spaces, the habitability requirements for "work proficiency" and "comfort" are more restrictive. S2.25 requirements for masts are more severe, as shown in the table:
Hull, Superstructure and Mast Vibration, Equipment Environment Criteria:

Broadband (1-100 Hz) Location rms velocity Hull and superstructure 10 mmls (0.4 inls) Mast 30 mmls (1.2 inls) In cases where different environmental vibration magnitudes are appropriate, the owner and builder should agree on acceptable criteria. It should be kept in mind, however, that more than 10 mmls on the hull or superstructure may indicate a major mechanical deficiency. An analysis that relates shake table amplitudes and equipment survivability vibration levels is given as Annex A in S2.25. That analysis indicates that a level of 10 mmls rms for hull vibration where equipment is to be mounted would be acceptable. The analysis makes a number of assumptions to establish the relationship. If it is felt that a safety factor is warranted because of the assumptions, or, for vulnerable equipment, a lesser limit could be used. The shake table

amplitudes for mast mounted equipment are about three times those required for the hull, so the 30 mmls rms limit for mast vibration is also reasonable for most cases. Some equipment is mounted directly on diesel engines or other reciprocating machinery, which places it in a severe environment. ANSl S2.25 has no requirement for the maximum allowable vibration of reciprocating machinery, but ANSl S2.27 (Ref. 2-9) has limits for diesel engines which, depending on size, the mounting, and what the engine is driving, vary between 10 and 18 mmls rms. As shown in the velocity plot of Figure 2-4 (Section 2.8 on Shake Table Tests), the shake table tests for mast mounted equipment have similar velocity amplitudes as equipment mounted on reciprocating machinery, but the latter is tested at higher frequencies. Consequently, it is felt that the diesel engine limits will ensure that there is minimal probability of problems with equipment that survives the shake table tests for equipment mounted on reciprocating machinery. S2.27 has no requirements for reciprocating machinery other than diesels, such as compressors or pumps, but the vibration magnitudes should be about the same or less. Very often equipment is mounted on local structures, and although vibration measurements on local structures are not required, measurements are often prudent. In general, the vibration magnitudes of local structures should not exceed twice the hull girder vibration magnitude in the same vicinity. If the natural frequencies of local structures are close to any hull resonances it may be necessary to modify the local structure to change its natural frequency. 2.3.3 Major Propulsion System Problems vs Shipboard Environment.

Propulsion system components are usually the only items that are massive enough to cause significant vibratory magnitudes of the hull girder or superstructure. Annex B of S2.25 relates two of the most likely shaft rate vibration sources to hull vibration magnitudes: propeller unbalance, and pitch variations in the propeller blades. Recommended limits for both sources are used to calculate the resulting hull vibration magnitudes at hull resonant frequencies of two ships for which vibration hull calculations have been performed. Marine propellers normally have a Balance Quality Grade of no worse than G6.3 as defined in I S 0 1940/1 (Ref. 2-10). That corresponds to an eccentricity of 0.5 to 0.9 mm for the particular propellers and hull resonant frequencies of the two ships. Annex B shows that the hull vibration generated by those unbalances are from 0.095 to 0.326 mmls rms. I S 0 Recommendation 484-1981 (Ref. 2-11) states that the error in mean pitch per blade should be no more than +I- 1 to 1.5 % for most merchant vessels with large propellers. Relating the worst case, a 1.5% pitch error in one blade to the propeller bearing forces, and then to hull vibration magnitudes of the two ships, gives magnitudes of 0.22 to 0.60 mmls rms. These magnitudes might be greater if more than one blade had that amount of error, but then forces due to pitch error from different blades would not be in phase. Even adding the two worst cases together, the shaft rate vibration is below 1 mmls, which would be a reasonable expectation for a new ship with no marine growth, etc. A bent shaft or damaged propeller could easily cause much more severe vibration. Blade frequency vibration magnitudes, at least near the stern, are usually greater than the magnitudes due to pitch error and unbalance, perhaps as much as 3 to 4 mmls rms. The combination would still be less than about 5 mmls. Marine growth could cause that to increase, but, at 10 mmls, one would certainly be well advised to check out the source of the vibration, particularly if it is predominately at shaft rate.

2.4 2.4.1

PROPELLER FORCES (Pressure and Bearing Forces) General

Alternating propeller forces are the result of the propeller operating in a nonuniform wake. The nonuniformity is caused by a boundary layer along the hull, as it is affected by various obstacles to the flow before entering the propeller, such as skegs, bossings, struts, shafts, etc. The flow, in general, is slower in the upper part of the propeller disk since there are more obstacles in that vicinity. Skegs also slow the flow in the lower part of the propeller disk. A large part of minimizing propeller forces involves selecting an afterbody shape and appendages that maintain as much unformity in the wake as possible. In addition, the interaction of the propeller with the wake is determined by the propeller characteristics, such as number of blades, blade area ratio, rotational speed, skew, and pitch. The propeller design process normally involves measuring the wake (inflow velocities) of a model at various points in the propeller disk and resolving the velocities into axial, tangential, and radial components. Most of the axial components, when plotted, show a more or less constant velocity over the 360 degrees of propeller rotation, except for a large dip near 0 degrees (top dead center), and, if a skeg is present, another dip near 180 degrees. See Figure 2-1. The axial components are predominant in generating the forces while the radial components have little effect. The tangential components are measured with respect to a rotating blade. In general, near the stern the flow has an upward component, resulting in a positive flow with respect to the propeller direction for half of its rotation, and a negative flow for the other half. The Fourier components of all the velocity components can be found, and the resulting forces on a single propeller blade with the anticipated blade geometry can be calculated, also in terms of the Fourier components. These forces are added to those of the other blades, resulting in the forces and moments on the whole propeller. Even though the greatest forces on a single blade are normally first and second-order forces, when all the blade forces are added together to get the propeller forces, those components add up to zero. The alternating propeller forces are resolved into three forces (thrust, and vertical and horizontal bearing forces) and three moments (torque, and moments about the vertical and horizontal axes). For the alternating thrust and torque, only the blade frequency or nth order (no. of blades = n) wake components, and its harmonics, generate any propeller forces. For the vertical and horizontal bearing forces and the moments about the vertical and horizontal axes, because of the shaft rotation, only the n-I, n+l, 2n-1, 2n+l, etc. components of the wake generate propeller forces, and those are also at blade frequency and its harmonics. For calculating hull vibration amplitudes, only the vertical and horizontal bearing forces are normally used, but other forces and moments are very important for propulsion system vibration considerations. The pressure forces generated by the propeller greatly affect the hull vibration. These forces are the result of the pressure fields that surround the propeller blades, and primarily affect the vertical vibration of the hull because the pressure acts on a mostly horizontal hull bottom. These forces occur at blade frequency and its harmonics. The pressure forces are of the same order of magnitude as the vertical bearing forces if the propeller is not cavitating. If serious cavitation is present, the pressure forces may be an order of magnitude higher. More detailed information on propeller forces can be found in Reference 2-12.
2.4.2 Computer Programs

There are several computer programs that calculate the forces and moments on the propeller, with the input being the wake survey velocities and the geometry of the propeller. These programs have existed for many years and have been used extensively. It should be remembered that these programs assume constant velocities in the wake, whereas, due to waves, rolling, pitching, etc., the velocities actually modulate, causing the propeller forces to modulate as well. For alternating thrust, as an example, extensive measured data show a crest

factor (ratio of peak magnitude to rms magnitude) of about 4.0. A lesser set of data shows hull vibration to have a crest factor of about 2.5. These diffferences must be considered when comparing calculations to measurements and acceptance criteria.

Figure 2-1. Axial and Tangential Wake Velocities for Single Screw Merchant Ship The development of programs to predict pressure forces is more recent, but they do exist, for cavitating propellers as well as non-cavitating. The results of these programs have not been extensively compared to measured data. Of course, the objective is usually to have a noncavitating design. Both bearing forces and pressure forces can be calculated with respect to phase so that, when finding the total force exciting the hull, account can be taken as to the extent to which they are additive.

2.4.3

Preliminary Estimates

Reference 2-13 suggests a simple method of predicting, for preliminary purposes, the propeller forces of a new ship design. The estimate of propeller forces (except pressure forces) is based on data from other, similar ships. To qualify as "similar," a ship should have a similar stern configuration with regard to struts, skegs, propeller clearances, and number of screws. It should also have approximately the same block coefficient. Past studies on several ships (Ref. 2-14) have shown that highly skewed propellers can reduce hull vibration levels by about 50% compared to conventional propellers, so the amount of skew should be similar also. When used to calculate hull vibrations, the results of this preliminary method are multiplied by the following factors: Shio Tvoe Single Screw Twin Screw

Force
Horizontal Vertical Horizontal Vertical

Hull Pressure Factor 1 2 1 2

Phase Factor
1 1

Total Factor 1 2

The hull pressure factor of 2 is applied to the vertical forces since the pressure forces are of the same order of magnitude as the calculated bearing forces, and it is assumed in the preliminary stages that the two are in phase. The "phase factor" is applied to twin screw ships, and assumes the forces from the two screws are in phase. Both factors apply to vertical forces of a twin screw ship where the total factor is 4. Ref. 2-13 assumes that the results of propeller force calculations are given as single amplitudes, and represent sinusoids. It also suggests the use of a modulation factor (ratio of peak to "average single amplitude") of 2 be used for all four forces in the above table. This differs from the crest factor mentioned above, which uses rms rather than single amplitude. For a modulating sinusoid, the crest factor would be nominally 1.4 times the modulation factor. The hull crest factor of 2.5 found from data mentioned above would correspond to a modulation factor of about 1.8. In any case, if the limiting criterion is expressed in peak vibration magnitudes, then that factor should be used. However, most of the existing standards, including S2.25, use broadband rms criteria, in which case that factor should not be used. Instead, the propeller forces should be expressed as rms quantities (divide single amplitudes by 1.416 if necessary). Note that the factors above apply to blade frequency forces and harmonics. If engine excited hull vibration is also to be considered, those vibration magnitudes should be added before comparing to a broadband criterion. The preliminary estimate suggested by Ref. 2-13 is based on the assumption that, for "similar" ships, the alternating thrust and torque and bearing forces are roughly dependent on the value of propeller advance ratio (J) at design speed.

Where:

VA

= speed of advance of prop relative to the water n = rotational speed of propeller D = diameter of propeller

Therefore, the calculated forces (as a percentage of mean thrust) for similar ships can be plotted against the various values of J at design speed, and, knowing the J of the new design, the forces

(as percentages of mean thrust) can be determined from the plot. Ref. 2-13 cites the data for three ships as shown in the following table for twin screw naval ships: Ship # I ,469 1.08 1.7 1.5 1.0 Ship #2 589 .839 .98 0.32 0.42 Ship #3 .480 1.13 1.79 1.43 1.O

Block Coeff. Adv. Ratio Alt. thrust, % mean thrust Hor. Force, % mean thrust Vert. force, % mean thrust

Even though the block coefficient of Ship #2 is significantly higher than the other two, the data are plotted against the advance ratio in Figure 2-2, and if the relationship is valid the points should fall close to a straight line. Then the percentage forces for a new similar design can be taken from the curve. The confidence in the method increases according to the number of ships that are included in the plot, and how close the results are to a straight line.
Propeller Forces VS Advance Ratio

Advance Ratio
-0--

Alternating Thrust

Horizontal Forces

-+ Vertical Forces

Figure 2-2. Alternating Propeller Forces VS Advance Ratio


2.5 2.5.1
DIESEL ENGINE EXCITATION Introduction

The forces and moments generated by a diesel engine are either internal or external. Internal forces and moments stress the engine parts, but, if the engine and its foundation are rigid enough to avoid significant distortion, these forces and moments are not transferred to the foundation or other supporting structure. Consequently, it is not normally necessary to consider these as excitations to the hull, but it must be a goal in the design process to ensure that such rigidity exists. The external forces include vertical and horizontal forces, and the external moments exist about all three axes. The forces and moments occur at engine rotational frequency and various multiples of that frequency, and if one of these exciting frequencies coincides with a natural frequency of a ship's structure or hull, it could cause a resonant vibration. Forces excite resonances most efficiently at antinodes of the mode of vibration excited, and moments excite

resonances most efficiently at the nodes of the mode. A diesel engine also exerts a torsional excitation on the shaft, but, at least with a direct drive, that only affects the rotating components of the propulsion system. In a geared drive, the bearing reactions in the reduction gear can be transferred to the foundation of the gear, and again, adequate rigidity of the gear case and foundation is necessary. Torsional vibration is discussed in Part Ill. Large, low-speed, two-stroke, direct drive diesel engines are the most likely to generate forces large enough, and in the frequency range, that could excite structural resonances. The external forces and moments that the engine generates can be calculated for any engine design, and are normally provided by the engine manufacturer.
2.5.2 lnertia Forces

lnertia forces result from the reciprocating motion of the internal parts of the engine. For a single cylinder engine, the largest inertia force is in line with the piston motion, opposite in direction, and at the first order. The lower end of the connecting rod also travels in the horizontal direction (on a vertical engine). If the connecting rod were infinitely long, the motion of the piston would be sinusoidal. Because the rod is finite in length, second and higher-order vertical forces are also produced. In a two cylinder engine, with the cranks 180 degrees apart, the first-order forces cancel, but the second-order forces add. With four cylinders and the cranks 90 degrees apart, the first and second-order vertical forces cancel. This is the normal configuration of a four cylinder two-stroke engine. A four-stroke engine, however, has two cranks at 0 and two at 180 degrees. This is because firing all four cylinders takes two revolutions, so the firing occurs every 180 degrees. This topic gets fairly complicated with various numbers of cylinders, V-engines, and other variables, but the engine manufacturers can provide the forces for each engine. As a general rule, most engines will be designed to cancel the first and second-order vertical inertia forces. In some configurations, such as four-cylinder, four-stroke in-line engines and six-cylinder, two-stroke in-line engines, rotating weights may be used to compensate for otherwise unbalanced forces or moments, as noted below.
2.5.3 Inertia Moments (Couples)

In the above example of a two cylinder engine, we noted that the first-order vertical forces were balanced when the cranks were 180 degrees apart. Since the two forces are displaced from each other along the crankshaft, the opposing forces will create a moment or couple about a horizontal axis, called a pitching couple. Depending on the number of cylinders and whether it is a 2- or 4stroke engine, the pitching couples may or may not be balanced. Pitching couples can be converted, wholly or partially into a couple about the vertical axis, or a yawing couple, by adding weights to the crankshaft in addition to those needed to balance the cranks and webs. They are arranged to generate vertical forces that compensate for the pitching moment, but they will have unbalanced components in the horizontal direction. In general, vertical forces can be countered without generating horizontal forces (or vice versa) by having weights on two shafts spinning in opposite directions, with the order of the correction being controlled by the rotational speed of the weights. Similarly, by locating weights near both ends of the engine, out of phase, couples are generated which can be used for balancing. Again, the subject is complicated with various numbers of cylinders, various V-angles, and different arrangements of compensating weights, so users should refer to the manufacturer's data to determine the couples for each engine.
2.5.4 Firing Forces

The gas forces in a cylinder peak within a few degrees after top dead center, and then trail off gradually. These forces are far from sinusoidal, and therefore, in addition to providing a firing rate excitation, will contribute to excitation at harmonics of the firing rate. The expanding gas in a cylinder pushes down on the piston, resulting in compression in the connecting rod, and a torque on the crank, with an opposing force exerted on the crankshaft by the crankshaft bearings. Since the connecting rod is at an angle, the piston is forced against the side of the cylinder, or, in a

at an order equal to half the number of cranks (114 the number of cylinders). 14-, 16-, and 20cylinder, 4-stroke engines may have moderate first and second-order pitching, yawing, and rolling couples. Two-stroke, V-engines have a large first-order pitching couple. Evenly distributed power strokes can be achieved with a 45-degree bank angle in V-8 and V-16 engines, but not in V-12 and V-20 engines, although the large number of cylinders and the rigidity of the engine tend to suppress the net effect.
2.5.6 Design Considerations

In the preliminary design stage it may be feasible to select an engine whose major excitations avoid the lower hull natural frequencies. As the design matures, and a more detailed hull model is developed, hull responses can be calculated, and higher hull modes can be considered. Which orders to examine in the calculations will depend on the orders of major excitation, the hull frequencies, and the operating speeds. Rolling and racking couples can be restrained by supporting the top of a tall engine with transverse struts tied in to the ships structure. Direct structural connections sometimes result in very large forces in the struts. Friction connections and hydraulic stays have been used, which can reduce the forces, provide damping, and provide adjustment to suit a particular situation. Det Norske Veritas (Ref. 2-16) describes a solution to vibration problems caused by first and second-order pitching couples exciting a vertical hull mode. A mechanical shaker is placed, not on the engine, which would be near a node of the hull mode shape, but at an antinode of the hull mode shape, in the aft part of the ship. It utilizes vertical forces to counter the pitching couple. Reportedly, this technique has been used to solve first and second-order vibration problems on more than 70 ships. Reference 2-13 gives samples of the external forces and moments provided by M.A.N. and Sulzer. Each engine can be rated for a range of speeds and powers, and forces and moments are given for the maximums and minimums of those ranges. The terminology and data given differ with the manufacturer. M.A.N. gives the following for each engine: External forces (0 for all engines) External moments, first and second orders (refers to pitching moments) Guide Force H-moments, first, second and third harmonics of firing (rolling moments) Guide Force X-moments, first through twelfth orders (racking moments) Sulzer gives the following for each engine: All first and second-order forces are balanced. First-order vertical (pitching) and horizontal (yawing) couples: with standard counterweights, with non-standard counterweights, with combined first and second-order balancer Second-order vertical (pitching) couple with and without second-order balancer. Reaction to torque variation, presumably at firing rate (rolling couple). All forces and moments described pertain to a properly operating engine. If an engine misfires, the forces and moments could be considerably higher at the first order and harmonics for a twostroke engine, and at the half order and harmonics for a four-stroke engine. The manufacturers will provide the forces and moments associated with misfiring as well as for normal operation. Reference 2-17 gives a more complete description of diesel engine excitation.

2.6 2.6.1

HULL VIBRATION CALCULATIONS Approaches

There are several different approaches to modeling the vibration of a ship's hull. The approach used will depend on the amount of information about the hull structure that is available and the resources available to the analyst. When only the displacement and general design of a ship is available, only the lowest modes of the ship can be estimated using simple methods. With a more developed ship design and sufficient computer resources and software, a more detailed analysis will give more accurate results. Model generators and computers are continually improving, usually making the more accurate results of a large finite element model easier to obtain than those of some simpler methods that may require extensive calculations to develop an adequate representation of an immersed hull structure. The modeling approaches available are: I. empirical formulas 2. beam models 3. plate models
2.6.2 General

The direction of vibration and number of nodes (locations of zero displacement) in the mode shape identify each flexible hull vibration mode. For example, the first vertical mode is the 2noded vertical. The hull vibration modes are not affected by the forward speed of the ship, although flow-induced forces may sometimes excite hull vibration. The rigid body modes can be affected by ship speed, but these modes, including heave and pitch, are distinguished from the hull vibration problem. In general, the ship should be designed such that the undamped natural frequencies of the hull are not too close to the exciting frequencies for the range of operating speeds. The lower frequencies can be passed without significant resonance as the ship comes up to speed. Machinery, hull form hydrodynamics, and alternating propeller forces can all be excitation sources for potential resonances. In the preliminary design stage the undamped frequencies are calculated and the hull design and/or exciting frequencies (e. g., shaft rpm and number of blades) modified as necessary to avoid potential resonances. At this stage of design information about the hull may be limited so that an empirical or crude beam model may be adequate. An added factor of safety should be included to reflect the uncertainties in the model. When more refined design details become available an updated model should be used to confirm the frequencies. Added (or virtual) mass has a significant effect on the predicted hull frequencies and mode shapes. The empirical formulas are based on statistical data from classes of ships and therefore inherently include estimates of the hull girder added mass. When the water depth is less than six times the draft, the added mass will increase because of the boundary effects. Shallow water can also affect the propeller forces. Damping has minimal effect on the natural frequencies but is the dominant effect controlling the magnitude of the response near resonance. A frequency response analysis, either direct or modal (see Section 2.6.6), is normally required to determine the magnitude of the vibrational response. Damping by both the structure and certain bulk cargo can be important, but fluid damping from the surrounding water is small at hull girder frequencies and amplitudes and can usually be ignored. Unfortunately, data and methods for accurately representing the damping in the hull models is usually a weak point in the analysis.

2.6.2

Empirical Formulas

There are several formulas (Ref. 2-18, 2-19) that use the basic dimensions of a ship and estimate the lowest modal frequencies. The Schlick (Ref. 2-20), Burrill (Ref. 2-21), Prohaska (Ref. 2-22), Todd and Marwood (Ref. 2-23), and Dinsenbacher and Perkins (Ref. 2-24) equations use the displacement, length, beam, draft, and midship section properties to estimate one or more of the lowest modes, often with factors that depend on the class of ship being modeled. These formulas tend to give results that are not either consistently higher or lower than the observed frequencies for a selection of ships. In general, the lowest vertical mode is calculated and the higher vertical modal frequencies are estimated by multiplying the 2-noded vertical frequency by two to get the 3-noded vertical, by three to get the 4-noded vertical, and by four to get the 5-noded vertical. Empirical formulas for vibration in the horizontal plane (Ref. 2-25) and for the first torsional mode (Ref. 2-26) are less common and more difficult to characterize accurately because of the added complexity of the horizontal modes including the added mass effect and the high degree of coupling with torsional modes.
2.6.3 Beam Models

The beam model approach represents the ship with a longitudinal series of beam elements and can adequately represent relatively simple monohulls. Destroyers, cruisers, and large tankers can be represented by beams. A short hull, an asymmetric hull, extensive hatch openings (as in container ships), or discontinuous longitudinal decks or bulkheads would all be difficult to model accurately using the beam model approach. A beam model may be used for natural frequency and mode shape calculations, perhaps in the preliminary design stage, but it may be better to do forced response calculations with a plate model. Beam models usually use finite element bar or beam sections. A twenty or forty element model is generally used. The grid points are often equally spaced at station locations, but they can be moved to positions of cross section changes or major equipment locations. Most beam element formulations are adequate, as long as the neutral axis and the shear center can be located independently. With asymmetric ship sections the more sophisticated formulations may be needed to account for warping. For example, the Nastran CBAR element (Ref. 2-27) places the neutral axis, the shear center and the center of mass in the same location, while the CBEAM element allows each of these locations to be distinct and also provides for a warping coefficient. In certain situations, sprung masses or additional beam elements will have to be added, for example to model a flexible mast with the same frequency range as the hull, to model a large unsupported deck area that can move independently of the outer hull, or to include a specific piece of heavy equipment. A beam model may not be adequate or appropriate if it needs much detail of this sort. To develop the model, the hull stiffness and mass properties for each beam element are required. In most instances these would be: The areal properties:
1. 2. 3. 4. 5. 6. 7.

The effective axial area (effective cross-section area of longitudinal beam elements) The effective area moment of inertia about the lateral axis The effective area moment of inertia about the vertical axis The effective shear area for vertical loading The effective shear area for lateral loading The effective area polar moment of inertia for torsion Offset from the shear center to the neutral axis

The mass properties:

8. 9. 10. 11. 12.

The effective mass The effective mass moment of inertia about the lateral axis The effective mass moment of inertia about the vertical axis The effective torsional mass moment of inertia Offset from the shear center to the center of mass (CG)

The hull stiffness and mass properties are assembled (Ref. 2-28) from the most detailed information available. Too often this consists of a few cross sections and a longitudinal strength drawing with center of gravity and mass distribution. If it is available, a Ship Work Breakdown Structure (SWBS) (Ref. 2-29) or weights database will give a more detailed weight distribution. The mass of internal fluids should be included. When data for a particular location are not available, they must be interpolated from the adjacent locations. Rather than a linear interpolation, quantities should often be weighted by the appropriate scaling factor. For example, the effective axial area used for the longitudinal vibration would increase in proportion to the perimeter (or width plus height) if the plate thicknesses do not change. The axial area and the moments of inertia can be calculated from the shell plating, decks, and longitudinal stiffeners. Some plating may not act as a fully effective flange for hull bending because of shear lag (Ref. 2-30), or because of their own bending out of plane, and the affected moment of inertia should be reduced. (Shear lag occurs with very thin or wide flanges on a beam. Due to the shear deformation of the flange, portions of the flange that are remote from the web are not fully loaded and so have a reduced contribution to the stiffness of the beam.) The torsional constants can be calculated for the closed section formed by the hull plating and the strength deck and increased to account for the additional closed sections formed by the other decks. Ships with large open hatch areas will not have closed sections in these locations, which will show sharp drops in torsional stiffness as well as sharp downward shifts of the shear center. The superstructure should be included in all of the areal properties if it will contribute significantly to the hull stiffness or if it exceeds 60 percent of the length of the hull. Many of the section properties can be calculated using small finite element models or the beam property tools of a graphical preprocessor. The shear area for vertical deflection can be calculated from the outboard shell plating and any continuous longitudinal bulkheads, and for lateral deflection from the continuous decks and bottom plating. The area of the vertical plate members will be effective in resisting shear in the vertical direction; the area of the horizontal plate members will be effective in resisting shear in the lateral direction. Inclined plates will have reduced effectiveness. The shear area can also be calculated using a small finite element model or using the section generator feature of some finite element preprocessors. The shear deflections can be significant because a ship can be characterized as a deep beam, so the shear properties should always be included in the model. The mass is modeled using either concentrated masses at the grid points, offset to the center of gravity of the segments, or using masses distributed along each beam element. Here a segment would be the length between the centers of the adjacent beam elements. The rotational mass inertias in all three directions can be important for a beam model, except for the very lowest modes. The rotational inertias can be estimated using a uniformly dense cylinder or brick shape approximating the shape of the section. A small finite element model of the section could also be used to estimate the rotational inertias, scaled up to match the full weight of the outfitted section. Both the light and heavy ship displacements should be used so that the hull frequencies can be enveloped. Equipment that exceeds one-half percent of the total ship weight should be included as a concentrated mass and stiffness (a sprung mass) if it is expected to have frequencies in the same range as the hull, since it will affect the normal modes. It is only necessary to model the equipment modes (and directions) that have frequencies in this range. A consistent (coupled) mass formulation accounts for the fact that, with distributed masses, the inertial force of any small element of mass will affect more than just the closest grid point. In general, this will be more

accurate than a lumped mass formulation for a beam model, with little increase in cost for such a simple model. The added mass for a beam model can be calculated using the Lewis forms (Ref. 2-31). Lewis found that a simple formula described many common ship hull cross sectional shapes and could be conformally mapped onto a semi-ellipse, i.e., transformed to a new coordinate system where the hull section becomes the lower half of an ellipse. The solution for the infinite elliptical prism (an infinitely long elliptical ship) could then be mapped back, giving the vertical added mass for the hull shape. A reduction factor was introduced to account for the longitudinal motion of the water due to the finite length of the ship, by comparing the infinite elliptical prism solution to that of an elliptical solid of revolution. The reduction factor is different for each mode. Horizontal added mass can be approximated using the added mass for a circular cylinder (Ref. 2-18) or by an extension of the Lewis method (Ref. 2-32). Longitudinal added mass is ignored. While the Lewis method is well established, the procedure can be tedious. The boundary element approaches described under plate models will give more precise results and transfer the labor onto the computer. The vertical, horizontal, and torsional vibration characteristics can all be calculated from the same beam model if all of the appropriate mass and stiffness parameters are included. Alternatively, the vertical can normally be calculated separately if the hull is symmetric with respect to a vertical plane at the centerline. The horizontal and torsional modes are often coupled to the extent that they should not be calculated separately, but could be separate from the vertical. 2.6.5

Plate Models

A plate model uses triangular and quadrilateral finite element plate elements to model the hull structure. Major stiffeners should be modeled with plate elements, but minor stiffeners can be modeled with beam elements or smeared into the plate properties. Typically, the plate elements are 2 to 3 feet in size (and up to 8 feet in the longitudinal direction), but this will depend on the size of the ship, the ship data available, the accuracy required, and the computing resources. The ship weight information is obtained as in the beam model. If the weights are given in sufficient detail, the large equipment can be modeled as concentrated or distributed masses at their physical location. The additional weight deficit can be distributed to the structure by increasing the structural density or adding nonstructural mass to the plate elements. Equipment that exceeds one-half percent of the ship weight should be included as a mass and stiffnesses in the appropriate directions if it is expected to have frequencies in the same range as the hull. It is only necessary to model the equipment modes that have frequencies in this range. Equipment on resilient mounts may fall into this category. The coupled mass formulation has no advantage over the lumped mass for the plate model. The added mass can be included using a number of methods. Fluid brick elements have been used, modeling a portion of the ocean surrounding the hull. "Infinite" finite elements placed on the hull plate elements have also been used to model the added mass (Ref. 2-33). These elements have the appearance of plate elements in the model but mathematically extend to infinity into the fluid. More commonly, boundary elements or the MFLUID elements mentioned below are used to create the added mass matrix for the finite element solution. These also are two-dimensional elements that are placed on the wetted surface of the structural hull model. The Doubly Asymptotic Approximation (DAA) (Ref. 2-34) is incorporated into some finite element codes. While created to solve the underwater shock problem, it is mentioned here because it uses the fully coupled added mass matrix in calculating the low frequency response. This added mass matrix is calculated using a boundary element approach and could in principle be extracted and used in a vibration model.

The Underwater Shock Analysis (USA) software (Ref. 2-34) implements the DAA and will calculate the undamped natural frequencies of the ship, using either a beam or a plate model, including the effect of shallow water. The Nastran MFLUID element (Ref. 2-27) also uses a boundary element approach to model the mass effects of the fluid and can model the interior tank fluids as well. This added mass matrix can be modified to be a banded matrix rather than a full matrix. While this adds a degree of approximation, it speeds the computations considerably.

2 . 6 . 6

Analysis

A modal analysis will reveal any hull frequencies that are in the range of the exciting frequencies. Forced frequency response analysis, using either a modal or direct formulation, can be used to predict the magnitude of the vibrational response. The modal formulation first calculates the modes of the ship up to several times the highest frequency of interest. It then uses the calculated frequencies and mode shapes to determine the response of each mode to a particular oscillating force or combination of forces for a large range of frequencies, usually equally spaced. The modal results are superimposed to get the total response of the hull. The direct formulation does not calculate the modes or frequencies but, instead, directly solves for the response of all grid points for each forcing frequency. For a frequency response analysis with a small number of frequencies, the direct formulation is more efficient, but for a large frequency range with small increments, the modal formulation is more efficient. A sufficient number of modes must be used with the modal formulation. The predicted bow response, for example, may be too large if not enough modes have been included. Modes must be included up to five or ten times the largest response frequency. A separate run with more modes will show similar answers if a sufficient number of modes have been used. Damping is needed in a frequency response calculation and will have a significant effect on the results. The loading for the frequency analysis consists of a table or a function of load versus frequency. The oscillatory forcing loads are generated by the propeller (bearing forces, pressure forces, and sometimes the alternating thrust should be considered), and the propulsion machinery (external forces and moments from diesel engines). The loads are imposed as forces at the point of application. For example, the load from the propeller might be modeled using two forces at the aft bearing in the vertical and athwart directions, perhaps with a third longitudinal load at the thrust bearing location. The relative phase of the loads can be specified, in addition to the relationship of amplitude with frequency. The loads can be applied simultaneously or individually in each direction to investigate their effects. The light and heavy ship ballast conditions should be modeled, plus various cargo layouts for cargo ships, to envelope the range of responses. The structural damping of the hull is largely hysteretic; the stress-strain curve describes a loop when the structure oscillates, expending energy in the process. This kind of damping is proportional to the stiffness of the structure and the damping ratio is therefore proportional to the frequency. (The modal viscous damping ratio is given by 5, = Cn/(2.o,.M,), so if C, = P.K,, then 5, = Pa,,/2 since 0 : = K,/M,.) When a structural damping coefficient is given, it is equivalent in the energy sense to twice the viscous damping ratio (2.5) at the given reference frequency. When the predominant weight of a ship is structural, for example in a combat ship, structural damping should control. Mass-proportional damping can also be used in the model. With mass-proportional damping, the damping ratio is inversely proportional to the frequency, becoming infinite at zero frequency. (Again the modal viscous damping ratio is given by 5, = Cn/(2.0, M,), so if C, = a M , , then 5, = al(2.%)). Equipment or cargo may produce some mass-proportional damping.

A linear combination of stiffness-proportional and mass-proportional damping is called Rayleigh damping and is efficiently solved with matrix methods. This formulation is offered in most finite element codes. By using both types of damping, two points on the damping ratio versus frequency curve can be chosen. The damping curve can then approximate a constant damping or other desired relationship between these two frequencies. As an example, suppose a uniform structural damping of 0.10 is desired for a structure. This is equivalent in the energy sense to five percent of critical damping. The Rayleigh damping equation would be

See Figure 2-3. If the frequency range of interest in this example is from 74 Hz to 250 Hz, the values of a and p can be estimated by first assuming that the values for (at 74Hz and 250 Hz are equal and somewhat higher than the desired 6 perhaps 0.055, and solving the two resulting equations. Then the undershoot at the bottom of the curve is checked to see if it is about equal to the overshoot at the end frequencies. Note that mathematically, the minimum of the curve is

and occurs at

If the undershoot is too low or too high, the initial assumed values are adjusted to get a better fit.

Percent of Critical Damping VS Frequency


15 -

.-

C 0 D.

n 0 m .+ .-

f._.---_I_..--.--------0 74

...-. ._.---...----"-$ only ..-. --...,.:.-_..---__..--------.._____


250

400

---------....._______ a only -----------.-_-._._.---------

Frequency, Hz

Figure 2-3. Example of Rayleigh Damping Curves If the modal method is used for the frequency response analysis, damping can often be entered in tabular form as a function of frequency. A good approach is to glean damping information from a similar ship. Actual damping values are hard to estimate and no consistent rules can be drawn from the existing hull data. Some data show that the damping is actually proportional to the 0.75 power of

the frequency. Damping data for a number of steel ships suggest that a stiffness proportional damping value of 2.5 percent of critical damping (damping ratio of 0.025) at 5 Hz might be used as a first cut if no other information is known (Ref. 2-35). If more low frequency damping is required, a linear combination of stiffness and mass proportional damping using from 1 to 4 percent damping in the range of 5 to 100 Hz might be tried. Published damping values for similar ships can also be used as a guide (Ref. 2-36). Wood or fiberglass hulls are significantly more damped than steel and aluminum. Answers should be thoroughly checked and compared to experience, test data, similar ships, and hand calculations. Odd behavior of any mode will often point to a modeling problem. 2.7 2.7.1 TESTING REQUIREMENTS Transducer Locations

ANSl S2.25 describes the minimum number of transducer locations considered to be essential for evaluting hull and superstructure vibration. However, local structures should also be considered. Any local structures (bridge wings, sections of deck or bulkheads, masts, foundations, etc.) and pieces of equipment that are excited by hull girder vibration, and appear to have excessive magnitudes, should be measured on sea trials. The location and direction of the transducers will depend on the situation, but often a triaxial array on top of the equipment is appropriate. 2.7.2 Test Report

ANSl S2.25 gives a comprehensive list of items to be included in the test report. While the guidelines are in terms of broadband rms amplitudes, there are some requirements for narrow band data. Narrow band data can be used to verify calculations that were done in the ship design process, most of which involve blade frequency. They can also be used for diagnostics and possible design changes if the vibration limit is exceeded. Item (h) of Section 9 requires the "results of vibration order analysis." This refers to plots of rms amplitudes (displacements, velocities, or accelerations) vs shaft speed for all major orders (shaft rate, blade rate, machinery frequencies, diesel firing rate, and any other major excitations). The amplitudes used are from averaged spectra obtained from each free-route run. Note that Section 8 on data processing requires such spectra to be obtained, but Section 9 on the report does not require the spectra themselves, just the order plots that are generated from them.
2.8

SHAKE TABLE TESTS

MIL-STD-167 (Ref. 2-37) was the first standard in this country that required shake table tests for shipboard machinery. In that document, such tests are designated Type I vibration. IS0 incorporated similar requirements in I S 0 10055, Mechanical vibration - Vibration testing requirements for shipboard equipment and machinery components (Ref. 2-38), which was first published in 1996 and differs from MIL-STD-167 in several respects. Variable frequency dwell tests are not required, and measures to prevent wrecking resilient mounts at resonances during testing are not included. The U.S. delegation did not agree with those changes, and ANSl subsequently issued ANSl S2.26 (Ref. 2-8) which does include the variable frequency dwell tests, and, for equipment tested on resilient mounts, requires hard mounting the equipment at frequencies in the vicinity of the mounting resonances to avoid destroying the mounts or the equipment. (There is also a requirement that the mounting frequencies do not coincide with any major excitation frequencies.) There are three general categories of equipment for testing: hull mounted equipment

mast mounted equipment equipment mounted on reciprocating machinery (diesel engines, compressors, etc.)

Figure 2-4. Shake Table Amplitudes from ANSI S2.26. The wide frequency ranges of the tests encompass typical excitation frequencies for propeller shaft rate, propeller blade rate, and engine firing rates for large diesel engines. It is recommended

that shipboard equipment be qualified to this standard to obtain a reasonably high degree of probability that the equipment will not fail or malfunction due to vibration during its service life. While the categories are the same in both I S 0 10055 and ANSl S2.26, the frequencies and amplitudes vary slightly. Figure 2-4 shows the required shake table amplitudes in ANSl S2.26 for all three types of equipment. They consist of constant displacements at low frequencies (up to 15 or 25 Hz, depending on the equipment category), and essentially constant accelerations at higher frequencies. Figure 2-4 expresses the amplitudes in terms of displacements, velocities, and accelerations. The fact that hull equipment is tested using a series of constant displacement steps at higher frequencies rather than a constant acceleration, such as the other two types of equipment, is a carryover from early shake table testing when mechanical shake tables required readjustment or resetting to change the displacement amplitudes. It would be easy to maintain a constant acceleration with modern control systems. For the other two categories, which were added at a later date, a constant acceleration is required at the higher frequencies. It is interesting that I S 0 10055, which was issued before ANSl S2.26, does use constant acceleration for hull-mounted equipment, as well as the other two categories. It is recommended that the I S 0 10055 requirement of 7 m/s2 acceleration could be used as an alternative to the ANSl steps. Figure 2-4 shows that 7 m/s2 is about the average amplitude of the steps above 15 Hz. Also of interest is the inclusion of allowances on the displacement amplitudes, which may be a result from normal tolerances on mechanical shakers. As percentages, they vary from 2% to 33%, but a note below the table says a 10% deviation is allowable. It is suggested that 10% be used. Note also that there is a discontinuity in the mast mounted equipment curve at 15 Hz, which was probably not intended. It is recommended that 2.54 mm be the maximum displacement used. 2.9
LOCAL VIBRATION

A local vibration is a vibration that occurs in various components on board a vessel or at discrete locations somewhere within the vessel's structure because of excitation by internal or external forces, with a magnitude significantly greater than the supporting hull structure. The potential for the presence of local vibrations on board vessels has grown substantially over the past few decades, primarily because of significant increases in vessel power, the location of machinery spaces and accommodations further aft (and thus nearer to the propeller), and the increased use of diesel engines. Vibration control is most effective when it is attempted in parallel with the design of a vessel. The configuration of the hull, the arrangement of the structure, the selection of the main engine, the characteristics of the propeller and propulsion train all have a profound influence on the exciting forces and the local (as well as global) responses of the vessel. Adequate attention to the effects of such parameters on shipboard vibration early in the design of a vessel can significantly lessen or even eliminate the presence of unwanted vibrations. A wide range of methodologies is available for assisting shipowners, builders, and designers with shipboard vibration control. These include: methods for identifying potential sources of vibration problems corrective measures to help eliminate such problems in the initial design stage analysis (finite element or otherwise), measurement, and diagnosis of vibration sources comparison of measured vibration with regulations, standards, and specifications It is good design practice for the foundations or resilient mounts of rotating machines to have stiffnesses such that the natural frequencies of the machines on their mountings are separated

from their rotational speeds or speed ranges by about 10 to 40%. The exact amount depends on the particular application, including the sharpness of the resonance curve, the ruggedness of the machine, and its state of balance. Without a good understanding of those factors, a generous separation should be used. It is also good design practice for all equipment and their foundations to avoid natural frequencies within the propeller blade rate range from 50% power through full power, with a similar separation in the frequencies. For resiliently mounted equipment, it is particularly important for the mounting system rigid body modes not to coincide with significant sources of excitation. Resonance avoidance is most effectively used early in the design phase when changes to structures andlor equipment can be more easily accomplished. It is important to identify all major sources of excitation and the natural frequencies of the equipment. The natural frequencies can be plotted against the exciting frequencies (Campbell plots) to identify resonances. In the event of a resonance, a forced response analysis can be performed to determine the magnitude of the response in order to assist in determining if the resonance is harmful. Typical exciting frequencies are the rotational speed of the shafting, blade rate, gear mesh frequencies, turbine blade passing frequencies, engine rotational and firing rates (including half order rates for four-stroke diesels), and the harmonics thereof. In addition to considering how the natural frequencies compare to the exciting frequencies, the manner in which the forces or motion are applied to the local structure should also be considered. Both forces and motion, if applied at an antinode of a local structure, are more likely to generate high vibration magnitudes than if they are applied at a node. Note that a point on a structure may be a node in one direction, but an antinode in another.
2.10 REFERENCES

(2-1) (2-2) (2-3) (2-4) (2-5) (2-6) (2-7) (2-8) (2-9)

ANSl S2-25-2001, "Guide for the Measurement, Reporting, and Evaluation of Hull and Superstructure Vibration in Ships," Feb. 22, 2001. International Standard I S 0 6954, "Mechanical Vibration and Shock Overall Evaluation of Vibration in Merchant Ships," Dec.15, 1984

- Guidelines for

the

ANSl S3.18-1979 (R1993), "Guide for the evaluation of human exposure to wholebody vibration." International Standard I S 0 263111, "Mechanical vibration and shock - Evaluation of human exposure to whole-body vibration - Part 1: General requirements." lnternational Standard I S 0 263112, "Mechanical vibration and shock - Part 2: Continuous and shock-induced vibration in buildings (1 to 80 Hz)." MIL-STD-14720, "Human Engineering Design Criteria for Military Systems, Equipment and Facilities," 14 March 1989. I S 0 8041, "Human response to vibration - Measuring instrumentation," 1990. ANSl S2-26-2001, "Vibration Testing Requirements and Acceptance Criteria for Shipboard Equipment," 20 Nov. 2001. ANSl S2.27-2002, "Guidelines for the Measurement and Evaluation of Ship Propulsion Machinery Vibration."

(2-10) IS0 194011-1986, "Mechanical vibration - Balance quality requirements of rigid rotors -

Part 1: Determination of permissible residual unbalance." (2-1 1) I S 0 Recommendation 484-1981, "Shipbuilding, ship screw propeller manufacturing tolerances, Part 1: Propellers of diameter greater than 2.5 meters." (2-12) E.V. Lewis et al, "Principles of Naval Architecture, Vol. II - Resistance, Propulsion, and Vibration," 2nded., Society of Naval Architects and Marine Engineers, 1988. (2-13) Ship Structures Committee Report SSC-350, "Ship Vibration Design Guide," 1990. (2-14) Hammer, N.O. and McGinn, R.F., "Highly Skewed Propellers - Full Scale Vibration Test Results and Economic Considerations," Paper R, Ship Vibration Symposium, Arlington, VA, Oct. 16-17, 1978, Society of Naval Architects and Marine Engineers, 1978.

(2-15) Harrington, R.L., "Marine Engineering," SNAME,1992. (2-16) "Prevention of Harmful Vibration in Ships, Det Norske Vertitas Guidelines," July 1983. (2-17) Taylor, Charles F., "The Internal-Combustion Engine in Theory and Practice," Vol. 2, Chapter 8, "Engine Balance and Vibration," MIT Press. (2-18) Lewis, E.V., "Principles of Naval Architecture," SNAME, Jersey City, NJ, 1988. (2-19) Noonan, Edward F., "Ship Vibration Design Guide," Ship Structure Committee, Washington, DC, 1990. (2-20) Schlick, O., "Ship Vibrations," Series of Articles on Ship Vibration in TINA (Technological Institute of Naval Architecture), 1884, 1893, 1894, 1901, 1911. (2-21) Burrill, L.C., "Ship Vibration: Simple Methods of Estimating Critical Frequencies," NECl (North East Coast lnstitution of Engineers and Shipbuilders) Transactions, Volume 51, 1934-35. (2-22) Prohaska, C.W., "The Vertical Vibration of Ships," Shipbuilder and Marine EngineBuilder, Volume 5, October and November 1947. (2-23) Todd, F.H., and Marwood, W.J., "Ships Vibration," NECl (North East Coast lnstitution of Engineers and Shipbuilders) Transactions, Volume 64, 1948. (2-24) Disenbacher, A.L., and Perkins, R.L., "A Simplified Method for Computing Vertical Hull Natural Frequencies and Mode Shapes in Preliminary Design Stage," DTRC (David Taylor Research Center) Report 3881, January 1973. (2-25) Johnson, A.J., "Vibration Tests on All Welded and All Riveted 10,000 Tom Dry Cargo Ships," NECl (North East Coast lnstitution of Engineers and Shipbuilders) Transactions, Volume 67, 1951. (2-26) Horn, F., "Horizontal and Torsions - Schiffschwingungen auf Frachtschiffen," WRH, Volume 6, Number 18, 1925. (2-27) "MSC Nastran V68 Reference Manual," (3 Volumes), MSC Software Corporation, Santa Ana, CA. (2-28) Bruck, H.A., "Procedure for Calculating Vibration Parameters of Surface Ships," NSRDC (Naval Ship Research and Development Center) Report 2875, October 1968.

(2-29) "Work Breakdown Structure," Department of Defense Handbook, MIL-HDBK-881, January 1998. (2-30) Roark, R.J., and Young, W.C., "Formulas for Stress and Strain," McGraw-Hill Book Company, New York, 1975. (2-31) Lewis, F.M., "The Inertia of the Water Surrounding a Vibrating Ship," SNAME Transactions, 1929. (2-32) Landweber, L., and de Macagno, M.C., "Added Mass of Two-Dimensional Forms Oscillating in a Free Surface," Journal of Ship Research, SNAME, November 1957. (2-33) Armand, J. and Orsero, P., "A Method for Evaluating the Hydrodynamic Added Mass in Ship Hull Vibrations," SNAME Transactions, Volume 87, 1979. (2-34) DeRuntz, Jr., J.A., "The Underwater Shock Analysis Code and Its Applications," 6oth Shock and Vibration Symposium, Volume I, 1989. (2-35) Robinson, D.C., "Damping Characteristics of Ships in Vertical Flexure and Considerations in Hull Damping Investigations," DTMB (David Taylor Model Basin) Report 1876, December 1964. (2-36) Woolam, W.E., "Research on Ship-Hull Damping Coefficients for Low-Frequency Vertical Flexural Modes of Vibration," NSRDC (Naval Ship Research and Development Center) Report 2323, May 1967. (2-37) MIL-STD-167-1 (Ships), "Mechanical Vibrations of Shipboard Equipment (Type I Environmental, and Type II - Internally Excited)," 1974. (2-38) I S 0 DIS 10055, "Vibration testing requirements for shipboard equipment and machinery components," 1994.

PART Ill - MAIN PROPULSION & AUXILIARY MACHINERY VIBRATION


3.1 INTRODUCTION

Part Ill supplements ANSl S2.27 (Ref. 3-1) with regard to vibration of the main propulsion system (propellers, diesel engines, gas turbines, steam turbines, electric drives) and ANSl S2.28 and S2.29 (Ref. 3-2 and 3-3) regarding auxiliary machinery. It also discusses ANSl S2.19 (Ref. 3-4) regarding balancing of shipboard machinery. These ANSl documents have been developed in recent years, and supersede SNAME Code C-5 (Acceptable Vibration of Marine Steam and Heavy-duty Gas Turbine Main and Auxiliary Machinery Plants; Ref. 3 4 , which was developed by the SNAME M-20 Panel in 1976. Part Ill provides some of the rationale for the ANSl standards, suggests processes for predicting vibration characteristics, and discusses measurement methods to verify compliance with the limits in the ANSl documents. ANSl S2.27 is the first major vibration standard to address some of the more modern propulsion systems, such as thrusters, cycloidal propellers, and waterjets. These are referred to as nonconventional systems. The thrust of a nonconventional system is often transmitted to the ship through structure close to the propulsor rather than through a long shaft, and this results in different vibration characteristics. Such systems are often installed as a unit rather than as separate components (engines, shaft, propeller, and gears). The design of the unit may be satisfactory, but vibration problems are still possible due to improper mounting, flow obstructions, or other factors having to do with the application or installation of the unit. The section on diesel engines suggests appropriate magnitudes for main propulsion diesel engines, but other types of reciprocating machinery, such as compressors, pumps, and diesel generators are not covered by ANSl documents. I S 0 10816-6 (Ref. 3-6) covers all diesel engines and compressors, shipboard or not, but the limits given are for different classes of machines, with no guidance as to how to classify particular machines. Many machinery vibration standards, particularly those published by ANSl and the ISO, include guidelines for monitoring machines during their useful life for maintenance purposes. ANSl S2.27 does not include condition monitoring provisions, and is intended primarily for the acceptance of new machinery. The criteria could be used during the life of the ship, however, to check for mechanical damage or the adequacy of a repair, as long as allowances are made for some increase in the vibration, particularly for rotating machinery, because of such problems as bearing wear and erosion of impellers. The following types of mechanical vibrations are addressed in Part Ill of these Guidelines: Longitudinal (fore and aft) vibration of the propulsion system Torsional vibration of the propulsion system Lateral vibration of shafting normal to the shaft axis Self-excited rotating machinery vibration Diesel engine vibration Gear case vibration Auxiliary machinery vibration

4. 5. 6. 7.

3.2 LONGITUDINAL VIBRATION

Longitudinal or axial vibration is any vibration of the propulsion system in which the various mechanical elements, such as the propeller, shafting, thrust bearing, gearing, turbines or diesel engines move in the direction defined by the axis of the shafting. Axial vibration is excited by the propeller rotating in a nonuniform wake caused by the hull and appendages, hydrodynamic cross coupling between torsional

and axial vibration at the propeller, and, in the case of diesel main propulsion units, flexing of the engine cranks caused by connecting rod forces (Section 3.2.3). There can also be a cross coupling with the torsional vibration through the elastic properties of the crankshaft, which may be important when a natural frequency of the torsional system is close to that of the axial system (Ref. 3-7).
3.2.1 Historical Background and Perspectives

Although there are a few cases of longitudinal shaft vibration before 1941, because of lower powers they were not serious. However, in 1941 on the builder's trial of the battleship NORTH CAROLINA, the axial vibrations were so serious that it was considered unwise to run at full power. Later battleships and the aircraft carrier MIDWAY experienced similar troubles (Ref. 3-8). All these cases involved high power and skegs, which replaced conventional struts. The shaft was long and the thrust bearings, located forward of the reduction gear, were relatively flexible. In all cases the vibration occurred at blade frequency and was corrected by changing the number of blades on the propeller. Using fewer blades would move the resonance to a higher RPM, hopefully above the operating speed, while using a greater number of blades would move the resonance to a lower RPM, hopefully below the operating speed. The greater number of blades would also produce less excitation because the thrust carried by each blade is reduced, and also because the thrust would be less at the lower resonant RPM. The M-20 Panel was formed at about this time to deal with longitudinal vibration problems. Modern commercial ships run at much higher power than earlier ships and so require more attention to longitudinal vibration. Generally, the thrust bearing is moved aft of the reduction gear and so can be more rigidly mounted. More attention is paid to propeller design to reduce the magnitude of excitation. Larger ships have lower structural resonant frequencies and so are more likely to have a local machinery space bottom structural resonance that corresponds to a blade frequency excitation. An example of this is presented in Ref. 3-9. Even where there is no resonance, the combination of a short shaft (where the machinery is placed near the stern) and flexible hull structure (perhaps where the hull cross-section is Ushaped rather than V-shaped) may make the structure supporting the thrust bearing the most flexible element in the vibrating system. The first reference to the coupling between torsional and longitudinal vibration in engine shafts was in 1938 (Ref. 3-10). In the early 19401s,in measuring the torsional vibration on an opposed piston engine with a crankshaft that was very flexible in the axial direction, it was found that there was a strong axial vibration at the same frequency as the torsional vibration. This relationship existed because the natural frequency of the crank in the axial direction was close to the natural frequency of the engine in torsional vibration. It can be shown that, where adjacent cranks are oblique to each other, the axial and torsional modes are coupled. The advent of multi-cylinder long-stroke engines has necessitated more attention to this coupling.
3.2.2 Conventional and Nonconventional Systems

There are no "longitudinal vibration" requirements, as such, for nonconventional systems, such as thrusters, waterjets, and cycloidal propellers. However, there are limits for vibratory velocities of the units in the three coordinate directions. For all systems, it is normal practice to require both design calculations and trial measurements. For conventional systems there are four different longitudinal propulsion system vibration criteria that are required by various guidelines, including the ANSI S2.27. The following quantities are limited: (a) the alternating thrust in the thrust bearings (b) the motion of the propulsion system components (c) the reduction gear tooth stresses (d) the bull gear motion Most of the following discussion is based on conventional systems. It is difficult to make generalizations about nonconventional systems, particularly about the prediction of vibration magnitudes because of the

variety of systems used and a lack of data. Even in conventional systems the difficulty in evaluating the parameters, especially exciting forces and damping, sometimes leads to significant differences between the calculated vibration and the measured. With this in mind, some propulsion systems are designed and built with the knowledge that there may be a problem, but with a solution in mind if the problem materializes during trials. Such solutions could include changing the rpm-pitch relationship on a controllable pitch propeller, changing the number of blades, relocating the thrust bearing, stiffening the thrust bearing foundation, adding blade skew to the propeller, adding fins forward of the propeller to change the flow into the propeller, and installing a resonance changer as described below. 3.2.3
Mathematical Model

The mathematical model for longitudinal vibration will represent the propulsion system as a mass-elastic system comprised of its mechanical elements, The evaluation of each element should be sufficiently accurate so as to adequately calculate a frequency band from zero to 200% of the maximum frequency of excitation. This includes the maximum blade frequency, and, for diesel engines, it will include firing frequency. The possibility of important harmonics of the excitation forces should not be overlooked.
Propeller. The propeller mass must include the mass of the entrained water which is usually taken to be about 50 to 60% of the propeller mass. That can be determined more accurately by the procedure given in Ref. 3-11. Propeller damping, which constitutes most of the damping in longitudinal vibration, must also be estimated. Rigby, in Ref. 3-12, suggests that the damping of a 3-bladed propeller is approximately 16.5 Ib-seclin times the developed area of the propeller in square feet, and that the damping of 4- or 5-bladed propellers is approximately 39 Ib-seclin for each foot of blade edge length. The magnification factor of the propeller at a longitudinal resonance is often about ten, and sometimes a value of damping is selected which produces that result. Propeller damping is discussed further in Ref. 3-1 1. Shafting. If the frequency range of the analysis is low enough, the number of masses and stiffnesses representing the shaft can be relatively few. However, if the excitations are at higher frequencies, or if higher modes are of interest, the range must be greater, and more masses are required as the mass distribution of the shaft sections become important. The simplest model, often used for short shafts, assumes that half the shaft mass is at the propeller, and half at the bull gear. Also fairly simple, one shaft mass can be used with half the actual shaft mass, and a quarter of the shaft mass at each the prop and the gear. For most purposes, a mass element for each physical section of shaft is sufficient. Thrust Bearing. In a system with a reduction gear, the thrust bearing is the mechanical element that transmits the propeller thrust to the hull of the ship. It ordinarily consists of a thrust collar, which is a large circular disk attached to the shaft, and a thrust block, which is a structural member rigidly mounted to the hull of the ship. The thrust is transmitted to the thrust block via a thrust bearing which usually consists of a number of lubricated shoes which bear upon the thrust collar. The shoes may be supported by a mechanical system to ensure that each shoe takes its share of the load. In most cases, the thrust bearing can be represented as a stiffness which, combined with the thrust bearing foundation stiffness, couples the thrust collar to the ship's hull. However, in some cases, such as ships with flexible hull structures, it may be necessary to model the thrust bearing foundation and ship bottom in more detail. Such models normally include the hull bottom between the engine room's forward and aft bulkheads.

In some thrust bearings, known as resonance changers, the shoes are supported on hydraulic pistons contained in the thrust-block structure. The pistons are supported by a volume of oil which equalizes the shoe force and, by changing the volume of oil, provides a convenient means to change the stiffness of the thrust bearing. Resonance changers require a completely different type of model, as explained by Goodwin (Ref. 3-13)
Couplings Frequently dental tooth couplings will couple the main reduction gear and pinion to the higher speed elements to allow for shaft misalignment and axial displacement of the thrust collar caused by steady-state propeller thrust and thermal growth. Although such couplings are valuable for

neutralizing static misalignment, this type of coupling usually does not decouple the longitudinal vibration because the coupling is locked together as a result of the steady-state torque, and this must be reflected in the model. Locked couplings may cause significant axial force to be transmitted by the gears. The flexibility of the bull gear web can be important when the couplings are locked, and this should also be included in the model. Diaphragm couplings can be used in place of dental tooth couplings to decouple the longitudinal vibration.
Low-Speed Diesel Engines - With low-speed diesel engines, there is no gear box and the propeller shaft is directly coupled to the engine drive shaft. The thrust bearing is usually integrated into the diesel engine itself. The axial dynamics of the crankshaft are important and must be considered.

Usually the thrust bearing is stiff enough to force a node so that the dynamics of the propeller and shafting are essentially independent of those of the crankshaft. If this is not the case then the dynamics of the propeller and shafting are coupled to those of the crankshaft with the thrust bearing stiffness acting between the shafting and the hull, and the whole system should be represented in the dynamic model. Longitudinal excitation of the crankshaft comes about as the result of the connecting rod forces acting on the crankshaft. An axial displacement of the shaft occurs as a result of bending of the cranks. Serious axial vibrations can occur when the axial natural frequency of the crankshaft coincides with the torsional natural frequency. This frequently occurs in long-stroke engines. The dynamic model consists of a series of masses (corresponding to the mass of the shaft between cranks and the mass of the cranks themselves) which are coupled together by springs which represent primarily the axial stiffness of the cranks. Although the masses are easily calculated, it may require a finite element analysis of the cranks to establish accurate axial stiffness values. Calculating the exciting forces requires estimating the combustion-gas force acting on the pistons less the amount of force required to accelerate the piston. From these and certain geometric considerations one can calculate the connecting rod force. The connecting rod force deflects the crank an amount that depends on its angular position, and produces longitudinal vibration of the crankshaft. Frequently, the axial natural frequencies of the crankshaft will lie in the operating speed range. If a longitudinal vibration damper is connected to the forward end of the crankshaft, as is sometimes done, it must be included in the math model. With such a damper, the resonant modes in the operating speed range can be controlled to 20 to 30% of critical. Both torsional and axial dampers are occasionally installed in long-stroke diesel engines. Another issue for diesel crankshafts is the axial motion tolerable between the connecting rods and the associated cranks. These constraints must obviously be honored at all speeds and torques. Since such motion cannot be easily measured except at the end of the crankshaft, analytical methods must be used to calculate that motion. Allowable motion must be supplied by the engine manufacturer.
3.2.3 Operational Factors

Modulation and other features of shipboard vibration data can be characterized by "operational factors" which must be considered in both the longitudinal calculations and measurements required by ANSI S2.27. In Ref. 3-14, the Naval Surface Weapons Center1 Carderock Div. (NSWCICD) evaluated measured data to determine three "operational factors." These result in the actual alternating torque and thrust at the propeller being more than that calculated from a wake survey, which is based on sinusoidal wake variations of constant amplitude. The data studied were longitudinal data, but torsional data seem to be similar. The factors are: The "crest factor" (the peak-to-rms ratio) reflects the variation in magnitude that is characteristic of alternating thrust, and, for conventional propulsion systems, is equal to about 4.0. The "hard turns factor" relates the propeller excitation in a hard turn to that on a straight course, and, for conventional propulsion systems, is equal to about 2.5 or 3.0, depending on the type of stern.

The "full power factor" is applied to the top 10% of the operating speed range to account for cavitation and/or other loading that occurs only at high speeds. It varies from 1.0 at 90% of full power RPM to an empirically derived factor at full power. These factors should be applied to the rms values of the alternating torque and thrust (as obtained from a wake survey) when calculating the expected peak values of alternating shaft stresses, torque across gears, and thrust in the thrust bearing. ANSI S2.27 requires the following operational factors be used for conventional propulsion systems unless data indicate other values are warranted: Crest factor: Hard turns factor: (if required) Full power factor: Single shaft, open stern* Single shaft, closed stern* Multiple shafts, open stern* 4.0 2.5 3.0 3.0 1.5

*Ships tested were Navy ships with these stern types. An open stern has a shaft strut, but no other appendages just in front of the propeller (destroyers, amphibious ships). A closed stern has a skeg that may support the bottom of the rudder as well as the stern tube (auxiliary ships). Note that the crest factor used here, which was derived from alternating thrust measurements, is greater than the crest factor used for hull excitation forces (Section 2.4.3), which was derived from hull vibration measurements. Estimates of alternating thrust can also be obtained from similar ships or typical excitations as given in the literature. Care must be taken to ensure that these estimates are comparable to rms amplitudes as would be calculated from a wake survey. For example, they may be expressed as single amplitudes rather than rms, or they may already take the modulation (crest factor) into account. The above operational factors apply to conventional propulsion systems only. Similar data is not available to obtain the factors for nonconventional systems, so the buyer and builder should agree on what factors to use, depending on the application. In ANSl S2.27, note that some quantities are limited by their rms values, and some by their peak values. Since the rms to peak ratio (crest factor) is about 4.0, it is essential to report the correct quantities.
3.2.4 Measuring Alternating Thrust in the Main Thrust Bearing.

It is common practice to limit the alternating thrust as a percentage of the mean thrust in the main thrust bearing, and to require a measurement of the alternating thrust near the thrust bearing. In ANSl S2.27, the maximum allowable alternating thrust for a straight course is 25% of the mean full power thrust or 75% of the mean at any speed. For turns, it is 50% of the mean full power thrust or 100% of the mean at any speed. The mean thrust increases in turns, but the percentages refer to the mean thrust on a steady course. Various methods have been used to make the measurements, which can sometimes be difficult. Three acceptable methods are discussed.
(1) Strain gage bridges are often applied to the shaft to measure thrust. The dynamic strain values are very low, and, at least with conventional foil gages, the noise in the measurement system may obscure the signal. If semiconductor gages are used, the signals are about 80 times higher, and are usually sufficient in magnitude. The bridge should normally be AC coupled so it does not register the mean thrust, which is much larger. If mean thrust is desired, separate foil gages can be used. To get the required sensitivity, ANSl S2.27 requires that a full bridge of semiconductor gages be used, unless it can be shown that another arrangement will have adequate output. Another difficulty in using strain gages has

to do with alignment. The magnitude of the shaft strains due to alternating torque on most ships is about 10 times that due to alternating thrust. In a thrust bridge, the thrust is measured and the torque is excluded only when the gages are exactly aligned. Calculations can show that, on a typical ship, a thrust gage with one degree of misalignment will register enough of the torsional strain to cause the thrust measurement to be about 10% in error. ANSl 52.27 requires thrust strain gages be aligned within one degree.

(2) Instrumented thrust bearings are installed for many ship trials to measure the mean thrust. If such a bearing is available, it can be used to measure the alternating thrust provided the associated electronics are adequate for the purpose. (3)Relative displacement across the thrust bearing can be measured, and the result multiplied by the stiffness of the thrust bearing to get the alternating thrust. The stiffness must be reliably determined either by analysis or by testing, and should be available from the manufacturer. 3.2.5
Turbine and Diesel Thrust Bearings.

Sometimes the alternating thrust in turbine thrust bearings is a concern. This thrust could be limited as a certain percentage of the mean thrust, but the mean thrust in some turbines is small, and other measures may be more appropriate. The rotor may float between the forward and reverse thrust shoes without enough impact to be harmful. The ANSl S2.27 requirement is to limit the combination of mean and alternating thrust to the continuous thrust rating of the bearing. Excessive alternating thrust may occur in turbine thrust bearings when there is a second longitudinal mode within or close to the blade frequency operating range. In that case, it is reasonable to require that calculations be made of the longitudinal turbine response, and that some measurement be made on the turbines. It is more likely that the second harmonic of blade rate will coincide with the second longitudinal mode within the operating speed range. In this case, as well, it is reasonable to calculate the longitudinal turbine response, and to make a measurement on the turbine. If the magnitude of the propeller excitation at twice blade rate is not available from a wake survey, an estimate of 25% of the blade rate excitation is recommended. For years, there has been a question as to whether or not dental couplings used on geared turbine drives stick and transmit an axial excitation to the turbines. Assuming that they do stick makes analysis more difficult because it is unclear exactly how sticking couplings should be modeled. If a negligible amount of the axial propeller excitation is transmitted through couplings, which would be the case with diaphragm couplings, the only axial propeller generated excitation on the turbine rotor is transmitted to the rotor through axial motion of the turbine casing and via the thrust bearing. For the mathematical analysis of the entire propulsion system, it can be assumed that the turbine rotor has the same motion as the casing. The rotor mass is small enough that it will not significantly affect the results. From that analysis, the motion of the casing is known, and a single degree-of-freedom model excited by the motion of the casing is adequate for the mathematical analysis of the turbine rotor, and for interpreting trial measurements. When making measurements, either the rotor motion, the casing motion, or the relative motion can be measured for any frequency component, and the other quantities, including the thrust in the turbine thrust bearing can be found from a single degree-of-freedom model, providing there are no thrust reversals. Thrust reversals will invalidate the analysis, and cause a lower natural frequency. Measuring the absolute vibration of the rotor has the advantage that the alternating thrust in the thrust bearing can be found by simply multiplying the acceleration of the rotor times its mass. Note that the above discussion does not apply when the couplings stick, in which case the entire longitudinal vibration model may be required to find the thrust bearing force.

The requirements for turbine thrust bearings in ANSl S2.27 apply to the thrust bearings of both gas and steam turbines. For diesels, measuring the crankshaft motion would not necessarily lead to the forces in the thrust bearing, due to the axial flexibility in the crankshaft and other complexities. For that reason, and since there seem to be few problems associated with the thrust bearings on diesels, there is no requirement for limiting the alternating thrust on the thrust bearings of diesels.
3.2.6 Vibration of Propulsion System Components.

Some criteria limit the vibration of propulsion system components. The propeller and a portion of the shafting can experience large vibration magnitudes without deleterious effects, so vibration limits should apply only to inboard propulsion components, and the part of the shafting internal to the thrust bearing and reduction gears. In limiting vibratory motion, there is the issue of which quantity should be limited. There are valid reasons for establishing vibration limits in terms of displacement and acceleration as well as velocity. Machinery degradation can generally be defined in terms of either fatigue of structures or wear of moving parts. Fatigue is caused by high alternating stresses, which are proportional to relative displacements in structures. The magnitudes of vibratory displacement that will cause fatigue on a machine may be unknown, but it is reasonable to establish an upper limit above which experience indicates that problems are likely. Excessive wear can be caused by excessive loads on bearings or other moving parts. Alternating loads are usually proportional to accelerations. Wear can sometimes be related to the relative vibratory velocity between machinery parts, particularly with dental couplings. Velocity limits are often used in the mid-frequency range, with corresponding displacement and/or acceleration limits at the lower and/or higher frequencies. However, velocity limits are also often used by themselves over a wide frequency range SNAME Code C-5 allowed the longitudinal acceleration of propulsion system components to be 0.39. This applies while the vessel is on a straight course, and to the peak value of a single frequency component. To compare guidelines in terms of broadband rms values with narrow band peaks, it may be reasonable to allow twice as much vibration due to the change from narrow band vibration to broadband, and 114 as much for the peak-to-rms conversion. Therefore, a comparable guideline would be a broadband limit of 0.15 g (rms). ANSl S2.27 uses a velocity criteria, 13 mmls (0.5 inlsec) rms, for propulsion system components, which is the same as 0.15 g rms at about 18 Hz. Code C-5 allowed the longitudinal acceleration of the bull gear to be O.lg while on a straight course. This limit is intended as a flag, and higher accelerations were allowed if calculations could show acceptable gear tooth stresses. ANSl S2.27 allows a bull gear vibration of 5mm/s rms, unless it can be shown that the gear tooth stresses are acceptable.
3.2.7 Gear Tooth Stresses.

Excessive gear tooth stresses can be generated by longitudinal vibration. Those stresses can be limited by allowing them to be only a certain percentage of the stress in the teeth due to full power torque (Q). It is not necessary to actually calculate the gear tooth stresses. All that is necessary is to relate the force on the teeth due to longitudinal vibration (Nv) to the force due to full power torque (N The letter N is used for the forces on the teeth, which, because of the lubrication, are normal forces. The gear geometry is shown in Figure 3-1. N has a tangential component (Ft) due to the driving torque, and q an axial component (Fa) due to the tooth angle (8). If R is the reduction gear radius, N can be found q from:

N = Q/(R cos 8) In a double helix gear configuration, the axial force from one row of teeth is opposed by a similar force on the other row, and the normal force is divided evenly between the rows. N is a constant force

perpendicular to the line of contact, but it is transferred to different teeth as the gears turn, making the loads seen by the teeth as alternating. The expression for N in terms of the axial component of the vibratory force across the gears (Fv) is: v

This force is split between the two rows, but they are opposite in sign. The ratio of N to N is: v q N IN = (Fvl sin 8)[(R cos 8)l Q] = (Fv R IQ) cot 9 v q For a particular gear R, 8, and Q are known, so the stress ratio becomes a given parameter times the axial force across the gears. For design, the axial force can be found from the longitudinal calculations. For acceptance trials, the axial force can be extrapolated from measurements, using the calculations. Normally, the measurement most closely related to the axial force is the bull gear shaft vibration, and, from that, the math model will allow the approximation of the force. It will account for the bull gear flexibility, for example, and other physical characteristics that affect the relationship between the bull gear motion and axial force in the gears. ANSI S2-27 allows the peak alternating stress due to longitudinal vibration to be 5% of the stress due to mean torque, unless a different limit is given in the ship specifications.
FOR TORQUE:

W
PINION

t ! /
i

FORCE

1
GEAR TOOTH ANGLE F O R LONG. VIBR:

1
3.3
TORSIONAL VIBRATION

RED'N GEAR

Figure 3-1 - Force Diagram for Gear Teeth

Torsional vibration of a rotating shafting system is an oscillatory rotational movement of the elements comprising the system. This vibration causes torsional stress and strain in shaft sections, possibly gear tooth hammer in reduction gears, variations in voltage and frequency of generated electric power and other undesirable effects which could either cause damage to the equipment or interfere with its proper

operation. Torsional vibration is excited by a periodic torque acting on one or more elements of the system or by a sudden load change in the shafting system. For commercial ships not intended for operation in ice, only continuous steady state vibration is normally considered, while transient vibration caused by sudden load changes in the shafting system is typically neglected. From a torsional vibration point of view, the shipboard shafting system is a free system, i.e. it is not connected at any of its points to the ship structure. Therefore, this vibration neither directly affects nor interacts with vibration of the ship structure or the installed machinery, except through bearing reactions in reduction gears. Also, torsional vibration of a propulsion system can cause significant variation in propeller thrust as a consequence of the variation in the rotating speed of the propeller. This oscillating thrust can then set in motion axial vibration of the shafting system, or, when transmitted through the thrust bearing to the ship structure, can cause vibration of the ship structure or installed machinery. Damage from torsional vibrations is usually associated with excessive alternating shaft stresses or torques, so these are the quantities that are usually limited in torsional criteria. Excessive stress may cause shafts to break, excessive vibratory torque may cause pitting and failure of gear teeth or may damage couplings, and excessive variation of the rotating speed of generator rotors may affect the voltage and frequency of generated electric power sufficiently to cause damage. These and other effects necessitate that torsional vibrations be carefully analyzed before the system is built, considering all operating speeds and all operating modes. Note that propulsion systems operate throughout a speed range, and the torsional vibration must be acceptable for the entire range, while most auxiliary machines operate at a constant speed, simplifying the requirements for torsional vibration. Of practical importance to ship designers are torsional vibrations of propulsion systems, side thrusters, diesel generator engines and other applications with diesel engines as prime movers. Torsional vibration of larger turbine and electrically driven pumps and compressors are not negligible. However, this equipment typically comes on board packaged as a unit, assembled and tested at the manufacturer's plant, and a ship designer normally is not concerned with its torsional vibration characteristics. In shipboard applications, torsional excitations typically originate at the propeller and the diesel engine. The variable torque produced by other machines such as gas or steam turbines, reduction gears, electric motors or generators normally does not affect in any significant way the torsional vibration of the shafting system because the frequencies involved are well above the operating speed range of the shafting system and the magnitudes are relatively low.
3.3.1
Torsional Calculations

Calculations are required by ANSI S2.27 during the design of a propulsion system to show that applicable criteria are likely to be satisfied, and measurements are required to verify that they are satisfied. Calculations include natural frequencies, mode shapes, and responses. For torsional responses, estimates of propeller and engine excitation must be made. In order to mathematically describe torsional vibrations of a shafting system, the system is represented by an equivalent dynamic system consisting of a series of rigid discs representing the inertial characteristics of the system, connected by massless shafting sections representing the torsional stiffnesses of the system. Damping in the system is introduced in the form of absolute damping proportional to the absolute velocity of the inertial discs, and relative damping proportional to the relative velocity between adjacent discs. The excitation of the system is represented by harmonic oscillatory torques of varying amplitude and frequency. The approach and considerations relative to creating an equivalent dynamic system are described in detail in References 3-15 through 3-17. Torsional vibration of an equivalent dynamic system can be represented by a set of differential equations, and torsional angular amplitudes of all the inertial discs in a given system are normally obtained by analyzing either free or forced damped vibrations of the system. The evaluation of other quantities (torque, stress, velocity, acceleration) is based on the calculated torsional angular amplitudes.

Calculations Based on Free Vibrations For many years, calculations of torsional vibrations were almost entirely based on the free vibration analysis. This method employs sets of tables, known as Holzer tables, to obtain natural frequencies of the system. To calculate the natural frequencies only inertias and stiffnesses need to be considered. The system of differential equations is, therefore, simplified and the solution gives the natural frequencies of the equivalent dynamic system and the corresponding relative amplitudes representing the mode shape for that particular natural frequency. Once natural frequencies and mode shapes are known, the actual amplitudes for each disc can be obtained by using either the dynamic magnifier or the energy balance method.

The dynamic magnifier method is characterized by its simplicity. It is not based on strict theoretical considerations, but rather uses known results of similar installations for assessing vibration amplitudes. The basic quantity to be calculated is the equilibrium amplitude, which represents the vibratory amplitude of mass No. I obtained when no dynamic magnification is present in the system. To obtain the actual amplitude on mass No. 1, the equilibrium amplitude must be multiplied by an assumed dynamic magnifier obtained from similar installations. Disadvantages of this method are that the algorithm does not consider the deviation from the undamped free mode caused by the damping and excitation, and the difficulty in estimating dynamic magnifiers. In general, this method is not sufficiently accurate for analyzing installations with modern diesel engines. It can be successfully applied only if the investigated system is similar to systems previously analyzed and measured.
I

The energy balance method is based on the fact that the energy delivered to the system must be equal to the energy that it dissipates. This method is more advanced than the dynamic magnifier method in that it uses actual damping and excitation factors of the system components in calculating vibration amplitudes. However, like the dynamic magnifier method, it does not consider the deviation from the undamped free mode caused by damping and excitation. This method generally gives satisfactory predictions of stresses in the propulsion shafting. The calculated torques in elastic couplings, gears, and PTO (power takeoff) shafts could vary significantly from the measured values. Typically, this method gives good shaft stress predictions in low-speed direct-connected diesel engine installations.
Calculations Based on Forced Damped Vibrations - This approach considers simultaneously the influence of the inertial forces, elastic forces, damping forces, and external excitation forces on the behavior of the system. The calculated amplitudes are not based on the assumed (undamped) vibratory mode shapes, but rather, these mode shapes are obtained as a result of the analysis. The method is capable of accurately predicting vibratory amplitudes at any frequency, at resonances or away from them. This is an advanced method for calculating steady state torsional vibration of shafting systems. The results obtained for different excitation orders at a given speed can be combined by geometrical synthesis to obtain total vibratory torque or stress in any part of the installation.

The computation methods employed in torsional vibration analyses are described in more detail in References 3-18 through 3-22.
Uneven Firing Forces. Due to machining and assembling tolerances, fouling and wear of parts and other effects, the combustion process in the cylinders of a diesel engine is not uniform, but varies from cylinder to cylinder. This difference may result in increased torsional vibration magnitudes when compared to an ideally adjusted diesel engine. This effect must be considered in torsional vibration calculations of all diesel engine propulsion systems. Typically, a 5% to 10% variation of a mean indicated pressure in each cylinder selected in a way to give most undesirable distribution for each considered vibration order would give reasonable assurance that the calculated magnitudes would not be exceeded in normal operation if the engine is kept in good condition.

Many propulsion diesel engines can be readily operated with one cylinder not firing. If a problem develops at sea necessitating fuel shut down to a cylinder, a significant amount of time may elapse before a ship reaches a port where the problem can be rectified. This could significantly increase torsional vibrations, especially the half order vibration in four stroke engines and the first order in two stroke engines. Therefore, ANSI 2.27 requires that, for diesel engines, torsional vibration calculations be performed to cover the engine operation with one cylinder not firing, and the results may indicate a barred

range is necessary for that condition. The ship specifications may require similar calculations for two cylinders not firing.
3.3.2
Torsional Vibration Criteria

Practically all major ship classification societies have limits for permissible torsional stresses in the propulsion shafting. Some societies also have limits for alternating torque, and some for shafts in generator-engine systems. These limits are based on the extensive experience that the societies have and can be taken as a valuable guide in designing propulsion shafting systems. Where a ship is not required to comply with any particular society's rules, the criteria in ANSl S2.27 can be used. With respect to torsional stress, which is the major concern, many of the society's formulas for maximum allowable stress do not clearly indicate the effects of each of the factors involved, such as material strength, size effect, speed factor, and stress concentration. The formula in S2.27 has the advantage that all of the factors are multiplied together, making their effects more obvious. Where the formulas of the societies are in agreement, the S2.27 formula is similar, and where the formulas are not in agreement, research and/or analysis were used to develop a reasonable and usable formula for S2.27. To develop the torsional criteria in ANSl S2.27, regulatory documents from the following authorities were reviewed: U.S. Navy American Bureau of Shipping (Ref. 3-23) Lloyd's Register of Shipping (Ref. 3-24) Germanischer Lloyd (Ref. 3-25) Nippon Kaiji Kyokai (Ref. 3-26) The Navy document cannot be discussed because it has limited distribution, but there are several significant points regarding the other documents: All of the formulas for allowable alternating stress depend on the tensile strength, or require a minimum tensile strength. Stress concentration factors, in all cases, are accounted for by assigning standard factors for different types of shafts and different types of couplings. All of the formulas allow for higher alternating stresses at lower speeds, where the mean stresses are lower. Allowable stresses vary in inverse proportion to speed squared, except for Germanischer Lloyd, where it varies in inverse proportion to speed. All the formulas make an allowance for the size of the shaft, recognizing that larger shafts have lower fatigue limits Less agreement is apparent in the alternating torque limits, where two authorities have quantitative limits, one cautions about gear chatter, and one has no requirements. Only Lloyd's has a requirement for misfiring. Only the Japanese have different limits for 2 and 4 cycle engines. Lloyd's and the Japanese allow much higher limits over 100% of full power rpm, but the Japanese do so for the crankshaft only. Only the Germanischer Lloyd has different factors for oil lubricated shafts and grease lubricated shafts.
3.3.3 ANSI S2.27 Criteria for Torsional Vibration

The ANSl S2.27 criteria for torsional shaft stress are based on the tensile strength as well as on factors for stress concentrations, speed, and size. In general, it is preferable to use tables or curves for stress concentrations rather than using factors based on the type of shaft. For ease of use, however, ANSl

S2.27 includes nominal stress concentration factors for different shaft types. Figures B1 and B2 of the standard give stress concentration factors as functions of the dimensions of a flange or keyway. Quantitative limits for vibratory torque are included in ANSl S2-27. Since cylinder misfires occur on occasion, a calculation of the torsional response during a misfire is required by ANSl 52.27. Also, allowable vibratory stress limits are given for pass-through conditions.
ANSI S2.27 Stress Formula. The allowed torsional vibratory shaft stress in ANSl S2.27 is:

S , < (UTS x K , x Kd x Ks)/(Ktrx Kt, x Ksfx Ksc)

(I)

where: UTS = Ultimate tensile strength K , = Alternatinglmean factor Kd = Size factor = d" Og3 , where d = diameter, in.(mm125.4) K , = Speed factor = 2 - A* (1.0 at full power) Ktf = UTSIFatigue strength Kt, = Tensile stresslshear stress Ksf= Safety factor K ,, = Stress concentration factor A = Actual RPMIFull power RPM A material can withstand a certain combination of mean and alternating stresses. If one is higher, the other must be lower. The "alternatinglmean factor" reflects what portion of the material's strength is "allocated" to resisting alternating stresses rather than mean stresses. For ship propulsion systems, a suggested preliminary design value is 0.5. For most steels, five of the above factors can be approximated by the following (Specific material properties are given in References 3-27 and 3-28): K , = Alternatinglmean factor = 0.5 Ktf = UTSIFatigue strength = 2.2 Kt, = Tensile stresslshear stress = 1.732 Ksf= Safety factor = 1.8 K , = Stress concentration factor = 1.3 for shaft with flange or thrust collar = 1.6 for shaft with spooned keyway (or a crankshaft) When available, the above factors should be derived from supportable test data such as References 3-27 and 3-28. Note that the stress concentration factors are for conventional flanged shaft, spooned keyway designs. Other design features, including radial holes in shafts for controllable pitch actuators should be considered separately. However, with the agreement of the buyer, the above nominal values may be used. With these values, the allowable alternating stresses are: S , < UTS x Kdx Ks/l8, for a shaft with a flange or thrust collar S , < UTS x Kd x Ks/22, for a shaft with a keyway (or a crankshaft) (2) (3)

AlternatinglMean Factor. Reference 3-29 requires Navy ships to have a 20% allowance above full power torque to allow for turns and rough weather, and indicates a further allowance should be made for alternating stresses. We will assume that the alternatinglmean factor is normally 0.5.

,, Figure 3-2 shows Soderberg's failure criteria for a material subjected to a combination of mean stress, S and vibratory stress, S , . It uses a straight line between the yield stress (YS) and the FL. Any point above the line is at risk of failure. Actual failure data points for steels usually fall near a curve above Soderberg's line, and that is especially true for torsional stresses, so the use of this failure criteria is conservative.

The equation for Soderberg's line is:

, I Y S Setting the alternatinglmean factor to 0.5 means that, on the failure line, the ratios of S,/FL and S are both 112. For design, the appropriate factors (Kt,, Ksf,and K , for FL, and Kt,, and Ksffor YS) must be included. In special applications, the manufacturer and buyer may agree to a different factor. For example, if the mean stress is known to be relatively low, this factor could be increased somewhat, and vice versa.

Mean Stress, S ,
Figure 3-2 - Failure Criteria Diagram
Speed Factor. The steady torsional stress increases as the square of the speed, so the allowable alternating stress should decrease as the speed increases, and the speed factor allows for this. , = 0.5, the failure condition is represented by the point designated FP in Figure 3-2. At less than full For K power, the stress limit can be represented by a point, designated <FP (to the left of the full power point). The mean stress at that point is h2 x YSl2. Substituting in the equation for Soderberg's line:

Solving for (S, at CFP): (S, at <FP) = FL ( 1 - h2/2) = (FL /2)( 2 - h2 ) Since (FLl2) is the full power stress limit, (2 - A*) accounts for speeds less than full power.
Size Factor. The reference standards do not agree on the magnitude of the size factor. However, a research report by General Electric (Ref. 3-30) established a size factor on the basis of experimental data. As Figure 3-3 shows, the factor that GE recommends is similar to the one given by ABS and Nippon Kaiji Kyokai, except that the GE factor is higher throughout. Some of the other factors in the ABS INippon formulas may account for the difference. It seems reasonable to "normalize" the curve to a one inch shaft, which is what GE does, and which approximates the size of normal fatigue specimens. This leads to the proposed size factor, d-' OQ3.

Shaft Diameter. inches

Figure 3-3 - Size Factor for Torsional Vibration


3.3.4 Comparing Criteria to Measurements

So far, the discussion applies to sinusoidal, single frequency stresses that can be directly compared to fatigue strength. The actual alternating stress modulates, and may contain several frequency (2), components. The unfiltered peak value of the stress must be compared to the limits in Equation (I), or (3). Limits on alternating stresses apply to free-route conditions, since the number of cycles associated with hard turns will be limited. The number of torsional measurements made on a propulsion system is usually limited by poor access to rotating parts. The measurements may be of torsional motion, or of shaft strain. Ideally, any measurements of the alternating stress, if they are to be measured, would be made near the location of maximum stresses for all major torsional modes of vibration. Unfortunately, the stresses for one mode are often greatest in one location and for another mode in a different location. Refer to the next section to see how to extrapolate measured data (using calculated mode shapes) to find the peak stresses to compare with the criteria. In geared systems, the alternating torque in the gears is limited by the criteria. If the alternating torque is measured near the gears, the broadband peak value can be compared directly to the criteria. However, it is often necessary to extrapolate to find the alternating torque in the gears, just as it is for the alternating stresses.
Extrapolating Torsional Measurements. When torsional measurements are not made at the locations of the highest stresses or torques, extrapolation is done with the calculated natural frequencies and mode shapes. The maximum vibratory torsional displacements, stresses and torques will occur at or near resonances. Each measured frequency component is measured separately, then related to the highest torques or stresses for the mode nearest the frequency component. It is common in the design calculations to relate torsional displacements (if these are to be measured) to the highest stresses for each mode in terms of stress per degree, to facilitate the extrapolation. Broadband measurements cannot be related to shaft stresses through the mode shapes because several different modes are involved.

After determining the highest rms stresses from the measurements, the crest factor must be taken into account. Each propeller excited frequency component of the stress, expressed as an rms value, must be multiplied by the crest factor, and then it must be combined with the other components. Engine excited components are multiplied by a crest factor of 1.414 since they do not experience the same modulation as propeller excited components. The sum is the broadband peak value that can be compared to the allowable stresses given in Equation (I), (2), or (3). If the phases of all components are known, they should be accounted for in combining them. If the phases are not known (such as between the engine

forces and the propeller forces), the combination can be done by taking the square root of the sum of the squares of the individual components. The procedure for alternating torque is similar, except that it need only be found at the gears. The alternating torque is limited by S2.27 to less than +I-75% of the mean at any speed during free-route runs, and less than +I-25% of the rated mean torque (+/-lo% for turbine systems). If turns or bollard runs are required, the alternating torque must be less than +/-loo% of the mean at any speed, and less than +I-50% of the rated mean torque. For systems incorporating generator drives, it is suggested that the generator rotor cyclic irregularity be checked against the permissible limits given in Reference 3-31. For other equipment comprising shipboard shafting systems like elastic couplings, torsional dampers, pump shafts, etc., it is suggested that the equipment manufacturer limits be observed.
3.4 LATERAL VIBRATION OF PROPULSION MACHINERY

If lateral vibration of a horizontal shaft is defined as athwartships or vertical vibration of the shaft, or some combination of the two, when the shaft is not rotating, then the lateral vibration may be analyzed by the same methods used for stationary structures. Whirling shaft vibration may then be defined as the lateral vibration of a shaft which is rotating, and whose behavior is significantly affected by that rotation. Consequently, specially developed techniques are generally needed to analyze whirling vibration, although there are some situations, described below, in which analytical techniques developed for lateral vibration may be useful. The gyroscopic moments generated by the rotation of the propeller are responsible for most of the differences between whirling vibration and ordinary lateral vibration. As a result of these gyroscopic moments, the free end of a whirling shaft generally describes a circular or elliptical orbit about the shaft centerline, whereas a shaft undergoing purely lateral vibration may oscillate in a single plane. The whirling may be in the "forward' direction (vibration of the free end in the same direction as shaft rotation) or in the "reverse" direction (vibration of the free end in a direction opposite to that of the shaft rotation). In general, the forward whirling critical speed is slightly higher than the lateral critical speed, while the reverse whirling critical speed is generally slightly lower. In shafts having long spans between bearings, lateral or whirling vibrations are generally excited by the centrifugal force associated with unbalance of the propeller andlor shaft. These vibrations can also occur in shafts which suffer from alignment problems, resulting in one or more unloaded bearings, and a long span between the bearings actually carrying the load. However, most cases of current interest deal with vibrations excited by hydrodynamic forces on the propeller blades as they rotate in a nonuniform wake. These vibrations generally occur at frequencies which are multiples of the blade rate (the shaft RPM times the number of blades). They are, strictly speaking, whirling vibrations. However, because the rate of rotation is only a fraction of the rate at which the shaft vibrates, the effects of rotation are much smaller than for shaft rate whirling. As a result, some analyses can be performed using standard finite element methods or other techniques which do not account for the effects of rotation. Lateral or whirling vibrations can be a cause for concern when the excitation frequency coincides with a natural frequency of the shaft system, especially if this condition (resonance) occurs at shaft speeds (referred to as critical speeds) which are close to the vessel's rated RPM. Lateral or whirling vibrations can cause excessive bearing wear, wear and leakage of oil seals, and premature formation of fatigue cracks in shafts or hull plates. In such cases, it is especially important to determine the forward and reverse whirling critical speeds as well as the lateral critical speed.

It is general practice in conventional propulsion systems to require the lateral vibration natural frequencies to be above shaft rotational frequency. A perfect ship will not have any shaft rate excitation, but this requirement prevents any misalignment, unbalance, bent shaft, damaged propeller, or other once per revolution excitation from inducing a resonance in the operating range. Other major excitations, such as blade and diesel firing frequencies do not usually provide enough lateral excitation to cause problems. Many shafting systems used in nonconventional propulsion systems, such as cycloidal propellers, thrusters, etc., are different from conventional shafts in that they may include cardan shafts with flexible couplings or universal joints, and they normally do not transmit thrust. They may rotate at a different speed than the propulsor. Cardan shafts are a special case because the Hooke's universal joints used on the ends generate a twice rotational frequency excitation in the lateral direction (as well as torsionally). As a result, the standard requires that fundamental frequencies of cardan shaft sections be above twice the rotational speed. The lateral vibration of diesels, nonconventional propulsors, and separately mounted gearboxes is addressed in the sections on those types of machinery.
3.4.1

Lateral Calculations

It is common to treat lateral or whirling vibration separately from longitudinal and torsional vibration. Since the coupling between lateral and other modes of vibration is generally very weak, this practice is generally justifiable. It is also common to treat lateral or whirling vibration of the shaft separately from the vibration of the hull, representing the shaft supports either as rigid points or simple springs. While this simplification is common practice in an industry that still separates machinery design from hull design, it often leads to significant differences between predicted and observed behavior. In general, as the combined shaftlhull system is modeled more accurately, the accuracy of the predictions improves. Traditional analyses begin with a calculation of lateral or whirling natural frequencies, assuming no damping. The simpler analyses treat the bearings as point supports. This approximation worked reasonably well when water lubricated stem tube bearings were predominant. However, oil lubricated bearings, particularly when subjected to relatively high-frequency excitations, are much more complex. Static or quasi-static analyses of oil film stiffnesses are not sufficient in these situations, since hydrostatic squeeze-film forces become very large at blade-rate frequency. Calculations of the squeeze-film effects in a typical propeller support bearing suggest that at typical blade-rate frequencies, an oil-lubricated white metal bearing acts more like a clamp than like a point support. This would suggest that the magnitude of vibration of the shaft, relative to the bearing, will be very small. However, transmission of vibratory motions and forces to the hull would be greatly increased. Additional research is needed to establish the best representation for this bearing.
A finite element analysis may be used to calculate natural frequencies. As indicated above, it is common for the shaft to be modeled separately from the hull. Best results are obtained by creating a mathematical model of the entire shaft line, up to the thrust bearing and/or main reduction gear. While this may be costly, the model will, in all likelihood, be useful for a variety of analyses, allowing the costs to be shared among them. If only lateral or whirling vibration is to be investigated, a model which includes the three or more bearings closest to the propeller will generally give results of acceptable accuracy.

Several special-purpose frequency calculation methods may also be used, such as those of Panagopulos (Ref. 3-32), Jasper (Ref. 3-33 and 3-34), and Woytowich (3-35). These may be carried out by hand, utilizing a calculator, or may be incorporated into a simple computer program. These assume that the aft most span of shafting has the lowest natural frequency, and only the last two bearings are generally considered in these calculations. Some of these methods require that a rough estimate be made of the flexibility of the inboard shafting, in order to increase the accuracy of the results. When assessing the results of the natural frequency calculations, it is necessary to consider the bias inherent in the chosen frequency calculation method. For example, calculations in which the propeller support bearing is assumed to be rigid, and the inboard end of the shaft is assumed to be clamped, will

generally tend to give predicted critical speeds higher than the actual values. If such a calculation results in a predicted critical speed at, say, 125 percent of rated RPM, it is still possible for the actual critical speed to be below the ANSl S2.27 requirement of 115 percent of rated RPM. Similarly, calculations which tend to model the inboard end of the shaft as simply supported often give results which are lower than the actual values. If a more detailed calculation, treating the bearings as flexible supports, and modeling the inboard end of the shaft with a degree of constraint which lies between fully fixed and simply supported, still predicts that the critical speed will be above 115 percent of rated RPM, then this design criterion can reasonably be said to have been satisfied. If the existence of a potentially troublesome propeller-excited lateral or whirling critical speed is predicted, the designer must evaluate the excitation forces and torques, and must then calculate the expected magnitude of vibration in order to decide whether this situation should be considered acceptable. This evaluation may be performed by rule-of-thumb estimate, or may be based on model test data, hydrodynamic calculations, or some combination of these methods. Calculations of propeller characteristics, such as excitation forces and damping constants, can be found by the methods discussed in Section 2.4. The propeller-induced forces on the hull and the vibratory forces or motions at the shaft bearings may then be used to determine whether excessive hull vibration will result. If the excitation forces and/or torques are large, the stresses in the shaft should also be estimated. These stresses can be compared with classification society limits or available fatigue strength data. ANSl S2.27 requires that the lateral fundamental frequency of each shaft "section" be calculated. "Section" refers to a part of the shafting separated from the other parts by gears, flexible couplings, universal joints, etc. The fundamental mode shape(s) should be calculated as well, and the flexibility of the bearing support pedestals should be considered. Since the shaft rate excitation is usually negligible, response calculations are not required.
3.4.2

Lateral Measurements

ANSl S2.27 requires that the vertical and athwartships vibration be measured at the main thrust bearing and at the line shaft bearing with the highest relative amplitude in the calculated fundamental mode shape. In cases where there is more than one section of shafting separated by gears, flexible couplings, universal joints, etc., the vibration of the bearing with the highest relative amplitude in each section should be measured. If the stern tube or another bearing seems to vibrate more, that should also be measured.
3.4.3 Lateral Criteria

ANSl S2.27 requires lateral natural frequencies to be more than 115% of the rated shaft speed, and that is generally accepted as a reasonable requirement. For cardan shafts, the calculated fundamental frequency must be greater than 230% of the top shaft speed, For vibration magnitudes, other existing criteria for line shaft bearings and stern tubes are given by Det Norske Veritas (Ref. 3-36) and Bureau Veritas (Ref. 3-37). Reference 3-36 gives the following criteria for peak values and for different frequency ranges, implying they apply to each frequency component: 1-2Hz 2 - 100 Hz 0.4 mm pk (16 mils pk) 5 mmls pk (0.2 inls pk)

Reference 3-37 suggests displacement and velocity limits for normal operation (rather than for new machinery) of propulsion shaft bearings as follows: Displacement limit for less than 150 rpm Velocity limit: 100 to 1,000 HP Over 1,000 HP 0.5 mm rms (20 mils) 4.5mmls rms (0.18 inls) 7.OmmIs rms (0.28 inls)

ANSl S2.27 includes a limit of 7 mmls (0.3 inls) rms for lateral vibration of the stern tube, and thrust and line shaft bearings during straight runs. For cardan shafts, the velocity limit is 10 mmls (0.4 inls) rms.

Most lateral vibration problems occur at shaft rate (twice shaft rate for cardan shafts), so the measurements must include those frequencies. If an analysis conforming to the above recommendations shows that shaft speeds up to 115 percent of the rated speed are free of significant critical speeds, the designer may be reasonably confident that the vessel will not experience unacceptable lateral or whirling shaft vibrations in service. On the other hand, predictions of large-magnitude whirling or lateral vibrations have not always been confirmed by actual measurements, particularly in vessels having oil-lubricated white metal stem tube bearings. This may be a consequence of the difficulties encountered in modeling the oil film in these bearings in a manner which accurately reflects their behavior at blade rate frequencies. In these cases, designers should proceed with caution, and make their final design decisions based on experience with similar designs, professional judgment, and any other calculations andlor tests which they believe to be necessary.
3.5 BALANCING

Balancing deals with internally excited vibration, also referred to as self-induced or self-excited vibration. Internally excited vibration is the motion and stress resulting from forces and moments generated within the equipment. Resonance avoidance and quality balancing are key parameters in controlling internally excited vibration. Externally excited vibration, also referred to as environmental vibration, involves motion and stress resulting from sources external to the equipment. Typical sources of external vibration are propellers, shafting, large reciprocating equipment, and ship motions due to operation in a seaway. External vibration can affect equipment that is mounted on a vibrating structure, or is mounted directly on a diesel engine, compressor, or other source, but external vibration is not affected by balancing. Balancing deals with internally excited vibration in rotating machinery. Forces and moments generated within the equipment can be transmitted to the structure that supports the equipment. The magnitude of vibratory motion and stress can be controlled by design methods and manufacturing processes. By identification of the exciting frequencies, control of natural frequencies, and proper balancing, the internally excited vibration forces can be minimized, resulting in improved equipment life. Balancing is a procedure by which the mass distribution of a rotor is checked and, if necessary, adjusted in order to ensure that the vibration of the journals andlor forces on the bearings at a frequency corresponding to operational speed are within specified limits. An improper mass distribution results in an unbalancing force proportional to the square of the speed. The acceptable magnitude of vibration resulting from unbalance is defined by balance quality grades (see next section). Guidance is available to select the appropriate balance quality grades for a wide range of different rotating and reciprocating equipment, rigidly and resiliently mounted. I S 0 194011 (Ref. 3-38) provides useful general information on the subject, and assigns balance quality grades to various types of machinery, including marine equipment. It does not address high quality equipment for naval applications, but ANSI S2.19 (Ref. 3-4) has been modified to encompass a wider spectrum of high quality naval applications. Balancing of rotors depends upon their classification. If a rotor is operating within 30 per cent of any critical speed, it can be considered to be a flexible rotor. Flexible rotors fall into five different categories depending on the details of the system. I S 0 11342 (Ref. 3-39) provides detailed information on flexible rotors. Flexible rotors must be balanced in multiple planes. Rotors that fall outside the flexible category are considered rigid. Balancing of rigid rotors is much easier to accomplish and can be performed either statically or dynamically. Static balancing occurs in a single plane normal to the axis of rotation, and is used for slower speed components and those that do not lend themselves to be dynamically balanced, such as propeller blades. Dynamic balancing occurs in more than one plane and requires the element to be rotated. The classification of a particular rotor may change depending on the application. It is good practice to measure and record the vibration characteristics of the assembled equipment after balancing. While the primary concern of vibration testing is to evaluate the balancing, testing can also

provide insight to the quality of assembly and workmanship by detecting, for example, alignment problems of gears, loose bearings, or poor flow conditions into a pump. For complex pieces of equipment, the significant sources of excitation should be recorded and evaluated.
3.5.1 Balance Quality Grade

The balance quality grade, G, defines the allowable unbalance of a rigid rotor. It is found from, G = w x e where w is the rotational velocity in radlsec and e is the maximum permissible eccentricity in mm. G has units of mmls and is given as G followed by the appropriate value. Balance quality grades are presented for all types of machinery in ISO-194011 (Ref. 3-38). However, its use is not recommended for shipboard machinery because it is too lenient in certain areas. ANSl S2.19-1989 (Ref. 3-4) has recently been revised to include types of shipboard equipment that need to have low noise levels. These include: G6.3: Drive shafts (marine applications requiring low noise) G2.5: Fans (marine applications requiring low noise) Flywheels (marine applications requiring low noise) Medium and large electric armatures (marine applications requiring low noise - over 10 HP) Pump impellers (changed from G6.3) GI: Small electric armatures (marine applications requiring low noise - 10 HP or less) Rotors (marine applications requiring low noise)
3.5.2 Balancing and Vibration Documentation

A balancing report should include the balancing machine used, its minimum sensitivity, the method of calibration, the rotor dimensions and weight, the operating and balancing speeds, the balancing method and planes of correction used, and the initial and final unbalance in each plane. If a vibration test is made in the shop to check the balancing, the vibration report should include the natural frequencies of the mounting system used in the test, transducer locations, rotational speed measured, and the vibratory displacements andlor velocities.
3.5.3 Shipboard Measurements

While balancing is done on a balancing machine, the actual vibration of the machine depends on its foundation as well as the amount of internal exciting forces, so vibration measurements are usually performed on board after installation. Such measurements are usually made on rigid structural elements such as bearing caps. Section 3.8 Auxiliary Machinery discusses ANSl and I S 0 criteria for auxiliary machinery of various types.
3.6 DIESEL ENGINES

Diesel engines have both rotating and reciprocating elements. They generate complex internally excited forces and moments. Engine designers control the phasing of these exciting forces to reduce the amount of vibration which can cause torsional vibration problems in engine crankshafts or in attached components such as shafting, couplings or generators. Additionally, the amount of vibration the engine produces should be considered in the design of components that are mounted to the engine directly or in close proximity, such as gauge boards, piping, pumps and motors. The primary source of excitation is the gas pressure acting on the pistons. The secondary sources are the inertia forces and moments generated by reciprocating and rotating components, namely, the piston, connecting rod, crank, and, if present, the crosshead and piston rod. In addition, there are components such as gears, cams, and pumps which can all contribute to generating vibration. Although the excitation is complex, it repeats as a function of engine rotation. By defining the excitation as a set of Fourier series and using the principles of linear superposition, these complex excitations can be considered in terms of their harmonic components, if the phasing relationships are maintained. Methods of analyzing gas pressures and inertia forces and moments are well developed in the literature (Ref. 3-18, 3-21, 3-40 and 3-41).

Diesel engine exciting torques are analyzed using Fourier series to define the gas pressure tangential effort and the inertia tangential effort for a single cylinder. Gas pressure tangential efforts are calculated by knowing the fundamental engine properties such as the ratio of the length of the connecting rod to the crank radius, and the gas pressure as a function of the crank angle and the number of strokes per cycle. Tangential effort curves can be developed for the different harmonic components and are given in units of Iblsq inch of piston area. These gas pressure forces tend to remain constant over a wide speed range but are affected by engine loading. The inertia tangential efforts are primarily generated by the reciprocating masses, the deadweight associated with the reciprocating and rotating parts and the connecting rod couple. These can also be described by Fourier series with the dominant term being the inertia of the reciprocating masses, which increases by the speed squared. The gas pressure and inertia tangential efforts from a single cylinder, along with the number of cylinders, their arrangement and firing order, are typically used to find the exciting forces for the torsional vibration analysis. The torsional vibration analysis is covered in detail in Section 3-3. While this section does not address torsional vibration, it is concerned with the same excitation forces and their transmittal to the diesel engine structure. A well designed engine achieves a maximum balance of the exciting forces by appropriate use of crankshaft counterbalance weights to balance internal forces and moments and the appropriate selection of cylinder firing order. A well balanced engine will result in less vibration transmitted to the bearing and cylinder wall or crosshead guides resulting in increased life. The vibration magnitudes of the diesel engine and adjacent structure can be quantified to define the operational environment of equipment mounted on or in close proximity to the engine.
3.6.1 ANSl S2.27 Diesel Engine Criteria.

ANSl S2.27 gives definitive criteria for the linear vibration of shipboard diesel engines used for propulsion. The criteria apply to either rigidly mounted or resiliently mounted engines, require calculations of the natural frequencies of the engine on its mounts or foundations, unless it can be shown that similar engines in similar applications have operated satisfactorily. ANSl S2.27 also requires trial measurements, and give criteria that compare to the measurements. These are in addition to the torsional vibration requirements for diesel engines, which are discussed in Section 3.3. The ANSl S2.27 criteria apply to broadband (1 to 1000 Hz) rrns velocities: Less than 1000 HP (High speed diesels) 13 mmls (0.5 inls) rms Bearings 13 mmls (0.5 inls) rms Cylinder head More than 1000 HP (Slow or medium speed diesels) 13 mmls (0.5 inls) rms Bearings 18 mmls (0.7 inls) rms Cylinder head Lower criteria are given for engines that are mounted separately from the propulsors and not subjected to the alternating thrust of the propulsor: Bearings: Cylinder head:
3.6.2 Other Criteria

10 mmls (0.4 inls) rms 13 mmls (0.5 inls) rms

There are several other criteria and perspectives that pertain to diesel engines. Guidance on the vibration magnitude of diesel engines in all applications (not just shipboard) is provided by I S 0 10816-6 (Ref. 3-6). The standard applies to reciprocating engines above 100 kW. It establishes seven categories of vibration severity, defining the magnitudes of displacement, velocity, and displacement that are equivalent in severity, and relies on the ship specifications to state which category is appropriate. These severity categories can be used for condition monitoring as well as for acceptance of new machinery. Over most of the frequencies of interest, velocity is the limiting criterion, and the values of velocity given vary from

about 1 to 100 mmls (peak). MIL-STD-2048 (Ref. 3-42) is more definitive, but applies to naval diesel generator sets, which, on average, will vibrate less than propulsion diesels. It sets the following criteria, as well as higher criteria for "alert" and "alarm" conditions. Broadband Velocity Limits At enaine tor, At shaft bearinqs New equipment <0.25 inls rms <0.34 inls rms Long-term operation 0.34 - 0.92 inls rms 0.25 - 0.6 inls rms Bureau Veritas (Ref. 3-43) recommends the following limits for normal operating conditions (straight course) for diesel engines: High speed engines, <I000 HP On foundation or bearing On cylinder head Slow & medium speed, >I000 HP On foundation or bearing On cylinder head
I Immls (0.43 inls) rms I Immls (0.43 inls) rms I Immls (0.43 inls) rms 18 mmls (0.7 inls) rms

Det Norske Veritas (Ref. 3-36) recommends the following limits for normal operating conditions. The vibration limits are given as "peak" values (single amplitudes), and for different frequency ranges. This implies that the limits are for individual frequency components rather than broadband limits. Corresponding rrns values would be lower than peak, but broadband values would be higher than those for individual components. The overall difference may be small, so the following may be roughly comparable to broadband rrns limits. Note that the limits are in terms of displacement or velocity, depending on the frequency. Large diesels: Transverse Vert & Long Medium speed: All directions 1 - 2.4 Hz 2.4 - 100 Hz 1 - 2.4 Hz 2.4 - 100 Hz 1 - 4.8 Hz 4.8 - 100 Hz
I . O mm pk (40 mils pk) 15 mmls pk (0.6 inls pk) 0.5 mm pk (20 mils pk) 8 mmls pk (0.3 inls pk) 0.5 mm pk (20 mils pk) 15 mmls pk (0.6 inls pk)

Comments from MAN B&W while ANSl S2.27 was being developed indicate that velocities of 25 to 45 mmls (presumably on cylinder heads) would be more appropriate for modern large direct drive diesels. One of the major reasons for diesel engine criteria is to ensure that the engine is properly mounted, whether on rigid or resilient mounts. If the engine is properly mounted and the recommended vibration levels are still exceeded, and if the manufacturer claims that those vibration levels are characteristic of that type of engine and feels comfortable in guaranteeing the engine in that application, it would be reasonable to consider a waiver of that requirement.
3.7 GEAR CASE VIBRATION

ANSI S2.27 contains vibration criteria for gear cases that are not integral units with the driver (engine or motor) or the propulsor (thruster, cycloidal propeller, etc.). Some criteria for gears, such as Refs. 3-44 and 3-45 are intended for vibration from internal sources such as unbalance in the gears, gear tooth excitation, and misalignment. These are normally checked in the shop before the gears are installed. ANSl S2.27 does not cover such shop tests, and the criterion for gears is generally greater than for internally excited vibration by itself. It is intended to apply to installed gearboxes, and to vibration caused by external forces such as propeller forces and engine excitations. It applies only to separately mounted gearboxes because gearboxes that are integral to the driver or propulsor are covered by the criteria for those components. The S2.27 criteria cover both calculations and measurements.

3.7.1

Calculations and Measurements.

Separately mounted gearboxes may not be subjected to alternating thrust forces, but they are subjected to the bearing reactions associated with alternating torque. If the input and output shafts are arranged one over the other, the bearing reactions will be horizontal. If the shafts are beside one another, the reactions will be vertical. If a torsional excitation frequency coincides with a mounting frequency of the gearbox, a resonance can result. To anticipate such problems, the natural frequency of the gearbox on its mounting should be calculated for the expected directions of the applied forces. There should be no natural frequencies in those directions that coincide with the major torsional excitation frequencies. These calculations may be waived if it can be shown that similar gearboxes in similar applications and with similar excitations operate satisfactorily. For measurements, a triaxial array of gages should be mounted on a rigid portion of the top or on a bearing housing near the top of a separately mounted gearbox.
3.7.2 Criteria.

For background, there are acceptance requirements for gears in an ANSIIAGMA standard (Ref. 3-44), in an IS0 standard (Ref. 3-45), and in a DNV report (Ref. 3-36). The first two are intended for shop tests before installation, when the only excitations are those affected by the quality of manufacture, such as gear tooth accuracy and mass unbalance. After the gear is installed, it should be tested again when it can be excited by external forces such as the propulsor and engines, and when the vibration will probably be greater. The requirement in ANSI S2.27 is motivated in part by a 1996 vibration trial, where a separately mounted gearbox had an athwartships vibration of about 2 inls. The standard invoked did not have any gearbox requirements, and the unit had to be accepted. ANSIIAGMA gives criteria for shop tests for individual frequency components, based on pitch line velocity (PLV): Class PLV Peak Velocitv (each freauencv) A <25 mls 12.5 mmls B >25 mls 7.5 mmls The IS0 standard is for broadband vibration (in shop tests), and the gears are classified as "Navy" and "Merchant" applications. The exact level within the following ranges depends on the power transmitted: Navy Merchant rms velocitv (broadband) 3.15-5 mmls 8-12.5 mmls

The DNV guidelines apply to installed gearboxes, and the limits include internally excited and externally excited vibration from 5 to 1000 Hz.They recommend that no individual frequency component be above 5 mmls peak, and indicate that damage is probable if it is above 10 mmls. ANSl S2.27 requires the broadband (1 to 1000 Hz) vibration of separately mounted gearboxes to be less than 10 mmls (0.4 inls) rms when installed, for internally and externally excited vibration together, although it is primarily for the purpose of avoiding excessive vibration due to external excitation. If the internally excited vibration is near or even more than 10 mmls, and still meets whatever gear standard is invoked, it may be reasonable for the buyer and seller to agree to a somewhat higher limit when installed.
3.8

AUXILIARY MACHINERY

Auxiliary machinery is not covered in ANSl 52.27, which deals with propulsion machinery. Rotating auxiliary machinery, which includes the majority of shipboard applications, is covered in two other ANSl

standards, S2.28 (Ref. 3-2) and S2.29 (Ref. 3-3). Diesel engines used for propulsion are covered in S2.27, but other reciprocating machinery, such as diesel generators and compressors, are not specifically covered by ANSI. There are two ANSl standards on the vibration of shipboard auxiliary machinery because ANSl divides guidelines for measurements into (a) those on non-rotating machinery structure, normally bearing housings, and (b) those of the shaft or rotor motion, usually relative to a non-rotating part of the machine. This section gives an overview of the two ANSl standards, which are largely self-explanatory.
3.8.1 ANSl S2.28 Measurements on Bearing Housings.

Guidance for rotating shipboard auxiliary machinery vibration is given in ANSl S2.28 "Guide for the Measurement and Evaluation of Vibration on Machine Structures - Shipboard Machinery," (Ref. 3-2). Main propulsion machinery is specifically excluded in S2.28. Guidance for machinery in general is provided in I S 0 10816 - Part 1 (Ref. 3-46). S2.28 is based on vibration measurements made on various types of machinery on Navy ships and reported in Ref. 3-47. Over 60 types of machines were divided into four groups according to their vibration characteristics as installed in the ship. Different vibration limits are given for each group, expressed as broadband velocities. Displacement limits are also given, in the event there is excessive low frequency vibration, but there is only one displacement limit that applies to all groups. S2.28 differs from S2.27 in that it can be used for machinery condition monitoring as well as for acceptance of new machinery. It establishes four zones, A through D, for monitoring: Zone A - New machines Zone B - Acceptable for unrestricted operation Zone C - Limited operation, repair when possible Zone D - Likely to cause damage to the machine The condition of a machine is reflected in changes in vibration characteristics as well as the actual magnitudes. Therefore, guidance is also given in S2.28 regarding the maximum change that should be tolerated. Recommendations for "alarms" and "trips" are given on the basis of both absolute magnitudes and changes in magnitude. Different guidelines are given for resiliently mounted machinery and rigidly mounted machinery. Guidance on how often measurements should be made for condition monitoring is limited. The frequency of the measurements varies considerably with the type of machinery and how it is used. More on condition monitoring can be found in SNAME T&R Bulletin 3-42 (Ref. 3-48) and many other references.
3.8.2 ANSl S2.29 Measurements on Shafts.

In some machinery, usually large machines with fluid film bearings, it is customary to measure the vibration of the rotating shaft, usually relative to a non-rotating part of the machine. There may not be any machines in this category on a particular ship, and the most likely shipboard application would be main propulsion turbines (which are included in the scope of this standard, even though it is S2.27 that deals specifically with propulsion machinery). In any case, if such measurements are to be made, the machinery specifications should require them, and ANSl S2.29, "Guide for the Measurement and Evaluation of Vibration on Machine Shafts - Shipboard Machinery," (Ref. 3-3) would apply. Guidance for shaft vibration measurements on machinery in general can be found in I S 0 7919 - Part 1 (Ref. 3-49). Shaft vibration measurements are most often taken with proximity gages mounted on non-rotating structure with the shaft as the target. Results are normally expressed in peak-to-peak displacement, and criteria are given for the same four zones as described in the last section. Separate criteria are given for "normal" applications and "special" applications which could include applications where low-noise or

minimal bearing reactions are desired. For condition monitoring, guidelines for changes in vibration magnitudes are given as well as absolute magnitudes. To minimize the probability of wiping bearings, guidelines are also given in terms of displacements as related to the diametrical bearing clearance. For example, it requires the relative peak-to-peak displacement to be less than 25 percent of the diametrical bearing clearance for new equipment, but, for purposes of condition monitoring, allows 50 percent for unrestricted operation. 3.9 REFERENCES (3-1) ANSl S2.27-2002, "Guidelines for the Measurement and Evaluation of Ship Propulsion Machinery Vibration." (3-2) ANSl S2.28, "Guide for the Measurement and Evaluation of Vibration on Machine Structures Shipboard Machinery." (3-3) ANSl S2.29, "Guide for the Measurement and Evaluation of Vibration on Machine Shafts Shipboard Machinery." (3-4) ANSl S2.19-1989, "Mechanical vibration - Balance quality requirements of rigid rotors - Part I : Determination of permissible residual unbalance." (3-5) SNAME T&R Code C-5, "Acceptable Vibration of Marine Steam and Heavy-Duty Gas-Turbine Main and Auxiliary Machinery Plants," September 1976. (Superseded) (3-6) IS0 1081616 - 1995, "Mechanical Vibration - Evaluation of machine vibration by measurements on non-rotating parts - Part 6: Reciprocating machines with power ratings above 100 kW." (3-7) D. vanDort, N. J. Visser, "Crankshaft Coupled Free Torsional-Axial Vibrations of a Ship's Propulsion System," International Shipbuilding Progress, vol. 10, September 1963. (3-8) Kane, J. R. and McGoldrick, R. T., "Longitudinal Vibrations of Marine Propulsion Shafting Systems," SNAME Trans. v 57, (1945) pp 193-252 (3-9) Burnside, 0. H., Kana, D. D. and Reed, F. E., "A Design Procedure for Minimizing Propeller-Induced Vibration in Hull Structural Elements," Southwest Research lnstitute Report to Ship Structural Committee SSC-291, 1979 (3-10) Dorey, S. F., "Strength of Marine Shafting," N.E. Coast Institute of Engineers and Shipbuilders, Trans. v 55, (1938) p. 203 (3-11) Parsons, M. G. and Vorus, W. S., "Added Mass and Damping Estimates for Vibrating Propellers," SNAME Symposium Propeller 1981 ,Paper # I 61 p 273. (3-12) Rigby, C. P., "Longitudinal Vibration of Marine Propeller Shafting," Institute of Marine Engineers, Feb. 1948. (3-13) Goodwin, A.J.H., "The Design of a Resonance Changer to Overcome Excessive Axial Vibration of Propeller Shafting," IME Transactions, Vol. 72 (1960). (3-14) NKF Report 9042-101002. "Evaluation of Longitudinal Operational Factors of MIL-STD-167-2, Type IV," April 1991. (3-15) The British Internal Combustion Engine Research Association, "A Handbook on Torsional Vibrations," Compiled by E.J. Nestorides, Cambridge, University Press, 1958

(3-16) S. Archer, "Torsional Vibration Damping Coefficients for Marine Propellers," Engineering, May 13, . 1955, P ~ s594-598 (3-17) L.C. Burrill and W. Robson, "Moment of Inertia of the Entrained Water by a Marine Propeller," Trans. N.E. Coast Institute of Engineers and Shipbuilders, Volume 78, 1961162 (3-18) J.P. Den Hartog, "Mechanical Vibrations," McGraw-Hill, New York, 1947 (3-19) F.M. Lewis, "Torsional Vibration in the Diesel Engine," Trans. SNAME, Vol. 33, 1925 (3-20) W. Ker Wilson, "Practical Solutions of Torsional Vibration Problems, Vol. 1, Frequency Calculations," Chapman and Hall, London, 1971 (3-21) W. Ker Wilson, "Practical Solutions of Torsional Vibration Problems, Vol. 2, Amplitude Calculations," Chapman and Hall, London, 1963. (3-22) S. Timoshenko, "Vibration Problems in Engineering," D. Van Nostrand Co., Princeton, NJ, 1968 (3-23) "Rules for Building and Classing Steel Vessels," Part 4, American Bureau of Shipping, 1994. (3-24) "Rules and Regulations for the Classification of Ships, Part 5, Main and Auxiliary Machinery," Lloyd's Register of Shipping, January 1986. (3-25) Germanischer Lloyd, "Rules for the Classification and Construction of Seagoing Steel Ships," Volume 1 1 , 1973 edition. (3-26) Nippon Kaiji Kyokai, "Rules and Regulations for the Construction and Classification of Ships, 1986. (3-27) "Standard Handbook of Machine Design," Shigley and Mischke (3-28) "Metals Handbook," 9th edition, Volume I, American Society for Metals. (3-29) Department of the Navy, BuShips, "Design Data Sheet DDS 243-1, January 1960 (3-30) Placek, Williams, Adams, and Klufas, "Determination of Torsional Fatigue Life of Large Turbine Generator Shafts," General Electric Company (Large Steam Turbine Generator Dept.) Report EL-3083, April 1984. (3-31) IEC Publication No. 92,"Electrical Installations in Ships," Part 301 (3-32) Panagopulos, E., "Design State Calculations of Marine Shafting," Trans. SNAME Vol. 58, 1950. (3-33) Jasper, N., "A Design Approach to the Problem of Critical whirling Speeds of Shaft Disc Systems," Taylor Model Basin Report 890, Dec. 1954 (3-34) Jasper, N., "A Theoretical Approach to the Problem of Critical whirling Speeds of Shaft Disc Systems," Taylor Model Basin Report 827, Dec. 1954 (3-35) Woytowich, R., "Calculation of Propeller Excited Whirling Critical Speeds," Journal of Ship Research Vol. 23, Dec. 1979. (3-36) Det Norske Veritas Report 95-0421, "DNV Guidelines for Vibration Evaluation," 4 Dec. 1995. (3-37) Bureau Veritas Guidance Note N1 138 A - RD3, "Recommendations designed to limit the effects of vibrations on board ships," June 1979.

(3-38) I S 0 194011, "Mechanical vibration - Balance quality requirements of rigid rotors - Part I : Determination of permissible residual unbalance." (3-39) I S 0 11342, "Mechanical vibration - Methods and criteria for the mechanical balancing of flexible rotors." (3-40) Harrington, R.L., "Marine Engineering," SNAME, 1992. (3-41) Lilly, L.R.C., "Diesel Engine Reference Book," Butterworth and Co (1984). (3-42) MIL-STD-2048 (SH), "Mechanical Vibration of Naval Diesel Generator Sets," 11 June 1993. (3-43) Bureau Veritas Guidance Note N1 138A - RD3, "Recommendations designed to limit the effects of vibrations on board ships," June 1979. (3-44) ANSIIAGMA 6000-B96, "Specification for Measurement of Linear Vibration on Gear Units," 1996. (3-45) IS0 8579-2, "Acceptance code for gears - Part 2: Determination of mechanical vibrations of gear units during acceptance testing." (3-46) I S 0 10816-1 "Mechanical vibration - Evaluation of machine vibration by measurements on nonrotating parts - Part 1: General guidelines." (3-47) Cautilli, A.J. and DeFriece, P.D., "New Vibration Criteria for Surface Ship Machinery: An Upgrade to Military Standard 167-1," NAVSSES Philadelphia Code 051 B Paper. (3-48) SNAME T&R Bulletin 3-42, "Guidelines for the Use of Vibration Monitoring for Preventive Maintenance," May 1987. (3-49) IS0 7919-1, "Mechanical vibration of non reciprocating machines - Measurements on rotating shafts and evaluation criteria."

Vous aimerez peut-être aussi