You are on page 1of 304

Dynamic behaviour of hydraulic structures

Part A Structures in flowing fluid

P.A. Kolkman T.H.G. Jongeling

Delft Hydraulics

The three manuscripts (parts A, B and C) were put in book form by Rijkswaterstaat in Dutch in a limited edition and distributed within Rijkswaterstaat and Delft Hydraulics. Dutch title of that edition is: Dynamisch gedrag van Waterbouwkundige Constructies. Ten years after this Dutch version Delft Hydraulics decided to translate the books into English thus making these available for English speaking colleagues as well. Because of in the text often referred is to Delft Hydraulic reports made for clients (so with restrictions for others to look at) we decided to limit the circulation of the English version as well. However, the books are of value also without perusal of these reports. The task was carried out by Mr. R.J. de Jong of Delft Hydraulics. Translation services were provided by Veritaal (www.veritaal.nl). Delft Hydraulics 2007

Delft Hydraulics

Preface Table of contents Part A Table of contents Part B Table of contents Part C List of symbols Part A CONTENT

Delft Hydraulics

ii

Delft Hydraulics

Preface The background for composing this book on Dynamic behaviour of hydraulic structures was the ending of the active career of one of its authors, mr. P.A. Kolkman, who was working at the Waterloopkundig Laboratorium (Delft Hydraulics). Rijkswaterstaat put a request to him, to write up his experiences in this field as comprehensive as possible. The request included providing a general introduction concerning the interaction between water and structures, naturally with special attention for those aspects that result in the amplification of the load and in vibrations. Next to flow-induced vibrations, it was decided to also discuss wave impacts in this book. The same applies here, that in order to determine the response, the interaction between water and structures needs to be known. Responsible for the chapters on wave impacts is the books second author, T.H.G. Jongeling, who has been involved in the research in this field at Delft Hydraulics over the last fifteen years. The target group of the book is primarily designers of Rijkswaterstaat, but this naturally does not exclude researchers and experts at Delft Hydraulics. As far as the first group is concerned, they are either designers of structures or clients commissioning the design. For this, it is necessary to have knowledge concerning the role that the design, the strength and the weight of the structure play in safety in relation to the dynamic loads and the response of the structure to these. It is exactly these elements that cause so much insecurity in designing hydraulic structures. Depending on the accepted levels of risk, the scale of the structure and the issue whether the design is an interpolation or an extrapolation based on experience, experts will be enlisted and further research may be required. The subject is treated from the standpoint of hydrodynamics and applied mechanics, in which a certain knowledge of both fields of study is presupposed. The interaction phenomena are discussed on that basis. The subject matter is not treated exhaustively, but the book does present fundamental principles, making the rather specialist literature in this field more accessible. A good overview of the available sources in the field of dynamic loading caused by flowing fluid and flow-induced vibrations may be found in Naudascher (1991 and 1994). Little attention is given to offshore structures; other manuals already extensively discuss the response of these to flow and waves (e.g. Blevins (1990), Hooft (1982) and Patel (1989)). In putting this book together, a choice was made for a subdivision into three parts, which may be used independently. The first part focuses on vibrations caused by flow, the second discusses wave impacts and the third deals with research and calculation methods. The third part goes into these matters more deeply than the first two; by nature, the first two parts constitute a general introduction. When designing structures it is of prime importance that not one single cause of (great) dynamic loads is overlooked. Though arbitrary to some extent, a systematic division into types of dynamic load has therefore been made. Provided the limitations of this are understood, these may be used as a checklist.

iii

Delft Hydraulics

As much as possible, the examples are concerned with structures, concerning which Delft Hydraulics and Rijkswaterstaat have obtained first-hand experience. As most of the structures are situated in the Netherlands, most Dutch readers will be familiar with these. While writing this book, support was given by many. After parts of the initial concept had been written, representatives of Rijkswaterstaat (G.J.M. Hertogh, J.S. Leendertz, H. Verwoert, H. van der Weijde, P.H.A.J.M. van Gelder and J.B.A. Weijers) have commented on the contents of the book and on the accessibility and applicability of the subject matter. At a later stage, an editorial committee was composed from this group, with the task of doing an editorial review. Finally, H.L. Fontijn of Delft Technical University was asked to provide a critical review of the equations and formulations. Delft Hydraulics also provided a lot of support. R.J. de Jong has provided critical comments, both in the initial phase, and in the final stages. In addition, Delft Hydraulics employees have taken care of the graphic and typographic aspects of the book. We thank all these employees for their cooperation and support. They have made it possible to work on the book with enthusiasm and see it through to the end. T.H.G. Jongeling P.A. Kolkman

iv

Delft Hydraulics

Table of contents Part A List of symbols Part A 1 INTRODUCTION ..............................................................................................................1 1.1 General ......................................................................................................................1 1.2 Specification of concepts ..........................................................................................2 1.3 Classification of causes of vibration excitation at gates and at objects in nonblocked flow, as well as self-excitation of fluid oscillations ..................................4 1.4 Summary diagram of excitation types 1 through 5 .............................................13 1.5 Cavitation ................................................................................................................14 2 VIBRATIONS OF STRUCTURES AND OSCILLATIONS OF FLUIDS.....................15 2.1 General ....................................................................................................................15 2.2 Vibrating structures: in dry condition and in water ...........................................16 2.2.1 The mass-spring-damper system equation .......................................................16 2.2.2 The natural frequency.......................................................................................17 2.2.3 The dimensionless damping .............................................................................18 2.2.4 Analysis in the frequency domain ....................................................................21 2.2.5 Calculating in the time domain.........................................................................23 2.2.6 General description of the mass-spring-damper system in water.....................24 2.2.7 Harmonic excitation due to flowing fluid and the response to it......................27 2.2.8 Response to noise excitation ............................................................................28 2.2.9 Structure behaviour in case of multiple degrees of freedom ............................32 2.2.10 Causes of vibrations .........................................................................................34 2.3 Oscillations of fluids ...............................................................................................34 2.3.1 Properties of a fluid oscillator ..........................................................................34 2.3.2 Survey of favourable situations for the existence of a fluid oscillator .............35 2.3.3 Communicating vessels compared to a mass spring system ............................37 2.3.4 Properties of the standing wave........................................................................40 2.3.5 Causes of oscillations .......................................................................................43 2.4 Systems with mechanical components and fluid components ............................44 3 PASSIVE FORCES: ADDED WATER MASS, ADDED DAMPING, ADDED STIFFNESS IN STILL WATER AND FLOWING WATER .........................................47 3.1 General indication of issues ...................................................................................47 3.1.1 Review of the influence of the free water surface ............................................48 3.1.2 Review of the situation with flowing fluid compared to the situation with still water .................................................................................................................50 3.1.3 Calculation of the added water mass ................................................................52 3.1.4 Added stiffness due to immersion and flow .....................................................52 3.2 Added water mass...................................................................................................52 3.2.1 Methods of calculation for two-dimensional situations ...................................54 3.2.2 Added water mass for various two-dimensional structure shapes....................55 3.2.3 The proximity of a wall ....................................................................................58 3.2.4 Polar mass inertia moment in case of rotational vibrations..............................60 3.2.5 Added water mass of a culvert gate..................................................................61 3.2.6 Assessment of the added water mass of a gate as part of a wall of a hemi-space ..........................................................................................................................61 3.2.7 Additional remarks about the added water mass ..............................................64 3.3 Added stiffness ........................................................................................................65 3.3.1 Vertical stiffness in case of a floating body .....................................................65 3.3.2 Stiffness due to flow .........................................................................................65 v

Delft Hydraulics

3.3.3 The sudden stiffness ...................................................................................... 66 3.4 Added damping ...................................................................................................... 67 3.4.1 Damping due to wave radiation ....................................................................... 67 3.4.2 Added damping due to flow in case of an object vibrating in flow direction.. 67 3.4.3 Added damping due to flow in case the vibration direction differs from the flow direction ................................................................................................... 68 3.4.4 Flow-induced damping of a gate...................................................................... 70 3.5 Inertia of flowing water ......................................................................................... 70 3.6 Water-induced coupling of natural vibrations .................................................... 71 4 EXCITATION SOURCES FOR VIBRATIONS AND FLUID RESONANCES AT GATES AND VALVES .................................................................................................. 75 4.1 General regarding gate vibrations........................................................................ 75 4.2 Excitation due to turbulence and due to periodic release of vortices................ 75 4.3 Excitation due to flow instability .......................................................................... 78 4.4 Excitation induced by the movement of the gate itself (self-excitation) ............ 83 4.4.1 General ............................................................................................................. 83 4.4.2 Theory of varying gap width applied to the bath plug valve ........................... 85 4.4.3 Theory of varying gap width in case of vertical gate vibrations...................... 94 4.4.4 Horizontal gate and plate vibrations; theory of the varying discharge coefficient......................................................................................................... 99 4.4.5 Galloping........................................................................................................ 102 4.4.6 Application of an instability indicator ........................................................... 103 4.4.7 Theoretical aspects of vibrations of a gate with overflow (nappe) ................ 104 4.5 Self-excitation in case of fluid oscillations (among which standing waves) .... 107 4.5.1 Instable fluid oscillations due to coupled discharge changes at gates ........... 107 4.5.2 Self-exciting fluid oscillations due to excitation by waves in case of floating flap gates (bottom-hinged gates).................................................................... 111 4.6 Review of small waves generated by the vibrating gates in the upstream water ................................................................................................................................ 115 5 FLOW-INDUCED EXCITATION IN CASE OF FLOW-SURROUNDED STRUCTURES (IN PARTICULAR RODS AND CYLINDERS) AND FLUID RESONANCE AT PUMPS, TURBINES AND DISCHARGE SLUICES................... 119 5.1 General regarding flow-surrounded objects ..................................................... 119 5.2 Angular rods: excitation due to flow-induced turbulence................................ 123 5.3 Flow instability of cylinders and rods with circular cross-section and other cross-sections with rounded corners .................................................................. 125 5.3.1 General ........................................................................................................... 125 5.3.2 The periodicity of the excitation .................................................................... 125 5.3.3 Great excitation perpendicular to the flow..................................................... 126 5.3.4 Excitation with double frequency in flow direction....................................... 127 5.3.5 Dependency of Strouhal number on Reynolds number and roughness ......... 128 5.3.6 Dependency on excitation of the vibration amplitude ................................... 130 5.3.7 Correlation of the excitation in the length direction ...................................... 131 5.3.8 Shift of excitation frequency.......................................................................... 131 5.3.9 Strong variation of forces in-phase with the vibration................................... 133 5.3.10 Quantification of vibration amplitudes .......................................................... 136 5.4 Self-excitation in case of flow-surrounded objects and rods ............................ 141 5.4.1 Galloping; a mechanism of self-excitation in case of vibrations perpendicular to the flow ...................................................................................................... 141 5.4.2 Self-excitation in flow direction .................................................................... 145 vi

Delft Hydraulics

5.4.3 Flutter .............................................................................................................146 5.5 Amplification of the excitation due to fluid resonance......................................147 5.6 Instable fluid resonance caused at pumps, turbines and discharge sluices.....148 5.7 Vibration danger at trash racks ..........................................................................151 5.7.1 Design in relation to sensitivity to vibration ..................................................151 5.7.2 The maximum permissible flow velocity at round and rectangular grid rods 157 6 EXPERIENCES IN PROTOTYPE AND SCALE MODEL INVESTIGATIONS .......167 6.1 Gates above a threshold (free-flow (critical) discharge) ...................................167 6.2 Low-head gates with drowned discharge conditions .........................................177 6.3 High-head gates and other culvert gates ............................................................196 6.4 Valves .....................................................................................................................202 6.5 Seals and leakage gaps .........................................................................................210 6.6 Overflow gates and flap gates ..............................................................................221 6.7 Stop logs .................................................................................................................230 6.8 Trash racks............................................................................................................233 6.9 Dynamic phenomena due to cavitation and partial aeration ...........................236 7 RECOMMENDATIONS FOR THE PREVENTION AND COMBAT OF VIBRATIONS ........................................................................................................................243 7.1 General ..................................................................................................................243 7.2 Overall design .......................................................................................................244 7.2.1 Gates ...................................................................................................................244 7.2.2 Rods and trash racks ...........................................................................................248 7.3 Gate edges and seals .............................................................................................249 7.4 Stiffness ..................................................................................................................250 7.5 Damping and friction ...........................................................................................253 7.6 Aeration to prevent vibrations and cavitation ...................................................254 8 REFERENCES ...............................................................................................................257 8.1 Delft Hydraulics Reports (in Dutch) ...................................................................257 8.2 Other reference material .....................................................................................265 APPENDIX ............................................................................................................................273 THE FREQUENCY AT WHICH, IN CASE OF A VERTICALLY VIBRATING GATE, THE FLOW-INDUCED DAMPING CHANGES INTO A FLOW STIFFNESS EFFECT INDEX BY SUBJECT (PART A) .........................................................................................281 Biographies of the authors......................................................................................................288

vii

Delft Hydraulics

Table of contents Part B (Structures in waves) List of symbols Part B INTRODUCTION WAVES 2.1 Wave phenomena 2.2 Wind-generated waves 2.2.1 Wave characteristics 2.2.2 Wind-generated waves as a stochastic process 2.2.3 Reflection, refraction and diffraction 2.2.4 Forecasting of wind-generated waves 3 WAVE LOADS 4 QUASI-STATIC WAVE LOAD 4.1 Analytical calculation 4.1.1 Slender structures 4.1.2 Non-slender structures 4.2 Numeric calculation 4.3 Scale model investigation 4.4 Influence of flow 4.5 Response of structures to quasi-static wave load 5 WAVE IMPACT LOAD 5.1 Wave impacts in relation to the structure 5.2 Wave impact characteristics 5.3 Influencing factors 5.4 Types of wave impacts 5.5 Analytical calculation of maximum impact pressures 6 RESPONSE OF STRUCTURES TO WAVE IMPACTS 6.1 Important properties of the structure 6.2 Response of a single mass spring system to an impulse load 6.3 Influence of the response on the impact load 7 EXPERIENCES WITH WAVE IMPACTS IN PROTOTYPES AND SCALE MODELS 7.1 Storm surge barrier Eastern Scheldt. Grid gate design (scale model investigation) 7.2 Storm surge barrier Eastern Scheldt. Caisson design (scale model investigation) 7.3 Storm surge barrier Eastern Scheldt. Design with piers and lifting gates (scale model investigations and prototype measurements) 7.4 Discharge sluice Haringvliet (scale model investigations and prototype measurements) 7.5 Cooling-water intake of Alto Lazio nuclear power station (scale model investigation) 7.6 Civitavecchia caisson breakwater (scale model investigation) 7.7 Asphalt slope (scale model investigation) 8 GENERAL DESIGN DIRECTIVES 9 REFERENCES 9.1 Delft Hydraulics Reports 9.2 Other reference material INDEX BY SUBJECT (PART B) viii 1 2

Delft Hydraulics

Table of contents Part C (Methods of calculation and experimental investigation) List of symbols Part C 1 2 3 INTRODUCTION ANALYSIS OF DESIGN AND CONDITIONS CALCULATION METHODS FOR DYNAMIC BEHAVIOR OF STRUCTURES IN WATER 3.1 Considerations for calculating in the frequency domain or in the time domain 3.2 Calculation of the added water mass for two-dimensional situations (ignoring wave radiation) 3.2.1 Assessment of the frequency range with the added water mass is not frequency-independent anymore 3.2.2 Added water mass calculated with potential theory 3.2.3 Assessment of added water mass based on schematized flow pattern 3.2.4 Examples of complete calculations 3.3 Calculations in the frequency domain 3.3.1 The single (degree of freedom) mass spring system 3.3.2 The single (degree of freedom) mass spring system in water 3.3.3 Response of a double (degree of freedom) mass spring system (direct method) 3.3.4 Modal Analysis in case of a double (degree of freedom) mass spring system 3.3.5 General formulation of a system with multiple degrees of freedom 3.3.6 Coupled systems with structure components and fluid components 3.3.7 Examples of coupled systems with structure components and fluid components 3.4 Calculations in the time domain using the indirect method 3.4.1 Modal Analysis and impulse-response method in case added mass and damping are frequency-independent 3.4.2 The impulse-response function in case added mass and damping are frequency-dependent 3.5 Calculations in the time domain with the direct method 3.5.1 General 3.5.2 Response of a single mass spring system to an external load 3.5.3 Coupled systems with structure components and fluid components 4 CALCULATION METHODS FOR WAVE IMPACTS 4.1 General 4.2 Impulse theory 4.3 The linear shock wave model 4.3.1 Wave impact against a rigid wall 4.3.2 Rigid wall and air-water mixture 4.3.3 Wave impact against a compressible wall 4.4 The non-linear shock wave model 4.5 The flow-pressure model (ventilated shocks) ix

Delft Hydraulics

4.6 The air compression model 4.6.1 The linear air compression model 4.6.2 The non-linear air compression model 4.7 Numeric calculation of the pressure function (in time) in case of a wave impact 4.8 Extrapolation of results from a scale model to prototype values 4.9 Influences on impact load due to a responding structure SCALE MODELS 5.1 Introduction 5.1.1 General 5.1.2 Investigation strategy for a real project 5.2 Scale rules and scale effects in case of vibration investigations and investigations of wave loads 5.3 Classification of scale models for vibration and wave impact investigations 5.3.1 Classification with respect to reproduction of geometry 5.3.2 Classification with respect to reproduction of dynamic properties 5.4 Possible critical aspects of dynamic models 5.5 Verification of the model technology 5.6 Measuring system and data acquisition 5.6.1 General 5.6.2 Instrumentation 5.6.3 Monitoring and registration of measuring signals 5.7 Processing of measuring results 5.7.1 General 5.7.2 Statistical processing 5.7.3 Elaboration in the time domain 5.7.4 Elaboration in the frequency domain EXAMPLES OF SCALE MODELS FOR DYNAMIC INVESTIGATIONS 6.1 Rigid model with flowing water, for vibration studies 6.2 Rigid model for investigation of wave load 6.3 Single (degree of freedom) mass spring system for vibration investigations; translatory 6.4 Single (degree of freedom) mass spring system for vibration investigations; rotating 6.5 System with multiple degrees of freedom in case of floating gate 6.6 Multiple degree of freedom mass spring system for response investigations in case of flow and waves 6.7 Continuous-elastic model for vibration investigations 6.8 Continuous-elastic models for investigations of wave loads and vibrations PROTOTYPE TESTS OF STRUCTURES 7.1 General 7.2 Vibration measurements 7.3 Wave impact investigations 7.4 Elaboration of measuring results 7.5 Experiences with respect to vibration and wave impact measurements REFERENCES 8.1 Delft Hydraulics Reports 8.2 Other reference material

Delft Hydraulics

APPENDIX I

DEDUCTION OF THE RESPONSE FUNCTION IN THE TIME DOMAIN FROM THE TRANSFER FUNCTION IN THE FREQUENCY DOMAIN SCALE RULES AND SCALE EFFECTS FOR INVESTIGATIONS OF DYNAMIC BEHAVIOR OF SCALE MODELS DESCRIPTION OF A CALCULATION PROGRAM FOR THE DETERMINATION OF THE ADDED WATER MASS FOR A FLAT ROD WITH RECTANGULAR SECTION (STRIP) IN WIDE WATER IN CASE OF TRANSLATORY AND ROTATING VIBRATIONS

APPENDIX II APPENDIX III

INDEX BY SUBJECT (PART C)

xi

Delft Hydraulics

List of symbols Part A a a A A Aa Ac c ce cw C C' Ci Ck Cw CL CL Cm Cm Cr Cs d d' D e E Ek Ep f fn fR fr F FL FW F perm Fy Fw1 Fw2 F g h h1 h2 k kw xii = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = half thickness of symmetric body (m) opening of gate or gate (m) amplification; ratio between spring force and external force = kY/F (-) surface area (m2) front surface area (m2) culvert cross-sectional area (m2) damping (Ns/m) equivalent linear damping (which equals the non-linear damping) (Ns/m) damping due to water, also called added damping (Ns/m) force coefficient (-) dynamic part of force coefficient (-) coefficient of flow inertia ratio between stiffness of the bath plug suspension and the sudden negative fluid flow stiffness (-) drag (resistance) coefficient (drag is in flow direction) lifting coefficient (lift is a force perpendicular to the flow) (-) length coefficient related to flow inertia (index u relates to the upstream part and index d relates to the downstream part) (-) coefficient of added water mass (-) ratio between mass of bath plug and water mass in tube (-) coefficient of linear flow resistance (resulting from wave radiation) (-) suction force coefficient (-) water depth, bar thickness, strip thickness, gate height, opening underneath gate (m) strip width projection perpendicular to the flow (m) diameter of cylinder (m) length of strip diameter (m) energy of a spectrum = time-average value of y'2 (m2) kinetic energy (Nm) potential energy (Nm) vibration frequency; dominant excitation frequency (Hz) natural frequency (Hz) resonance frequency of standing wave (Hz) reduced frequency = fnL/V (-) (external) force (N) lifting force (perpendicular to the flow) (N) drag (resistance) force (in main flow direction) (N) flow force at permanent conditions (N) force component in y-direction (N) force from water, not influenced by vibration (N) force from water, coupled to vibration (N) amplitude of periodical force component (N) gravitational acceleration (m/s2) water depth, culvert height (m) upstream water depth downstream water depth stiffness (spring constant) (N/m) hydrodynamic stiffness, or added water stiffness (N/m)

Delft Hydraulics

K0 L LC Ld Lu Lw m m mw n p p0 q qp qv Q Qp Qs r R Re RMS s S S' Sc Sn t T u U0 v V Vr Vr' w W W(f) x y y' Y Yn z z Z

= = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = =

compression module (distortion per unit of pressure) (m2/N) length, length of culvert, basin etc., also reference length (m) perimeter of the bath plug gap (m) downstream length of the (fictive) tube which is a measure for fluid flow inertia (m) upstream length of the (fictive) tube which is a measure for fluid flow inertia (m) representative length of added water mass (m) mass (kg) discharge coefficient (-) added water mass (kg) whole number (-) pressure (N/m2) reference pressure (N/m2) discharge per unit of width (m2/s) tube discharge per unit of width (m2/s) feeding discharge per unit of width (m2/s) discharge (m3/s) tube discharge (m3/s) gap discharge (m3/s) radius (r is a variable) (m) radius (m) Reynolds number = VL/ (-) Root Mean Square= mean time value of squared signal distance between bars (m) Strouhal number = fL/V (-) Strouhal number related to the strip width d' (-) projection perpendicular to the flow Scruton number (defined in A5.7) reduced frequency (Strouhal number related to resonance frequency) = fnL/V time (s) period of harmonic movement or wave period (s) velocity component in x-direction (m/s) reference velocity in x-direction (m/s) velocity component in y-direction (m/s) upstream fluid flow velocity or reference velocity (m/s) reduced (fluid flow) velocity = V/(fnL) or V/(fnD) reduced velocity related to the strip width d' (-) projection perpendicular to the flow velocity component in z-direction (m/s) drag (resistance) force (N) spectral density (dimension depends on magnitude on which spectrum is based) distance in main direction of a coordinate system distance in the second (horizontal) direction of a coordinate system dynamic part of y (sometimes y' is also indicated by y) (m) amplitude of vibration y (m) amplitude of nth vibration peak (m) distance in the third direction of a coordinate system (vertical) (m) dy/dt (m/s) or water level variation in case of oscillation (m) amplitude of water level oscillation (m) xiii

Delft Hydraulics

d E H
h t v s n

= = = = = = = =
= = = = = = = = = = = = = =

angle of approach flow (radials) proportion constant between feed discharge qv and water surface variation z (m/s) relative damping: = c /(2 km ) wall thickness, gap width (m) logarithmic decrement = e log(Yn / Yn +1 ) displacement thickness of boundary layer at the wall surface (m) increase of energy per period (Nm) hydraulic head difference over total structure, local hydraulic head difference of energy level (m) water level difference (m) time interval between two peaks, time increment (s) vibration velocity or difference in water velocity (m/s) change of angle of approach flow (radials) wave length (m) discharge coefficient or contraction coefficient kinematic viscosity (m2/s) density of a fluid (kg/m3) density of steel (kg/m3) density of normalised spectrum (i.e. the total surface area of the spectrum = 1) flow potential (m2/s) phase angle (radials) angular frequency of the vibration (radials/s) angular natural frequency of the vibration (radials/s)

Symbols used to assign a special property to a variable: (subscript) initial value (before vibration starts) 0 ~ indicates periodical variation of a quantity ' part of magnitude with movement in time ^ amplitude of movement in time _ (underlining) matrix presentation (overlining) mean value of quantity over time first derivative with time second derivative with time

xiv

Delft Hydraulics

INTRODUCTION 1.1 General

When designing structures in flowing fluid, it is desirable and sometimes even a prerequisite to pay attention to the possible occurrence of vibrations. Experience shows that building cost-effective structures often means: applying lighter and more slender structures that are more sensitive to dynamic loads and vibrations. Special attention for dynamic behaviour is particularly necessary when dealing with a new type of design or in case of extrapolating dimensions or circumstances of flowing fluid. Right from the start of the design process, it would be advisable to pay close attention to the possible danger of vibrations, because then at that point, important choices may still be possible. As regards vibrations, as an example, it is of great importance whether a choice is made for a gate in an open or in a closed culvert. The same applies to the choice for a large-scale gate that is introduced afloat into a situation of flowing fluid, or for gates with an open-plate structure and heavier hoisting equipment. The scale is also important: one large gate or a gate composed of several smaller ones. There isnt a single structure in flowing fluid that is vibration-free: therefore a guarantee is required, ensuring that acceptable limits are not exceeded. The latter will depend on the extent of excitation, the response behaviour and the strength of the structure; all this, combined with fatigue reviews. An understanding of flow-induced vibrations may lead to a considerable improvement of the safety of the design and the maintenance, because: A good design is safer and requires less maintenance; the safety relates to the structure itself, but also to its surroundings. An increase in scale relative to existing designs is more sound in case the consequences in the field of dynamics may be assessed as well. For structures that are very deep under the water surface, or otherwise difficult to access, there may be unfortunate consequences, when vibrations occur that are detected too late.

Next to flow-induced or wave-impact-induced vibrations, dynamic phenomena may also be generated by other causes. The flow may be instable, as an example when the flow of nearby water shows tendencies of separation. Air enclosed in a closed pipe may also cause instable flow. Moreover, the elasticity of air, together with the water mass in the part of the pipe that is full of water, forms a mass spring system with a certain resonance frequency. This system may become resonant. In all those cases in which a wave reaches a structure and cannot move away, the water is suddenly slowed down, resulting in a wave impact. Wave impacts are not only caused by wind-generated waves; because of its movement, any moving water surface that hits a parallel wall (a water level variation caused by a passing vessel may hit the ceiling of the culvert), may result in very high pressures. Other causes of dynamic loading are: cavitation, hammer shock phenomena and the transition between flow with a free surface and flow in which the ceiling plays a role. A moving hydraulic jump may also cause a dynamic load. If an aeration pipe becomes clogged 1

Delft Hydraulics

up with ice, it may considerably change the flow elsewhere in the system and, as an example, cavitation may occur. For the assessment of the dynamic behaviour of structures in flowing fluid, a number of aspects are essential: It may be the case that the dynamic load that causes the vibrations has an external cause, but there are also vibrations that are caused by accidental circumstances, as it were, and subsequently intensify by themselves, reaching very high amplitudes; in case of flow-surrounded structures, both types of vibration occur. Vibrations that intensify by themselves may be extremely strong, especially in case of valves and gates (consider as an example the vibration of a pipe, when the washer of a leaking tap needs replacing). In case of a structure in water, the structure behaves like a system of which the mass, stiffness and damping differ considerably from those of a structure in air. The term virtual mass (stiffness and damping, respectively) is used for the total of the mass of the structure itself and the so-called added water mass. Both the added mass, the added damping and the added spring stiffness may be far greater than that of the structure itself. In case of a thin strip or plating that vibrates perpendicular relative to its plane, the added water mass equals the water mass of a flow-surrounded cylinder. In case there is a wall near the strip, this mass may even be considerably greater.

1.2 Specification of concepts


Added water mass. A measure for the force exerted by the water on a vibrating structure as far as this force is in-phase (see below) with the acceleration of the vibration. The added water mass then equals this force divided by the (vibration) acceleration. Anti-phase. The signal is proportional to the harmonic reference signal, but with opposite sign. Cavitation. Boiling of a fluid substance at normal temperatures as a consequence of very low pressure. As this low pressure may occur short-lived and very locally (turbulence), bubbles continuously grow (explosion) and diminish (implosion) again. If implosion occurs in the proximity of a wall this results in cavitation damage. Drowned discharge. Discharge influenced by the downstream water level. Drowned nappe. Situation in which a nappe issues into wide surrounding water with no or relatively weak flow velocities. Dynamic pressure. Defined as 1 2 V 2 , with representing the density of the fluid and V the (undisturbed) approach flow velocity. Free boundary layer. Transition zone between an area with continuing flow and an area with no flow or an eddy current. Free-flow discharge. Situation in which the discharge is not influenced by the downstream water level. Free runoff. Situation in which the water is discharged into the air. Frequency synchronisation (or lock-in). In case of a greater amplitude of the vibration, the frequency of the flow-induced excitation shows a tendency to shift toward the resonance frequency of the structure, resulting in an even stronger vibration. Froude number. Dimensionless number V / gL (V = velocity, L = length, g = gravitational acceleration) indicating the relative influence of the hydrostatic pressure differences on the

Delft Hydraulics

flow relative to the dynamic pressure; in case of a high Froude number the influence of dynamic pressure is great. Froude scale. The scale on which, in a scale model with free water surface and dominant influence of gravity, time and celerity are reproduced. Galloping. This excitation mechanism is named after a phenomenon observed with electricity wires enveloped with frozen rain, where very large amplitudes (of several meters) occurred as a consequence of the wind. The mechanism is related to the change in flow forces due to a change in the angle of approach during the vibration. Hydraulic jump. A sudden water level difference, at which fast flowing (super-critical flow) water changes into normal flow (sub-critical flow). Super-critical flow is referred to when the Froude number V/gd is greater than one (V = water velocity and d = water depth). In-phase. The signal is proportional to the harmonic reference signal. Movement-induced excitation. Excitation in the course of which the alternating force results from the vibration and intensifies with the amplitude thereof. In that case, self-excitation occurs (see below). Natural frequency. The frequency of the vibration that remains after excitation no longer occurs, assuming there are no damping elements. Noise excitation. Dynamic load, which is not purely periodic, but for which the excitation is spread out across a certain frequency band. Out-of-phase. The signal is proportional to the harmonic reference signal, but running at a time period faster or slower (or: phase difference is + or 90). Phase synchronisation. Process in which, due to the vibration of a structure, a strongly correlated and synchronic flow-induced excitation is generated. Reduced frequency. Dimensionless presentation of the natural frequency, fn, by multiplying with a length measure, L, and dividing by the approach flow velocity, V. So: f r = f n L / V . Resonance frequency. With this frequency, the excitation gives the maximum amplification of the system. Reynolds number. Dimensionless number, defined as VL / , with V = approach flow velocity, L, a representative measure of length and , the kinematic viscosity of the fluid. The Reynolds number is a measure for the influence of the viscosity on the water movement; a greater Re value means a smaller influence of the viscosity. Self-excitation. Situation in which the vibration or the fluid oscillation itself is the cause of the generation of a periodic excitation. Spectrum. A method of representing a signal if this is spread out across a certain frequency band (as is the case with noise excitation). It is customary to work with an energy density spectrum. Each little piece of the spectrum with bandwidth f (f = frequency) and value E then equals a periodic signal with amplitude Y, in such a way that the time-average value of Y 2 equals E f . Strouhal number. Dimensionless number representing the dominant frequency, f, of the flowinduced excitation. The Strouhal number (S) is defined as fL / V , with L as a representative measure of length and V representing the approach flow velocity. Turbulence. The extent to which flow properties such as celerity and flow pressure deviate from mean values, expressed in a relative standard deviation and a frequency distribution. Vibration mode. The offset from the zero position in case of maximum deformation of the structure in case of a free vibration, made dimensionless by dividing it by the vibration amplitude at the location of the maximum.

Delft Hydraulics

1.3 Classification of causes of vibration excitation at gates and at objects in non-blocked flow, as well as self-excitation of fluid oscillations
In order to clearly identify all possible types of causes for the assessment of the danger of vibration, Naudascher (1991) has compiled a checklist of possible causes that may generate dynamic flow excitation and vibrations. At a later stage he has added considerable refinements to this arrangement (Naudascher, 1994), but these have not been used in the following. Before fully presenting the summary diagram (Paragraph 1.4), we will first discuss the different sources of excitation separately on the basis of the same diagrams that appear in the summary diagram. The enumeration of types of flow-induced excitation as presented by Naudascher, is completed with a type called self-excitation in case of fluid resonance, to be able to differentiate between vibrations in the resonance frequency of the structure and lowfrequency movements in the resonance frequency of a standing wave or of communicating vessels. For each type of vibration excitation, a distinction is made between structures in free flowing fluid and gates that more or less block the flow. The most important causes of vibrations may differ considerably for both types. Below, the excitation sources are further explained on the basis of the examples of the diagram, in which the numbers 1.1 through 1.5 and the letters a and b of the summary diagram are used. The summary diagram is presented in Figure A1.6. In Chapters 4 and 5, in which each of these types of excitation is further discussed, the arrangement of Figure A1.6 is used. An advance warning seems appropriate: first of all, each arrangement includes random elements. Moreover, using the summary diagram in no way guarantees the discovery of all causes of flow-induced vibrations. Type 1. Excitation by initial turbulence and by turbulence in the wake (including the vortex trail behind the structure) The initial flow may be very turbulent, causing a dynamic load upon the structure. For the structure it is important what magnitude the amplitudes of the pressure fluctuations are, and whether the dominant frequency of the excitation is close to one of the resonance frequencies of the structure. When the initial turbulence is only generated due to bottom and wall friction, there is little danger to think of; the amplitude of the pressure fluctuations only amounts to a few percent of the dynamic pressure and the frequencies are spread out across a large bandwidth. The presence of the structure itself also causes turbulence, but this occurs more specifically in the wake downstream. The excitation caused by this is greater and, because a so-called vortex trail is generated, may be almost periodic. 1a The turbulence in the approaching water is very irregular and there is little if any periodicity in the load. In the wake, however, this is different. In case of a prismatic object or a cylinder in flowing fluid, the diameter is normative for the dimension of the vortices generated in the wake. Therefore, the excitation frequency is determined by the approach 4

Delft Hydraulics

velocity and the width of the diameter. The excitation frequency is expressed in the Strouhal number (S = fD/V, with f = dominant excitation frequency, D = prism width or cylinder diameter and V = approach velocity). The Strouhal number is a constant, but it does depend on the shape of the diameter of the object and of the Reynolds number (a measure for the influence of the viscosity of the water on the flow pattern). The latter dependence only applies to situations with an instable flow pattern, such as those that occur with circular cylinders. The amplitude of the pressure variations is proportional to the dynamic pressure 2 (V ). In case of circular cylinders, the flow excitation is almost periodic and in case of more extended rods or cylinders, the excitation frequency may easily approach the resonance frequency.

Figure A1.1a: (a) rod / beam Excitation due to initial turbulence in a wake. 1b In case of gates, the excitation due to turbulence essentially is no different from that of prisms and cylinders. A gate however is a non-prismatic, complex structure with elements that strongly vary in shape and dimension. In case of a trottled gate, strong turbulence is generated, because the water has high flow velocities locally. In case of gates, the excitation due to turbulence usually is of little significance, because as a rule the resonance frequency is high in relation to the frequency of the flow-induced excitation and because the excitation is spread out across a wider frequency band.

Figure A1.1b: (b) regulating gate or valve Excitation induced by turbulence in the wake of a culvert gate.

Delft Hydraulics

Type 2. Excitation due to flow instability This category, not always clearly distinguishable from Type 1, includes those instabilities that are coupled to the specific shapes of the object. Flow instability is related to the instability that occurs at the separation of flow and in cases when the flow attaches again further on. In case of round shapes, the point of separation is undetermined, so that the flow is very instable in those places. As the flow field (and thereby the pressure) strongly depends on this point of separation, small effects (such as a boundary layer development that differs, depending on the Reynolds number, on the wall roughness and on fouling) may cause big differences in the force of the flow. Also, a vibration of the structure may cause the point of separation to shift periodically, generating a change of force, which is triggered exactly at the point of the resonance frequency. Because the separation process is instable, the issue whether the flow attaches or not may be undetermined as well. As attachment in the zone between the points of separation and attachment generates a low(er) pressure, a strongly alternating flow causes great variations in force. There are very many situations in which flow instability may occur. A special type of flow instability occurs at a gate or a fixed gate edge with overflow, in which an air cushion is enclosed between the wall and the falling nappe. This occurs in case of a two-component system of water/air that may behave in an instable manner. 2a In case of a cylinder with a circular diameter, the separation of flow is instable on both sides. Because of this, a situation is caused in which the flow on one side also influences the flow on the other side. Flow is generated with very strong, alternatingly separating left and right vortices, the so-called Von Krmn vortex trail. This one causes, also in case of a stiff circular cylinder, a big and largely periodic load. In case of a vibrating cylinder, the vibration itself may still amplify the excitation, due to the fact that it becomes more synchronous across the length of the cylinder. Moreover, the excitation frequency starts running in line with the resonance frequency (frequency synchronisation). In the direction perpendicular to the flow, the amplitude of the force excitation (per unit of surface) in case of a circular diameter, nearly equals the dynamic pressure. This is referred to as a dynamic lifting coefficient equal to 1. The Strouhal number is, only in cases of a circular diameter, approximately 0.2. Paragraph 5.3.5 further discusses the flow-induced excitation due to flow instability of circular cylinders with other diameters. In case of long rods with circular diameters, amplitudes of vibrations of one time the diameter may occur. As regards the force in the flow direction, this is referred to as a dynamic drag (resistance) coefficient. In case of a high Reynolds number, often occurring with prototype structures, the boundary layer at the cylinder becomes so turbulent, that the excitation is no longer purely periodic and spreads out across a certain frequency band. In case of extremely high Reynolds numbers, the turbulence boundary layer becomes thinner again, generating an almost periodic excitation.

Delft Hydraulics

Figure A1.2a: (a) rod / beam Flow-induced excitation due to flow instabilities resulting from flow separation at a circular cylinder. 2b In case of a gate with a protruding lip (used in the U.S., but also elsewhere) the flow is separated at the lip in case of a small lifting height: this point of separation however moves to the rounded side upstream in case of a greater lifting height. Situations of very instable flow may also occur: in that case the flow either separates upstream of or at the place of the lip. Initially, the point of separation is still undetermined due to the rounded shape. Moreover, the flow may separate at the curve, and more or less attach at the location of the lip. In case the flow attaches, this may generate an underpressure in the zone between the point of separation and the point of attachment: the alternating attachment and separation of the flow creates great force fluctuations.

Figure A1.2b: (b) regulating gate or valve Flow-induced excitation due to flow instabilities resulting from flow separation at the bottom edge of a gate. Type 3. Self-excitation (excitation generated by the moving structure itself) A vibration is amplified again and again, in case it gets a push in the back (as known from children on a swing). In case both the vibration and the exciting force are harmonic and, moreover, the magnitude of the exciting force is coupled to the vibration amplitude, then selfexcitation occurs as (part of) the force in-phase with the velocity of the vibration. The excitation is then referred to as movement-induced excitation. This is also referred to as feedback vibration, self-excitation vibration or negative damping. In case the structure does not vibrate, then this type of excitation is absent. Measurements of the spectrum of the load in case of a stiff structure or a stiff scale model also give no indication whether self-excitation may occur.

Delft Hydraulics

In case of movement-induced vibrations, both the amplitude of the vibration and the excitation coupled to it, grow exponentially, until a boundary is reached in which the structure hits the wall or bottom, as an example, or when the flow pattern changes. As this type of vibration may reach very high amplitudes, the prevention of self-excitation vibrations is the first requirement when designing a structure. For gates and valves very many mechanisms are known that cause a movementinduced excitation. Chapter 4 will discuss these in greater detail. 3a Long rods with rectangular diameters appear to be very sensitive to vibration. In this situation (in case of perpendicular flow), there is no strong excitation when the rod does not vibrate. But when there is a (small) vibration in the direction perpendicular to the flow, the approach flow direction changes, from the point of view of the rod. If, in case of a horizontal rod, the movement is directed upward, then the flow that initially came from the left, will now come transversely from top left. This means that the flow at the top will tend to attach towards the back, while the flow at the bottom will bend away from the rod. At the top, in the area between the point of separation and the point of attachment, a certain underpressure is generated, which amplifies the upward movement. At the bottom, a potentially present underpressure however decreases. For the downward movement the reverse pattern is generated. The upward and downward force appears to be proportional to the angle of the approach flow, which therefore is also proportional to the velocity of the vibration. The preconditions for self-excitation therefore are present. The sensitivity to this type of vibration may be determined in a scale model. By approaching the object both at right angles and crosswise, it may be assessed whether the force of the flow varies with the angle of approach in such a way that excitation may occur. This quasi-stationary approach is not suited for high-frequency vibrations (the Strouhal number, now relative to the resonance frequency, may not be greater than 0.1). The mechanism of this type of vibration of rods and wires was first described for icecovered electricity wires in the wind, which were moving with great amplitudes in a kind of galloping movement (see Den Hartog (1956). The ice formation causes the profile of the wire at the bottom to become longer, with a sharp edge. The profile takes on the shape of the wing of an airplane. Actually, here too galloping may occur, i.e. when the wing is so steep that with an even greater angle the lifting force strongly decreases. At a later stage, the word galloping also took hold for situations in which the cause of the vibration excitation is the same, but the vibration is much less strong.

Figure A1.3a: (a) rod / beam Excitation in case of a rod with rectangular section, induced by the movement of the rod itself. 8

Delft Hydraulics

3b In case of gates and valves, the movement-induced vibration may be very strong. This relates to the blocking of the flow. The example shown represents a situation in which a rubber seal is closed due to the water pressure. In case there is a leakage flow or the gate is lifted slightly, then the vibrating sealing strip causes a periodic blocking of the discharge through the gap. In case the discharge is trottled during the closing of the gap, the slowing down of the upstream water causes a pressure increase at the gate. The slowing down of the upstream water coincides with the decrease of pressure downstream of the gate. The pressure drop on the gate and on the rubber edge increases proportional to the flow inertia. This is proportional to the pipe length for the gate and the decreasing discharge. The decreasing discharge is linked to the speed at which the gap is closed. This means that the force exerted on the rubber edge is proportional to the speed of movement of the edge, which again meets the precondition of self-excitation. When widening the gap, the forces of inertia also amplify the movement again. Although it is only the strip that vibrates, a heavy load is exerted on the entire gate, especially when this is a gate in a culvert of great length. Another well-known phenomenon is the vibration of a washer of a tap, causing the entire pipe to vibrate. Flow inertia as a cause of self-exciting vibrations is also present at gates in open water, though in that case it is smaller.

Figure A1.3b: (b) regulating gate or valve Excitation induced by the movement of the flexible rubber seal of a gate. Type 4. Amplification of the excitation due to the resonant rise of the water Many examples are known in which the water has a clear resonance frequency. In case of communicating vessels (this includes a surge shaft together with the outer water), a standing wave, an air bubble enclosed in water, again and again there is an inertia and a stiffness component of the water (or of the air bubble) and during the amplification, kinetic energy is alternatingly converted into potential energy and vice versa. When this fluid resonance system is periodically loaded in the resonance frequency, strong oscillations may be caused that generate yet more load on the structure. 4a This example concerns a bridge pier in a canal with flowing water. In case the frequency of the flow-induced excitation of the water (dependent on the Strouhal number) exactly equals the period of a transverse oscillation of a standing wave, then this wave becomes higher and higher. It must be taken into account that, when the water periodically hits the pier, this also means that the pier periodically hits the water! The flow-induced excitation may therefore be amplified by the oscillation of the water.

Delft Hydraulics

Figure A1.4a: (a) rod / beam Amplification of the excitation due to fluid resonance in case of a single bridge pier in a canal. Figure A1.4b: (b) regulating gate or valve Amplification of the excitation due to fluid resonance in case of a culvert gate, in a situation in which air is enclosed in the shaft of the emergency gate. 4b Enclosed air, combined with a culvert of a finite length filled with water, causes the water-air system to have a resonance frequency. Because of this or because of the vibration, the turbulence downstream of the gate may cause an amplification when frequencies approach each other. A standing wave is also a type of fluid resonance. In case of gate vibrations, interference with a standing compression wave has been observed. Type 5. Instable fluid resonance with self-excitation Instable fluid resonances are caused by the interaction between a structure and water. In case of permanent flow, the instability is not visible. But when a small, more or less periodic water movement is generated due to a random water disturbance, the movement may become purely periodic. Amplitudes may become very high; in scale models, as an example, amplitudes of surface waves have been observed up to one tenth of the water depth. Although the cause of self-excitation of fluid resonances essentially resides in the structure, for the understanding of the concept it may be better to take the fluid as a starting point. The periodicity of the resonance is also completely determined by the natural period of the resonant rise of the fluid. This may be the period of a standing wave or the period of the resonance of communicating vessels. To understand the phenomenon of self-excitation in fluid resonance, a number of abstractions is introduced. A standing wave may theoretically be seen as the total of a wave that travels forward and backward. And each wave may be seen as an interplay of a large number of small travelling pulse waves. When conditions are such, that each small pulse wave is amplified when reflected from the structure, then the standing wave is instable as well.

10

Delft Hydraulics

There appears to be a parallel between the behaviour of a standing wave and the resonant rise of the water levels in communicating vessels (see Paragraphs 2.3.3 and 2.3.4). Below, only the self-excitation of a standing wave is further discussed. Amplified reflection is caused when there is a coupling between the water surface level or the water pressure and the discharge: when an extra pressure causes a discharge (so, in case there is a positive coupling), then a positive incoming wave is amplified when reflected. By reverse, in case of positive coupling, a negative wave will also cause a negative discharge, as a result of which this wave too is amplified when reflected. The extra discharge generated causes a wave that radiates from the structure and therefore coincides with the reflection of the incoming wave. An amplifcation of the discharge of a gate, pump or turbine generates a positive wave downstream and a negative wave upstream. The test criteria for the assessment whether instable fluid resonances occur or not are therefore different at the upstream side as compared with the downstream side.

In order to assess whether there is any danger of instable fluid resonances, the test criteria are as follows: The situation needs to be such, that a standing wave may be expected or such, that there is a system of communicating vessels. Downstream of a gate, turbine or pump, danger may exist when a change of pressure (at the downstream side) is positively coupled to a change of the discharge, i.e. higher discharge and higher pressure. In such a case with a small positive incoming pulse wave downstream, an increase of the discharge results, which generates itself an extra positive wave downstream and therefore amplifies the phenomenon. Upstream of the structure danger may exist, if there is a negative relation between the pressure and the change of the discharge. In such a case a with a small positive incoming pulse wave upstream, a decrease of the discharge results, which generates an extra positive wave upstream and therefore amplifies the phenomenon. Incidentally, the standing waves may also be translated as pressure waves within a closed pipe system. 5a When a beam (this may also be a gate) is hanging just above the water level, any initial wave originating upstream will hit the beam, which immediately causes a damming up (raising of the water level) that coincides with a contraction of the main flow. This damming up may be greater than the initial wave height, for the kinetic head is also converted into an increase of the water level. The flow resistance increases, the discharge decreases, and upstream a positive translatory wave is transmitted; downstream this is a negative translatory wave. As long as the damming up causes the reflected wave to be bigger than the incoming wave, this is considered to be amplified reflection. Before an instable water level oscillation may be generated, the total situation needs to be such, that a low-damped resonant rise or a standing wave is indeed possible. In a scale model it was observed, that in case there are two gates side by side just above the water level (in which case naturally there is also an intermediate pier), water level oscillations are generated perpendicular to the flow direction. Their period is that of the standing wave between the sides of the drain or the abutments. At each of the gates, there is an antinode in the proximity of the abutments. In case there are more than two gates, a system of transverse oscillations may be generated simultaneously. This kind of oscillation may of course also be generated in the absence of piers. 11

Delft Hydraulics

Figure A1.5a: (a) rod / beam Instable fluid resonance due to self-excitation in case of a sluice with an upper beam or a gate that is suspended just above the water surface. 5b At the upstream (incoming) side of a gate, a valve (or even a pump or a turbine), an instable fluid resonance may be generated. The resonating system that amplifies itself may be a standing wave analogous to, as an example, 1.5a, but it may also be a system of communicating vessels, in which case it consists of a gate and a surge shaft with a connecting culvert to the upper water. The upper water referred to here is considered to be an infinitely large vessel. Here too the precondition for an instable fluid resonance is similar to that of a small incoming pulse wave that is amplified when reflected. In this case, in which the shaft is upstream of the gate, it also means that self-excitation will occur when an increase of the water level results in a decrease of the discharge. The reason for this may be similar to that occurring with the rubber seal in example 1.3b, in which an increased hydraulic head in the water level impresses itself on the rubber, causing the discharge to decrease. It depends on the situation whether self-exciting gate vibrations (in the resonance frequency of the gate) occur, or whether self-exciting fluid vibrations (in the resonance frequency of the fluid basin) are generated. The criteria therefore apply in a very general way.

Figure A1.5b: (b) regulating gate or valve Instable fluid resonance due to self-excitation in case of an intake with a gate, pump or turbine.

12

Delft Hydraulics

1.4 Summary diagram of excitation types 1 through 5

Figure A1.6: Summary diagram of possible causes of excitation of flow-induced vibrations (Naudascher 1991, completed with type five).

13

Delft Hydraulics

1.5 Cavitation
Cavitation is a source of dynamic load on structures and of material damage. Cavitation is caused due to the fact that, at normal temperatures, water begins to boil: bubbles are generated that become bigger and implode again. This is caused by the fact that in case of turbulent flow, the low pressure occurs instantaneously and very locally, so that the fluid only experiences this pressure briefly. The implosion causes strong dynamic pressures (impacts) because the radius of the bubbles does not become any smaller, but in a sense they are flattened. It was decided that cavitation falls outside the framework of this publication (see for a general introduction Kolkman, 1984). There are however relations with vibrations: when circumstances give rise to a tendency of generating vibrations, that tendency is amplified by cavitation. And in case of strong vibrations, the pressure fluctuations in the fluid may be amplified, causing yet more cavitation to occur. The treatment of a few practical examples may suffice. See Chapter 6, Paragraph 6.9.

14

Delft Hydraulics

VIBRATIONS OF STRUCTURES AND OSCILLATIONS OF FLUIDS 2.1 General

This chapter presents the equations of the single mass spring system, in dry condition and in water. It is assumed that the principles as they are known in dry mechanics, apply (Paragraph 2.2.1). Characteristic quantities of this system are mass, spring stiffness and damping. From this, the resonance frequency (Paragraph 2.2.2), the relative (dimensionless) damping (Paragraph 2.2.3) and the response behaviour during periodic excitation (Paragraph 2.2.4) may be calculated. By calculating in the time domain, the response to a random, non-periodic load may be determined as well (Paragraph 2.2.5). When this involves a structure in still or flowing water, then (additional) terms need to be introduced to represent the forces at play due to the water (Paragraph 2.2.6). The response of the system may be calculated both for situations in which the flow excitation is periodic (Paragraph 2.2.7) and for situations that involve so-called noise excitation (excitation spread out across a certain frequency band) (Paragraph 2.2.8). In both cases the presentation may be made dimensionless with components that represent both the behaviour of the water and that of the structure. The causes of the forces exerted by the water are not really discussed in this chapter. In the following chapters the diverse aspects relating to flowing water are discussed in greater detail. Although a real structure is usually more complicated than that represented by a single mass spring system, the single system suffices to understand the interaction between the water and the structure. It may demonstrate that the dynamic behaviour of a complex structure can be described as an interplay of vibrations, each behaving like a single system (see Part C, Chapter 3, in which the Modal Analysis is discussed). Only the water may be a complicating factor here (Paragraph 2.2.9). Paragraph 2.2.10 enumerates all elements that may give cause to unacceptable vibrations or to dynamic loads that are too great. Completely analogous to the mass spring system, resonances may also occur in fluids; examples of this are freely oscillating communicating vessels or a standing wave. Situations in which oscillations may occur are discussed in Paragraph 2.3.1. Paragraph 2.3.2 demonstrates that the behaviour of a mass spring system shows a strong resemblance to the behaviour of a system with vessels connected through pipes. A standing free-surface wave has a lot in common with a standing compression wave (Paragraph 2.3.3). The latter occurs both in water in a closed pipe, and in a continuous-elastic structure. The causes of vibrations as summarized in Paragraph 2.2.10 and the causes of fluid oscillations, as found in Paragraph 2.3.5, have a lot in common. For the safety of structures it is important to know that both mechanical systems and fluid systems may be caused to resonate, generating great amplitudes. Moreover, both the mechanical system and the fluid system may demonstrate instable behaviour. When a beginning oscillation has been caused by a small random hit, a vibration or fluid oscillation is generated, of which the amplitude increases more and more. This is referred to as self-excitation and, in case of a mechanical system, as negative damping. The mass spring system demonstrates instable behaviour, when forces work upon it that are induced by the vibration itself, and these forces are proportional to the vibration velocity. 15

Delft Hydraulics

The communicating vessel fluid system becomes instable when a periodically varying discharge is introduced, again and again increasing and decreasing with the extent of upward and downward movement of the water level in the vessels. In case of instability of the mechanical system and of the fluid system, there is, on average, a net import of energy during one period of the vibration or of the oscillation. There may also be a possibility of coupled structure-water systems (Paragraph 2.4). In a number of cases these have been successfully described as one total system with a number of degrees of freedom. It is exactly this coupling of a fluid system with a mechanical system, that appears to result in instability.

2.2 Vibrating structures: in dry condition and in water 2.2.1 The mass-spring-damper system equation
The classic equation for a mass spring system is: m d2 y dy + c + ky = F (t ) 2 dt dt (A2.1)

In this, y is defined as the displacement from the zero position, m as mass, c as damping, k as stiffness, F as the external force and t as time. The terms on the left are called the inertia term, the damping term and the stiffness term respectively, while on the right there is the external force that here does not depend on the movement of the structure. Figure A2.1 is a schematic representation of an elastically suspended gate in a flowing fluid, illustrating the symbols used. The most basic assumption is that the flowing water causes the external force.

Figure A2.1: A gate in flowing fluid: symbols used 16

Delft Hydraulics

About linear damped systems as in Equation A2.1, almost everything is known. A few striking characteristics are: the natural frequency, fn, the dimensionless damping, and the response characteristic in the frequency domain.

These will be deduced in the following paragraphs of this chapter, followed by the calculation in time. Remark: The damping force (in which c is positive) is in anti-phase to the vibration velocity, dy/dt; in magnitude it is proportional to this velocity, but it operates in the opposite direction. When a force like this operates in-phase with the vibration velocity, then the term negative damping is used (c is negative in that case). In that case, a vibration is generated with an ever increasing amplitude, which is also referred to as self-excitation. Negative damping can only occur if a source is included in the system that produces energy.

2.2.2 The natural frequency


The natural frequency is defined as the frequency of a freely vibrating system in the absence of damping. The equation for this is: d2 y m 2 + ky = 0 (A2.2) dt This results in a harmonic vibration:

y = Y sin(t )

(A2.3)

In case the natural frequency occurs, the inertia force and the spring force are in balance and together result in a zero force. Introduction of the expression for harmonic resonance in Equation A2.1, based on the extra precondition that F = 0 and c = 0, immediately results in the value of the angular frequency and consequently of the frequency fn:
fn =

n 1 = 2 2

k m

(A2.4)

There is yet another approach for determining the natural frequency. This one starts on the basis that during the vibration, the total amount of potential and kinetic energy is constant. This may be deduced from Equation A2.2. By introducing z = dy/dt, it is first deduced that:

dt =

dy and z

(A2.5)

d 2 y d z dz dy dz = = =z 2 dt dt dy d t dy

17

Delft Hydraulics

Now Equation A2.2 may be presented as:


mzdz + kydy = 0 After integration, this results in: 1 2 1 2 mz + ky = constant 2 2 or:
Ek + E p = constant

(A2.6)

(A2.7)

(A2.8)

As both the kinetic energy (Ek) and the potential energy (Ep) vary periodically from zero to the maximum value, this also means that:
E pmax = Ekmax

(A2.9)

The natural frequency may be determined by calculating the potential energy of the spring deformation at the highest value of y, at which the vibration velocity equals zero, and by determining the kinetic energy. (This is determined by the maximum vibration velocity relative to Equation A2.3, which is Y.) As:
E pmax =
Ekmax

1 2 kY and 2 1 = m 2Y 2 2

(A2.10)

the Relation A2.4 is found again. This method of calculating the natural frequency offers several advantages in systems with multiple degrees of freedom: the lowest natural frequency may be calculated fairly accurately by first making a rough estimate of the type of vibration, followed by calculating the corresponding potential and kinetic energy and next the corresponding natural frequency. Working with the energy (see next paragraph) may also be useful for determining the damping.

2.2.3 The dimensionless damping


The dimensionless damping is represented by . This is the relation between the damping and the critical damping. Physically, the concept of critical damping may be described as follows. In case of increasing damping, and in case of a free vibration, not only the amplitude decreases faster over time, but also the frequency of the vibration decreases. In case of critical damping the frequency has become equal to zero, and in case of a free vibration the deflection relative to the zero position is reduced over time, but it does not change its sign. The critical damping follows from the equation:

18

Delft Hydraulics

m The solution may be presented as:

d2 y dy + c + ky = 0 2 dt dt

(A2.11)

y = Ya ent e+ in
n results from A2.4, while is:

1 2 t

+ Yb ent e in

1 2 t

(A2.12)

=
This equation may also be presented as:

c 2 km

(A2.13)

c c = n 2mn 2k

(A2.14)

The deduction of the equations above may be found in Part C, Chapter 3. On the basis of registering a free vibration, may be determined from the logarithmic decrement of two consecutive peak values of the amplitude:

Y 1 ln n = 2 2 Yn +1

(A2.15)

In this, is defined as the logarithmic decrement of the free vibration. Both and may therefore be determined from the amplitudes of two consecutive peak vibration values. (Yn = amplitude of peak vibration value n). It may be calculated how much energy per period is dissipated by the damping. To keep things simple, it is assumed that the vibration between two consecutive peak values is purely harmonic (this is possible, as will be explained below, when the damping does not exceed = 0.2). The energy transfer per unit of time is force multiplied with distance, so for one period we find: T dy dy E = (c ) dt (A2.16) 0 dt dt In case of a harmonic vibration, this may be presented as:
T 1 2 2 2 2 E = cn Y cos 2 n tdt = cn Y T = cnY 2 0 2

(A2.17)

The energy dissipation is related to the potential or kinetic energy from A2.10 (both based on the ne peak vibration value): En cnYn2 2 cn = = 1 2 En k kYn 2 (A2.18)

19

Delft Hydraulics

or, using A2.14:

E = 4 E

(A2.19)

This equation is used to determine the damping out in case of non-linear damping; often the energy dissipation per period may be determined, resulting in , and the relation of consecutive amplitudes may be calculated then.
Example: Suppose a vibrating object in still water experiences a damping force equal to:

W = Cw

1 dy dy Aa 2 dt dt

(A2.20)

(Aa is the frontal surface area of the vibrating object, Cw = drag term, = fluid density) It may be deduced that the energy transfer per period equals: 4 E = Cw AaY 3 2 3 (A2.21)

The linear damping ce that is equivalent to this, therefore now equals (compare Equations A2.21 and A2.17): 4 ce = Cw AaY (A2.22) 3 The damping effect of still water therefore decreases when the amplitudes become smaller. Similar calculations may also be carried out in case of Coulomb friction (i.e. a friction force of constant magnitude, that operates again and again in the direction opposite to the velocity). By deducing the logarithmic decrement as a function of the vibration amplitude from a measured signal of a free vibration, an understanding may be obtained concerning the nature of the energy dissipation caused by the damping. Explanation concerning the admissibility of the assumption of a constant amplitude in the time interval between two peak values in determining the energy dissipation. The real energy dissipation may be calculated by determining the potential energy at two consecutive peak values under conditions of a free vibration. This is proportional to Y2. Therefore:
Yn2 Yn2+1 4(1 Yn +1 / Yn ) E = = 2 E {(Yn + Yn +1 ) / 2} 1 + Yn +1 / Yn From Equation A2.12 it follows that the time interval between two peak values is:

t =

n 1 2

20

Delft Hydraulics

and that the amplitude ratio is:

Yn +1 2 / e Yn

1 2

When comparing E/E with what was found in Equation A2.19, it appears that at = 0.2, A2.19 has a value that is 11% higher; in case = 0.1, this is 3%.

2.2.4 Analysis in the frequency domain


In case of periodic oscillation, the amplitude of the internal force (or spring force) is represented by a factor A (for amplification) multiplied with the amplitude of the external force. A is a function of the relation of the excitation frequency, f, relative to the natural frequency, fn. This results in the normalized response diagram of Figure A2.2. The phase diagram also relates to this. The phase angle, , indicates how much the in-phase movement is behind the excitation. So, if:

sin(t ) F=F then the fundamental solution for the movement is:

(A2.23)

y = Y sin(t )
with:
( ) and = phase shift kY = FA The solution may be found in Part C, Chapter 3.

(A2.24)

(A2.25)

In case of resonance (defined as the response at the frequency at which A is maximum) and in case of damping that is not too great, the resonance frequency approximates the natural frequency. As mentioned above, the mass force and the stiffness force (these are the forces in anti-phase and in-phase with the vibration) together equal 0 at that frequency. So:

d2 y + ky = 0 dt 2

(A2.26)

The equilibrium amplitude that is reached follows from the condition that the residual forces in A2.1, i.e. the forces that are out-of-phase with the movement (defined as = 90 or 270), together also cause an equilibrium. The amplitude of the external periodic force therefore equals the amplitude of the damping force. The amplitude of c.dy/dt, in case of a periodic movement, equals cY, so that:

Y=

F c

(A2.27)

21

Delft Hydraulics

In the response diagram, Figure A2.2, this last equation translates into the magnitude of the resonance peak.

Figure A2.2: Response curve and phase diagram of a single (degree of freedom) (lumped) mass spring system in the time domain. Using A2.14 we find:

A=

ky 1 = 2 F

(A2.28)

The presentation in the frequency domain directly applies to the periodic load and response. For non-periodic loads it is necessary to convert data from the time domain to the frequency domain. Next, the response characteristics in the frequency domain may be used, but the results need to be translated again to the time domain afterwards. All this is possible when using Fourier and Laplace transformations, though that is complicated and requires specialist knowledge. It is however also possible to calculate directly in the time domain. This is discussed in the following paragraph.

22

Delft Hydraulics

2.2.5 Calculating in the time domain


Equation A2.1 may be solved through a relatively simple calculation model (for this, see Part C, Paragraph 3.2). With this, the responses in the time domain have been calculated and graphically represented (Figure A2.3) for a number of characteristic types of load, in case of linear and non-linear damping. All calculations shown relate to a mass spring system with m = 10 kg, k = 1000 N/m, c = 10 Ns/m (for as far as these relate to calculations with linear damping); the spring force and the external force are represented on the same scale. The natural frequency (following Equation A2.4) is fn = 1.59 Hz and the vibration period therefore is T = 0.628 s.

Figure A2.3: Response of a single (degree of freedom) (lumped) mass spring system in the time domain. Example A represents the free vibration after a pulse load. This also agrees with a free vibration with initial conditions that do not equal 0. The vibration is damped and the amplitude progresses according to a negative e-power. Example B represents the response in case of a gradual, stepped load, while Example C relates to a sudden load. 23

Delft Hydraulics

Example D shows what may happen, the load remaining the same as with C, in case of a Coulomb friction with a magnitude of 0.04N. The magnitude of this friction is independent of the vibration velocity, but always operates in the opposite direction of the velocity. The enveloping curve of the vibration amplitudes that decrease over time, now is not a negative epower, but a straight line. When an enveloping curve like this is found in investigations of a dynamic scale model, then there is friction and the moving model runs out of true (against something). Example E shows what kind of response is generated by a periodic excitation in the natural frequency of the system. The amplitude initially increases proportional to time. At a later stage the increase decreases and the enveloping curve approximates the equilibrium amplitude via a negative exponent (1-et) (with the equal to that of the free damping vibration). Example F relates to the response in case of an excitation frequency that is 1.5 times greater than that of the natural frequency. The irregularity at the beginning is due to the fact that two vibrations interfere with each other, one of which relates to an initial oscillation that damps out at a later stage. The last Example, G, shows what happens in case of negative damping; the vibration progresses according to an e-power, until the moment, in this case, when it hits a wall. This is included in the equation as a spring stiffness ten times greater than the normal spring. The linear wall damping that is introduced (1000 Ns/m) has little effect, because the vibration velocity in the proximity of the greatest deflection is small.

2.2.6 General description of the mass-spring-damper system in water


Here we further discuss the formal approach that needs to be followed when a load is exerted by (still or flowing) water in case of a structure that moves. It appears to be useful to rewrite Equation A2.1 as follows: (m + mw ) d2 y dy + (c + cw ) + (k + kw ) y = F (t , y, dy / dt etc.) 2 dt dt (A2.29)

The index w indicates that the magnitude concerned relates to the water. On the right, sometimes the term F1w(t) may be distinguished for those forces that are not coupled to the movement of the mass, while the other terms, F2w(y, dy/dt etc.), in that case represent coupled forces for as far as these are non-linear. The linear terms coupled to the movement, have been transferred to the corresponding terms on the left. A2.29 may therefore be represented as: d2 y dy (m + mw ) 2 + (c + cw ) + (k + kw ) y = Fw1 (t ) + Fw 2 ( y, dy / dt etc.) dt dt (A2.29a)

The mw is the so-called added water mass, cw is the added water damping and kw is the added spring stiffness. These magnitudes will be further discussed in the following chapter. Although here they are considered to be constants, it is possible that they are frequencydependent. As long as this is an abstract discussion, each of the magnitudes mw, cw etc. may be both positive or negative. But in reality too, cw and kw may turn out to be both positive or negative. 24

Delft Hydraulics

Below, these linear terms, as well as F1w(t), are discussed in further detail. The omission of non-linearities appears to be permissible for practical situations: as long as the vibration has a small amplitude, it only causes a small disturbance on an average situation. Mathematics also demonstrate that, in case of small variations in a non-linear force, the first term of a Fourier series may suffice. The limitation of small amplitudes usually is sufficient for practical purposes, because vibrations with greater amplitudes should be prevented in any event. There are also exceptions: long slender structures, such as risers at drilling platforms may have fairly large movements without exceeding permissible tensions. When first focusing on the flow-induced excitation F1w(t) = Fw(t), this may be separated into a permanent part and a dynamic part:

(t ) Fw (t ) = Fw (t ) + Fw

(A2.30)

The accent indicates the dynamic part of the load, and an overlining indicates the time average. In the flow, a dominant excitation frequency may be distinguished, coupled to the Strouhal number, S. S is a dimensionless number, defined as:

S=

fL V

(A2.31)

In this, V is de flow velocity and L is a reference length relative to the geometry of the flow pattern. Provided the relation of the different length measures remains the same, the Strouhal number has a more or less constant value for each type of object that is in the flow, and it is independent of the scale of the length and of the value of the approach flow velocity. In case the value of S is given, then also the frequency f of the flow-induced excitation is fixed, given a length measure and approach flow velocity. In case of the situation in Figure A2.1, in order to define the Strouhal number it naturally needs to be agreed which value of L is chosen: L1, L2 or L3, or rather the lifting height of the gate. Also the location of the reference velocity needs to remain the same, if the Strouhal numbers need to be compared. For the reference velocity, it is obvious to choose the approach flow velocity, V, as indicated in Figure A2.1. For this, the flow velocity beneath the gate may also be used. The concept of excitation frequency is completely determined when the excitation occurs within a narrow frequency band. But in case of excitation due to flow, it is also possible that the excitation is spread out across a certain frequency band; in that case, S is only a measure for the dominant frequency. The bandwidth may then be represented again by a percentage of S. Whether S will always remain constant, depends on the situation; if the flow pattern changes, S also changes. In case of Figure A2.1, the geometry of the flow pattern will depend somewhat on the water depth, but the form of the bottom edge of the gate, where the water moves fast, is more important. When the flow is instable (this especially occurs in case of round cylinders), then small causes may alter the flow pattern considerably and the roughness of the surface and the influence of viscosity of the water also come into play. From the response diagram of Figure A2.2 it appears that, given the dynamic load, the most important factor for the response of the structure is the relation of the excitation frequency, f, and the natural frequency of the structure, fn. Therefore it is sensible to represent the 25

Delft Hydraulics

resonance frequency dimensionless as well, in the same way as is done with the excitation frequency (of which the Strouhal number is the dimensionless representation). This results in the concept of reduced (natural) frequency:

fr =

fn L V

(A2.32)

Instead of fr, Sn is also referred to, the Strouhal number relative to the resonance frequency. Therefore:

Sn =

fn L V

(A2.33)

Also, the inverse of the reduced frequency, defined as the reduced velocity, is used often:

Vr =

V fn L

(A2.34)

Vr indicates the relation between the excitation frequency (if S is constant, then it follows from A2.31 that f is proportional to V/L) and the resonance frequency.
Remark: Which representation of fr, Sn or Vr is chosen depends on the situation. The authors prefer Sn when the aim is to determine the resonance frequency necessary for staying away from the critical area of the Strouhal number. In describing the response to noise excitation, Sn also appears to be the most useful magnitude1. When presenting test results, in which the resonance frequency is not varied, but the flow velocity is, then Vr seems to be the obvious choice. When dealing with test results in which the resonance frequency was varied with one value of the flow velocity, then Vr or fr seems to be the obvious choice.

In case the excitation frequency approximates the resonance frequency, the following applies:

S =1 Sn
and therefore also: 1 S

(A2.35)

Vr =

(A2.36)

This relates to the custom of presenting the energy density spectrum dimensionless. For this, the frequency, the variable on the horizontal axis therefore is made dimensionless in the form fL/V, and this is indicated with S. S now is a variable and it is no longer a constant magnitude like it was in Equation A2.31. The spectrum at the value Sn then indicates the energy density at or in the proximity of the resonance frequency.

26

Delft Hydraulics

2.2.7 Harmonic excitation due to flowing fluid and the response to it


Purely periodic excitation due to flow occurs in case of circular cylinders. The assumption that the force is independent of the vibration itself is not realistic here. For the time being, however, this pattern is followed. As appears from Figure A2.2, the amplification factor A, i.e. the relation between the spring load and the external load, especially in case of periodic excitation, depends on the relation between excitation and natural frequency. But, certainly in the proximity of resonance, the damping factor also plays a very important role, expressed in this Figure in the dimensionless factor . In case the excitation is caused by flow, the representative amplitude of the external force needs to be expressed in the dynamic pressure V2, multiplied with the surface area on which it is exerted, L2. This applies to both the permanent force and the dynamic force:

Fw = C
and:

1 V 2 L2 2

(A2.37)

1 Fw = C V 2 L2 2

(A2.38)

The proportionality factor C is the force coefficient; when this relates to the amplitude of the dynamic load, then it is indicated by C, the dynamic force coefficient. This coefficient is also more or less a constant. On top of this, it is customary to refer to lifting coefficient and drag term when dealing with the part of the force that operates perpendicular to the flow or in the direction of the flow respectively. The excitation frequency follows from the Strouhal number, and the response depends on the relation between the excitation frequency and the natural frequency. The amplitude of the response (amplitude Y, multiplied with the spring stiffness) may be directly represented as the (internal) spring force divided by the dynamic pressure multiplied with the load surface area. This relation is a function of the relation between excitation and resonance frequency and of the damping. As the Strouhal number remains constant, the reduced frequency or the reduced flow velocity appears to be a measure for the frequency relation. All this results in:

kY = f (Vr , damping ) V 2 L2

(A2.39)

Now, a number of elaborations is still possible, when the spring stiffness, k, and the mass, (m+mw), are introduced to calculate the resonance frequency. For the purpose of this discussion however, that would be beside the point. Figure A2.4 shows a presentation of the behaviour of a circular cylinder in air flow. The parameter D2/m is not varied, while two values for the damping have been tested. The results for a structure in water would most probably strongly differ, because the mass factor m/mw (or, more generally speaking, m/L3) would be much smaller then. It is interesting to note, that there is a sharply defined value of the approach flow velocity during which the vibration starts. This follows from the coincidence of the excitation frequency following from 27

Delft Hydraulics

the Strouhal number of the non-vibrating cylinder (here, 0.2) and the resonance frequency, fn (in which case Vr = S-1, as in Equation A2.36). After this, in case of increasing flow velocity, the frequency remains stationary around the resonance frequency for a while. The excitation frequency also remains constant, so that the Strouhal essentially changes. This, however, is not completely the case, because, due to the fact that the vibration amplitude is very great here, it increases the reference length (in this case the cylinder diameter D) on which the Strouhal number is based. In case the flow velocity increases even further, then the vibration amplitude does not increase any further and the vibration frequency shifts again to that corresponding to the Strouhal number of 0.2. Figure A5.10 shows similar results for water flow.

Figure A2.4: Vibration amplitude and frequency in case of a circular cylinder in air flow as a function of the reduced velocity (Blevins, 1977).

2.2.8 Response to noise excitation


The character of the excitation due to turbulence is that it does not occur in only one frequency, but it is spread out across a certain frequency band. This is referred to as noise excitation. Generally speaking, excitation due to turbulence is not influenced strongly by (small) movements of a structure. The excitation is described in the form of a frequency spectrum, in which the frequency, f, is indicated on the horizontal axis, and the spectral density, W(f), is indicated on the vertical axis. The total surface area under the curve of a frequency spectrum corresponds to the time average of the squared dynamic force. The RMS value (Root-Mean-Square value) is chosen as a representative value for the amplitude. 28

Delft Hydraulics

Equations A2.40 and A2.41 represent this in mathematical form:

W ( f )df = {F (t )}
0

= ( FRMS ) 2

(A2.40)

in which the RMS value of the force is defined as: FRMS 1 = lim F 2 (t )dt T T 0
T

(A2.41)

Here too, in the same way as with the periodic excitation, a dynamic force (or lifting or drag) term is used, which may be considered as a constant in a certain flow pattern. For C the same applies as for the force: it is the root of the time-average value of the squared instantaneous force coefficient. 1 Fdyn = FRMS = C V 2 L2 2 (A2.42)

Explanation: The squaring of the signal has the aim of obtaining a representation, in which parts of the spectrum (with a limited frequency band) may be added together again. Here, this is indeed possible. Suppose that the dynamic part of a force signal consists of sinus and cosinus functions with increasing frequency, f1, f2 and so on: F = Am sin(2 f mt ) + Bm cos(2 f mt )
m =1 m =1 n n

(A2.43)

Calculation of F'RMS2 following A2.41, provided T is large enough, results in: F 2 = 1 n 2 1 n 2 Bm Am + 2 2 m =1 m =1 (A2.44)

The double product terms do not reappear in the end result. Now it is easy to see, that when the amplitudes are only considered up to a certain frequency (so m runs from 1 up to n1) and, after that, all amplitudes above this frequency (so m runs from n1+1 up to n), the FRMS2 values may simply be added together. Also when the force signal is considered as a constant force, F, with a variation on this, F, and first the signal is considered that has been filtered in such a way that only the part below a certain frequency is admitted, and thereafter the part above this frequency, then it may be statistically demonstrated that the standard deviations (equivalent to RMS values) may be squared and added together. Also, a dimensionless spectral density function is often used, in which the excitation frequency is replaced by the dimensionless frequency (or the Strouhal number S = fL/V).

29

Delft Hydraulics

This is referred to as a uniform spectrum or unit spectrum (S) (below which the surface equals one); so:

(S )dS = 1
0

(A2.45)

Both spectra are equal as far as their form is concerned. To convert the unit spectrum into the energy density spectrum again, the magnitude of the excitation needs to be introduced (squared) and because of the extension of the horizontal axis with V/L (because f = S(V/L)) the vertical axis needs to be shrunk with the same factor. Thus it may be demonstrated that: L 1 W ( f ) = (C V 2 L2 ) 2 ( S ) V 2 (A2.46)

Here too, (S) and C no longer depend on the dimension of the structure and the flow velocity. The values however may depend on the flow pattern and there may be a certain influence of the Reynolds number (Re = VL/), the wall roughness etc. If the excitation spectrum in case of noise excitation is known, the response of the structure may be calculated by multiplying the square of the excitation amplitude (i.e. the function W(f) multiplied with a piece of the frequency band) per piece of bandwidth with the square of the response function. Provided it is small enough, the magnitude of the bandwidth that is chosen in itself is not important for the end result.

Figure A2.5: Excitation spectrum, response function (squared), and the squared response deduced from it. Figure A2.5 illustrates how the response may be calculated from the given spectral distribution of the excitation and the squared response (the magnitude A2 in Figure A2.5 refers to the amplification factor A of the response function in Figure A2.2). The multiplication of the excitation spectrum and the squared response curve may be made graphically or numerically. Based on an earlier publication of Fung (1960), Kolkman (1976) has proposed a simplified analysis. This is based on the fact that the multiplication results in a dominant peak at the resonance frequency that is completely determined by the value of the spectrum at the location of the resonance frequency, therefore of W(fn). The magnitude of the response has 30

Delft Hydraulics

been calculated by Fung, albeit assuming that across the whole frequency band, the spectral density has the same value corresponding with W(fn), although that hardly offers any difference with a more realistic approach. The response that is calculated is represented as an internal force (spring force): ( Fnternal ) 2 = k 2 y 2 =

f nW ( f n ) 4

(A2.47)

This deduction may also be found in Kolkman (1976). If the spectral density W(f) is expressed again in terms of the unit spectrum multiplied with the squared amplitude of the excitation force divided by the value V/L, then Equation A2.47 may also be represented as: ( Fnternal )2 =

1 Sn ( Sn )(C V 2 L2 ) 4 2

(A2.48)

In this, (Sn) is the value of the energy density spectrum in (the proximity of) the resonance frequency (represented dimensionless). If the resonance frequency is higher than the dominant excitation frequency (which naturally, in case of a good design, should be the case), then the response spectrum (the multiplication curve as shown on the right in Figure A2.5) will show a second bump in the low-frequency area. The content of this low-frequency bump is more or less the same as the content of the original excitation spectrum, because the response of the single mass spring system does not generate much amplification in that frequency area. To obtain the amplitude of the total internal force, the surface areas of the low-frequency bump and the resonance peak may be added together. By taking the root of this product again, the standard deviation of the internal force is found to be:
2 2 Finternal = F internal resonance + Fexternal

(A2.49)

This analysis was first carried out in this manner in a vibration investigation at a sector gate, which may be found in one of the appendices of Kolkman (1976).

31

Delft Hydraulics

Figure A2.6: Vibration mode of the visor gates of the Hagestein Weir in dry condition and partly immersed. Response measured during vertical mechanical excitation in the centre of the gate (Kolkman 1976).

2.2.9 Structure behaviour in case of multiple degrees of freedom


A system with multiple degrees of freedom also has multiple natural vibrations, each with its own natural frequency and natural vibration mode. Depending on the degree of complexity, for a dry structure these may be calculated analytically or with the Finite Element Method (FEM). To assess or calculate the dynamic behaviour in flowing fluid, it is of prime importance to know the vibration mode(s) and the natural frequencies of the structure in water. The added water mass has the effect of causing the natural frequency to decrease considerably relative to the dry condition. Moreover, in some cases the vibration mode may alter considerably. Figure A2.6 shows the vibration mode in dry and partially immersed condition of the visor gates of the Hagestein Weir in case of periodic vertical excitation in the middle of the gate (see also Paragraph 6.2). The Figures show the vibration mode at one of the lower resonance frequencies, in which the vertical vibration (in which the elasticity of the suspension plays an important part) appears to be accompanied again and again by 32

Delft Hydraulics

considerable torsion of the gate. As big horizontal movements are generated at the dry gate, the effect of the added water mass is also great. This appears to result in a strongly altered vibration mode, in which the amplitudes at the bottom of the gate are much smaller. When using the Finite Element Method it is not possible to include the influence of the added water mass just like that. The added water mass is a more complex magnitude than the mass of the structure itself. Even in case of a single structure with two degrees of freedom, the added mass in each of the vibration directions has a different value. This is discussed in Chapter 3, Paragraph 3.6, on the basis of a gate with an L-shaped diameter (Figure A3.14). This gate may vibrate both horizontally (in direction x) and vertically (in direction z). In case of vibration in direction z, the added water mass will strongly depend on the width of the horizontal strip, while the added water mass in case of vibration in direction x will depend on this much less. Also, in case of an acceleration in direction z, a horizontal force will be exerted (in direction x) that is proportional to the acceleration d2z/dt2, because the gate is very asymmetric. By reverse, an accelerated movement in direction x will also generate a force in direction z, that is proportional to the acceleration of the gate. This coupling often is very important. In case calculations of the natural vibrations are carried out with a FEM program, the program naturally needs to be suitable, in case of a system with n degrees of freedom, to include the added water mass as a complete matrix, including the coupling terms. The matrix includes all terms m11, m12, m21, m22 etc. The first index provides for the degree of movement that is exerted; the forces that are generated by this operate on another degree of movement that is represented by the second index. In case the two indices are equal, this means an increase of the natural mass of the structure; in case they are different, they are coupling terms. Some notes concerning the calculation of the added water mass matrix may be found in Part C, Paragraph 3.3. The added water mass may only be calculated when certain schematisations have been introduced. In the first place, small vibration amplitudes are assumed. Furthermore a moving wall is replaced by a fixed wall, while the water displacement is realized by modelling with sinks and sources on the fixed wall. The vibration velocities are also small; thus, pressures that are proportional to V2 may be ignored. Water level differences may be introduced as a limitation of the flow field, but the influence of the corresponding permanent flow on the added water mass may not. This method of schematization causes the flow, generated by the vibration of the structure, to behave as potential flow. About this a lot is known and the Finite Element Method may be adapted for their calculation. Experience shows that the added water mass does not change much by the flow; vibration modes and frequencies therefore may be determined sufficiently accurate in this way. Also the damping and the excitation forces as a consequence of flow are not so great that they alter the vibration mode or the vibration frequency. The diverse excitation sources, as discussed in Chapter 4 and 5 for a single mass spring system, also apply to a system with multiple degrees of freedom or a continuous-elastic structure. It is also necessary to see whether situations occur locally that are similar. There will be places where self-excitation may occur, and other places where damping may be expected. Theoretically speaking, it should be possible to translate this self-excitation and flow damping to local positive or negative energy transfer from the water to the structure. The aggregated energy transfer may then be a measure of possible resonant rise or damping of the entire structure. Before calculations like that become useful, a lot needs to be known about this energy transfer as a function of the frequency and the vibration amplitude.

33

Delft Hydraulics

2.2.10

Causes of vibrations

From the analysis of the basis for the mass-spring-damper system in water, the following possibilities for big dynamic loads may be formulated, in order of decreasing risk: A. The total damping (c+cw) is negative, in other words: self-excitation occurs. Because of this, the system becomes instable and a small random disturbance causes a vibration, the amplitude of which increases again and again (exponentially). The increase of the vibration amplitude only ends when non-linear terms of the Equation A2.29 water load (the terms Fw2, y, dy/dt etc. in Equation A2.29a) are introduced, or the structure behaves in a non-linear fashion. In case of gate vibrations at a small opening, the amplitude may be limited due to the fact that the vibrating gate touches the bottom2. The system starts to resonate due to a purely periodic flow load. Figure A2.4 shows an example of this. Even when the force amplitude is small, an extended periodic load in case of small damping may still generate big internal forces. Noise excitation can only be dangerous when the damping is small. The vibration response in this case is much smaller than in case of resonance with periodic excitation. Quasi-static load (this is a dynamic low-frequency load relative to the lowest natural frequency of the structure) is only dangerous when the load is very big. This appears to occur in case of great fluid oscillations, heavy wave loads or wave impacts. Because of the fact that in case of fluid oscillations the frequencies are usually much lower than the resonance frequency of the structure, often these are no longer referred to as vibrations, but as deformation (or displacement) due to dynamic load.

B. C. D.

2.3 Oscillations of fluids 2.3.1 Properties of a fluid oscillator


A fluid oscillator is here defined as a limited fluid area, in which a resonant rise or a standing wave may occur, without kinetic or potential energy radiating outward from that oscillation or being added externally. Damping out of oscillations only occurs due to damping factors within the system. A fluid oscillator may start to resonate in one or more resonance frequencies, depending on the number of its degrees of freedom. Just like the mass spring system, here too especially the system with one degree of freedom will be discussed.

In Paragraph 2.2.3 it has been deduced that in case a vibration resistance equals the square of the vibration velocity, the equivalent linear damping increases with the vibration amplitude (Equation A2.22). Starting from a certain amplitude, it is possible that the total damping becomes positive again (c + cw > 0), which means that the amplitude is limited by the water itself.

34

Delft Hydraulics

2.3.2 Survey of favourable situations for the existence of a fluid oscillator


The different modes of appearance of a fluid oscillator are presented in Figure A2.7. These are: Aa: Ab: Ba: Bb: Ca: Cb: System with two mutually connected basins: anti-symmetric oscillation. Half-closed Aa system; one basin, connected to the outer water by a pipe. Basin closed on all sides: standing anti-symmetric wave. Half-closed basin in open connection with large outer water area. Pipe closed on all sides; standing anti-symmetric pressure wave. Half-closed pipe in connection with a wide pipe or large outer water area.

As there is, in case of complete systems as shown here, anti-symmetry as regards the resonant rise or wave movement, there exists a constant pressure equal to the initial situation on the central axis. (As may be seen in Figure A2.10, symmetric waves are also possible). Should the system be cut in half at the central axis, and be connected with a space where the pressure is also constant, nothing changes in the dynamic properties; the resonance frequency does not change either. The precondition for constant pressure is met, when the outer water is very large. In that case there is no wave radiation there and at the occasion of the transfer from the basin or the tube to the outer water, a complete (negative) reflection of the wave in the basin (or tube, respectively) occurs. Figure A2.8 shows situations in which, in case of a double half-closed basin or double half-closed pipe system, a complete standing wave occurs, but in a folded-over way. Because of the fact that there are no pressure variations in the middle (this is also the location where the outer water begins) as a consequence of the reverse symmetry of the standing wave, there is no wave that radiates outward; consequently there is no requirement upon the outer area to be large. The fact that no radiation occurs with this wave, moreover is exactly the reason why standing waves may be generated. The system of communicating vessels may be extended with many more vessels and pipes. The number of resonance frequencies equals the number of pipe elements. Within a closed basin or pipe, many standing waves may be generated (see following paragraphs).

35

Delft Hydraulics

Figure A2.7: The six possible modes of appearance of a fluid oscillator.

36

Delft Hydraulics

Figure A2.8: Folded situations of a standing wave. K indicates the location of the node of the wave.

2.3.3 Communicating vessels compared to a mass spring system


Figure A2.9 shows a communicating vessel in case of an oscillation. The oscillation follows from the solution of two equations: the continuity equation and the movement equation. The following applies to the continuity, per unit of width:
q p = hpV p = B

dz dt

(A2.50)

in which qp = discharge in the pipe per unit of width.

37

Delft Hydraulics

For the meaning of the (geometric) symbols, see Figure A2.9. When ignoring the wall friction and other losses, the movement equation results in:

gz = L

dV p dt

(A2.51)

Figure A2.9: Basin connected with a pipe to a large outer water area. Eliminating Vp results in:
LB d 2 z + gz = 0 hp dt 2

(A2.52)

This equation is very similar to Equation A2.2 of the undamped mass spring system. The solution of Equation A2.25 may be presented as:
z = Z sin nt

(A2.53)

with:

n =

ghp LB

(A2.54)

The equation with the single mass spring system may be extended further, if the water mass in the pipe is considered as the mass and the horizontal displacement of the water relative to the condition of rest is defined as y. The second law of Newton, K = ma, is now directly applied. The spring stiffness follows from the opposing force, K, which the water in the pipe experiences due to the displacement y:
K = z ghp = y hp B

ghp

(A2.55)

because zB = yhp

38

Delft Hydraulics

From this follows the spring stiffness:


2 K ghp k= = y B

(A2.56)

while the mass (per unit of width) equals:


m = hp L

(A2.57)

Now that spring stiffness and mass are known, the natural frequency follows from Equation A2.2.
SYSTEM OF MULTIPLE BASINS In case of several inter-connected basins, the number of pipes determines the number of degrees of freedom and thereby also the number of natural frequencies of the system. The outer water may also be considered as a basin (a basin with an infinitely large surface area also leads to a constant water level). Even in case of a pipe that connects two areas, each with a constant water level, the equation for the acceleration of the water column in the pipe applies, but the resonance frequency is zero and the oscillation is super-critically damped. This is a relevant situation in case of a long pipe in which a tap is suddenly opened. The system may be compared to a single mass supported by a damper, without a spring. RESONANCES In case the basin is fed with a periodically alternating discharge, with a frequency that equals the resonance frequency of the basin-pipe system, resonance will occur. The equilibrium amplitude will depend on the discharge amplitude and the flow losses in the pipe. No examples are known of situations in which resonances of a basin-pipe system have caused any problems. INSTABLE OSCILLATIONS As it is found with the mass spring system, that negative damping causes selfexcitation and consequently instable vibrations, the fluid oscillator therefore may demonstrate an instable behaviour. In case a feeding discharge qv is present at the basin, coupled to the water level variation, z, then instability is also possible. Suppose: qv = z (A2.58)

in which represents a proportionality constant (not time-dependent). Now A2.50 changes into:
B

dz = qv q p = z hpV p dt

(A2.59)

In combination with Equation A2.51, this results in:


LB d 2 z L dz + gz = 0 hp dt 2 hp dt

(A2.60)

39

Delft Hydraulics

This equation strongly resembles the vibration equation A2.11 and this directly shows that a positive factor means that there is a negative damped system. Paragraph 4.5 further discusses an example, in which a positive factor is generated near a floating gate (Figure 4.23), which results in an instable fluid oscillation.

2.3.4 Properties of the standing wave


In a basin that is closed on all sides like the closed pipe in Figure A2.7 Ca, the natural frequencies of the standing wave are determined by the basin length, 2L, and the celerity of the wave, c. The wave shape in the first instance is a sinus or a cosinus. A better approach in case of a shorter wave length is the trochoid, but that is not relevant for the following. In case of the lowest harmonic, the basin length, 2L, corresponds to the half wave length (the wave length is referred to as ). Generally speaking, for a closed basin: 1 2 L = n( ) 2 (A2.61)

n is a whole number. n = 1 corresponds with the lowest harmonic. The lowest and the higher harmonics may occur simultaneously. The geometry of the wave is symmetric or anti-symmetric relative to the middle. This means that there is a node in the middle (in case of symmetry) or an anti-node (in case of antisymmetry). When analysing a two-dimensional closed system, the half system will therefore suffice. In case there is an anti-node on the central axis (at even values of n), then the central axis may be replaced by a wall when performing the analysis of the half system. In the other case, the condition of the central axis is such, that the pressure always keeps the value of the zero position. Figure A2.10 shows the symmetric and anti-symmetric wave patterns in a closed basin with a free water surface. In case of a closed pipe, the wave patterns are the same, but in this case these only refer to the pressure distribution of the compression waves. As, in Figure A2.7, the length L in case of the half-closed situation keeps the same meaning as the not-half-closed anti-symmetric situation, Equation A2.61 remains valid in both cases.

40

Delft Hydraulics

Figure A2.10: Symmetric and anti-symmetric waves in closed basin (with free water surface). To determine the period of the oscillation when is known, the celerity c needs to be known. For the free surface wave, this is:
c= gT 2 d tanh 2

(A2.62)

In this, T is the wave period and d is the water depth. In case the water depth is great relative to the wave length, the maximum value of the celerity in A2.62 becomes: gT (A2.63) c= 2 In case the water depth is small relative to the wave length (long or shallow water wave), the maximum value of the wave celerity is:
c = gd

(A2.64)

41

Delft Hydraulics

By using Equation A2.61 (if n is known), first the wave length may be calculated, after which the natural period of the oscillation may be found, using:
T=

(A2.65)

By using Equations A2.62, A2.63 or A2.64, the corresponding period may then be determined. From the general equation for c, A2.62, it then follows that: 2 g T= 2 d tanh

(A2.66)

The maximum values for the deep water and shallow water situations may be deduced directly from this. The celerity of a pressure wave in a closed, infinitely stiff, pipe, is:
c= K0

(A2.67)

In this, K0 is the compression module of the water. In case the tube is not infinitely stiff, then:
c= 1

D + K0 E

(A2.68)

In this, D is the internal pipe diameter, is the wall thickness of the pipe and E is the elasticity module of the wall material. The determination of the period of the standing wave, here too follows from Equation A2.65.
RESONANCE PHENOMENA Standing waves, when hit in one of their natural periods by a periodically varying discharge, may generate resonant rise. The discharge exerts the strongest influence when it approaches or runs off in an anti-node of a standing wave. The latter does not need to be taken literally: it may also involve a constant discharge on which the fluctuating discharge is superimposed.

Another way of excitation is the stirring effect. When an object is moved back and forth in such a way that a periodic force is exerted upon the water that operates in the direction of the movement again and again, then energy is transmitted from the object to the standing wave, and the wave movement becomes stronger and stronger. Something similar may occur at bridge piers in a canal (Chapter 1, Paragraph 3, excitation type 4). This is, however, a stationary object that exerts influences in the node of the wave, exactly at the location where the back and forth movement of the water is greatest.

42

Delft Hydraulics

INSTABLE STANDING WAVES Self-excitation may occur, not only in case of communicating vessels, but also in case of standing waves. Here too, the discharge is greatest in case it is coupled to the water level variation in such a way that the discharge is greatest at the wave crest (or the peak value of the pressure, in case of a compression wave). This phenomenon is especially well known in case of floating flap gates, beneath or over which water flows. The flap gate movement coupled to the water level also causes discharge variations. For more information, see Chapter 1, Paragraph 3, excitation type 5. The question regarding which wave is excited, n = 1 (the first harmonic), n = 2 (the second harmonic) or even higher values of n, depends on the location where the fluctuating discharge makes contact. It is essential that this occurs in the proximity of an anti-node of the wave. In case there are multiple possibilities for the value of n, then the instability is in fact always generated near the lowest possible value. The fact that the location where the discharge makes contact is so important, may be illustrated by the fluid oscillations that were found during an investigation (in a scale model) of the storm surge barrier Rotterdam Waterway (see Chapter 6, Example 6.2g). A resonant rise of the water level occurred that was perpendicular to the main flow direction. In case of flow from the sea toward the river this was the n = 1 wave, in case of flow from the river toward the sea a n = 2 wave. This relates to the fact that these were sector gates, of which the vertical axis resided in the abutments, so that in case of flow from the sea toward the river, the discharge would be directed toward the abutments (which results in n = 1), while in case of flow from the river to the sea, the discharge with flow-through underneath concentrates on the middle of the river. At n = 2, the anti-node of the standing wave is also in the middle of the river.

2.3.5 Causes of oscillations


In the preceding paragraphs, the main causes of fluid oscillations have been discussed. To begin with, there needs to be a system that is capable of oscillation: i.e. a system consisting of communicating vessels, whether connected with outer water or not, or where there is a possibility of standing waves occurring without energy radiating away (Figures A2.7 and A2.8). The causes of fluid oscillations, in order of decreasing risk of generating big oscillations, are: A. Self-excitation occurs when there is a discharge supply that is coupled to the water level or pressure variation in such a way, that at the wave crest or at the peak pressure the discharge is also at its maximum. An oscillation is generated in which the amplitude increases exponentially, until it is so big, that the phenomenon changes its character or strong damping is generated. Experience shows that in case of selfexcitation, oscillations of the free water level may occur that are several meters in height. In case of closed pipes, sometimes great pressures are generated as well. A well-known example is the broken washer of a tap in a residential home, in which pipes may start to move. Resonance occurs when there is a periodicity in a discharge supply that corresponds with one of the natural periods of the oscillator. Also when the system is stirred, or when forces periodically operate on the water, resonance may be generated. In case of a moving basin (or a moving lock chamber, such as used in large high-head gates) the 43

B.

Delft Hydraulics

C.

periodicity of the shaking may cause resonance as well. A pier placed in a canal may cause a standing, perpendicular wave in the canal by interference of the frequency of periodic separation of the flow and the frequency of the standing wave. Non-periodic disturbances may cause wave or resonant rise movements, but these are limited in amplitude.

2.4 Systems with mechanical components and fluid components


Next to the single mass spring system, there are mechanical systems with multiple degrees of freedom. In case of multiple mass spring systems, the number of degrees of freedom depends on the number of masses and the number of degrees of freedom of movement that each mass possesses. In case of communicating vessels, systems with one or more degrees of freedom may occur as well. The number of degrees of freedom equals the number of connecting pipes. The inertia of the water in the pipe demonstrates a similar behaviour to that of the mass of a mass spring system (see Paragraph 2.3.3).

Figure A2.11: Diagram of a system with gate (two degrees of freedom), shafts and connecting culverts. In case there is a combined system with mechanical components and fluid components, the number of degrees of freedom grows. Figure A2.11 represents such a system. In this example, the number of degrees of freedom amounts to five: two degrees of freedom of the gate, because it is assumed that these are only capable of performing horizontal and vertical movements, and three connecting culverts. Five resonance frequencies may therefore occur. 44

Delft Hydraulics

Now it is possible to put together one combined set of equations, in which the gate vibrations and the water level oscillations, and the interactions between them, are represented. The interaction consists of the fact that the water column in the culvert (or pipe) in which the gate is located, is accelerated as a consequence of the hydraulic head minus the loss (expressed in the hydraulic head) at the gate. The latter depends on the position of the gate at that moment in time, and that in its turn is influenced by the hydraulic head of the gate. Until now, the experience with combined water structure systems remained limited to a system with two grades of freedom in total. Part C, Chapter 3, discusses a number of examples, in which the combined system and the time domain have been taken into account. One of these is the so-called bath plug vibration (also discussed in Chapter 4, Paragraph 4.4.2). In those examples, it is exactly the interaction between the mechanical system and the fluid system that causes the self-exciting vibrations and the self-exciting fluid oscillations respectively.

45

Delft Hydraulics

46

Delft Hydraulics

PASSIVE FORCES: ADDED WATER MASS, ADDED DAMPING, ADDED STIFFNESS IN STILL WATER AND FLOWING WATER 3.1 General indication of issues

The dynamic equation of the single mass spring system, A2.1, has been completed in Chapter 2, Equation A2.29 with the terms added water mass, added water damping and added (spring) stiffness, magnitudes that may vary considerably and that may possibly be frequency-dependent. A calculation of these added terms, based on the complete NavierStokes equations, is very complex. Preconditions are also more complex in case of vibrating structures as compared with calculations in case of normal flow. In reality, calculations are always carried out in a very simplified way. Added water mass, as an example, is always calculated for an object in still water. The equations are linearised, which in reality means that the outcome is only valid for vibrations with a small amplitude. As far as calculations are carried out in situations with flowing water however, the superposition principle is taken as a starting point: a flow pattern coupled to the vibration of the structure is superimposed on the flow pattern that is present permanently. This results in added damping, on top of the added mass. This procedure has not been verified. Experience however shows, that generally speaking, the added water mass is hardly influenced by the strength of the flow. An exception is the circular cylinders in the flow, when the vibration frequency approximates the frequency of periodic separation of vortices. The preconditions for fluid limitations are: in case of a fixed wall, the velocity component perpendicular to the wall equals zero. This precondition applies to the permanent flow condition, but also to superimposed flow, generated by the vibrating structure; in case of a wall that vibrates in its own plane, the same precondition applies; in case of a wall that vibrates perpendicular to its plane, the water velocity perpendicular to the wall equals the vibration velocity. The simplification introduced in all reviews and calculations is the following: the moving wall is considered as a fixed wall, modelled with sinks and sources in such a way, that the amount of water displaced as a result equals the amount of water that is pushed away by the structure. The fluid limitation therefore does not change, which does not fully agree with reality. This schematization is only permitted in case the amplitude of the movement is small relative to the geometric length measures. In case of the free water surface, a similar schematization is used as that in case of a vibrating wall; a fixed horizontal surface is used as a flow limitation, in which the pressure variations are the same as the hydrostatic pressure of the small waves caused by the vibration of the structure. The effects of this result in the classic surface condition belonging with the linear wave theory. This simplification is only possible in case of a small wave amplitude and, as far as these waves are generated by the vibrating structure, therefore applies to small vibrations. An analysis of the dynamic behaviour of a structure limited to small vibration amplitudes in this review is, generally speaking, not critical, because greater vibration amplitudes actually only occur in case of a non-acceptable design. Exceptions to this are slender offshore structures like pipes and cables, as a result of which greater amplitudes (relative to the diameter) do not necessarily cause material tensions that are too great.

47

Delft Hydraulics

Below is a qualitative discussion of the properties of the free water level and of the flow on the added terms (in still water conditions). And finally, assuming yet more simplifications, it is indicated how the added water mass and the damping may be calculated.

3.1.1 Review of the influence of the free water surface


In case of radiating surface waves, the added water mass and the added damping may become frequency-dependent. This is caused by the fact that the wave length and the wave celerity depend on the wave period. The pressures relating to longer waves penetrate more deeply. In case of higher frequencies, such as occurring again and again in case of vibration problems, the wave length becomes small and the depth of penetration of the waves becomes small as well. Consequently, the influence of the wave radiation on the added terms is also small. What applies here, approximately, as condition for the free water surface, is that the pressure is constant (i.e. there is atmospheric pressure). Therefore there is no frequencydependency anymore of both the added water mass and the water damping; the latter rapidly approximates zero. The water surface is not in zero position, but it now moves up and down with the movement of the vibration. The corresponding pressure variation at the location of the original horizontal water surface is ignored.

Figure A3.1: Results of a two-dimensional calculation of the added water mass ( CL d V / dt ) and damping ( Cr d V ), taking into account the wave radiation, for a horizontally vibrating gate with water on one side, Kolkman (1976) and Delft Hydraulics Report W254.

48

Delft Hydraulics

Figure A3.1 shows a situation with a vertical wall, of which the bottom part (the gate) vibrates horizontally. The average dynamic pressure across the gate height, p, that operates on the gate, has a component that is proportional to the vibration acceleration (this generates the added water mass), and a component that is proportional to the vibration velocity (which generates the added damping). The vibration velocity is represented as:
v (short for V sin t )

(A3.1)

and the acceleration as:


v (short for

dv ) dt

(A3.2)

The coefficients CL and Cr are represented as a function of h/d and the vibration frequency. The vibration frequency is expressed in the angular frequency, , that is represented dimensionless as 2h/g.

The added water mass per unit of width is represented by the coefficient CL; dividing the relevant pressure component, multiplied with the gate height, by the acceleration, results in the added water mass, mw. So: (C d v)d = CL d 2 (A3.3) mw = L v The added damping, cw, per unit of width, here too is represented by the coefficient Cr (radiation coefficient) and equals the relevant pressure component, multiplied with the gate height, divided by the vibration velocity V:

cw =

(Cr d v)d = Cr d 2 v

(A3.4)

It is clearly visible, that for greater values of the frequency, the added water mass becomes independent of the frequency. The damping decreases strongly in case of higher frequencies, due to the fact that the pressure corresponding to the short waves hardly penetrates downward. Due to the fact, that the fluctuating pressure at the free water surface remains close to zero, the added water mass in that case is much smaller than, as an example, in a culvert. If it is possible to ignore the wave radiation, it may be demonstrated that all pressures within the entire fluid area are in-phase relative to each other; this also applies to the water velocities that have been generated by the vibration of the object. Because of this, in case of a vibrating object in still water, there is potential flow, and both vibration related flow and pressures may be calculated as a potential flow, such as, for example in case of permanent groundwater flows. In case the vibrating structure intersects the free water surface, then a vertical movement of the structure experiences not only an added water mass, but also an added spring stiffness (due to immersion). This stiffness equals gA, in which A represents the plane
49

Delft Hydraulics

of section of the structure with the water level. In case of gates with floating water compartments, this magnitude may be significant.

3.1.2 Review of the situation with flowing fluid compared to the situation with still water
For a situation with flow, the added water mass, water damping and spring stiffness may not be calculated easily. To obtain an impression of the influence of flow on the added water mass and water damping, we will review the flow at each point as a superposition of the flow in case of permanency and a flow pattern that periodically fluctuates as a consequence of the vibrating structure. Similar to the case of still water (without wave radiation), it is assumed that at the point (x, y, z) in the periodically fluctuating part of the flow there is no phase shift of the flow velocities relative to the velocity of the vibrating structure. We will further review the so-called Navier-Stokes equations, in which the flow pattern at permanency is supposed to be known. The velocities are separated into the components u, v and w in respectively the directions x, y and z, belonging with the permanent flow field and the dynamic components u, v and w coupled to the vibration. We will leave out the viscous terms in the equations. For the direction x we then see that: ( p + p) (u + u) (u + u ) (u + u) (u + u ) = (u + u ) (v + v) ( w + w) x t x y z

(A3.5)

Explanation: the terms on the right are to do with the acceleration of a water particle. This is the local acceleration ( (u+u) /t) and the convective accelerations, due to the fact that the particle moves in the velocity field of one location with a certain velocity toward another location with another velocity. Equations similar to A3.5 may be written for the directions y and z. In case of nonpermanent phenomena like waves, normally the first term on the right plays an important role; this then gives rise to the linear wave theory. In case of permanent flow (so when u' = 0 and u'/t = 0) the other terms on the right are determining. These also cause the stationary forces operating on a structure. When now considering the extra pressure field, p', that is generated by the vibration, this may be represented as: u p u u u u = u + u etc. v + v x t x y x y (A3.6)

There are also double product terms, in which u' appears twice; as long as the vibration velocities, u', however are small relative to the initial velocities, u, these dont play a significant role. There are terms that are proportional to u'/t, and therefore also proportional to the acceleration of the vibrating structure, and there are terms equal to u, and therefore also equal to the velocity of the vibration. Both terms are linear with the amplitude of vibration. The first terms cause forces that are proportional to the acceleration of the vibrating structure, and that again generates the added water mass. The second type of terms, proportional to the vibration 50

Delft Hydraulics

velocity, generates a damping force, the so-called flow damping. From Equation A3.6, approximately the following conclusions may be drawn: The added water mass does not depend on the initial velocity field, and may therefore be calculated on the assumption that the water does not flow. The water damping is proportional to the initial flow velocity.

When reviewing the relation between the two types of terms, the first type (added water mass term) appears to be dominant when the Strouhal number is high. This may be understood from the magnitude of the periodically varying dynamic velocities, u', to be described with the amplitude of this velocity: (u' = sint). For the derivative with time this now means that: u U t and for the velocity gradient the proportionality is: u U 0 x L (A3.8) (A3.7)

U0 is the reference velocity that is a measure for the flow velocities in the initial flow field and L is a reference length. When the first term on the right of Equation A3.6 is dominant, it follows from these relations that:

U L
U0

UU 0 or L 1

(A3.9)

By definition, the term L/U0, except for a factor 2, is the Strouhal number, fl/U0. For damping calculations it is usually assumed (in case of a flow-surrounded object), that a quasi-stationary calculation is possible; this means that at each moment, the flow forces are assumed to be proportional to the square of the flow velocities relative to the vibrating object, therefore with (U0-dy/dt)2 (y = displacement of the vibrating object). This subtraction must, in fact, occur vectorially. This calculation of the water damping is plausible in case the vibration is lowfrequency. This actually means that this approach is correct, in case the Strouhal number is low. In case the Strouhal number is high, then the added water mass flow field as calculated in still water, may be superimposed on the initial flow field. After that, the pressures may be determined, using Equation A3.6.

51

Delft Hydraulics

By the way, in both cases a damping force is generated proportional to the initial flow velocity. No experience has as yet been obtained using this method of calculation. Equation A3.9 only indicates in qualitative terms at which Strouhal number one or the other method is appropriate. There are indications that, in case of flow-surrounded cylinders, the quasi-stationary calculation is possible, as long as the frequency is lower than 0.1 V/D (V = approach flow velocity, D = cylinder diameter).

3.1.3 Calculation of the added water mass


The calculation of the added water mass in still water, which on the basis of the above also remains more or less valid in case of flowing water, assumes a rotation-free flow and ignoring the viscosity of the water (Lam, 1932). This assumption of potential flow actually replaces the movement equation; in addition, only the precondition of continuity ( = 0) needs to be met ( = velocity potential). If also the assumption of the absence of wave radiation is introduced, then it means that everywhere across the generated flow field the velocities are in-phase with each other, and therefore also in-phase with the velocity of the vibration of the structure. The flow pattern behaves like a periodically varying, quasipermanent flow field, and consequently all solution techniques, such as electric analogon, magnetic field, and conformal mapping, may be used.

3.1.4 Added stiffness due to immersion and flow


Next to the added stiffness as a consequence of the buoyant force, there is also a stiffness component, generated by the flow. This stiffness may be assessed by calculating the stationary flow force for each position of the object. An example of stiffness due to flow is the weather vane. When it is not in the right position, there is a retrogressive moment that is proportional to this, with a small angular deflection. The moment is also proportional to the square of the velocity. When the weather vane is situated wrongly in the wind, a negative added stiffness due to flow is generated!

3.2 Added water mass


In its most simple form, the added water mass is the water that is present at the top of a piston in a cylindrical vessel (Figure A3.2). If the piston moves, the water has to move along, as it is locked in sideways, and the acceleration force results in an extra pressure on the piston surface equal to Lw times the (upward) acceleration of the piston. Lw could be referred to as the added water length; in case the piston surface is not horizontal, this length varies across the surface. The value of Lw multiplied with the acceleration of the piston is a measure of the local pressure. If, in case of a vibration, the kinetic energy is considered in relation to the potential energy of a spring for the determination of the natural frequency, then the entire water mass is involved in the kinetic energy. In case of a vibrating object in water that is not flowing, a periodically varying potential flow is generated.

52

Delft Hydraulics

Figure A3.2: Added water mass in case of a piston. See as an example Figure A3.3, an infinitely long vibrating strip in an unlimited fluid, in which the flow pattern is two-dimensional. In the absence of a free water surface, the added water mass is completely independent of the frequency (However, see remarks in Paragraph 3.2.7). In case of the infinitely long strip of Figure A3.3, the added water mass equals the mass of water in the imaginary surrounding circular cylinder of this strip. This however is a different situation from the one of the vibrating piston. The flow pattern is such, that the water velocities further away from the strip are smaller, and the contribution to the kinetic energy is proportional to the square of the water velocity. The aggregated kinetic energy determines the eventual magnitude of the added water mass. As mentioned above, the magnitude of the added water mass may also be found by considering the pressures at the surface of the strip in relation to its acceleration. The added length, Lw, varies considerably across the width of the strip. Here too, the pressure distribution corresponds to the shape of the surrounding cylinder. Lw depends on the vibration direction, and is always coupled to the dimensions of the vibrating body.

53

Delft Hydraulics

Figure A3.3: Potential flow in case of a vibrating strip (Lamb, 1932).

3.2.1 Methods of calculation for two-dimensional situations


When calculating the added water mass, as discussed in the previous paragraph, the influence of the waves generated by vibrations may be ignored. It is assumed that the flow has no influence on the added water mass. In that case, a periodically varying potential flow is generated due to the vibration, in which the water velocities across the entire fluid area are synchronous and in-phase with the vibration velocity of the structure. As already discussed, the potential flow therefore is similar to that in case of permanent flow and may therefore be calculated with the available classic methods. By definition, for potential flow the following applies: u=

, v= x y

and

w=

(A3.10)

and with this the continuity equation may be presented as:

u v w + + =0 x y z

54

Delft Hydraulics

and therefore:
2 2 2 + + =0 x 2 y 2 z 2

(A3.11)

This is often written as:


= 0

A3.12)

This is the so-called Laplace equation. The pressures in the fluid generated by the vibration after linearising and ignoring the external forces may be written as: p=

+ constant t

(A3.13)

As the water velocities are in-phase with the vibration velocity of the object, the potential is also in-phase with the vibration velocity again and again, and therefore proportional to it. And as Equation A3.13 also holds for the pressures on the wall, these are proportional to the acceleration of the vibration. If the relation between the velocity potential on the vibrating wall and the corresponding vibration velocity have been calculated, the aggregate of these pressures, when divided by the vibration acceleration, results in the added water mass. In Part C, Chapter 3, the method of calculation is further discussed for a vibrating strip (translatory and rotational vibration). Kolkman (1988) proposed a relaxation calculation, using a spreadsheet program. The boxes of the spreadsheet correspond with the grid in the x-y plane. In each box, the equation is entered that is used for the adjustment of the potential value. The wall equations are different for the walls that vibrate and the walls that do not vibrate. The equation only needs to be introduced once for each type and may then be copied for the other boxes. As this calculation may be put into the program fast, it is very well suited to be used at gates in which there are many gate positions and water level combinations. (See Part C, Appendix III)

3.2.2 Added water mass for various two-dimensional structure shapes


Figure A3.4 shows the added mass for different values of the threshold height of the bottom for a horizontally vibrating gate with water on one side. The maximum value, in case of a flat bottom (h1 = h2), corresponds with what Westergaard (1931) found for (floodcontrol) dams that experience a horizontal acceleration due to an earthquake. For a gate under water, Figure A3.1 may be used. The pressure at a vibrating gate has a component that is proportional to dy/dt (indicated in Figure A3.1 as the velocity V) and a component proportional to d2y/dt2 (indicated in Figure A3.1 as dV/dt). The latter is representative of the added water mass.

55

Delft Hydraulics

Figure A3.4: Added water mass (in case of high vibration frequencies) of a horizontally vibrating gate with length L on top of a threshold with water at one side (Schoemaker, 1971). For vibrating structures the vibration frequency generally is so high, that also the term 2h/g is very great. In that case the added water mass is independent of the frequency. The term proportional to V is representative of the added damping due to wave radiation; this usually is small. Figure A3.5 gives an overview of the added water mass of a number of stiff bodies in unlimited water and in water with a free water surface. In the latter case, these are calculations relating to ships, but the results may also be used in other situations.

Legend of lower part in Figure A3.5: 2 strips, distance a = mw / d 2 L d = width of strip L = lenth of strip

2 square discs = mw / d 3 d = chord

56

Delft Hydraulics

Figure A3.5: Overview of coefficients for the added water mass of various bodies (Wendel, 1950 and Sarpkaya, 1960).

57

Delft Hydraulics

3.2.3 The proximity of a wall


Figure A3.6 shows the situation of a vertically vibrating, wide body in the proximity of a horizontal bottom. As long as the gap height, d, is small relative to the gap length, 2a, a one-dimensional flow calculation may be carried out. As the flow profile narrows due to the proximity of the wall, the water velocities and the kinetic energy increase; therefore the added water mass increases when the volume of the enclosed water decreases.

Figure A3.6: The added water mass of a wide body in case of a small distance to a wall. The calculation is as follows. Taking symmetry as a basis, the vibration velocity of the body, dy/dt, causes a horizontal water velocity, V(x), beneath the body. This (across the vertical average) water velocity follows from the continuity equation: V ( x) = At the lower end the exit velocity is: V (a ) = a dy d dt (A3.15) x dy d dt (A3.14)

As the vibration velocity, and therefore also the water velocities, are small, only the local acceleration of the water particles is of importance, and therefore this means that (F = ma, used in the direction x on an infinitesimal volume):

p V x d2 y = = d dt 2 x t
p = pressure change relative to the permanent situation Integrating this across distance x results in: p= 58 1 d2 y x2 + constant 2 dt 2 d

(A3.16)

(A3.17)

Delft Hydraulics

but as x = a, this means that the pressure equals that of the permanent situation, = 0, and therefore this results in: p= 1 d2 y x2 a2 d 2 dt 2 (A3.18)

The force exerted by the water in the gap, if the body has a length L, is:

1 d 2 y a 3 a 3 2 a3 L d 2 y K = 2 L pdx = 2 L 2 = 2 3 d dt 0 2 dt 3d d
a

(A3.19)

The added water mass is calculated by dividing this by d 2 y / dt 2 : mw = 2 a3 L 3 d (A3.20)

Taking the data of Figure A3.1 as a basis, a refinement of this is possible, by entering a fictive extension of the pipe length. This fictive pipe length is a way of taking into account the flow inertia of the water in the area outside the pipe. This fictive extension on each side amounts to CLd. This means that at the outer end of the pipe, the pressure is not 0, but (see Figure 3.1): p (a ) = CL d dV (a) dt (A3.21)

For the deduction of A3.21 use was made of the fact that the pressure at the water level remains constant, so that the hydraulic head necessary to accelerate the flow in the water below coincides with a pressure increase at the outer end of the pipe. After introducing Equation A3.15, the pressure increase of Equation A3.21 results in a pressure proportional to d2y/dt2. This pressure operates as an extra pressure across the entire length 2a. All this eventually results in the expression below for the added water mass: 2 a3 L mw = + 2CL a 2 L 3 d (A3.22)

59

Delft Hydraulics

3.2.4 Polar mass inertia moment in case of rotational vibrations


Figure A3.7 shows the added polar mass inertia moment for a number of bodies, in case of a rotational vibration, using the data of Wendel (1950) and Kolkman (1988). The method of calculation used is discussed in Part C, Paragraph 3.2.4.

Figure A3.7: Added polar mass inertia moment, Ip, in case of rotational vibrations of structures with length L, ignoring wave radiation (two-dimensional situations).

60

Delft Hydraulics

3.2.5 Added water mass of a culvert gate

Figure A3.8: Situation of a valve. In case of a horizontally vibrating gate, in particular in case of a closed culvert (Figure A3.8), the added water mass strongly depends on the position of the gate. The added water mass also depends on the vibration frequency. The complete calculation of the forces that operate on a culvert gate that vibrates in the flow direction may be found in Paragraph 4.4.2.

3.2.6 Assessment of the added water mass of a gate as part of a wall of a hemi-space
When ignoring the wave radiation, the added water mass of a gate that is immersed in water on one side, may be assessed by assuming that the water approaches and runs off on all sides. The situation in Figure A3.1 may serve as an example. The water domain is divided into an area enclosed by the surroundings of the gate, the so-called inner area (with r < R), and an outer area (with R < r < R0), see Figure A3.9. For the inner area it is assumed that this water completely operates together with the added water mass, so per unit of width: mwbi =

d 2

(A3.23)

For the outer area it is assumed that the water approaches and runs off radially with a velocity V, so: V= d dy / dt 2 d dy = r / 2 r dt (A3.24)

61

Delft Hydraulics

Figure A3.9: Schematized representation concerning the calculation of the added water mass of a vibrating gate, as part of a wall. For the pressure gradient (in linearised shape): dp dV 2 d d2 y = = dr dt r dt 2 (A3.25)

Through integration across r, the pressure jump, p = p(R)-p(R0), may be calculated between the outer area, where the pressure remains constant (at distance R0), and the contour at distance R. This eventually results in: d 2 y R0 p = d 2 ln dt R 2 (A3.26)

The force exerted on the gate is obtained by multiplying the pressure drop p with the gate height, d. Dividing this by the acceleration, d2y/dt2, results in the influence of the outer area on the added water mass per unit of width, so: mwbu = 2

d 2 ln

R0 d

(A3.27)

It appears that, in case of an infinite outer radius, R0, the added water mass also has no maximum value. Now, the rough assumption is introduced, that the outer radius R0 (at which place no influence of the vibration on the pressure may be noticed) equals the water depth, h. This results in: mw = mwbi + mwbu =

d 2 +

d 2 ln

h d

(A3.28)

62

Delft Hydraulics

Because, by definition mw = CL d 2 (equation A3.3), this results in:

CL =

ln

h d

(A3.29)

In the diagram of Figure A3.9 this method of assessment is compared with the complete solution, using the potential theory (see Figure 3.1 and Equation 3.3). The equation indicates that this method of assessment results in a useful initial approach.

Figure A3.10: Diagram for the calculation of the added water mass of a vibrating round disc, as part of a wall. A similar approach may also be used for a round vibrating disc as part of a flat wall. The outcome of this is used in the bath plug vibration, which will be discussed below (Paragraph 4.4.2). Here too, an inner and an outer area are introduced. For the inner area the following applies (hemisphere):

mw =

2 R3 3

(A3.30)

For the outer area the velocity is obtained by dividing the discharge displaced by the vibrating disc by the surface of the hemisphere. This results in:

V=

R 2 dy / dt 1 R 2 dy = 2 r 2 2 r 2 dt

(A3.31)

63

Delft Hydraulics

Now the following may be found for the radial pressure gradient (in linearised shape): dp dV = dr dt and therefore: dp 1 R2 d2 y = 2 2 dr 2 r dt (A3.32)

(A3.33)

The pressure jump is again obtained by integrating the pressure gradient. This time it is not necessary to define the radius where the outer area ends, because also for infinite water a finite hydraulic jump p = p(R)-p(), with p() = 0, is found: 1 R2 d2 y 1 d2 y p = 2 2 dr = R 2 = p ( R) 2 r dt 2 dt
R

(A3.34)

This pressure again operates on the surface of the disc, R2. After dividing this by the first acceleration d2y/dt2, the added water mass that corresponds with the outer area is found again:

mwbu =
The total added water mass now is:

R3

(A3.35)

mw = mwbi + mwbu = R 3 (

2 + ) 4 R3 3 2

(A3.36)

3.2.7 Additional remarks about the added water mass



As stated above in Chapter 2, modern reviews of the added water mass start from the basis of the concept of added mass matrix. About this, a number of things are remarked in Paragraph 3.6 and it is discussed in more detail in Part C. The concept of added water mass also plays a role in case of stationary objects, on which forces are exerted due to waves (the Morison equation); the value used for this however is greater than the one that is found in reviews about vibration. An extra term is added, that is equal to the content of the object multiplied with the density of the fluid. This is to do with the initial situation of the calculation, i.e. the undisturbed wave, in which also the water inside the contours of the object are in movement. To cause the water surrounding the object to stand still, the water in the inner area also needs to be brought to a standstill. According to Figure A3.1 the added water mass is greatest in case of high frequencies. This however only applies to a gate around which water cannot flow. As soon as water is present beneath or on top of the vibrating object, in case of vibration a reverse flow will be generated. At very low frequencies, the amplitude of the radiating wave decreases and the reverse flow proportionally increases. At low frequencies even situations occur, in which the water surface behaves like a fixed horizontal wall. The

64

Delft Hydraulics

added water mass then again is independent of the frequency and equals that at an object placed inside a closed tunnel. As the wave radiation is coupled to the damping and the reverse flow to the added water mass, the added water mass is relatively great again in case of low frequencies. It is always greater than in case of the high frequencies, in which case there is no wave radiation anymore, but in which the condition of atmospheric pressure at the water surface remains valid. Consider as an example a vibrating object immersed in water in Figure A3.11. Similar results have also been found in calculating the added water mass of a horizontally accelerating or decelerating ship (Part C, Figure C3.13 in Fontijn, 1978).

Figure A3.11: Added water mass and added damping of an immersed sphere (Hooft, 1972).

3.3 Added stiffness


Essentially, there are three types of added stiffness, kw. These are the stiffness due to immersion, the stiffness due to quasi-stationary flow forces and the stiffness that is generated when an object suddenly experiences a change of position (the so-called sudden stiffness). The latter is a special case and particularly occurs at culvert gates, where the discharge in the culvert has such an inertia, that it does not immediately adapt to the changed position of the gate.

3.3.1 Vertical stiffness in case of a floating body


The stiffness of an object that is vertically immersed may be described as:

kw float = gAcross section

(A3.37)

The surface of the cross-section with the free water level is normative for this stiffness.

3.3.2 Stiffness due to flow


The flow force (i.e. the force coefficient) may vary relative to the position. In Paragraph 3.14 already a weather vane was mentioned, that experiences a retrogressive moment in case of a deviation from the leeward position of the wind. In case the weather vane is accidentally situated the wrong way round, then the reversing moment is zero at one 65

Delft Hydraulics

particular position, but in case of a deviation from this, the moment will increase proportionally to the angular deflection. This is referred to as a negative flow stiffness. Now, suppose an object, dependent of its position, experiences a flow force in the direction y, then the stiffness in the direction y may be described as follows:
k w flow = Fy y = CFFy 1 1 CFy V 2 L2 = V 2 L2 2 2 y y

(A3.38)

In this, Fy is the component of the flow force in the direction y, which is again expressed in a force coefficient CFy, multiplied with the flow pressure, multiplied with the surface area that is excited by the approach flow. The flow stiffness is generally proportional to the squared flow velocity or, in case of gates (where the flow velocity is proportional to the root of the hydraulic head), proportional to the hydraulic head.

3.3.3 The sudden stiffness


The so-called sudden stiffness is a stiffness that only occurs at gates. Suppose the situation is as indicated in Figure A3.8. When the gate is suddenly put in another position, the discharge in the tube will not immediately increase or decrease, as a consequence of flow inertia; especially when the tube is long, this will take a while. The same discharge that at first passed through the gap , now passes through a gap (+); the flow velocity in the gap decreases and the gap of the valve will decrease with the square of this velocity. The force operating on the gate, which is proportional to the squared velocity, therefore changes from Fpermanent to (the force coefficient remaining the same):
2 Fnew = Fperm. 2 ( + y )

(A3.39)

This results in the added sudden spring stiffness by dividing the change of force by (-y):
k w sudden = Fnew Fperm. y Fperm. 2 1 = Fperm. 1 2 2 ( + y ) y

(A3.40)

The simplification in the second part of the equation is possible, assuming that y is very small, even relative to the opening . Also, it is implicitly assumed that the discharge coefficient of the opening does not change. Equation A3.40 indicates that the smaller the gap becomes, the more the sudden spring stiffness increases. If, as in Figure A3.8, y is chosen in the direction in which the gap is increased, then the Fperm operates in the direction of y, in case the gate is positioned downstream of the culvert. The sudden stiffness is positive when, as in Figure A3.8, the gate is positioned downstream of a tube or a culvert. In case the gate is positioned upstream, then the sudden stiffness is negative. As this stiffness also increases, the smaller the gap becomes, it will always be possible for it to become bigger than the mechanical spring stiffness of the 66

Delft Hydraulics

gate itself or of the connection. The gate is then closed by the suction. Whether this will actually happen depends on various factors; in the end, it is assumed in this calculation, that the gate changes position so fast, that the discharge has no time to adapt. The sudden stiffness is frequency-dependent again. It occurs in case of a highfrequency vibration, as in that case the discharge tends to remain constant.

3.4 Added damping


The added water damping is mainly caused by the flow. As already mentioned in Paragraph 3.1, the energy loss as a consequence of the runoff of small waves generated by vibration only plays a subordinate role in case of vibrating structures.

3.4.1 Damping due to wave radiation


According to Kolkman (1976), the damping due to radiation of the waves in case of a horizontally vibrating deep gate that penetrates the water level, may be characterized as:

cw =

2 g 2

(A3.41)

This equation represents the damping per unit of gate width. As the gate is immersed in the water on both sides, this damping still needs to be doubled, in order to obtain the total added damping. It is clearly visible that the added water damping at higher frequencies is small.

3.4.2 Added damping due to flow in case of an object vibrating in flow direction
The flow forces, that are square with the flow velocity, change due to the velocity of movement of the vibrating structure. As a consequence of this vibration velocity, both the magnitude of the relative approach flow velocity and the angle of approach change. By observing the instantaneous situation across the whole vibration period, the energy transfer to the water may be calculated. As long as the relative velocity only deviates a little from the approach flow velocity in the stationary situation, linearising techniques may be used. Greater movements require the complete approach. Suppose an object in flow experiences a flow resistance of:

F = Cw

1 V 2 L2 2

(A3.42)

In this, L2 is a measure of the frontally approached surface and Cw is the coefficient. When reviewing a vibration in the flow direction with vibration velocity dy/dt, then V changes relative to the object, and therefore F also changes. V becomes:

67

Delft Hydraulics

V = V perm.

dy = V perm. + dV ; V perm. = undisturbed flow velocity. dt

The change of force may be calculated using Equation A3.42, by differentiation and then replacing dV with dy/dt, the value by which the flow velocity, relative to the moving object, decreases: dF =

F dy dV = Cw L2V dV dt

(A3.43)

As, in case of positive dy/dt, the drag force decreases, it is as if on top of the permanent resistance, a force operates that is opposite to the vibration direction. By definition, the added water damping, Cw, is the opposing force (which is now positive), divided by the vibration velocity, therefore:
cw = CW L2V

(A3.44)

Knowing that Fperm = CwL2V2, this may also be written as:


cw =

2 Fperm.
V

(A3.45)

The flow damping, Cw, is therefore always coupled to the flow velocity, V, in a linear way.

3.4.3 Added damping due to flow in case the vibration direction differs from the flow direction
In case of a circular cylinder or a sphere, the damping may also be calculated when the vibration direction is not the same as the flow direction. An example is the situation presented here, in which the vibration direction is perpendicular to the flow direction (Figure A3.12).

Figure A3.12: Diagram for damping calculations in case the vibration direction is perpendicular to the flow.

68

Delft Hydraulics

In case the cylinder vibrates with an instantaneous velocity dy/dt, then the angle of approach flow changes with a value , for which:

tan =

dy / dt V

so: sin =

dy / dt V + (dy / dt ) 2
2

(A3.46)

The relative approach flow velocity becomes:


Vrel . From this it may be calculated that:
2 1 2 2 1 2 2 dy F = CW L Vrel . = CW L V + 2 2 dt

dy = V + dt
2

(A3.47)

(A3.48)

L2 is again the measure for the frontally approached surface. The force component in the direction y is obtained by multiplying F with sin . If this force is divided again by the vibration velocity, dy/dt, then this results in the flow damping. As long as dy/dt is small relative to V, this results in: F 1 2 cw = CW L2 V 2 + ( dy / dt ) perm. V 2 (A3.49)

The damping is therefore bisected relative to that of the vibration in the flow direction; on the one hand this reduction is caused due to the fact that the relative velocity changes less because of the vibration, and on the other hand because the change in drag force only partially operates in the direction y. For an arbitrary object that vibrates perpendicular to the flow direction, a more general relation may be deduced, when the stationary lifting force coefficient, CL, at a varied angle of approach flow, , as well as the coefficient for the flow resistance, Cw (which therefore relates to the force in the flow direction) are given. This deduction, which is based on the method of Den Hartog (1956) for the determination of galloping (negative damping), with A = frontally approached surface, results in: cw = 1 C VA( L + CW ) 2 (A3.50)

The second term follows directly from what was deduced in Equation 3.49 for the circular cylinder (in which CL/| = 0 because of symmetry). In both cases, it may occur that CL/| obtains a negative value that is greater than the positive drag term Cw. The total hydrodynamic damping then is negative and self-excitation will be generated if the mechanical damping is not big enough. This will be further discussed in Paragraph 5.4.1.

69

Delft Hydraulics

3.4.4 Flow-induced damping of a gate


The flow-induced damping of a gate will not be calculated explicitly in this paragraph. In Chapter 4 however, the self-exciting vibrations at gates will be discussed in more detail, based on negative flow damping. In Paragraph 4.4.3, the values for the negative damping are calculated for a gate with flow beneath it that vibrates vertically. For gates, the quantified positive water damping is of little significance, because strong vibrations actually only occur when the water damping is negative.

3.5 Inertia of flowing water


In Figure A3.1 the added water mass may be read in terms of an added length CLd, in which d is the height of a horizontally vibrating gate. The total pressure build-up p, on average across the height of the vibrating gate, is both determined by the added water mass (to be calculated using the coefficient CL), and the added damping (to be determined using Cr). p = Cr d V + CL d dV dy d2 y = Cr d + CL d 2 dt dt dt (A3.51)

In case the gate is shifted back a little, the inertia influence of the water increases and the added length needs to be increased with a piece of culvert length L; the total added length therefore becomes L+CLd. In a culvert, the flow inertia is determined by the culvert length, L; this inertia determines whether at the opening of a gate the discharge immediately corresponds to the gate opening, or whether there is some degree of lag. In case the culvert opens into free water, the culvert length needs to be increased with the added length CLd; this therefore implies a very different use of Figure A3.1.

Figure A3.13: Flow inertia for both a vibrating gate and a varying discharge. (a) vibrating gate (b) velocity variation due to discharge variation, and (c) coefficient CL

70

Delft Hydraulics

Figure A3.1 however may also be used for the inertia of the water that flows beneath a gate. It has been used for the composition of Figure A3.13. In this, the inertia line of Figure A3.1 has also been included again; as the frequency in case of vibrating structures is relatively high, the line for 2h/g = may suffice. In case water flows beneath a gate and the discharge changes due to a change of the gate position in vertical direction (vibration), then the total flow inertia (a kind of fictive culvert length) is represented by the thickness of the gate, increased with an additional culvert length CLd upstream and downstream. The discharge change is translated into an average velocity change of the water, dV/dt, across the lifting height, while the gate position remains unchanged. This discharge change causes a similar pressure build-up as in the acceleration, d2y/dt2, of the horizontally vibrating gate. As the water levels upstream and downstream are different, the CL value on both sides are also different; in reality however this does not make a lot of difference. This method assumes that it makes no difference whether the change of the discharge is caused by altering the hydraulic head or by altering the gate position (as an example, as a consequence of a gate vibration). This assumption is reasonable, as the inertia operates throughout the flow area and not only at the gate itself.

3.6 Water-induced coupling of natural vibrations

Figure A3.14: Periodic potential flow in case of a vibrating L-shaped gate in an added water mass. Delft Hydraulics Report M1322. Figure A3.14 shows the flow pattern in case of vertical vibrations of an L-shaped gate. This Figure has been determined on the basis of measurements using an electric analogon. It is clear that the flow pattern is strongly asymmetric. In that case, the pressures at the sides of the vertical plate are different as well. During the acceleration upward on top of the horizontal plate extra pressure will be generated, which results in a force toward the left on the vertical plate. By reverse, a horizontal acceleration will cause a vertical force. Contrary therefore to 71

Delft Hydraulics

structures, in which all mass terms are not coupled as regards the acceleration forces, these are coupled in case of structures in water. In case of a system of multiple masses and springs that generate a vibrating movement in a fluid, the masses are also coupled due to the accelerations; the movement of one mass causes pressures throughout the fluid, which also operate on the other masses.

Figure A3.15: Properties of added water mass, translated into a mechanical system (Kolkman, 1976). To clarify this, Figure A3.15 once again represents how this would translate into a mechanical system. Diagram a shows the situation in case the added mass for horizontal and vertical vibrations would be the same. In diagram b, in which masses are attached to hinges, the mass only operates when accelerated in the direction of the securing rod. Therefore, the added mass is different for horizontal and vertical vibrations. In diagram c also the coupling terms in the added water mass are generated. In case of continuous-elastic structures, it is found that due to the influence of the water, not only do the resonance frequencies decrease considerably, but also the vibration mode may change considerably. A striking example of this may be found in Figure A2.6, in which the vibration modes in dry conditions and in water are indicated for the visor gates of the Hagestein Weir (for further details concerning these gates, see Paragraph 6.2, example a). In order to forecast such an effect of the added water mass on the vibration mode, the 72

Delft Hydraulics

couplings mentioned above need to be included in the calculation. This kind of calculation is very complex and necessitates the use of calculation models. An interesting case, in which the coupling of vibration modes due to the added water mass caused (strong) vibrations, may be found in Paragraph 6.1, example e. As the horizontal and vertical vibrations both change into transverse vibrations in the same direction due to the coupling, a combined vibration is generated that is of an instable nature.

Figure A3.16: Influence of culvert length on coupling. In case of a gate in a culvert, the added water mass is much greater, and also the coupling between the different vibration modes of a gate may be extra strong. In case of Figure A3.16, a horizontal acceleration of an almost closed gate will accelerate the entire water mass in the culvert, which is accompanied by extra pressures near the gate that operates on the protruding upper lip. Therefore, a vertical force is caused, which also generates a vertical vibration component. Although so far this has not been explicitly proven, there are strong indications that the coupling factors in case of the added water mass are symmetric: a unit acceleration in horizontal direction generates a vertical force equal to the horizontal force that is generated in case of a unit acceleration in vertical direction. It is not only the added water mass that causes a coupling; for if the gate in Figure A3.16 vibrates vertically, the flow-through discharge will also vary periodically (the acceleration of the water is (d2y/dt2) multiplied with (a/h), and this again is accompanied by the pressure variations at the gate that result in a horizontally operating force. These days, there are calculation programs for the determination of the passive behaviour of elastic structures in water (without the influence of the flow therefore), in which the coupling due to the added water mass is fully taken into account. These have already been used for the gates in the storm surge barrier Eastern Scheldt. The DIANA calculation package developed by TNO, and also used by Rijkswaterstaat, includes a water module as well (see also Part C, Paragraph 3.3.5). In a dynamic Finite Element Method, as presented in Equation A2.1 for a system with multiple degrees of freedom, the dynamic for the single mass spring system is converted into:

73

Delft Hydraulics

d2 y dy + c + k y = F (t ) 2 dt dt

(A3.52)

(in which the underlining means that these are matrices of the entire system, which may have a large number of degrees of freedom). In A3.52, y is the displacement vector. Thus, the equation of the system in water (A2.29, while ignoring the terms that are not coupled to the movement in a linear way) is converted into: (m + mw ) d2 y dy + (c + cw ) + (k + k w ) y = F (t ) + Fw (t ) 2 dt dt (A3.53)

Considering the above, it is necessary that both elements, added water mass and added stiffness, are introduced into the equation, including the couplings. This is expressed in the matrices in the non-diagonal terms, which then become unequal to zero.

74

Delft Hydraulics

EXCITATION SOURCES FOR VIBRATIONS AND FLUID RESONANCES AT GATES AND VALVES 4.1 General regarding gate vibrations

There are various causes that may generate vibrations at gates. As indicated above, in the classification of Chapter 1, these are: excitation due to turbulence; excitation due to flow instability; self-excitation (excitation generated by the movement itself); amplification of the excitation due to fluid resonance (this, therefore, is not a real cause of vibrations); instable fluid resonance (self-exciting resonant rise of fluid). At gates, self-excitation is by far the most occurring cause of dangerous vibrations. The self-excitation may relate to either the gate vibration in the resonance frequency of the suspended gate, or to fluid resonance, such as a standing wave (a compression wave or a surface wave) with the corresponding wave frequency. In case of self-excitation the vibration or the standing wave is such, that the greater the vibration or wave amplitude, the stronger the excitation. Only in case of great amplitudes, extra damping forces are generated, but acceptable limits regarding safety may already have been transgressed by then. This is why the mechanism of self-excitation is discussed in great detail. As far as the vibrations are concerned, this will be done using the characteristic example of the so-called bath plug vibration, as this is easy to understand. But the mechanism is also representative of vibrations at many other gates, at which heavy vibrations have been observed, such as: the sector gate, inclined placed lifting gates and many others. This chapter discusses the various excitation sources in the order given above.

4.2 Excitation due to turbulence and due to periodic release of vortices


For the description of excitation due to turbulence, two characteristic values are important: the Strouhal number and the dynamic force coefficient. The Strouhal number is a measure for the excitation frequency. Often, in consulting practice, the Strouhal number suffices, in which case it is advised that the resonance frequency of the structure needs to be a factor higher than the excitation frequency. For, in that case the dynamic excitation operates like a quasi-stationary force which, in case of excitation due to turbulence, only amounts to a few percent of the stationary force (up to 10%). In case of turbulence there is a non-periodic force that may be described with a spectral distribution. For further details, see Paragraph 2.2.8. Excitation due to turbulence may be caused by turbulence in the approach flow, or by what happens around the part of the structure concerned. Turbulence in the approach flow is very variable; at gates it is not taken into account, as the approach flow velocities relative to the normative velocities in the gate opening are small in any event.

75

Delft Hydraulics

Definition of the Strouhal number: Just like in Paragraph 2.2.6, the Strouhal number is defined as:

S=

fL V

(A4.1)

In this, for as far as this concerns periodic excitation, f is the excitation frequency due to the flow, L is a measure for the size of the gate or the component of the gate that experiences the excitation, and V is the approach flow velocity. S is a constant, provided the flow pattern is univocally coupled to the shape of the gate or the component concerned: only then are all vortices and eddy currents within the turbulent flow field, as far as their dimensions are concerned proportional to the size of the object, and is their velocity of movement proportional to the approach flow velocity. In case there is noise excitation or excitation that is spread out across a certain width of the frequency band, then f may relate to the dominant frequency or f is a variable (see Paragraph 2.2.8). If S is to be a constant, then the following preconditions need to be met: a. b. the gate or the component concerned is sufficiently immersed in the water, so that an increasing flow velocity does not cause a deformation of the water level that influences the local flow pattern, the Reynolds number Re = LV/ (in which is the kinematic viscosity of the fluid) needs to be large enough, in order for there to be no more influence of the viscosity on the flow field. With the exception of the situation at round cylinders, in case of prototype structures this precondition is met in every respect. In scale model investigations it is always necessary to verify that the Reynolds number (a measure therefore of the influence of the viscosity) is still large enough for representation of the prototype situation.

At gates, excitation due to turbulence is rarely or never an issue. References that do discuss this are mostly dated sources, in which it was assumed that the vibrations that in reality relate to self-excitation, were in fact caused by turbulence. In general it is recommended to take a resonance frequency, fn, that is a factor 2 or 3 (or more) higher than the dominant excitation frequency, f, due to flow. So:

fn L > (2 3) S V

A4.2)

In case the factor is 3, the amplification factor of the response still remains very small. Figure A2.2 shows that the dynamic amplification factor, A, is of an order 1.1 in case of a purely harmonic excitation. In case of noise excitation it is even smaller. It is also not expected that the response of the structure itself influences the excitation frequency again. Circular rods are very sensitive to this when they are low-damped, but even here, it appears that no shift of the excitation frequency occurs. This follows from studying Figure A2.4 (an example, moreover, that relates to instable flow). In this, Figure Vr = 5 corresponds to a resonance frequency that coincides with the dominant excitation frequency. This shows that, in case of a higher resonance frequency (in which Vr < 5), the frequency of the response continues to coincide with the line S = 0.2, corresponds with the excitation frequency without
76

Delft Hydraulics

frequency shift. In case Vr > 5, the frequency does stay close to the resonance frequency, which points to the fact that the excitation frequency in that case does shift toward the resonance frequency (see also Paragraph 5.3.8). Most gates consist of elements of which the normative length, L, for the Strouhal number, is very small (beams, a thickness of a bottom edge, and so on), so that the excitation frequency in that case is very high. It is certainly not possible to build the entire gate with suspension in such a stiff way, that also the lowest natural frequency exceeds the highest excitation frequency. There is also no need for that, for those modes (vibration modes) that are not caused to vibrate due to self-excitation, experience a strong flow damping. It is therefore recommended to only sufficiently reinforce those elements that are loaded due to turbulent excitation, to meet the criterion of 2 to 3 times the dominant excitation frequency.

Figure A4.1: Strouhal numbers for horizontal turbulence excitation of culvert gates (Abelev, 1959). Although the excitation due to turbulence is not normative for the design of gates and hoisting equipment, for the sake of completeness a few measuring results concerning Strouhal numbers at gates are given here. Figure A4.1 represents measurements of Abelev (1959), which at a later stage have been referred to by many other authors.

77

Delft Hydraulics

Figure A4.2: Strouhal numbers for vertical excitation of the bottom edge of a gate (Naudascher, 1964). In Figure A4.2 the dominant Strouhal number is presented for vertical excitation at the bottom edge of a gate.

4.3 Excitation due to flow instability


Flow instability generates a strong periodic excitation and that, in case of gates, occurs in the following cases: 1. 2. 3. 4. 5. if the point of separation is not fixed by a sharp edge, but consists of a rounded shape, or; if, after release, the flow attaches again (further down) against a wall of the structure, in which the point of attachment is not fixed by a spoiler or edge, or; if the flow attaches again and the location of this is fixed, then a free boundary layer with undulations is generated. Periodic pressure fluctuations with discrete frequencies are generated within the enclosed vortices, or; at a gate, in case of overflow, and a flow pattern that is between a situation of diving nappe and a situation of nappe deflected toward the water surface, or; at a gate with a an overflowing nappe, beneath which there is an enclosed volume of air.

In all these cases a certain periodicity in the flow load is generated. Under certain conditions, a small gate vibration is already sufficient to cause the separation of the flow or 78

Delft Hydraulics

the periodic attachment to vary with the frequency of the gate vibration. The shift of the excitation frequency toward the resonance frequency of the structure is called frequency synchronisation or lock-in. In case of a non-fixed point of separation (at a rounded shape), there is a much lower pressure in the area where the flow has attached, as compared with the area past the point of separation, and also in the area between the point of separation and attachment, there is a lower pressure, as compared with the area of flow further down. This means that the periodic shift of these points of separation and attachment due to the movement of the structure results in periodic excitation. If the structure does not move, then the length of the zone between the points of separation and attachment is a measure of the dominant excitation frequencies. This relates to the instability in the free boundary layer (the transition between continuous flow and the enclosed vortex). A small periodic pressure variation in the fluid volume locked in between boundary layer and structure, in the proximity of the point of separation of the flow, may cause an undulating boundary layer, as a result of which the amplitude downstream increases. As the locked-in volume between boundary layer and structure remains more or less constant, the volume variation as a consequence of the undulations downstream are converted into a pressure variation that also exerts its influence in the proximity of the point of separation. Because of this, the undulations in the boundary layer are amplified again. In cases 3 and 5, in which there is a fixed length of the free boundary layer or of the falling nappe, vibrations of gates occur in a number of discrete frequencies. The undulations in the boundary layer and the falling nappe are caused and/or amplified by the vibrations of the gate. The available length needs to correspond exactly with a number of times the wave length of the undulations.

Figure A4.3: Gate geometry of the investigation of Martin et al (1975). Figures A4.3 and A4.4 show the results of pressure measurements at a type of gate that may generate an instable point of separation, in combination with a flow that attaches 79

Delft Hydraulics

again. Here, the length of the free boundary layer is completely determined by the geometry of the gate. The situation is critical at that particular lifting height, at which the flow upstream separates and the free boundary layer is still influenced by the downstream edge. According to Marin et al (1975), more frequencies are possible; one criterion is that the gate width b = (n0.33) is (n = 1, 2, ). The wave length of the undulations in the boundary layer is estimated using = V/f (V is the flow velocity in the contracted nappe, f is the excitation frequency of the water). The factor 0.5 results from the estimation that the boundary layer is located in the middle between the area of flow and the locked-in vortex, so that the velocity in the boundary layer is between both velocities. The factor 0.33 results from the fact that, in case of periodic undulations of the boundary surface there is no periodic variation of the volume that is enclosed between the boundary surface and the bottom of the gate. This volume variation therefore must be 0, because the water is not compressible. A tendency of volume and pressure variation however is generated, that generates a feedback with the instable separation of the flow. Naudascher (1994) further indicates that the flow instability may be prevented if (r+e)/b is greater than 1 or when the bottom of the gate is inclined at an angle greater than 45.

80

Delft Hydraulics

Figure A4.4: The frequency spectrum of the pressure beneath the gate and the dominant Strouhal numbers of the investigation of Martin et al (1975).

81

Delft Hydraulics

Figure A4.5: Some situations in which a free boundary layer may be generated. Figure A4.5 shows some more situations in which it was observed that a periodicity is present in the pressure fluctuations: a gate recess, the rectangular bottom of a gate and the underside of a random structure with straight angles. The length L on which the number of wave lengths is based, is indicated in the Figure. At the gate, the Strouhal number follows from Figure A4.2. In case of very small openings, the distance between separation and attachment of the flow is approximately proportional to the lifting opening. There, the Strouhal number based on the gate opening, should be constant. This however was not found; apparently there is no dominant frequency present anymore in case of a gate opening that is smaller than 0.3 times the gate thickness. In case of lifting heights greater than half of the gate width, the nappe separates from the gate completely and no locked-in water volume is left. In that case, flow instability plays no role. For a gate recess the following holds (Ethembabaoglu, 1978): S= fL 1 = n V 2 (n=1, 2, ....) (A4.3)

The instable situation in case of an overflowing nappe with enclosed air is discussed in Paragraph 4.4.7. It was found that instability occurs when (n+) wave lengths fit in the drop height. Analytically, this means that first the drop height is converted into a falling time, and then the following holds:

1 fT f = (n + ) 4

(with n=1, 2, ....)

(A4.4)

There are of course many situations, in which the flow first separates and then attaches again, and in which the free boundary layer may become instable in a similar way. This may happen, as an example, in the situation of two partially closed gates that are situated behind
82

Delft Hydraulics

each other in a culvert. If the concentrated water flow that is released from the upstream gate approaches the proximity or the edge of the downstream gate, a fluctuating flow is generated that results in great alternating forces. Fortunately the excitation frequency is relatively low in those cases, so that this does not always necessarily cause vibrations. Another situation is when a culvert gate is lifted and completely stored away in the gate shaft. The flow at the ceiling separates from the upstream edge of the gate shaft and will attach again downstream. The free boundary layer then has a limited length again, which causes a periodically varying load. A number of cases are known, in which the lifted gate was caused to vibrate as a result of this, or in which the water in the gate shaft started to move considerably. In case of instable flow situations, especially at slender structures, strong vibrations may be generated. About this, see Chapter 5. This may occur at gates with frame girders. This again, is called frequency synchronisation. The high vibration amplitude is also responsible for the fact, that the excitation across the length of the structure is no longer in-phase (phase synchronisation).

4.4 Excitation induced by the movement of the gate itself (selfexcitation) 4.4.1 General
Self-excitation at a structure occurs when the excitation force is caused by the movement of the structure itself and the stronger the movement, the more amplified it becomes. Because of this, an initial vibration gradually becomes stronger and stronger. In Paragraph 2.2.6, in which Equation A2.29 is discussed, it was already shown in the discussion concerning the added damping that, mathematically speaking, the water damping cw and consequently sometimes also the total damping may very well be negative. The consequence of this is that, after an initial disturbance, the free vibration does not damp out according to a negative e-power, but instead it grows according to a positive e-power. This growth continues for as long as there is a constant damping value (c+cw), until the vibration amplitude has become so great, that new elements start to come into play (the generation of positive damping, an impulse load against a wall, etc.). For as long as it is not known how, and at what amplitude, the boundary is reached, a negative total damping may never be permitted in the design process. In Paragraph 3.4 the added damping of an object in flowing fluid has been discussed. Equation A3.50 showed that the damping depends on (CL/ + Cw). This means, that when a greater angle of approach flow (oblique and from below) in case of steady flow generates a greater lifting force (this is normally the case), then the damping increases as well, and the added damping therefore is positive. This does not always have to be so, however. As an example, in case of a gate underneath which water runs off, and that has a rectangular underside for smaller openings, the flow at the upstream edge will separate and attach again further down at the underside of the gate. Between separation and attachment, there is an area with underpressure. If the gate moves up during a vibration, the apparent approach flow direction changes (relative to the gate, the flow in this case approaches more from above), because of which the flow tends to attach again less, and the pressure beneath the gate increases. A negative added damping may be generated. This is further discussed in Paragraph 4.4.5 under galloping. 83

Delft Hydraulics

Another example is that of the bath plug, which is hanging just above the seating during the draining of the bath. If the plug is lowered, then the discharge has to decrease: if the pipe behind the plug is completely filled with water, the flow needs to be slowed down, which is done as an example by holding the plug by the hand. The inertia force is proportional to the discharge decrease and this again is coupled to the velocity by which the plug is lowered. Therefore a suction force is generated that is proportional to the drop velocity. In case this concerns a vibration, a downward force is generated during the lowering, an upward force during the upwards movement. This, by definition, is a negative damping! The negative damping is proportional to the pipe length L (between the plug and the place where there is air enclosed in the pipe), and to the discharge change. The latter is proportional to the water velocity in the gap, with the root of the hydraulic head, therefore. In the above, it was assumed that the discharge adapts to the instantaneous position of the bath plug. This will be more or less so in case of a low-frequency vibration. If however the frequency is higher (because there is a stiffer suspension spring) and in case of a relatively great inertia (greater when L is greater), the discharge will tend to lag behind; in the maximum case of a very high frequency, the discharge even remains constant. This has been discussed in Paragraph 3.3.3, where it gave rise to the concept of sudden added stiffness. This added stiffness, kw, is positive when the plug is downstream of the pipe, and negative when the plug is placed upstream, as in Figure A4.6.

Figure A4.6: Diagram of the vibrating bath plug. In Figure A4.7 this has been represented in a qualitative way: in case of low vibration frequencies, the negative added damping is approximately constant, but it decreases in case of higher frequencies because the discharge then no longer follows the vibration.

84

Delft Hydraulics

Figure A4.7: Qualitative representation of the development of the added negative damping, of the added negative sudden stiffness and of the extra damping due to reverse flow as functions of the frequency, in case of a vibrating bath plug. As the discharge at higher frequencies tends to remain constant, a negative added stiffness is generated, that at a later stage also tends toward a maximum value. Moreover, something else happens at high frequencies that is coupled to the constant pipe discharge. The movement of the bath plug causes water to be displaced and when the pipe discharge remains constant, a local flow around the plug appears. There already was an initial gap flow; the extra surrounding flow causes an extra hydraulic head and therefore an opposing force that is proportional to the velocity of the plug and proportional to the initial water velocity in the gap (proportional to the square root of the hydraulic head). The situation is somewhat comparable to the cylinder that vibrates in the flow direction, resulting in an added damping proportional to the approach flow velocity (Paragraph 3.4.2, Equation A3.45). In case of the bath plug however, as a consequence of the movement of the plug, the counterflow increases more strongly than at the cylinder in free water. All this means that there is a new element of added damping that manifests more strongly, the higher the frequency. If however, the positive and the negative added damping neutralize each other, the socalled critical point of stability/instability of an undamped system has been reached. An initial disturbance then causes a harmonic vibration without damping out or resonant rise.

4.4.2 Theory of varying gap width applied to the bath plug valve
This paragraph discusses in detail the equation for the bath plug vibrations. Not only the outcome of the calculation is important, but also the equations themselves. The circumstances in which this type of vibrations occurs may differ widely. In reality, calculations have been used for:

85

Delft Hydraulics

a plug that moves inside a tapered pipe. Here another coupling is formed between gap width and vibration; a sector gate (Figure A4.11). In case of vibration, a variation in the gap width, both at the top and at the bottom of the gate, is generated. The plunger effect of the gate also deviates considerably; instead of a bath plug, a one-sided locked-in spring-like element that vibrates.

For each of those cases the equations had to be elaborated anew. The nature of the outcome however has remained the same. In order to make the theory useful, all magnitudes that vary with time are considered as the aggregate of the time-average value and a small periodic fluctuation that is a response to it. Thus it has become possible to consider only the first coefficient of these magnitudes and to linearise the non-linear magnitudes. The time-average conditions are also the conditions for the stationary, vibration-free situation. We start by introducing the fact that the discharge coefficient for the gap between bath plug and seating is constant and independent of the size of the opening. This makes it possible to calculate critical openings in which selfexcitation may occur.

Figure A4.8: Symbols used in case of a bath plug vibration. Figure A4.8 shows the symbols used. The added water mass upstream is added to the mass, m, of the bath plug. This was estimated in Paragraph 3.2.5 by assuming that the displaced water runs off radial-symmetrically (Equation A3.36). For the determination of the discharge that flows through the gap, the volume that is displaced by the bath plug is taken into account. 86

Delft Hydraulics

It is assumed that the gap is so small and creates so much flow resistance, that the pipe losses may be ignored.
Symbols The index 0 refers to the initial condition of the time-average value. The magnitudes that vary with time are: the pipe discharge, Q, the flow velocity in the pipe, V, the gap width, , and the hydraulic head of the bath plug, H. The value at which these magnitudes vary with time, around their respective average values, are indicated by accents. These therefore are Q', V', ' and H', in which ' equals the vibration displacement y again.

Figure A4.9: Both discharge components: the gap discharge and the plunger effect.
Discharge equations The discharge consists of two terms, the gap discharge and the plunger effect of the bath plug. This is clarified in Figure A4.9. Q = Qgap R 2

dy dt

(A4.5)

and so, with Qgap = (2 R) 2 g H :


Q = (2 R) 2 g H R 2

dy dt

(A4.6)

87

Delft Hydraulics

The flow velocity in the pipe, V, is:


V= Q 2 dy = 2 g H 2 R R dt V = V V0

(A4.7)

Furthermore:

(A4.8)

And so:

dV dV = dt dt

(A4.9)

Using the chain rule, and knowing that and H are time-dependent, this results in:
dV 2 d = dt R dt Note that: d dy = dt dt and: dH d(H 0 + H ) dH = = dt dt dt (A4.11) 2 1 dH d 2 y 2 g H + 2g 2 dt dt R 2 2 g H

(A4.10)

(A4.12)

therefore:
dV 2 dy = + 2 g H dt dt R R 2 g dH d 2 y 2 H dt dt

(A4.13)

To obtain an analytical solution, another simplification is introduced. In reality everything revolves around the question whether the system has a negative damping or not, i.e. whether in case of a disturbance, self-excitation may be expected. The analysis is limited to the area for which dy, dH and dV' are small relative to the average values of , H and V. In Equation A4.13 therefore, the instantaneous may be replaced by 0, the H by H0 and the V by V0. This results in:
dV 2 dy 0 = + 2 g H 0 R R dt dt 2 g d H d 2 y 2 H 0 d t dt

(A4.14)

Pipe equations The hydraulic head of the pipe is defined as the difference in pressure height just below the bath plug and the pressure height at the knee of the pipe of Figure A4.8. The latter is a constant. The hydraulic head of the pipe varies with time; this variation is coupled to a minus sign to the variation with time of the hydraulic head H' on top of the bath plug, for the pressure height on top of the bath plug is also assumed to be constant. The variation of the

88

Delft Hydraulics

hydraulic head of the pipe in its turn is also coupled to the acceleration of the water in the pipe. This results in: H = L dV g dt (A4.15)

The dynamic bath plug equations As the hydraulic head of the bath plug is defined in the direction of the flow and the direction of the displacement, y, operates exactly in the opposite direction, we find for the dynamic bath plug equation: m d2 y dy + c + ky = g R 2 H 2 dt dt (A4.16)

The part on the right represents the external force that operates on the bath plug. Equation A4.16 may also be written as:
H = m(d 2 y / dt 2 ) + c(dy / dt ) + ky g R 2

(A4.17)

Combination of equations A4.14, A4.15 and A4.17 By introducing the dV'/dt of Equation A4.14 into Equation A4.15, a connection is made between H' and y. As there already existed a second relation between these two magnitudes (Equation A4.17), H' may be eliminated. All this results in a third-order differential in y: d3 y d2 y dy 3 + 2 + + y = 0 (A4.18) dt dt dt The coefficients , , and are:

L 0 m 2 g g 2 R 3 H 0

0 Lc 2 g m L + + 2 gR g g 2 R 3 H 0

2 2 L H 0 0 Lk 2 g c + 2 gR g 2 R 3 H 0 R g

k gR 2

Now, the point of departure is a harmonic vibration. This results in a condition right between a condition in which the vibration causes resonant rise, and a condition in which this damps out. If in reality, in a certain situation and in the absence of mechanical damping, there is a vibration that will damp out or that will cause resonant rise, then by way of a thought

89

Delft Hydraulics

experiment, a mechanical damping, c, may be introduced that is of such a magnitude that the vibration becomes purely harmonic again. By comparing this fictive damping value with the actually present value, it may be determined whether the real situation is stable or not. The harmonic vibration may be represented as: y = Yeit Because of this, Equation A4.18 changes into: i 3 2 + i + = 0 From this, it follows that: (A4.20) (A4.19)

2 =

(A4.21)

As the sum of both the real and the imaginary terms must be zero, after equalization of / with /, after some deduction it may be found what magnitude the combination of mechanical damping and spring stiffness needs to be, in order to prevent vibrations (see Kolkman, 1976). In Part C, Paragraph 3.3.7 yet another way of elaboration is presented, that also results in an analytical solution (Kolkman, 1984). The harmonic vibration as in Equation A4.19 is introduced as a forced vibration. Next, it is calculated what magnitude the hydrodynamic forces are that operate on the bath plug. These forces are proportional to the vibration y and include a term that is in-phase with this movement, as well as an out-of-phase term, therefore: Fw = YA( ) + iYB( ) No water damping here means B = 0. For more details, see Part C, Paragraph 3.3. For the description of the solution resulting from Equation A4.21, two auxiliary magnitudes are introduced, the dimensionless mass factor, Cm, and the dimensionless stiffness factor, Ck. The mass factor is defined as: (A4.22)

Cm = m / Ap L

(A4.23)

Cm therefore is defined as the mass of the bath plug, divided by the water mass in the pipe. A number of corrections apply. First of all, in the above no mention was made of the fact that above the bath plug, in case of vibration of the bath plug, a certain quantity of water needs to be stirred into movement. Therefore it is desirable to consider m as m+mw. Also, it has been ignored that the periodic variation of the gap discharge experiences a certain inertia, both outside and inside the pipe. The effective pipe length therefore is somewhat greater than the real length. This results in insecurities concerning the outcome of the calculation. Also, it needs to be understood that if the pipe has a local cross-section that, as an example, is greater than the surface of the bath plug, the inertia of the flowing water decreases and does not increase. This also means that the mass of water per unit of length of the pipe that needs to be included in the calculation for the pipe section concerned, becomes smaller relative to a pipe with a prismatic diameter.
90

Delft Hydraulics

Moreover, there is an extra pipe length that needs to be included in the calculation, in case it is a pipe that discharges below the water level. In case the discharge varies with time, the out flowing water is influenced by a certain flow inertia. This will be discussed in Paragraph 3.5. The second auxiliary magnitude is:
Ck = k k w sudden = k 0 k 0 = 2 g H 0 Ap 2 Fperm.

(A4.24)

Ck is defined as the stiffness of the suspension, divided by the sudden added negative stiffness as defined in Paragraph 3.3.3 (i.e. with a reversal of the sign, the flow stiffness, assuming that the discharge through the gap remains constant). As the bath plug is situated upstream of the pipe, the sudden stiffness is negative, but the Ck is introduced as a positive number. Ck may be easily calculated and equals half the spring force necessary to push the bath plug onto the seating, divided by the stationary force that the bath plug experiences due to the hydraulic head. In case of a twice higher stiffness of the spring, the same Ck value is generated with a twice smaller gap size. It appears that Equation A4.21 results in: Ck = 1 + C m The precondition for a stable situation is: Ck > 1 + C m (A4.26) (A4.25)

The explanation of A4.21 included that both the mass m and the pipe length L needed a correction. It appears from Equation A4.26 that in analyzing the vibration sensitivity it would be safe to give a high estimate of Cm. This is only the case when an extra mass is included in the calculation. Paragraph 3.2.6 presents an estimate of the added water mass of the bath plug. Using test results of scale model investigations concerning the dynamic behaviour of gates, it was more or less possible to verify Equation A4.26. Verification I: this was an investigation of culvert gates in the Kreekrak shipping lock for the situation in which the gate was still closed, but due to the relief installation it was already detached from the seating. The situation is as shown in Figure A4.10 (see also Paragraph 6.3, example a). In the investigation it was determined what magnitude the hydraulic head was in which vibrations occurred, given a certain gap width and stiffness of the relief installation.

91

Delft Hydraulics

Figure A4.10: Cross-section of a culvert gate in the Kreekrak shipping lock (see example a in Paragraph 6.3). Verification II: this was a sector gate in a high-head shipping lock, as shown in Figure A4.11 (see also Paragraph 6.3, example c). Here, during the vibration, the gate does not move in the direction of the culvert, but perpendicular to it. Nevertheless, it may be demonstrated (Kolkman, 1976) that, due to the protruding lip at the top of the gate, a similar situation may be generated as in case of the bath plug. There too, when the gate vibrates a volume of water is displaced. The pressure required to accelerate and decelerate the water in the culvert, again influences the lip in the direction in which the gate is vibrating. As the lip is small relative to the cross-section of the culvert, to determine Cm it is not the volume of water of the culvert, but a strongly reduced water volume that needs to be taken into account. This investigation also looked into the question: at what hydraulic head did vibrations start to occur? As soon as vibrations were generated, the amplitude became so great that the gate touched the bottom; the forces in the lifting system doubled relative to the stationary situation! Again and again, these were vibrations generated at small openings.

92

Delft Hydraulics

Figure A4.11: Sector gate (reversed Tainter gate) in case of a high-head shipping lock. A summary of the results of the verifications is given in the table below. The experimental Ck values relate to the boundary situation in which the vibration initially started. Verification I: Verification II: Cm = 0.3 Cm = 26.5 Ck in theory 1.3 27.5 Ck observed 1 < Ck < 2 29

The results in theoretical calculations closely match those observed in reality.

Figure A4.12: Cross-section of a pressure relief valve with brake (a) and test results on vibration sensitivity (b) (Weaver, 1980).

93

Delft Hydraulics

At a later stage Weaver (1980) carried out yet another verification on a pressure relief valve with brake. The situation is presented in Figure A4.12. The Cm factor here is approximately zero, as it is a valve in a long culvert. It may be observed, that Equation A4.26 is rather safe. Weaver has provided some suggestions for refinements of the calculation, as a result of which the observations match the calculations more closely, but this requires calculating in the time domain. The upper limit for the damping required to compensate the negative flow damping, may be analytically determined. For it appears that this is maximum in the low-frequency area and, for the maximum case ( = 0), an analytical solution for the damping may be found. This is a quasi-stationary flow, for which it may be calculated how great the acceleration and deceleration of the flow in the pipe is (during vibration). In case of quasi-stationary flow, the Gap Equation A4.6 for the bath plug is:
Q = Lc 2 g H

(A4.27)

From this (in case of a very slow variation of the gap width), it follows that: dQ dy = Lc 2 g H dt dt (A4.28)

with Lc = circumference of the gap. For the coefficient of the gap width the following applies: d/dt = dy/dt. As the pressure that is exerted on the bath plug downstream is coupled to the acceleration of the flow in the pipe: dV p = L (A4.29) dt and (the force is defined as opposite to the flow direction):
F = pAp

(A4.30)

for the negative water damping cw = F (dy / dt ) , this results in:


cw = L Lc 2 g H

(A4.31)

4.4.3 Theory of varying gap width in case of vertical gate vibrations


Here, a similar mechanism as in the case of bath plug vibrations is found, but it may also be used to explain vertical vibrations of a gate with flow-through discharge underneath, as in the situation of Figure A4.13. First, it is assumed that the shape of the gate is such, that due to the flow suction forces occur that are proportional to the local hydraulic head beneath the gate; in case of the rectangular gate this is applies to a small lifting height, because in that case the flow separates and attaches again downstream. As a consequence, an area with underpressure (relative to the downstream water level) is generated, that is proportional to the 94

Delft Hydraulics

hydraulic head. At a later stage it will be shown, that self-excitation is also possible without suction force. The situation is schematically represented in Figure A4.13.

Figure A4.13: Diagram and symbols for the vibrating gate with flow-through discharge underneath. As indicated in Figure A4.13 with dotted lines, in case of flowing water with a free water surface there is also flow inertia. If the discharge that flows beneath the gate changes with time, a fictive tube may be imagined upstream with a length Lu and a length Ld downstream, see dotted lines in Figure A4.13. The values of Lu and Ld may be written as CLu and CLd, and both coefficients may be read from Figure A3.1 (This Figure holds for both Clu and Cld as well). The assumption that, while calculating the CL values in Figure A3.1 the wave radiation is ignored, has already been discussed. The total tube length now equals:
L = Lu + Ld + b = b + (CLu + CLd )

(A4.32)

This again is represented in a (flow) inertia term Ci:


Ci = CLu + CLd + b

(A4.33)

If the gate shows a harmonic vibration y = Yeit and the discharge coefficient remains constant, then the variation of the discharge may be calculated with time. In case of low frequencies, the discharge will simply follow the gap variation; the inertia of the flow plays no role then. This means that dq/dt (q is the discharge per unit of width) varies with dy/dt. Because: q = m 2 g H (A4.34) in case of a quasi-stationary situation, this means that: dq dy = m 2 g H dt dt (A4.35)

95

Delft Hydraulics

with

m = discharge coefficient, H = hydraulic head and d/dt = dy/dt.

Assuming that the water level difference at the gate (the global hydraulic head) does not vary, the force to accelerate or decelerate the water in the tube must originate from the gate. The local hydraulic head of the gate therefore is greater if, as an example, the discharge is slowed down; the pressure increases upstream and decreases downstream:
H local = H global L pipe dV L pipe dq = H global g dt g dt

(A4.36)

If dy/dt is positive, then the discharge increases as well, and therefore the hydraulic head decreases. The suction force also decreases in that case, and may be considered as a dynamic suction force that is directed upward. This force may be written as:
F = Cs gbH

(A4.37)

with

Cs = suction force coefficient.

H' is defined as the dynamic part of the hydraulic head: H' = -(Ltube/g)dq/dt. This dynamic hydraulic head is negative in case of a positive value of dq/dt (and therefore of dy/dt). As the negative water damping equals this force again, divided by the vibration velocity dy/dt, after some elaboration using Ci = Ltube/ (Equation A4.33), for the negative flow damping it is found that:
cw = mCi bCs 2 g H

(A4.38)

This reasoning is similar to that of determining the maximum negative damping of the bath plug. That too was found at very low frequencies. The higher the frequency, the more the discharge tends to remain constant. Then also the extent of self-excitation, in case of vertically vibrating gates, decreases. The addition that the suction force variation coupled to the vibration is not only generated due to the hydraulic head variation, but also due to the pressure variation itself that is present as a consequence of the flow inertia downstream of the gate, is important. The latter is caused by the fictive tube downstream with length CLd. This may be represented as a proportional part of the total flow inertia. In case of a gate with a lip protruding at the downstream side, the pressure variation partially operates at the bottom and partially at the upper side of the lip (though less). In that case the reduced influence of the downstream inertia (due to the influence at the top of the lip) is indicated by a reduction factor . All this results in the fact, that a term Ld/(Lu+b+Ld) needs to be added to the local suction coefficient Cs. The total suction force coefficient now becomes:
Cs = Cs + CLd /(CLu + CLd + b / )

(A4.39)

The factor has been introduced by Kolkman (1976). For the situation of Figure A4.13, the equals 1 as the lip is absent there. When the lip is present, 0 < < 1. 96

Delft Hydraulics

From Equation A4.39, in combination with A4.38, it appears to be clear that also a gate edge that normally does not experience suction force, may be sensitive to vibration. The complete elaboration of the theory for the entire frequency area may be found in Kolkman/Vrijer (1977). From this, Figure A4.14 has been taken; in this also experimental values for the extent of amplification may be found. The self-excitation coefficient (the definition of this is included in the Figure) has been plotted out as a function of a kind of Strouhal number, based on the gap width. If this Strouhal number had been based on the length of the fictive tube (calculated with the length coefficients of A4.33), then the theoretic lines would have converged on one line. That has not been elaborated further here.

Figure A4.14: Calculated extent of self-excitation, compared with test results, from Kolkman and Vrijer (1977). The conclusion of the experimental verification is threefold. In the first place, the order or magnitude of the maximum negative flow damping is reasonably well represented. Secondly, it is found that the upper limit of the frequency area in which the vibrations may occur is situated too far to the right in relation to the theory. For the theoretic deduction of the tube length, it might have been better to replace the lifting height by m (m is the discharge coefficient, but in this case also something like the flow contraction). That would almost neutralize the horizontal shift. And thirdly, it appears that, below a certain Strouhal number, in reality there is no self-excitation whatsoever; no exact explanation for this has been found. It may be that the assumptions that wave radiation plays no role and that the potential flow may be superimposed by the normal flow pattern, in this case no longer apply. The values of the Strouhal number at which self-excitation occurs, are rather low: suppose S = 0,05, h = 10 m, = 0,2 m, then the resonance frequency that is most sensitive is fn = 3.5 Hz. 97

Delft Hydraulics

It is important that the separation is situated between downstream and upstream. If the point of separation of the flow is upstream, as in Figure A4.13, then the pressure decrease in the flow directly downstream of the gate is immediately noticeable beneath the gate. This resulted in = 1 in A4.39. If the point of separation is downstream, then the pressure beneath the gate is coupled to the upstream dynamic pressure. As this operates in the opposite direction to the downstream pressure, this results in a contribution to the positive water damping.

Figure A4.15: Comparison of various shapes of bottom edges related to vibration sensitivity (Vrijer, 1977 and Delft Hydraulics Report M1322). Vrijer (1977) has investigated various shapes of bottom edges regarding vibration sensitivity, based on both theory and testing, see Figure A4.15. Shapes A and B generate selfexcitation (B somewhat less, as the factor of Equation A4.39 is smaller), while shapes C and D are stable and result in a positive flow damping. Shape D is the strongest positive damping, because here too the pressure that is exerted on the top of the protruding lip downstream still is an additional influence. For the sake of completeness, a shape of the edge is given here that is most sensitive to vibrations: Figure A4.16. In this case, the suction force coefficient is maximum and also the flow inertia downstream fully cooperates. In modern gate structures, this edge is no longer used; no test results are available.

98

Delft Hydraulics

Figure A4.16: A very unfavourable shape of the bottom edge regarding vibration sensitivity.

4.4.4 Horizontal gate and plate vibrations; theory of the varying discharge coefficient
In case of the self-excitation as found in case of the bath plug vibration, the discharge is periodically pinched off, because the gap width periodically varies. In case of horizontal vibrations of the plating, induced by excitation at the bottom edge of the gate, such as observed at the new gate of the storm surge barrier at Krimpen (Paragraph 6.2, example b), this however is not the case. For this vibration mechanism another theory was devised, i.e. the theory of the fluctuating discharge coefficient (Kolkman, 1980). Figure A4.17 shows the flow pattern in case of the non-vibrating situation at the hemispherical bottom edge. The flow relative to the vibrating bottom edge in still water is on the right. This last type of flow is potential flow, as may also be found at the vibrating strip in Figure A3.3. Provided the vibration amplitude is small and the Strouhal number is not too small, the second flow may be superimposed on the first, as the relative pressure gradients and flows in that case only cause a kind of ripples on the streamlines of the permanent flow. Now, suppose that at a certain moment of the vibration the gate edge moves to the right (image b of Figure A4.17), then water around the edge is forced to the left and it is plausible that in that case also the streamlines of the permanent flow will separate at a certain point that is more to the left than in case of the permanent situation. The contraction of the main flow becomes stronger (the dotted lines in image a), and the flow-through capacity of the main flow decreases, because of which also the discharge will decrease. Due to the flow inertia, an extra hydraulic head builds up at the gap. In this case, the hydraulic head is in-phase with the vibration velocity, i.e. the precondition for negative flow damping.

99

Delft Hydraulics

Figure A4.17: Elements of simultaneously occurring gap flow and added water mass flow (Kolkman, 1980). Figure A4.17 relates to an edge at which the point of separation in itself is already hardly stable. This however is not a necessary precondition for this type of vibration. Experiences with the storm surge barrier at Krimpen have, at a later stage, also occurred by coincidence during a scale model investigation, in which a plate of 20 mm was fastened between the walls of a 1 m wide flume (Paragraph 6.5, example d). At a later stage it appeared that yet other cases of vibration that occurred, must be ascribed to this vibration mode. Kolkman (1980) initially concluded that the area of critical Strouhal numbers (based on the edge thickness) is between 0.2 and 1, and when based on the gap width, between 0.5 and 2. The frequencies at which these vibrations may occur are much higher than in case of the vibrations in which the gap width varies with the vibration. It is therefore harder to prevent these by making the structure stiffer. The weir at Krimpen was purposefully provided with a sharp gate edge (thin plate), which in case the gate was closed would still remain detached from the bottom, because the gate may deflect almost horizontally with the tide, without wear and tear of the rubber edge. It was expected, that a relatively sharp edge would function vibration-free. As it was observed that the vibrations that occurred were very strong and were situated at a depth at which it would be very difficult to conduct verification measurements, extensive scale model investigations were carried out. Jongeling (1987 and 1988) has measured flow velocities beneath the gate (using laserDoppler anemometry) and local pressures downstream of the vibrating plate, simultaneous with the vibration registrations. In Figure A4.18 the results of the vibration measurements that were obtained for the rectangular edge are summarized. It was observed that in one particular situation vibrations were generated in multiple frequencies. This is analogous to what was found in case of an instable free boundary layer (Paragraph 4.3 and Figure A4.4).

100

Delft Hydraulics

Figure A4.18: Maximum horizontal vibration amplitudes for the first, second and third harmonic as a function of the gap width (Jongeling, 1987). Vr = S-1. The critical area of Strouhal numbers based on the edge width:
S = Vr1 = fb 2 g H

(A4.40)

is between 0.1 and 1.4. In addition, a number of other edges have been investigated, see Figure A4.19, to assess whether these vibrations may also occur in case of edges at which the flow attaches completely or separates completely at the upstream edge, because the edge downstream is bevelled. As might be expected, edge 2 behaved in a similar way to edge 1 (Figure A4.19). Edge 3 was completely stable (i.e. for the investigated area of openings of up to 2.5 times the edge thickness). Edge 4 was stable up to /b = 0.75 and somewhat instable up to the maximum investigated opening of /b = 1.75. Edge 5 was completely stable.

101

Delft Hydraulics

Figure A4.19: Edge shapes that were investigated (Jongeling, 1988).

4.4.5 Galloping
The vibrations that are generated by forces resulting from the periodically varying angle of approach flow due to the vibration, are referred to as galloping. This expression was first used by Den Hartog (1956) for galloping movements of telephone wires, the crosssection of which had changed to a kind of lateral wing profile, due to being ice-covered. The property of such a profile is that, when by an instantaneous upward movement, the apparent angle of approach flow of a normal horizontal flow is crosswise from above, then the flow at the upper side attaches better and generates an upward suction force there. In case of a downward movement, the flow appears to come transversely from below, and the flow of the (rounded) upper side separates earlier. Various profiles are sensitive to this, for more details see Paragraph 5.4.1. This could also happen at gates with a hemispherical or a rectangular edge, see Thang and Naudascher (1986). These authors base themselves on an investigation with a somewhat artificial schematization, in which the oblique approach flow is simulated by bevelling off the rectangular underside of a gate. In this, it is indeed observed, that in case of an altered angle, the flow tends to attach or separate more, in such a way that the gate with a rectangular crosssection meets the precondition for galloping. Moreover, at a later stage it was verified again by Kanne et al (1991) that this phenomenon, next to the mechanism of the varying gap width as discussed in Paragraph 4.4.3, is an auxiliary cause of vibrations. In case of a rectangular shape of the gate, in which the flow separates at the upstream edge, in case of a downward movement of the gate, the zone between the free boundary layer and the underside of the gate is reduced, which makes it more difficult for the concentrated water flow to suck in water, and an underpressure is generated. Moreover, it may also occur that this suction is only caused in case of somewhat greater vibration amplitudes, as a result of which a gate vibration of a small amplitude damps out after a small initial vibration; and in

102

Delft Hydraulics

case of a vibration with a greater amplitude, it is maintained. This is an aspect that deserves close attention when carrying out scale model investigations.

4.4.6 Application of an instability indicator


In various examples that have been discussed above, it may be shown that for those situations in which the sudden flow stiffness (or the flow stiffness that is generated at high frequencies) is negative, the flow damping is also negative. This flow damping is generated specifically at low vibration frequencies. This was found both for the bath plug and for the vertical vibrations. Figure A4.7 illustrated how, in case of the bath plug vibration, the negative flow damping that occurs at low frequencies, changes into a negative stiffness at high frequencies. The term low frequency means that the discharge follows the variation at the gate opening as if there was a stationary situation again and again. This negative flow damping is represented in A4.31. The concept high frequency means that the discharge remains almost constant as a consequence of the very great flow inertia. Naturally, in case of a constant discharge, the hydraulic head of the gate is directly coupled to the gate opening. The gate load relative to the hydraulic head therefore is in-phase with the vibration. This eventually results in a negative flow stiffness. This stiffness has been quantified in Equation A4.24. The turnover frequency at which the flow damping changes into a flow stiffness, in case of a vertically vibrating gate with flow-through discharge underneath, appears to depend on a kind of Strouhal number, but this time it is based on the (fictive) pipe length. Also the discharge coefficient plays a role. The deduction of the turnover frequency is shown in Appendix I for both a culvert gate and a gate in a situation with a free water surface. The turnover frequency is defined as the frequency at which both the negative flow stiffness and the negative flow damping are bisected relative to their maximum values (respectively at = and = 0, in which = angular frequency of the vibration). The following equation was found for the turnover frequency:
S= f 1 = 2 g H 2 mCi

(A4.41)

H is the hydraulic head of the gate. In Figure A4.13 and in Equations A4.32, A4.33 and A4.34 the remaining magnitudes have been defined. Equation A4.41 matches the results of Figure A4.14. In Appendix I, the turnover frequency for vertically vibrating gates in a closed culvert has been deduced as well. Because of the relation between the negative sudden flow stiffness and the negative flow damping, the instability indicator formulated by Kolkman (1976) is defined as:
Assuming that the discharge remains constant, if a gate movement is such, that it reduces a leakage gap and the extra forces that are generated in that case are such that these amplify the gate movement, there is a plausible probability for self-exciting vibrations to occur.

103

Delft Hydraulics

In addition, suction forces (forces that tend to reduce the gap) generally are a first indication of self-exciting vibrations. A shape such as in Figure A4.16, as an example, is extremely unfavourable. In case the lip would be deflected in downstream direction, a much better situation would already be achieved; it would be much better if the stiffness was situated higher.
Remark: Various sources provide criteria concerning the question whether certain types of gate vibrations occur or not. If these relate to the required stiffness or the minimally required mechanical damping, then one should be prepared for the fact that often the change to use dimensionless magnitudes, such as a dimensionless damping of only the dry gate, is made too quickly. A complete presentation should always be based on a resonance frequency in water. Next to this, the damping, which in any event should compensate the flow damping, should have the dimension of a stationary flow force, divided by the water velocity. Such a dimension may also be found in the expression for flow damping (Equation A4.31) or the negative flow damping (Equation A4.38), only now V is replaced by 2gH.

4.4.7 Theoretical aspects of vibrations of a gate with overflow (nappe)


There are various mechanisms as a result of which the overflowing nappe, which closes off the air beneath it like a curtain, may become instable. In some cases the dynamic properties of the flap gate (bottom-hinged gate) play an important role, in other cases the flap gate is only loaded dynamically due to the pressure variations of the enclosed air. In the following considerations, the point of departure is a situation as schematically presented in Figure A4.20.

Figure A4.20: Diagram of elements that are of importance in case of a vibrating bottom-hinged gate with overflow. 104

Delft Hydraulics

The following three phenomena or mechanisms may occur. A. The enclosed volume of air periodically varies, which generates pressure fluctuations. In this, it is assumed that the pressure is the same throughout the air cushion. The pressure fluctuation excites the nappe, and travelling waves are generated in the nappe. The excitation at the top has the greatest effect on the volume variation, which in turn generates the pressure fluctuations again. The volume variations of the air cushion are caused in particular by the strongly alternating flow at the lower side of the nappe. Here the amplitude is greatest. The undulations also have a greater wave length; this again relates to the greater water velocity at the location of the nappe. All these aspects are illustrated by the wave pattern in Figure A4.21. These are the results of a calculation of the behaviour of a vertical water curtain, influenced by a one-sided air pressure variation. Partensky and Sar Khloeung (1971) have found that, in case of a non-vibrating structure, the critical overflow height above which no vibrations of the water curtain occur, is 5 % of the drop height. In case it is a gate capable of vibrating, then this may increase up to 15%. Another mechanism for the generation of vibrations is the turbulent boundary layer, B. that is generated in the air due to the falling nappe. The air pressure variations in turn cause strong alternation of the flow of the nappe. Moreover, the volume variations in the air excite the gate and this also causes extra air pressure variations. This phenomenon has been investigated by Binnie (1972). Suppose, the bottom-hinged gate vibrates after an initial disturbance. The gate C. vibration then also generates a strong alternation of the nappe; this again causes a volume variation of the air, which in turn results in pressure fluctuations, which again causes the vibration of the gate. Naturally, the nappe itself is also excited again. Although these phenomena are often described separately in various sources, they often occur simultaneously. Phenomena A and B may however also occur in case of a nonvibrating overflow edge. Various sources also discuss investigations that review resonances in the air cushion itself. This may play a role in case of very long flap gates. A measure for the vibration frequency at which the self-excitation occurs, is Tf, as a result of which Tf is the falling time of a water particle in the nappe. The parameter Tv determines how many wave lengths fit in the nappe. The volume variation of the air beneath the nappe is maximum if the number of wave lengths is (n+) or (n-), in which n+ results in a volume variation and air pressure variation, which itself generates self-excitation (in this, n is a whole number). This results in the Relation A4.4 as described in Paragraph 4.3: 1 fT f = (n + ) 4 (with n=1, 2 .....)

f = vibration frequency, Tf = falling time of a water particle. The phenomenon has been both theoretically and experimentally investigated extensively by Schwarz (1964), Kolkman (1972) and Treiber (1972). Kolkman carried out theoretical and experimental investigations of a vertical falling water curtain (see Figure A4.21), both in case of a stiff and in case of an elastic limitation of the air cushion. Treiber carried out investigations of an inclined water curtain. 105

Delft Hydraulics

Figure A4.21: Shape of a vertical water curtain with one-sided periodic variation of air pressure (Kolkman, 1972). The most important remedy for all these types of self-excitation is the same for all mechanisms described above: i.e. making a connection between the enclosed air cushion and the outer air, so that the air pressure variations are equalized. Because of the possibility of the occurrence of phenomenon B, which is very local, the nappe splitters that are used for this 106

Delft Hydraulics

should not be placed too far apart. A distance of about a quarter of the drop height seems adequate. Aeration from the side usually is not sufficient. Phenomenon B, which is particularly known in case of thin nappes, may also be combated with an irregular top. Even if the nappe is still not interrupted because of that, the difference in falling time of the various parts of the nappe may suffice to prevent periodic volume variations of the air cushion. Figure A4.21 shows that a staggered top is used.

Figure A4.22: Situation with only water under the nappe (Kolkman, 1980). Kolkman (1980) has published about this, following Delft Hydraulics investigation M1300; at a later stage Van Rhee (1984) has also written a thesis about this. In case of this vibration, the coupling is not caused due to an air cushion, but due to a water cushion. This coupling is much stronger than in the case of air; consequently, very strong vibrations have been observed in scale model investigations. In this case the aeration from the side is sufficient, as the aim is not a connection that should leak away the pressure fluctuations in the air cushion, but to introduce air, in order to prevent the presence of a water volume underneath the entire nappe.

4.5 Self-excitation in case of fluid oscillations (among which standing waves) 4.5.1 Instable fluid oscillations due to coupled discharge changes at gates
Instable fluid oscillation due to self-excitation as a consequence of the elastic properties of a gate or valve is a phenomenon that was (re)discovered and analyzed following the investigations at the storm surge barrier Rotterdam Waterway. These concerned periodic gate movements that are coupled to water level oscillations or standing wave movements, in

107

Delft Hydraulics

which also the frequency of the oscillating gate equals the resonance frequency of the fluid. The elastic properties of floating gates is the immersion stiffness.

Figure A4.23: Plan view and vertical cross-section of the storm surge barrier Rotterdam Waterway (Bakker et al, 1991). During the closing of the sector gates, which are afloat at that moment in time, in the flow at high tide there appeared to be very strong vertical oscillations of the gates, specifically in the situation in which the gates had already been brought close to each other and it had already been possible for a hydraulic head to be formed. In case of a hydraulic head of the order of magnitude of 3 m, fluctuations of the water level occurred between the abutments, crosswise to the flow, near the gates at the riverside. In case of a hydraulic head of approximately 3 m, the vertical (crest-trough) amplitude could rise up to 5 m. As the gates are afloat, these were also moving up and down, with almost the same amplitude. If one gate moved up, the other moved down. Also in case of flow in the other direction, a similar phenomenon occurred, but in that case it was the second harmonic of the standing wave; see Bakker et al (1991) and Example f in Paragraph 6.2. In this publication it has been made plausible that the cause of the standing waves is in fact that the gates are afloat, as a result of which, in case of a rising water level at the gate downstream, the gate was lifted again and again an extra discharge flowed through beneath the gate. It may be demonstrated that because of this, any initial small wave that might reflect from the gate due to the extra generated discharge, returns in an amplified way. In case of a standing wave, in which essentially large numbers of small pulse waves move back and forth, this results in a continuous resonant rise. Bakker et al (1991) have shown that the system of communicating vessels experiences resonant rise as well, when the vessels are fed by a discharge that is positively correlated with the instantaneous water level. In Paragraph 1.3, at the introduction of the concept of instable fluid resonance with self-excitation (see under type 5), two preconditions have been given that need to be met, if instable fluid resonances are to occur.

108

Delft Hydraulics

Precondition 1 The situation needs to be such, that a low-damped resonance of a fluid oscillator or of a standing wave is possible. In Chapter 2, Paragraph 2.3.1, an overview of this type of situation is given. Precondition 2 In case of increased pressure, the floating gate generates an extra discharge (superimposed on the normal discharge) in the direction of the basin or the pipe in which the standing wave may be caused. This matches the precondition that was deduced in Chapter 2 for an instable oscillation in communicating vessels (Paragraph 2.3.2, Equations A2.58 and A2.60). In reality this means that, when the basin is situated upstream of a gate or gate, the gate discharge decreases in case of a pressure increase upstream. If the basin is situated downstream of the gate, instability occurs when the gate discharge increases in case of a pressure increase downstream of the gate. In Paragraph 5.6 it will be further discussed that this does not only occur at gates or doors, but also at valves, pumps or turbines.

The instability of the lower reach is expressed in the equation:


dq >0 dh2 For instability of the upper reach this is:
dq <0 dh1 with q = gate discharge per unit of width. In certain cases A4.42 may be analytically elaborated even further. Point of departure in case of floating gates is the discharge relation: q = a 2 g (h1 h2 ) with = discharge coefficient. Suppose, in case of a situation with a hydraulic head, that the upstream water level remains constant. In case of a small variation of the gap beneath the gate, a, and of the downstream water level, h2, the following applies: dq = 2 g (h1 h2 ) da with is assumed to be constant. (A4.43) (A4.42b)

(A4.42a)

ag
2 g (h1 h2 )

dh2

(A4.44)

109

Delft Hydraulics

If it is known that the gate completely floats on the downstream water level, it means that the change in the gate opening equals the change of the downstream water level, therefore: da = dh2 (A4.45)

Now dq/dh2 may be directly calculated using Equation A4.44. The boundary between stability and instability follows from the precondition dq/dh2 = 0. After some elaboration, the following is found for the borderline case: h1 h2 = 1 a 2 (A4.46)

Also, in case the gate only partially floats on the downstream water, the borderline case may be analytically determined. Investigations of the storm surge barrier Rotterdam Waterway have shown that such a calculation closely resembles the outcome of scale model investigations (Delft Hydraulics Report Q1271).

Figure A4.24: Situations with possibilities for low-damped fluid oscillations or a standing wave. Figure A4.24 shows a number of other situations, in which a low-damped fluid oscillation or a low-damped standing wave may occur. The standing wave, moreover, may also be a pressure wave in a closed pipe or a standing wave in an open pipe. The first example in Figure A4.24 is a basin (a gate shaft or a surge shaft) in a pressure pipe. Here, the water 110

Delft Hydraulics

reacts as incompressible, and the resonance frequency is determined by the storage surface of the basin combined with the inertia of the feeder pipe. The second example is a closed pipe, in which a standing pressure wave may be generated. The vibration time of the resonance vibration here is four times the pipe length, divided by the celerity of the pressure wave. The third example is a situation with several gates side by side, separated by intermediate piers. Standing waves may be generated that travel from one gate to one of the adjacent gates. Naturally, standing waves may also be generated in the situation of Figure A4.23. The type of gate that may generate this kind of self-exciting fluid oscillations, due to its discharge-pressure relation, is: The floating gate, as shown in Figure A4.23; The bath plug (but this time as part of a very long pipe, in which pressure waves are generated) or a variation of this, such as a closed gate with an improperly used music note rubber seal (see also Chapter 6, Examples 6.4a and 6.4c); The pressure relief valve with brake (Figure A4.12).

For all these gates and valves, an increasing hydraulic head will generate less discharge and a decreasing hydraulic head will generate more discharge. The boundary between stable and instable vibrations exists in those situations when the hydraulic head increases, while the discharge remains constant. This appears to match exactly the precondition that the dimensionless stiffness factor Ck that was introduced in Paragraph 4.4.2 (compare A4.26) equals 1; the precondition for stability therefore is: Ck > 1 (A4.47)

This precondition therefore is less severe than the one that applies to the bath plug in case of a non-compressible fluid, for which it was found in Equation A4.31 that Ck > Cm + 1 (considering that the mass coefficient Cm is always positive). Precondition A4.47 has also been formulated already by Streeter and Wylie (1967) for valves in long pipes. Considering how these aspects have been presented above, distinguishing the preconditions required for generating instabilities downstream or upstream, the phenomenon of instable fluid oscillation has been discussed in a general way.

4.5.2 Self-exciting fluid oscillations due to excitation by waves in case of floating flap gates (bottom-hinged gates)
During an investigation of floating flap gates that are hinged to the bottom, scale model investigations have been carried out concerning the response to waves. In case the flap gate is not performing its retaining function, it is situated in a niche on the bottom, and it is filled with water. If it does however need to function as a retainer, the water in the flap gate is replaced by air; the flap gate derives its retaining capacity from buoyancy. This way of retaining means that a (limited) hydraulic head may be observed; waves however may pass through almost unhindered. Figure A4.25 shows a cross-section of the flap gate in retaining position. 111

Delft Hydraulics

Figure A4.25: Cross-section of a floating flap gate, hinged to the bottom. Investigations into regular waves have shown, that in case more than one flap gate was placed in a laboratory flume, next to regular fluctuations in the frequency of the incoming wave, also out-of-phase movements of the flap gates could be generated in the double period relative to the wave period. The period of the low-frequency movement was found again in still water as the natural vibration period relative to the same out-of-phase movement. This period depends on the width of the flap gate (which again determines both the mass inertia moment and the stiffness due to buoyancy), and of the water depth. But in case of a system consisting of more than two flap gates, multiple movements are possible as well. It is therefore a system that has multiple natural vibrations. Either the wave length of the flap gate movement is twice the length of the flap gate (the flap gates move alternatingly), or the wave length of the flap gate movement is three or more times the length of the flap gate. For each movement there is another period. Figure A4.26 shows some out-of-phase movements of the flap gates, as found in a four-flap scale model investigation (in which flap gates of half the initial width have been mounted on the wall). In case of still water, it appeared that at the correct period of movement, particularly little energy was required to maintain these flap gate movements. Due to the alternating movement, there is no net resulting discharge flowing in the flume, and at some distance from the barrier there is also no observable wave radiation in the flume. The resonance time does not fully match the wave length of the (double-period) flap gate movement. In case of a situation with two flap gates in the flume it was expected that the period would correspond to twice the flume width, divided by the celerity relative to that wave length. A longer period was found however; obviously the standing wave does not travel according to a crosswise straight line, but takes a somewhat longer route. For all harmonic components the still water natural period may be determined quite accurately. If next the wave board is put into operation with a period equal to half of the one mentioned above, then after some time the corresponding flap gate movement will be generated again and again.

112

Delft Hydraulics

Figure A4.26: The out-of-phase movements of the bottom-hinged flap gates, seen from above. Mode 1 is the ordinary movement induced by waves. By registering the difference signal of two flap gate movements during measurements, a signal is generated, at which the movement of the flap gates in the wave frequency is not noticeable, but only the low-frequency movement is registered. This movement is caused at a random moment in time and appears to grow exponentially with time, after which, some dozens of periods later, an equilibrium amplitude is reached. The subharmonic flap gate movement in case of purely periodic wave excitation is so strong, that a considerable leakage gap is caused between the flap gates. This phenomenon is similar to the so-called transverse waves that are generated when a long wave board is put into operation. After a while standing transverse waves may be generated with amplitudes greater than the normally generated waves. The wave length of such a standing transverse wave corresponds to the double period of the wave board movement. To obtain more understanding of this phenomenon, it has been investigated with a calculation model which wave board movement may cause resonance of the vessel system in a situation with two communicating vessels. The communicating vessels are situated in such a way relative to the wave board, that relative to the wave board there is a transverse oscillation as well. The calculation model is discussed in Part C, Paragraph 3.5.3 (Example c). Resonance also appears to occur in calculations in case of a wave board movement with the half period relative to the vessel oscillations. Resonant rise only occurs when there already is a small oscillation in the vessel system. During the exponential growth of the amplitude of the oscillation in the vessels, the wave board appears to have its maximum forward velocity (which means in the direction in which the water surface in the vessel decreases), when the water level difference in the vessels is maximum (as shown in Figure A4.27) as well. The energy transfer from the flap gates that are moving back and forth with the wave movement to a standing subharmonic transverse wave or to an up and down movement of the water level in the vessels, may also be calculated in an energy review. The growth of the amplitude in the calculation model corresponds to the analytically determined energy transfer. For more details see Paragraph 6.2, example h, and Jongeling and Kolkman (1995).

113

Delft Hydraulics

Figure A4.27: Diagram of communicating vessels with synchronous wave boards. The energy transfer is calculated as follows: Suppose, the water depth in one of the vessels during the oscillation (in the resonance frequency of the vessels) may be described as (see Figure A4.27): h = h0 + z = h0 + Z cos(t ) (A4.48)

In case of an assumed hydrostatic pressure distribution, the force that the wave board now exerts per unit of width, is:
2 F=1 2 gh

(A4.49)

Now, the wave board velocity is introduced with the double frequency, and with the phase condition that this velocity is maximum, if the deflection of the still water surface, z, is also maximum:

114

Delft Hydraulics

Vwave board = V0 cos(2t ) In principle, the energy transfer of one period may now be calculated as follows:
T

(A4.50)

E=

t =0

wave board

Fdt

with T = 2 /

(A4.51)

Elaboration of this equation shows, that the term that in total generates a net energy transfer in one period, is the component of the force that is proportional to z2, or proportional to Z2cos2(t). Therefore:
T

E=

t =0

2 gZ

1 cos 2 (t )V0 cos(2t )dt = Z 2VT 8

with T = 2 /

(A4.52)

All other terms eventually do not contribute to the energy transfer. As there are two basins, the total energy transfer needs to be determined from the total width of the wave boards of both basins together. By relating E to the potential or the kinetic energy of the oscillations in the vessels, the extent of resonant rise may be calculated. The interesting aspect of this review is, that exactly the part of hydrostatic force that is proportional to z2, which in case of a normal linear wave review is ignored, here is responsible for the generation of the subharmonic transverse waves and the related parasitic flap gate movement. The above theoretical review does offer an explanation for the occurrence of transverse waves, but does not clarify the wave amplitudes that may be expected. The excitation strength, and with that the exponent of the resonant rise, is proportional to the wave height of the incoming wave. If there is a certain (linear) damping present in the system, then the instable oscillation will only occur at a certain wave height. The equilibrium amplitude of the subharmonic wave is only reached when the damping factors are no longer linear, but have a strength that increases with the amplitude of the fluid oscillations. Jongeling and Kolkman (1995) have further elaborated the theory mentioned above.

4.6 Review of small waves generated by the vibrating gates in the upstream water
If a gate vibrates with a horizontal component, then small waves are generated in the relatively smooth upper water and radiated out. The gate is a kind of wave board and it may be shown that the wave amplitude is twice the vibration amplitude (Kolkman, 1976), as this relates to deep water waves. The radiated waves, with their front parallel to the gate plating, may possibly also be generated due to fact that in case of the vertical vibration component of gates with flow-through, the discharge beneath the gate periodically varies. Experience shows, that after some time transverse waves are generated, perpendicular to the flow direction (see Figure A4.28, the second picture). After some time (the third picture) the waves that run off at right angles (the first picture) can hardly be noticed. The vibration direction of the plating on the pictures is horizontal (from left to right). At some distance from the gate the transverse waves start to move somewhat oblique relative to the vibrating gate plating.

115

Delft Hydraulics

This concerned a scale model investigation at the gates of the intake sluice at the Volkerak (see Example 6.2.g in Chapter 6).

Figure A4.28: Initial wave radiation and small transverse waves observed at a later stage in a scale model of the gates of the discharge sluice in the Volkerak dam (Delft Hydraulics Report M1129-II). There are indications, but only few measurements, that this mostly concerns transverse waves with a wave length relative to a double period in relation to the period of the vibrations. The small transverse waves usually become much stronger than the initially radiated waves with the front parallel to the plating. For one particular case, at a gate with overflow, prototype data about the wave length are available. The pictures of Figure A4.29 show the wave pattern in the overflowing nappe and in the upper water. The small plank in the upper water was placed there to facilitate the determination of the wave length. The explanation for the small waves, transverse to the flow direction, is similar to what was found in Paragraph 4.5.2 for the subharmonic flap gate movement in case of wave load.

116

Delft Hydraulics

Figure A4.29: Small waves at right angles to the flow direction, observed at a vibrating (bottom-hinged) gate in prototype (on the left, waves in the nappe, on the right the small waves upstream of the flap gate). Vibration frequency 15 Hz, wave length of the small transverse waves 5 cm. Delft Hydraulics Report R251. The equilibrium amplitude is not theoretically explained. Sometimes the transverse waves are so high, that the normal maximum steepness that may be reached without the breaking of waves, is exceeded. Sometimes the water starts to splash considerably, see Petrikat (1980). This probably generates still more damping, due to which the growth of the wave height does not continue.

117

Delft Hydraulics

118

Delft Hydraulics

FLOW-INDUCED EXCITATION IN CASE OF FLOWSURROUNDED STRUCTURES (IN PARTICULAR RODS AND CYLINDERS) AND FLUID RESONANCE AT PUMPS, TURBINES AND DISCHARGE SLUICES 5.1 General regarding flow-surrounded objects

Flow-surrounded objects are divided according to causes of excitation, as indicated in Paragraph 1.3. These relate to turbulence, instable flow, excitation due to vibration itself, amplification due to fluid oscillation and instable fluid oscillation. Finally, a separate review is given on trash racks. In case of flow-surrounded objects, naturally the added water mass, the added water damping and the added water stiffness play a role as well. These elements however are strongly interconnected and interwoven with the excitation, as they all relate to the same local flow pattern around the structure. Therefore, separating these forces is not very useful, with the exception of the added water mass. The data about this are given in Chapter 3, Figure A3.5. It is not possible to give a more or less full overview of what is known about flow excitation at flow-surrounded objects in just one chapter. Therefore reference is made to manuals such as Blevins (1990) and Naudascher (1992) and (1994). Remedies for vibrations of rods and tubes are given in Chapter 7. The problems regarding rods and other flow-surrounded objects are somewhat different from those of gates and valves: In case of flow-surrounded structures there are, except for the object itself, no flow limitations. The flow pattern has a great degree of freedom. Therefore the flow pattern may vary strongly, depending on the circumstances. Roughness of the surface and the Reynolds number, as an example, often exercise great influence on the stationary and dynamic flow forces. Also the initial turbulence of the water is influential. If there are more objects close to each other, then the flow pattern is again more fixed and the sensitivity to the factors mentioned above decreases. In case of flow-surrounded objects the (vibration) movement of the object does not have the effect that the discharge periodically varies. Therefore the flow inertia plays a much less important role than in case of gates (where an entire part of a culvert may cooperate, such as in case of the bath plug vibration in Paragraph 4.4.2). Also the plunger effect, such as occurring at gates (illustrated in Figure A4.9), does not occur at flow-surrounded objects. Therefore the added water mass also remains limited (Figure A3.5). In case of longer rods and cylinders, vibrations with higher harmonics may also be generated. As vibrations occur in a limited Vr area (Vr = V/fnD, with V = approach flow velocity, fn = resonance frequency and D = cylinder diameter), vibrations are generated in a number of velocity ranges. In case of slender structures and loose elements in flowing fluid, in general some vibration is permissible, provided it may be forecasted how great that movement will

119

Delft Hydraulics

be. Therefore, for the description of flow excitation and the vibration behaviour often a presentation is looked for, that includes the amplitude magnitude. The influence of the viscosity of the water (expressed in the Reynolds number) plays a major role, particularly in case of rods with a more or less rounded shape. Therefore objects with a sharply defined point of separation of the flow (the angular rods) and the round shapes are discussed in separate paragraphs. As in case of flow-surrounded rods and tubes, round cross-sections are often discussed, below a number of issues are further explained. In case of round or rounded objects, the point of separation of the flow is not fixed. Therefore even in prototype (varying with the water velocity), the Reynolds number, however great it may be, still influences the Strouhal number and the drag term. In case of a round cross-section (or in case of bevelled shapes in general) the flow pattern is so instable, that small influences such as the thickness of the boundary layer along the wall and the roughness of the surface, have a great influence on the point of separation of the flow and on the periodic variation of the point of separation. The place where the flow separates determines the width of the wake, and this again is the determining factor for the stationary and dynamic forces operating on the object. The drag term and the Strouhal number both depend on the width of the wake and therefore are correlated. This may be best illustrated on the basis of the characteristics of a circular cylinder. For this, Figure A5.1 shows the drag term Cw and the inversion of the Strouhal number S (the dimensionless representation of the frequency of the flow excitation). The width of the wake is a function of the Reynolds number VD/ (V = approach flow velocity, D = cylinder diameter and = dynamic viscosity of the fluid). In case of a broader wake, greater vortices are generated that also move at a greater distance from each other. The velocity as a result of which these vortices move away from the object however is not influenced by the width of the wake, so that the separation frequency of the vortices is inversely proportional to the width of the wake. The flow resistance of the object however is exactly proportional to the width of the wake. The width of the wake in case of a circular cylinder depends on the place where the flow separates. The more upstream the point of separation is located, the broader the wake of the flow. The influence of the Reynolds number on the width of the wake may be further clarified by reviewing the mechanism as a result of which flow separates from a curved surface. Here, this is limited to the flow in the proximity of the cylinder surface. At the upstream half of the cylinder, there is an accelerating flow. Along the streamlines there is a negative pressure gradient in the area of acceleration (the flow accelerates because water particles move from higher to lower pressure). The pressure gradient operates across the entire boundary layer.

120

Delft Hydraulics

Figure A5.1: Drag term and Strouhal number for a cylinder with circular cross-section (Bishop and Hassan, 1964). The distribution of the acceleration of water particles due to the pressure gradient does not match the velocity distribution in the boundary layer. The velocity distribution progresses logarithmically or parabolically (at the wall, the velocity is zero), the distribution of the velocity however is almost block-shaped. Subsequently a point arrives, at which, as long as the flow remains attached, the flow will slow down again. There, the pressure gradient is positive. The water particles along the wall slow down just as much as the particles further away from the wall. The particles in the proximity of the wall, where the velocity was very low due to the boundary layer, even start to flow back (vortex formation)! The thicker the boundary layer, the greater the vortex formation, and the quicker the flow will separate. Because of this, the wake becomes wider. So, the thicker the boundary layer, the wider the wake. A wider wake means that the resistance of the cylinder is greater and that the frequency of the flow excitation decreases. In Figure A5.2, from Harrisson (1967), the thickness of the boundary layer is presented (expressed in the so-called displacement thickness d, in which U0d is the shortfall of the discharge due to the boundary layer) as it develops in case of a flat plate. x is the distance from the beginning of the plate and k is the roughness. Assuming for the cylinder that x is given (and depends on the cylinder diameter and the place where the observation is made), then it may be observed how, in case of increasing velocity (at the cylinder with an increasing Reynolds number therefore), the boundary layer thickness develops. This explains for a major part the course of the lines in Figure A5.1 (at least, starting from Re = 105). The bundle of lines on the left in Figure A5.2 represents the values in case of small initial turbulence of the water and on the right in case of turbulence-free water (therefore, when the plate is dragged along). More turbulence in the water has a similar influence as compared to that of a higher Reynolds number. 121

Delft Hydraulics

In case of greater Re values (starting from Re = 2 106), the value of S-1 in Figure 1 decreases again. From other data it also appears that the Cw tends to decrease in that area. This decrease does not follow the boundary layer thickness. Below is the current explanation. In case of a turbulent boundary layer, albeit the boundary layer thickness becomes greater at increasing Reynolds numbers, but also other factors come into play that influence the separation of flow at a circular cylinder. The increasing turbulence causes water particles further away from the wall to exchange more intensively with particles that are closer to the wall (turbulence exchange between the different layers) and therefore the velocity distribution in the boundary layer becomes more block-shaped (filled), which has the same effect as a thinner boundary layer. Therefore the resistance becomes lower and the Strouhal number becomes higher. But the dynamic flow excitation is also much more irregular (noise-like) due to the turbulent boundary layer. In case of even greater Reynolds numbers (approx. 5,106), the boundary layer thickness further increases, and a more stable situation is generated again with a more regular periodic load. More about circular cylinders may be found in Paragraph 5.3.

Figure A5.2: Thickness of boundary layer in case of a flat plate (Harrison, 1967).

122

Delft Hydraulics

5.2 Angular rods: excitation due to flow-induced turbulence


For as far this is caused by the turbulence in the wake, the excitation frequency of the flow excitation may be expressed by the dimensionless Strouhal number, which depends on the shape of the cross-section. If the flow pattern is stable, the roughness of the surface and the viscosity of the flowing fluid (expressed in the dimensionless Reynolds number) hardly play any role. In this case it especially concerns angular rods or other kinds of objects with an univocally defined point of separation. Profiles, in which the flow may attach again after separation, deserve special attention. This again has consequences for the vibration behaviour, see Paragraph 5.4.1 (regarding galloping) and 5.7.2 (regarding grid bars). The turbulence in the wake of an object with a sharp point of separation is irregular and the excitation is spread out across a certain bandwidth. The Strouhal number relates to the dominant frequency. In case of this type of rods, usually no distinction is made between the Strouhal number based on the excitation in the flow direction and the one that is perpendicular to the flow. Sometimes the frequencies are only determined on the basis of pressure measurements in the wake directly behind the rod. Care must be taken that the natural frequency or natural frequencies of the structure do not coincide with the frequency of the flow excitation. The Strouhal number is determining for the excitation frequency. In case of flow-surrounded objects, it is advised to choose a natural frequency above the excitation frequency, and not below it, as in case of lower flow velocities the excitation frequencies also become lower. Sometimes a factor 2 to 3 is recommended, but as the excitation is noise-like, 1.5 also appears to be sufficient. Systematic data about this however are not available. Figure A5.3 offers an overview of Strouhal numbers for different kinds of rods.

123

Delft Hydraulics

Figure A5.3: Strouhal numbers for various steel profiles with sharp edges (University of Washington, 1952). 124

Delft Hydraulics

5.3 Flow instability of cylinders and rods with circular crosssection and other cross-sections with rounded corners 5.3.1 General
As mentioned in Paragraph 5.1, there are various causes that may generate flow instability. In this paragraph attention is only given to the instability that relates to round or bevelled shapes of the object in an approach flow. Therefore it discusses objects with a crosssection that is circular or elliptic, rectangular or triangular with rounded corners. For these shapes the point where the flow separates has not been defined, and that goes for both sides. Therefore an instable flow is generated, in which the flow at one side influences the flow on the other. A flow pattern is generated in the wake, with a more or less regular pattern of left and right turning vortices that shifts, and at which new vortices separate from the cylinder again and again. Characteristic for the excitation as a consequence of a two-sided instable separation of the flow is that: the excitation is almost periodic; the excitation is strongest in the direction perpendicular to the flow; in the flow direction there is dynamic excitation with the double frequency of the excitation perpendicular to the flow; the Strouhal number strongly depends on the Reynolds number and on the roughness of the surface of the rod; the excitation is strongly influenced by the vibration; the vibration amplitude itself causes a kind of phase synchronisation (across the length of the cylinder) of the separating vortices, because of which the correlation of the excitation across the length of the cylinder increases; the excitation frequency may pull somewhat toward the resonance frequency (also referred to as lock-in); the forces in-phase with the vibration (added water mass and/or added spring stiffness) in the proximity of the dominant excitation frequency strongly depend on the frequency. The out-of-phase forces are also frequency-dependent.

These points are subsequently discussed on the basis of the data regarding circular cylinders, about which, all things considered, a lot is known.

5.3.2 The periodicity of the excitation


In Paragraph 5.1 the cause of the periodicity has already been discussed. As an exception to the strong periodicity in the excitation, the super-critical Reynolds area of 5 105 < Re < 5 106 was mentioned, in which the excitation across a certain frequency band is spread out due to the fact that turbulence dominates the formation of a regular vortex pattern. The periodicity of the excitation is coupled to the alternating separation of left and right turning vortices, as indicated in Figure A5.4. The vortices turn around each other in the wake in such a way, that at that location there is not much energy loss. The distance between the vortices is completely coupled to the width of the wake.

125

Delft Hydraulics

Figure A5.4: The wake, with alternating vortices.

5.3.3 Great excitation perpendicular to the flow


The fact that the excitation is strongest crosswise also relates to the separation of vortices that alternatingly turn left and right. As within a flowing fluid the total amount of rotation does not change, the formation of vortices is accompanied by a counter rotation of the water around the cylinder, which accelerates the flow velocity on one side, while it decelerates the flow velocity on the other side. This results in great pressure differences on both sides (the Magnus effect).

126

Delft Hydraulics

5.3.4 Excitation with double frequency in flow direction


The fact that there is an excitation with a double frequency in the flow direction, that is twice the magnitude of that of the excitation perpendicular to the flow, follows from a review of the anti-symmetry. Left or right turning flow makes no difference to the force in flow direction!

Figure A5.5: Character of flow, turbulence and flow-induced excitation in the area of larger Reynolds numbers (based on the diameter of the cylinder), after Wootton et al (1972) and Morkovin (1964).

127

Delft Hydraulics

Figure A5.6: Spectra and registrations of the excitation perpendicular to the flow direction for various values of the Reynolds number (Jones et al, 1969).

5.3.5 Dependency of Strouhal number on Reynolds number and roughness


Figure A5.1 shows clearly for circular cylinders how much the drag term and the Strouhal number depend on the Reynolds number. This has already been extensively discussed in Paragraph 5.1. Figure A5.5 provides the names of the various flow regimes that may be distinguished in case of circular cylinders. If the Reynolds number is in the sub-critical or in the trans-critical area, then the vortex pattern develops with alternating left and right turning vortices (the Von Krmn vortex trail). In the sub-critical area the Strouhal number is approximately 0.2, in the transcritical area it tends toward 0.3 to 0.4. In the super-critical Reynolds area the turbulence dominates, which results in an excitation that is spread out across a broader frequency band. Above Re > 5 106 a more regular periodic excitation is generated again. Figure A5.6 shows frequency spectra and the character of the matching registration for different Re values. Figure A5.7 also shows the S values for other shapes with rounded cross-sections. In Paragraph 5.1 it has been discussed that the influence of the Reynolds number on the excitation frequency relates to the boundary layer that is caused for that part of the cylinder where the flow is still in a condition of being attached. Also the roughness of the surface has a big influence on this. An increasing roughness has a similar influence as a shift toward higher Reynolds numbers, but at the same time the variations of Cw and S become less pronounced, which means that the Reynolds sensitivity decreases. Memorable tests were those of Sarpkaya (1976 and 1982) and Aschenbach and Heinecke (1981).

128

Delft Hydraulics

Figure A5.7: S values of rounded rod profiles as a function of the Reynolds number (Delaney and Sorenson, 1953).

129

Delft Hydraulics

5.3.6 Dependency on excitation of the vibration amplitude


The flow instability results in the fact that the excitation is strongly influenced by the vibration amplitude, especially because strong vibrations occur in the proximity of the dominant excitation frequency. The amplitude of the excitation force increases with the vibration amplitude. The causes may be different, see also Paragraph 5.3.7 concerning the correlation across the length. Here too, the vibration amplitude plays an important role. Moreover, self-excitation may occur: in that case the excitation increases proportional to the vibration amplitude. Figure A5.8 shows the magnitude of the excitation perpendicular to the flow as a function of the vibration amplitude (measured during a forced movement). The excitation strongly increases relative to the vibration-free situation in case of increasing vibration amplitude, and next (but only at the very high vibration amplitude of 2/D > 1) decreases again. Of course, the excitation force increases again when the equilibrium amplitude is approximated more closely. The different lines are the results of different researchers. The spread indicates that it is difficult to obtain exact data from measurements.

Figure A5.8: Dynamic excitation, perpendicular to the flow direction, as a function of the vibration amplitude (Griffin, 1980).

130

Delft Hydraulics

5.3.7 Correlation of the excitation in the length direction

Figure A5.9: Correlation of the excitation across the length of the cylinder, as a function of the vibration amplitude (Toebes, 1969). Figure A5.9 shows the increase of the correlation of the excitation for a cylinder, across the distance in the length direction, in case of increasing vibration amplitude. The measurements were taken at a cylinder that was artificially caused to vibrate in the frequency in which the excitation is greatest (S = 0.2). This concerns forces and vibrations perpendicular to the approach flow direction. The definition of the correlation, r, between signals at two locations (z1 and z2) is as follows: 1 lim T F ( z1 , t ) F ( z2 , t )dt T 0 1 im T F ( z1 , t ) F ( z1 , t )dt T 0
T T

r=

(A5.1)

5.3.8 Shift of excitation frequency


The fact that the frequency of the excitation may move somewhat toward the resonance frequency of the rod (the so-called lock-in effect) has already been discussed in Paragraph 2.3 and shown in Figure A2.4. This concerns vibration investigations at a circular cylinder in air flow. In Figure A5.10 similar results are shown for situations with flowing water. The mechanical damping here is varied; this is expressed in the Figure in the Scruton number. The significance of this number, defined by Equation A5.7, will be discussed at a later stage in this paragraph. It appears from Figure A5.10 that also in case of flowing water there is a tendency that the vibration frequency moves toward the resonance frequency, but to a much lesser extent than in the tests in air, as indicated in Figure A2.4. 131

Delft Hydraulics

In case of purely harmonic excitation of a mass spring system, the frequency of the vibration equals the frequency of the excitation. Normally the excitation frequency, and consequently also the vibration frequency, increases proportional to the flow velocity (see Paragraph 2.2.6 at the introduction of the Strouhal number). However, in case of a flow velocity that is equal to or greater than the value at which the excitation frequency, as follows from the Strouhal number, equals that of the resonance frequency, it appears that the vibration frequency remains in a certain velocity range around the resonance frequency. In this area the vibration amplitude also remains high. These aspects may be explained by assuming that in case of a greater vibration amplitude, the frequency of the periodically separating vortices due to the vibration are influenced in such a way, that the separation frequency becomes equal to the vibration frequency (the lock-in effect). As in case of lock-in the excitation frequency remains equal to the resonance frequency, there are stronger vibrations in the frequency area concerned. Another explanation for lock-in is that the great vibration amplitude causes the width of the wake to be greater and consequently the excitation frequency to be smaller. The fact that the wake is caused to be greater in case of very great amplitudes of the vibration is very likely, and this is somehow confirmed by the fact that in Figure A5.10 in case of greater damping, and therefore also smaller amplitudes, the vibration frequency stays less close to the resonance frequency. It is not clear why beyond Vr = 7, the line S = 0.2 is not reached again, which however is the case in Figure A2.4.

132

Delft Hydraulics

Figure A5.10: Response perpendicular to the flow direction of a cylinder in flow; density factor of the added water mass LD 2 / m = 0.272 . Meier-Windhorst (1939) (taken from Naudascher (1994).

5.3.9 Strong variation of forces in-phase with the vibration


The periodically varying forces that operate on a vibrating cylinder in flowing water may be distinguished in periodic excitation forces and coupled forces. The latter may be inphase with the vibration and the out-of-phase forces. The in-phase forces, also referred to as the forces due to the added water mass and added water stiffness, influence the natural 133

Delft Hydraulics

frequency (Paragraph 2.2, Equation A2.3, in combination with A2.4, shows that mass and spring stiffness are determining for the resonance frequency). The out-of-phase forces determine, together with the mechanical damping, the extent of damping out or amplification of the vibration. The coupled in-phase forces are referred to in various sources only in terms of added water mass and not in terms of added spring stiffness. As a result, there is the risk that the physical reality is not reflected accurately. Positive added water mass in this kind of presentation means that the frequency becomes lower, and if negative, the frequency becomes higher. By reverse, it may also be stated, that when the vibration frequency is higher than the resonance frequency, then consequently there is a negative added water mass. And if the vibration frequency is lower, then there is a positive added water mass. (Actually, a negative added water mass has no physical significance, as it does not relate to inertia.) Below, the lock-in phenomenon discussed in the above sub-paragraph, will be discussed as if it were an effect of added water mass and added water damping. From Figure A5.10 it also appears that the lock-in effect mostly occurs in the frequency range in which the resonance frequency is equal to or somewhat lower than the dominant Strouhal frequency. This means that the added water mass is negative in the lock-in area.

/ 1 V 2 DL in-phase forces; Cmh = F 2

'/ 1 V 2 DL out-of-phase forces; Cdh = F 2

Figure A5.11: The in-phase forces and out-of-phase forces, made dimensionless with the flow pressure and the cylinder diameter (Sarpkaya, 1978). As regards the water damping, it can only be said that in case of the vibration amplitudes that are reached during a free vibration, the water damping is negative with an amplitude that equals the (positive) mechanical damping. These results are compared with the force measurements at a cylinder in flowing water during artificially generated vibrations (Sarpkaya, 1978). The latter only distinguishes inphase and anti-phase forces, leaving aside whether in the Vr area where the lock-in effect occurs, there are also excitation forces of which the frequency is determining of the vibration 134

Delft Hydraulics

frequency. In these tests both the flow velocity, the vibration frequency and the vibration amplitude have been varied. As in case of the free vibration, as shown in Figure A2.4 and A5.10, the vibration frequency f does not strongly deviate from the resonance frequency (for 0.8 fn < f < 1.2 fn), the parameter V/fnD that is on the horizontal axis, is also more or less representative for the value V/fD. Thus the results of Figure A5.10 may be globally compared with the test results of Sarpkaya. In Figure A5.11 both the in-phase and the out-of-phase forces are represented dimensionless by dividing the amplitude by the dynamic pressure, multiplied with the surface excited by the approach flow. It may be observed that when V/fD > 5, the in-phase forces are negative. This globally matches what was found in Figure A5.10, i.e. that the vibration frequency f, in the corresponding area of V/fD, is higher than the resonance frequency fn.

'( / 4) D 2 Ly  in-phase forces; Cml = F

/ 1 2  y  y DL out-of-phase forces; Cdl = F

Figure A5.12: The same results as in Figure A5.11, plotted out as added water mass and as drag force (Sarpkaya, 1978). The in-phase forces have been plotted out by Sarpkaya (1978) in terms of added water mass, by dividing them by the acceleration and by the mass of the water that is displaced by the cylinder (Figure A5.12). The negative added water mass hardly changes for V/fD > 6.5 and that also corresponds to Figure A5.10, where it appears that the vibration frequency in the corresponding V/fD area tends somehow to remain constant. The fact that here again the added water mass may become negative, means that the vibration frequency at a freely vibrating cylinder may even become higher than the natural frequency in air! In case of low velocity the added water mass coefficient equals 1, which corresponds to the values for non-flowing water (see Paragraph 3.2.2, Figure A3.5). From Figure A5.12 it further appears, that in the area 2 < V/fD < 5, the (positive) added water mass is greater than in case of non-flowing water. In Figure A5.10, in the corresponding Vr area, a jump in the vibration frequency may be observed, albeit only in case of the low-damped vibration. Only 135

Delft Hydraulics

in case of an even lower Vr, the vibration frequency no longer tends toward the resonance frequency in still water, though it seems this should occur, following the tests of Sarpkaya. In Figure A5.12 also the out-of-phase forces are represented in another way as in Figure A5.11, i.e. by dividing these by the amplitude of the dynamic pressure as it would occur in case of vibrations in non-flowing water, and due to the surface of the cylinder (perpendicular to the vibration direction). This is not logical for the situation with flow: as deduced in Paragraph 3.4, in case of flow the differential velocity between flow and vibrating object is determining. The flow damping then tends toward a damping force that is proportional to the water velocity, multiplied with the vibration velocity (see Equations A3.43 and A3.44). Qualitatively however, Figure A5.12 is significant, as it indicates how strongly the out-of-phase force varies with V/fD. The area of negative flow damping largely corresponds to the area of strong vibrations in Figures A2.4 and A5.10.

5.3.10

Quantification of vibration amplitudes

Below, the vibrations perpendicular to the flow and in-flow vibrations are discussed separately. Vibrations perpendicular to the flow Figure A5.13 shows the vibration amplitudes as a function of a dimensionless damping for a two-sided freely supported prismatic rod with a circular cross-section.

Figure A5.13: Amplitude in the critical Strouhal area, as a function of the mechanical damping, in case of a two-sided, freely supported rod with a sinusoidal vibration pattern (Blevins, 1977). The vibration amplitude has been calculated by taking a certain vibration mode (in Figure A5.13 it was half a sinus) as a point of departure. In case of an estimated amplitude it 136

Delft Hydraulics

may be assessed from Figures A5.8 and A5.9 how great the flow excitation is at each location of a rod. As there is excitation in the resonance frequency, it may be stated that the excitation force creates an equilibrium with the damping. As the influence of the excitation and damping are not distributed equally across the length, an energy review is applied. The total energy added by the excitation (per period) needs to equal the energy dissipation due to damping. If this is not the case, then a new estimate of the amplitude needs to be made. The method of presentation of the damping is made plausible below. The amplitude of a vibration in the resonance frequency follows from Equation A2.28: kY 1 = F 2 Because, in case of resonance (compare Equation A2.28): (A5.2)

k = 2m
Equation A5.2 may be presented as: Y F = D 2 m 2 D This, with F = CL 1 V 2 DL , may also be presented as: 2 Y CL V 2 DL = D 2 m 2 D

(A5.3)

(A5.4)

(A5.5)

As the lifting coefficient CL again is a function of Y/D (Figure A5.8) Equation A5.5 may now be presented as: Y m 2 D2 = f( ) (A5.6) 2 2 D L V D As = 2f and S = fD/V, this is the same expression as that which is presented in Figure A5.13. For the estimation of the influence of the length of the rod, Blevens used Figure A5.9. As the correlation of the excitation decreases with the distance, the vibration amplitude also decreases in case of a greater length of the rod. In order to make use of Figure A5.13, it is necessary to know the vibration frequency to determine S. To make a safe estimate (i.e. an estimate that results in a high damping), a low estimate of the vibration frequency needs to be made. Both in air (Figure A2.4) and in water (Figure A5.10) it appears that when great vibration amplitudes occur, the vibration frequency is at maximum a factor 1.35 below the natural frequency in case of non-flow conditions. From Figure A5.13 a kind of limit value for the damping may be found, above which the vibrations are negligible. This concerns the value, as defined in Equation A2.14. In the absence of an externally applying damper, the is a material constant. As the vibrations are again and again in the proximity of the Strouhal frequency relative to a Strouhal number of 0.2 (the measurements concern the sub-critical Reynolds area), the value of is fixed. For the mass, the mass of the cylinder together with the added water mass (in this case a mass equal to the cylinder volume, multiplied with the density of water) needs to be introduced. The magnitude on the horizontal axis must be 120 in order to 137

Delft Hydraulics

prevent vibrations (this is estimated, based on extrapolation from Figure A5.13), and thereby the value is fixed. In case of long rods and pipes, is a constant that depends on the structure material. In general it is so low (0.002 to 0.02) that because of this, the amplitude hardly reduces relative to an undamped system. Often sources give the Scruton number Sc as a measure for the minimally needed damping: Sc = 2m(2 ) LD 2 (A5.7)

As long as this concerns the vibration behaviour of circular cylinders with a more or less constant Strouhal number, this number is as useful as the magnitude used in Figure A5.13 and Equation A5.6. In more general terms, the limit value of the Scruton number (at which vibrations are prevented) is a function of the Strouhal number, based on the resonance frequency Sn (Sn = fnD/V) or the reduced velocity Vr (Vr = V/fnD). Scruton himself has also introduced it that way (Scruton, 1963; see also Figure A5.17). Remark 1: In Figure A5.13 and Equation A5.6 the mass m equals the mass including the added water mass. From Figure A5.13 it may be observed that a low estimate of the mass results in a high estimate of the vibration amplitude. Remark 2: The data of Figure A5.13 relate to rather low Reynolds numbers. It may be, that in prototype situations a Reynolds number is expected that is no longer within the sub-critical area. About this, very few data are available. If the vibration amplitude is greater, the influence of the Reynolds number on the excitation is less great than would follow from previous reviews (as an example presented in Figure A5.1); the freedom-of-place of the point of separation of the flow, is less great in case of a strong vibration. Probably the results of Figure A5.13 are still valid to some extent regarding the greater amplitudes. Blevins (1977) has also created diagrams such as Figure A5.13 for other vibration modes. These do not deviate much, although the maximum amplitude may deviate up to 20% more or less, relative to that in Figure A5.13. Vibrations in the flow direction (in-flow vibrations) As stated above, in-flow vibrations have a frequency (or Strouhal number) that is twice higher than in case of crosswise excitation. The consequence of this is, that at a given structure, resonance may already occur at a twice lower flow velocity. Therefore the possible in-flow vibrations are often determining for the design. A number of reported problems therefore only relate to in-flow vibrations. Instability in the flow direction not only relates to vibrating rods, but may also retroact on the flow itself. In case of flow through a grid, this may therefore result in a pulsating discharge. Figure A5.14 shows amplitudes measured in the flow direction, found in the proximity of piers that were driven into the soil, for an oil terminal under construction in a tidal area (Wootton et al, 1972). In this situation there are Reynolds values in the critical area. Two frequency areas may be distinguished, one of which is close to the doubled Strouhal number (relative to the excitation perpendicular to the flow direction).

138

Delft Hydraulics

Figure A5.14: Measured amplitudes in flow direction, in case of a bottom-attached (cantilever) pier in prototype (Wootton et al, 1972). A value of the inverse Strouhal number at large Reynolds numbers (order 106) is in the area 2.2 to 3.8, according to Figure A5.1; for the excitation in flow direction this number would therefore be bisected, which corresponds with the left bump of Figure A5.14. In case of measurements in the sub-critical area such as done by Walshe (1967), a second bump was not found. It is not clear whether initial turbulence in the approach flow played a role in Woottons measurements. In reality the right bump is not so important, because, taking the requirement that the natural frequency needs to be high enough to avoid the area on the left as a starting point, the area on the right certainly meets that requirement. Just like there is a limit value for the damping that may be indicated for vibrations perpendicular to the flow direction, at which there is hardly any vibration occurring (compare A5.7), this also holds for in-flow vibrations. King and Prosser (1972) indicated that, if for the damping:
Sc > 1.2

(A5.8)

then no vibrations occur. This concerns laboratory measurements at a one-sided bottomattached pier, for which the Reynolds number was sub-critical. The Scruton number Sc is defined by Equation A5.7. King and Prosser indicate that when the resonance frequency is high enough, vibrations do not occur, and not in case of low Sc values either. The minimum value for the resonance frequency (for in-flow vibrations) is at Sn = 0.55 (Sn = fnD/V). As Sn = Vr-1, a parallel may be drawn with Figure A5.14. This Figure shows that it is only safe from Sn > 1 onward.

139

Delft Hydraulics

Circular cylinders close to each other A criterion as in Equation A5.8 is also mentioned in references regarding circular cylinders that are positioned close to each other (such as rod bundles in reactors); see Padoussis (1980) and Connors (1970). No distinction is made between in-flow vibrations and vibrations perpendicular to the flow direction. The criterion used for vibration-free functioning is:

1 D2 L Sn 2 m

0.5

<K

(A5.9)

Padoussis indicates that according to experiments, the exponent 0.5 is between 0.3 and 0.5. For K, in case of rod bundles, 3.3 seems to be a safe value (referring to Pettigrew, 1978). The criterion of Equation A5.9 is again related to the minimum required damping. Naturally, safety may also be achieved by choosing the resonance frequency high enough relative to the dominant excitation frequencies. Padoussis (1980) provides references concerning the Strouhal number in the summary diagram of Figure A5.15. He states that in case of rod bundles it is difficult to make a distinction between excitation in the flow direction and excitation at right angles to the flow. Usually this too is not important, as the resonance frequency of circular rods and tubes is the same for both directions. The lowest Strouhal number always determines the design.

Figure A5.15: Strouhal numbers for various configurations of circular cylinders and rods (Padoussis, 1980). Conclusion: In this paragraph three different ways have been presented as a result of which it may be calculated which damping, in the shape of the relative damping , may be necessary to prevent vibrations at circular cylindrical tubes or rods:

140

Delft Hydraulics

the data from Blevins (1977) presented in Figure A5.13, for vibrations perpendicular to the flow; Equation A5.8 from King and Prosser (1972), for in-flow vibrations; Equation A5.9 from Padoussis (1980) for rod bundles, regardless of the direction of the vibration.

Although the data are not fully comparable, it appears that, when applied, Padoussis and King and Prosser roughly result in the same required damping, and that Blevins is one order higher. It is plausible that in case of vibration perpendicular to the flow, more damping is needed. The amplitudes that may be reached in case of undamped cylinders, are greater in case of vibrations perpendicular to the flow. At a given cylinder with a certain mass and stiffness, the transverse vibration occurs at a lower Strouhal number (and also at a lower Sn value) that is reached at a higher flow velocity. The amplitude of the excitation force therefore is also greater. It could have been expected that rod bundles also need more damping, as there too, vibrations will occur that are at right angles to the flow.

5.4 Self-excitation in case of flow-surrounded objects and rods 5.4.1 Galloping; a mechanism of self-excitation in case of vibrations perpendicular to the flow
As already discussed in Paragraph 4.4.5, galloping occurs because during oscillation of a wire or vibration of a rod, the instantaneous flow velocity relative to the moving body varies. This applies to vibrations perpendicular to the flow direction. The galloping of telephone wires or cables of high-voltage lines has already been described and explained in the classic manual of Den Hartog (1956). It appears that in case of ice-covered wires a kind of curtain may be formed at the underside, making the cross-section of the wire round at the top and sharp at the bottom. In case of a vertical sway motion, the place of the point of separation at the upper side varies strongly. In case of the (instantaneous) situation in which the flow strongly attaches (i.e. when the flow relative to the wire approaches obliquely from above), locally a great underpressure is generated that sucks the wire upward, exactly at the time when the wire is already moving upward. In case the wire moves downward, then the wire experiences, as it were, an oblique approach flow from beneath; the point of separation at the top shifts upstream, and the strong underpressure disappears. Relative to the stationary situation, there is a downward force, exactly when the wire moves downward. At the underside where the ice curtain is situated, there is a stable point of separation and nothing changes. Galloping is all about low-frequency oscillation with very great amplitude. A vibration phenomenon like this is also known to occur in case of airplane wings in situations in which the wing is positioned so steep that the lifting force may be lost (referred to as stall-flutter). The front of the wing is rounded and at the sharp back side the flow fully separates. In case of galloping it nearly always concerns a low-frequency vibration perpendicular to the flow direction. The vibration component in the flow direction is always positively damped. Den Hartog deduces an analytical expression for the extent of self-excitation. In Paragraph 3.4.3, the following flow damping was found (Equation A3.50) for a vibration perpendicular to the flow: 141

Delft Hydraulics

cw =

1 C VA( L + Cw ) 2

In this, cw is the water damping as it occurs in the dynamic, CL is the lifting coefficient, CW is the drag term and A is the surface of the vibrating object that experiences the approach flow. Both CL and CW are functions of the angle of approach flow. Now, the following holds for the self-excitation, that may be considered as a negative water damping:
cself exc. = C 1 VA( L Cw ) 2 (A5.10)

and it is clear that, if CL/ is strongly negative and great relative to the drag term, selfexcitation may occur (certainly in case of the absence of mechanical damping). This has been investigated in more detail by Novak (1969). Figure A5.16 concerns galloping of a rod with a square cross-section. In case of an upward movement, the lifting force occurs, because in case of oblique approach, the flow at the top attaches again more, while the separated flow at the bottom however moves further away from the rod. Due to the tendency of attachment, an underpressure is generated at the top, while the underpressure at the bottom, if present, disappears. Therefore this situation is instable as regards vertical vibrations. Whether these vibrations will indeed occur depends on the question whether there is sufficient mechanical damping to compensate for the self-excitation.

Figure A5.16: Lifting coefficient as a function of angle of approach flow load and both the calculated and measured equilibrium amplitude for various values of the (dimensionless) mechanical damping and flow strength (Novak, 1969). The diagram on the left concerns the dimensionless component of the flow force (in the direction perpendicular to the flow direction), CFy, as a function of the angle of approach

142

Delft Hydraulics

flow . From the diagram it appears that the greatest value of the coefficient (dCFy/d(-) only occurs when the (-) has reached a certain value. This means that the vibration first needs to be pushed a little before the strongest negative damping is reached. By carrying out a calculation with small time intervals, it appears to be possible to calculate the amplification and the equilibrium vibration for a single mass spring system from the lifting force and resistance characteristics of a body. At each moment, the flow forces and the spring and damping forces are introduced in the dynamic equation. Figure A5.16 shows both the calculated and the observed equilibrium amplitudes, and this represents a good correspondence. On the vertical axis the quotient of the dimensionless vibration velocity (Cy = Y/V with Y = eventually reached vibration amplitude and V = approach flow velocity) and the dimensionless structure damping (c/aLV, with c = damping of the structure, a = chord of the square cross-section and L = rod length) have been plotted out. The fact that the vibration is sometimes only generated after a push is also expressed in the response diagram (on the right in Figure A5.16). The explanation of this phenomenon is that the flow that has separated at the corners and that was still far removed from the walls in the initial position, only approaches again because of oblique approach flow (albeit on one side), as a result of which suddenly a greater suction force may be generated. This is expressed in a strong negative value of CL/. From Figure A5.16 is appears that at certain values of the mechanical damping c multiple equilibrium amplitudes are possible, depending on whether the flow velocity is greater or smaller. Except for the fact that a pushed vibration may be stronger than a nonpushed vibration, and also a greater damping is required to prevent the vibration, this also has the effect that, in case of an increasing flow velocity and then again decreasing flow velocity, (in case of corresponding velocities) different vibration amplitudes are found.

Figure A5.17: Stability limit for vibration of a square rod in the direction perpendicular to the flow, in case of approach flow at right angles (Scruton, 1963) (taken from Naudascher, 1994). 143

Delft Hydraulics

In Figure A5.30 an indication is given of the Vr area in which galloping occurs. Naturally, the calculation method in the time domain only applies in case of lower Strouhal numbers (or high Vr), for if the frequency is too high, the coefficients determined for stationary circumstances are no longer valid. Before Novak published his research, Scruton (1963) had already indicated what damping is needed to prevent these vibrations of a square rod; see Figure A5.17. From Figure A5.16 it may also be read what damping is sufficient to suppress the amplification. From the Figure on the right this appears to be the case when:

LV
c

< 0.6 , or if

c > 1.67 LV

(A5.11)

It now appears that this may also be presented differently: Sc > 1.67Vr (A5.12)

The results of Scruton (1963) partially correspond to this, but there are also differences with Novak (1969). The deviation at greater Vr can hardly be explained. In case of a greater (wind) velocity, the Reynolds number (Re = V.a/) was also greater. That may result in a different degree of turbulence, but this explanation is not satisfactory. In case of the smaller Vr values however, the peak in the required damping is explainable according to Naudascher (1994), for in that case there is extra excitation due to flow instability (resulting in periodic separation of vortices).

Figure A5.18: Cross-sections of rods that are sensitive or insensitive to galloping (Blevins, 1977). 144

Delft Hydraulics

Other shapes of objects that are sensitive to galloping are mentioned in Figure A5.18. These all have in common that beyond the point of release, the flow remains rather close to the object, because of which the suction of fluid toward the passing flow is obstructed. Because of that, an underpressure is generated that becomes stronger the closer the flow approaches the wall. For those shapes that are insensitive to galloping, the approach flow results in positive damping (as in the case of a circular cylinder, Paragraphs 3.4.2. and 3.4.3).

5.4.2 Self-excitation in flow direction


In Paragraph 4.4.4 the in-flow vibrations of gates have been discussed. Figure A4.17 shows the mechanism of self-excitation that is caused by the simultaneous occurrence of the gap flow and the added water mass flow. Due to the latter, the extent of attachment of the main flow (i.e. the gap flow near the almost closed gate) is very strongly influenced. If the gate is instantaneously moved in downstream direction due to the vibration, then the flow of the added water mass flow is opposite to the vibration, which results in an earlier separation of the main flow. This again results in a stronger flow contraction, and therefore in a higher resistance of the gate. If during the vibration, the gate instantaneously moves along with the flow, the hydraulic head therefore increases, which amplifies the vibration. During the gate movement opposite to the flow, the hydraulic head however decreases, which yet again amplifies the vibration. This results in a self-exciting gate vibration, of which the amplitude increases exponentially. This hypothesis was developed by Kolkman (1980). Experience shows that the flow contraction may be influenced especially strong if it concerns an instable point of separation (at a hemispherical bottom edge or a quarter-elliptic edge that is curved too much) or if the separated flow is hesitant with regard to attachment further downstream or not (at a gate with a rectangular underside, in case of a lifting height of a quarter part up to twice the edge thickness).

Figure A5.19: Self-excitation in case of a circular cylinder as a result of the simultaneous occurrence of the normal flow pattern and the added water mass flow pattern (Naudascher, 1994). 145

Delft Hydraulics

Naudascher (1994) also thinks this hypothesis applies to vibrations due to flow of circular cylinders. Circular cylinders may vibrate considerably in flow direction at a Strouhal number that is approximately twice the value of excitation perpendicular to the flow. See Paragraph 5.3, Figure A5.14, with accompanying explanation. Figure A5.19, following Naudascher (1994), and referring to the flow pattern measurements of Aguirre (1977), shows a) the flow pattern relative to the added water mass flow, b) the average flow pattern and c) the flow pattern relative to the maximum negative acceleration, the maximum negative velocity, the maximum positive acceleration and the maximum positive velocity during the vibration respectively. It may be clearly seen, that when the velocity is in the direction of the flow, the flow radiates out strongly. The wake is wider in that case, so that exactly then, the flow resistance will be at maximum.

5.4.3 Flutter
Generally, flutter or flapping does not occur at hydraulic structures, but it is nevertheless discussed here for the sake of completeness. First of all, the phenomenon has become widely known in aircraft construction, where wings may indeed start to flutter. Next to this, it is also a well-known failure mechanism with regard to suspension bridges. The phenomenon is also known in daily life in case of Venetian blinds that rattle in the wind; even the carrier straps on a bicycle are known to demonstrate this flapping phenomenon. The mechanism is fairly complex and relates to the fact that there are two degrees of freedom, i.e. the vibration perpendicular to the flow in combination with a rotational vibration. The following may serve as an example: if, in case of horizontal flow a horizontally placed strip or lamella moves upward, while the point of fixation remains at the same place, then a turning moment proportional to the displacement is generated due to the horizontal resistance. Meanwhile the lifting force will decrease because the apparent velocity approaches oblique from above, and therefore the turning moment will decrease due to the lifting force as well. These two opposing tendencies do not compensate each other just like that, because one is proportional to the deviation and the other is proportional to the vibration velocity. These forces therefore are out-of-phase relative to each other. The turning moment causes a rotation of the structure, which in turn influences the lifting force again. This whole mechanism may be described in the form of two coupled equations for the twist, , and the vertical displacement, y. These equations also include the first and second time coefficients. Written in a very general way these are: d2 y dy d2 y d d 2 m 2 + ky = Ly y + Ldy / dt + Ld 2 y / dt 2 2 + L + Ld / dt + Ld2 / dt 2 2 dt dt dt dt dt and: Ip d 2 d d 2 dy d2 y k M M M M y M M + = + + + + + (A5.14) p y d / dt dy / dt d 2 / dt 2 d 2 y / dt 2 dt 2 dt dt 2 dt dt 2 (A5.13)

146

Delft Hydraulics

In the equations it is indicated that both the lifting force L and the flow moment M are dependent of all movement components; both the displacement that occurs at the vibration, as well as the velocity and the acceleration play a role. By filling in these magnitudes, and the general solution for a negatively or positively damped vibration, the extent of amplification or the extent of damping out may be calculated. It is also possible to introduce yet another mechanical damping into the equations and to perform a calculation in the time domain. The coefficients themselves are determined by a flow calculation or by scale model tests, in which the forces are measured during a forced vibration.

5.5 Amplification of the excitation due to fluid resonance


Each natural oscillation of the fluid may be excited by a periodically operating force with a frequency equal to the resonance frequency or close to it. The excitation force may be caused by an object in the flow that experiences periodic forces due to the flow. As action equals reaction, the force that the flow exerts on the object also determines the reactive force that the object exerts on the fluid. Once the fluid has been caused to oscillate, then this naturally generates an extra dynamic load of the object. It is possible that a periodically alternating flow pattern is now generated, in which the already present periodic separation of vortices is amplified. The previous paragraphs provided an overview of periodic excitation of objects in flow. These same excitation sources may therefore also cause fluid resonance. In the following enumeration various situations are indicated, in which the fluid forms a fluid resonator that may generate amplification. In order to determine the natural period of the resonator, the stiffness and inertia components need to be known. An energy review may also be useful: equalization of potential and kinetic energy. a. One vessel, connected with the outer water by a pipe, or two vessels that are mutually connected by a pipe. The stiffness is determined by the retrogressive force relative to a difference in water levels. That difference again is coupled to the displacement of the water in the pipe. The inertia is caused by the mass of the fluid in the pipe. A twovessel system also constitutes a fluid resonator. A closed vessel in which there is a standing wave. As shown in Figure A2.7 the situation is somewhat comparable to example a, provided that the water level difference now relates to the water level that changes across the distance, and what initially was the constant velocity across the length of the tube, now is a velocity that varies with time. This means that the natural period needs to be found on the basis of the equalization of the potential and the kinetic energy. Also a half-closed vessel or two half-closed vessels may generate a similar oscillation, see Paragraph 2.3.1 and Figures A2.7 and A2.8. A closed gate shaft. Now the compressibility of the enclosed air, together with a possibly occurring water level difference, generate the spring stiffness. The inertia is again determined by the mass of the water in the culvert parts between the shafts. An enclosed air bubble. The stiffness again follows from the compressibility of the enclosed air or gas bubble. The inertia must be calculated from the kinetic energy of the radial approach flow and runoff.

b.

c. d.

147

Delft Hydraulics

e.

f.

A closed tube, in which there is a standing wave. This is a similar situation as in b, only here the potential energy is not caused by gravity, but by the compression of the fluid and the elasticity and deflection in the pipe wall. A half-closed pipe or two halfclosed pipes may also be caused to oscillate, see again Paragraph 2.3.1 and Figures A2.7 and A2.8. In a canal or an outer harbour a transverse oscillation may be generated, although this oscillation is more strongly damped than in the previous examples due to the fact that in this case wave energy radiates away toward the open end.

A striking example of fluid resonance is the transverse oscillation that may be caused at a bridge pier in a canal. A certain flow velocity is critical, in which the frequency following from the Strouhal number of the pier corresponds to the natural period fR of the transverse oscillation. Figure A5.20 shows scale model investigation results. The results are very similar to the response of a mechanical oscillator due to flow, as presented in Figure A2.4.

Figure A5.20: A standing wave in transverse direction in a canal, due to flow along a pier (Crausse, 1939) (from Naudascher, 1994).

5.6 Instable fluid resonance caused at pumps, turbines and discharge sluices
Paragraph 4.5 discussed the preconditions at which valves or gates caused a risk for instable fluid resonances. These may be resonances in pipes, in gate or surge shafts, in closed or half-closed basins and in the outer water. Risk-bearing circumstances relate to the relation between the discharge and the hydraulic head of the gate. These relations between discharge and hydraulic head may also occur at pumps and turbines.

148

Delft Hydraulics

A first precondition for the occurrence of instable fluid resonance is a situation in which low-damped standing wave movements may occur, regardless whether these are freewater surface waves or compression waves. In that case, similar to Paragraph 5.5, there is a fluid resonator. Figure A4.24 also indicates a number of situations, in which instabilities may occur due to the presence of valves (these may also be gates). Figure A2.8 also shows situations in which low-damped standing waves may occur. Resonators at hydraulic structures are caused in case of: 1. 2. a situation in which upstream or downstream of a valve, gate, pump or turbine, there is a basin or shaft, connected with a pipe to the outer water, in such a way that the shaftpipe system has its own natural resonance frequency; a situation in which upstream or downstream a pipe with finite length is present, that is connected with the outer water, and at which a standing compression wave may exist. A similar situation may also be generated in case of an open canal; in that case these are surface waves; a situation with two or more gates, pumps or turbines side by side, in which, crosswise to the direction of the main flow, standing waves may be generated; if there is a trough at one gate, then there is a crest at the other and vice versa.

3.

The second precondition for instability relates to the discharge-head relation. As regards a downstream instability, the precondition is that a lowering of the downstream water level causes a decreasing discharge, and a raising of the downstream water level causes an increasing discharge. The precondition for an upstream instability is that a raising of the upstream water level results in a decreasing discharge. Pumps and turbines For situations occurring at pumps and turbines, the precondition for instability may be formulated more simply: Instability occurs when an augmentation of the hydraulic head generates a decreasing discharge and a reduction of the hydraulic head generates an increasing discharge. This precondition may also be found in Streeter and Wylie (1967). At a pump, this situation occurs when it hardly generates any discharge, due to the fact that a valve is almost closed or the rotational speed is such that it has to function far beyond its maximum output. Given a certain rotational speed, there is in the Q-H diagram always an area in which Q/H is positive and an area in which this magnitude is negative. In case of hydraulic turbines it may also occur that the Q-H relation enters the critical area. Often the situation in which instability is generated matches the separation of flow from the blade, a situation that is comparable to the stall-flutter mentioned in Paragraph 5.4.1. Siphons and automatically operated gates The oscillations observed above all relate to a gate or machine at which the discharge immediately reacts to a changed hydraulic head. In case of siphons however the priming and depriming takes some time. If this time shift is such, that the standing wave has already travelled half a period further before reaction occurs, then an inverse discharge relation is also 149

Delft Hydraulics

the cause of instability (a greater hydraulic head therefore causes a greater discharge). There are also experiences in which several siphons react to each other in such a way, that oscillations with great amplitudes are generated in the upper reach. Something similar may also occur at automatically operated water level control gates (widely known in irrigation techniques), in which float indicators are used where water comes in and runs off. Here too instable oscillations of the water level have been observed. In case of electrically operated controls, delays may be caused when the water level that needs to be regulated is at some distance from the gate, unless the regulation takes account of the transit times of the waves.

Figure A5.21: An instable fluid oscillation at a discharge sluice (Delft Hydraulics Report M1711). The registration shows the water level oscillation in the upper reach. Remaining situations Despite the available theoretical knowledge about the instable fluid resonator (Paragraph 2.3.3 and 2.3.4), situations also occur for which it is not possible to formulate a mathematical description of the observed phenomena. In Figure A5.21 a situation is represented (discharge sluice Zoommeer, Delft Hydraulics Report M1711), in which during a scale model investigation important water level oscillations were observed. One of the situations that was instable, was the one in which a discharge was withdrawn through six openings, and the other openings (especially the outer one) was completely closed. In case of a hydraulic head of approximately 2 m, in the upper reach crosswise to the flow oscillations could be generated with a double amplitude of approximately 2 m (everything translated to prototype value). For this oscillation movement of the water it has never worked to formulate a mathematical description. It is however very plausible that the instability relates to a variation of the discharge capacity of the openings with either the water level upstream or the obliqueness of the flow. Considering the 150

Delft Hydraulics

exponential growth of the amplitude, this is a typical instable fluid oscillation (selfexcitation).

5.7 Vibration danger at trash racks 5.7.1 Design in relation to sensitivity to vibration
In the field of vibrations of trash racks Delft Hydraulics has no experience of its own. Next to other sources that were consulted, in this paragraph use is made of the chapter about trash racks that J.J. Cassidy (Bechtel Engineering, San Francisco) wrote for the ICOLD 1996 bulletin, concerning vibrations of gates and trash racks. (dr P.A. Kolkman was actively involved in this bulletin).

Figure A5.22: Possible design of a trash rack (Syamalarao, 1989). Dimensions in mm. Quantitative data about the vibration behaviour of trash racks are very rare. Naudascher and Wang (1930) presented data about racks with circular rack rods and long strips. These data have also been included in Naudascher (1994). Many cases are known in which trash racks were severely damaged due to vibrations. Syamalarao (1989) provides an overview with an enumeration of damages, from which it appears that also for the racks that were (often severely) damaged, the rod profile of the rack rods differed widely in shape. In Paragraph 6.8 this overview has been included. It is remarkable that, next to flow excitation, also the excitation due to periodic pulsation in the flow generated by pumps and turbines (frequency here is rotational speed multiplied with the number of blades) is mentioned as a source of vibration. All kinds of damage have been observed, but these are mostly concentrated at the underside or side of the rack. Vertical rods were sometimes bent, broken, and in some cases they had even completely disappeared. Anchoring bolts had sometimes been pulled out and nuts had been loosened due to the vibration. 151

Delft Hydraulics

It appears that at trash racks different vibration mechanisms may occur. A complete analysis therefore is very complex. A very interesting introduction for this may be found in Naudascher (1994). In addition, in ICOLD (1996) it is mentioned that racks at pump-turbines may experience approach flow from both directions, which makes the system even more complex.

Figure A5.23: Possible details of rods with a coupling rod (Syamalarao, 1989). Dimensions in mm. In this paragraph it may suffice to discuss the design criteria. These relate to the minimal flow velocity at which vibrations may occur, respectively the minimally required natural frequency to avoid these vibrations (and that always relates to the first harmonic vibration). Naturally the design, especially the rod profile, the distance between the rods and the extent of obliquity with which the flow may approach the rack, play an important role. Figure A5.22 shows an impression of a trash rack as it may be used for an intake of a pump or a turbine. Figure a shows the vertical cross-section perpendicular to the surface of the trash rack, Figure b shows the front view and Figure c shows a detail in horizontal crosssection. Figure A5.23 shows a number of applied rack rod dimensions in combination with coupling rods. Figure a shows, in horizontal cross-section, the presence of a rectangular strip as coupling rod, positioned outside the rack surface. Figure b shows a coupling rod that is approximately in the plane. Figure c and d show, in vertical cross-section, a round coupling rod. Figure e shows, in horizontal cross-section, staggered round coupling rods and Figure f shows a narrow coupling rod (relative to the rack rod). Figure A5.24 shows a number of shapes of applied cross-sections of rack rods.

152

Delft Hydraulics

Figure A5.24: Rack rods or strips as used in trash racks (ICOLD, 1996).

Figure A5.25: Profiles of rods that may have a risk of self-excitation vibrations due to galloping (ICOLD, 1996).

153

Delft Hydraulics

For the possible vibration behaviour of a trash rack there are many possibilities. Suppose, as an example, there is a supporting structure on which panels are mounted; both the rods themselves and the panels, and sometimes the panels together with the beams on which they are mounted, may then start to vibrate. The rods of the trash rack are usually connected by coupling rods, in order to increase the natural frequency in the transverse direction to the flow. Because of this, vibrations of separate rods do not often occur. As regards vibrations of the rods and the bearing girders, special attention needs to be paid to the design. The design needs to be such, that the point of separation of the flow is stable; 1. therefore, preferably no circular rods or rods with a rounded cross-section are used. Naturally this may be considered in case of elements that are sufficiently stiff anyhow. The shape and the cross-section of rods or beams need to be such, that self-excitation 2. or galloping will not occur; i.e. after the separation of the flow, the streamlines need to remain sufficiently far removed from the side of the rod profile. Next to this, of course care needs to be taken that the resonance frequency is high enough relative to the main frequencies in the remaining flow excitation. Figure A5.25 shows, in addition to what was mentioned in Paragraph 5.4.1 (Figure A5.18), some more shapes of the cross-sections that are sensitive to galloping. To give an impression of the area of excitation frequencies, Figure A5.26 offers the Strouhal number for various rod profiles that relate to excitation perpendicular to the flow direction. Experience shows that vibrations of trash racks in total especially occur in the flow direction. In that vibration direction it is not possible to support the rods. Moreover, the frequencies of flow excitation are generally higher, because of which vibrations already occur at low water velocities. The excitation may be very strong and the amplitude of these vibrations is such that there is a real probability of collapse due to fatigue.

154

Delft Hydraulics

Figure A5.26: Strouhal numbers of various rod profiles (Levin, 1957). Figure A5.27 shows a trash rack at which vibration measurements have been carried out. This example was contributed by H. Makarechian (Stone and Webster, San Francisco) for the above-mentioned ICOLD bulletin (1996). Vibrations occurred at approach flow velocities in the order of magnitude of 1.5 to 2 m/s. Rod vibrations perpendicular to the flow had double amplitudes of 2.6 mm; in vertical direction this was 0.12 mm. Considering the damage of the welded joints, the most critical were the in-flow vibrations with a double amplitude of 0.24 mm. As indicated in the Figure, this concerned the second harmonic vibration with a frequency of 32 Hz. Damage was observed at the welded joints that connected the horizontal stiffening girders (cross-section 19.5 x 152 mm2) with the circular tubes of the frame. 155

Delft Hydraulics

Also, these round tubes themselves showed fractures and part of the tube of 1.8 m in length was even disconnected from the panel. The trash rack rods showed no fractures. When looking at the Strouhal numbers based on the 32 Hz and on the rod thickness perpendicular to the flow of 16 mm and an approach flow velocity of 1.5 to 2 m/s, this results in an order of magnitude of 0.25 to 0.3.

Figure A5.27: A trash rack panel, with an indication of the vibration mode in case of in-flow vibrations, and the entire trash rack (ICOLD, 1996). 156

Delft Hydraulics

5.7.2 The maximum permissible flow velocity at round and rectangular grid rods
This paragraph presents data concerning the vibration behaviour of trash racks with circular cylindrical and rectangular rods. The data are based on Naudascher and Wang (1993) (the Figures were already included in Naudascher, 1994). The data are based on tests under laboratory conditions. Normative for the permissible flow velocity are the resonance frequency of the rod and the excitation frequency following from the Strouhal number (a measure of the dominant excitation frequency in case the rod would not vibrate). Observed amplitudes are generally presented as a function of the reduced approach flow velocity, Vr (Vr = V/fnd, in which V = approach flow velocity, fn = resonance frequency of the rack rod and d = rod thickness, measured perpendicular to the flow). The significance of the rod thickness d is similar to that of the cylinder diameter D that was used above. Investigation results regarding circular cylindrical rods indicate again and again that, when the natural frequency of the rod exceeds the excitation frequency of the flow more than a factor 1.3, then no vibrations occur. See for vibrations that are crosswise to the flow Figures A2.4 (air flow) and A5.10 (water flow). In case of air flow it may be observed that vibrations do not occur when Vr < 5, while during the tests S = 0.2. This means that no vibrations occur when Vr < S-1. In case of water flow the already mentioned factor of 1.3 applies. Also for square rods and rectangular strips up to a length (in the flow direction) of 4.5 times the thickness, vibrations only occur in the area Vr > S-1. For this, see Figure A5.30. Using these data, the Strouhal number may be converted into a maximum permissible approach flow velocity for the vibration-free condition of the rods, i.e.: Vmax = fnd 1.2 S (A5.15)

By reverse, assuming a given value of the approach flow velocity, the minimum required natural frequency may be calculated. This is important for the design. The minimally required frequency in this case always concerns the lowest natural frequency. Naturally, regarding these data, in reality it is advisable to always keep a certain safety margin in mind. The approach flow velocity as such follows from the discharge, the flow distribution, the blocking by girders and coupling rods and the extent of fouling of the rack. For the vibration behaviour the shape of the cross-section and the angle of approach flow of the rod are important. For the calculation of the natural frequency, the stiffness and the mass of the rod (or of the entire rack) and the added water mass need to be known. The latter may be estimated from the data of Paragraph 3.2.2. In case of racks that are placed oblique (in the vertical plane), the velocities that are used in the calculation are obtained by assuming that the discharge approaches the plane of the rack at right angles. As the plane of section of the rack with the water is greater than the corresponding vertical dimension, the arithmetic approach flow velocity is smaller than the real velocity.

157

Delft Hydraulics

Below, Vr and S values are given for the following four cases: Ia: circular rack rods, vibrations transverse to the flow; Ib: circular rack rods, vibrations in the direction of the flow; IIa: rectangular rods, vibrations transverse to the flow and IIb: rectangular rods, vibrations in the direction of the flow. Case Ia: round rack rods, vibrations transverse to the flow The Strouhal number at circular rods is a function of the Reynolds number Vd/ ( = kinematic viscosity). In case of rack rods, the Reynolds number is sub-critical (according to Figure A5.5 that is at a value between 300 and 500,000). To illustrate this: in case of a rod thickness of 5 cm and an approach flow velocity of 1 m/s the Reynolds number appears to be 45,000.

Figure A5.28: Influence of the blocking factor and/or mutual rod distance on the Strouhal number of circular rods (flow excitation perpendicular to the flow). From Naudascher (1994), referring to Levin (1957) and Crausse (1939). For the circular rod and for long, rounded strips, Figure A5.28 gives an indication of what the influence of the rod distance is on the Strouhal number. The S value in case of long strips is higher than in case of a circular cross-section. As the flow attaches in the first case, the width of the wake equals the strip thickness, while at the circular cylinder the width of the wake in the sub-critical Reynolds area is greater. No data are given for a situation of oblique approach flow on the rack.

158

Delft Hydraulics

Figure A5.29: Response in flow direction of circular rods. From Naudascher and Wang (1993), referring to Callander (1988) and Naudascher (1987). Case Ib: round rack rods, vibrations in the direction of the flow For the single circular rod, Figure A5.14 already gives an indication for the maximum permissible Vr value. In case of a rack, also the rod distance plays a role; for this, see Figure A5.29. The relation of the vibration amplitude relative to the rod thickness is indicated in the /d . Figure as X Case IIa: rectangular rods, vibrations in transverse direction to the flow In case of rectangular rods that are approached by the flow in longitudinal direction (therefore with the short side perpendicular to the flow), Naudascher (1994) distinguishes various mechanisms that cause vibrations. Which mechanism is normative depends on the length/thickness ratio e/d of the profile (in this, e is the length in the flow direction). In case of a short strip, the width of the wake is only determined by the strip thickness d (up to e/d = 2 to 3). The flow separates at the edges of the upstream side and the wake is not influenced by the rest of the strip. It is however possible, as already discussed in Paragraph 5.4.1, that galloping occurs. In the area 2 < e/d < 15.5 the strip experiences excitation when the (more or less) free boundary layer is capable of containing a total number of undulations (n = 1, 2, 3 and 4), the wave length of these undulations is proportional to the approach flow velocity and inversely proportional to the vibration frequency. The excitation mechanism is more or less comparable to what was discussed regarding the free boundary layer in 159

Delft Hydraulics

Paragraph 4.3. The Strouhal number for n = 1 was expressed in Figure A5.30 as Sh1. For n = 1 and other n values: Sh1 = 0.6 d e and Shn = n * Sh1 (A5.16)

If the strip length is more than 15 times greater than the thickness, then the flow at the side walls is completely stable again; only the wake at the back determines the excitation frequency. These aspects are presented in Figure A5.30, with matching Vr values. For oblique approach flow of the single strip, the Vr values are represented in Figure A5.31. In this, S' is the Strouhal number based on the effective strip width. The latter is defined as:
d = d cos + e sin

(A5.17)

while for S' and Vr': S = and:

fd V

(A5.18a)

Vr =

V fn d

(A5.18b)

It is probable that d' is a representative value as a measure of the thickness that determines the width of the wake. Therefore Vr' and S' may also be based on d'. For the sake of completeness however it is pointed out that the test results only relate to e/d = 10. Naudascher and Wang (1993) state that in case of rectangular rack rods, the e/d relation preferably needs to be greater than eight and that in case of a strip, sharp edges are preferable above rounded shapes.

160

Delft Hydraulics

Figure A5.30: Areas of several types of excitation perpendicular to the flow direction (transverse direction), in case of a strip at right angles to the flow. From Naudascher (1994), referring to results of Naudascher and Wang (1993).

Figure A5.31: Response of a strip, excited by an oblique approach flow, due to undulation in the boundary layer flow, and to vortices in the wake. From Naudascher and Wang (1993).

161

Delft Hydraulics

Figure A5.32: Vibration response, in flow direction, in case of a rectangular rod and approach flow at right angles (Callander, 1987). mr = m/mw, d/B = blocking (of the flow).

162

Delft Hydraulics

There are no data available concerning the influence of the rod distance on the excitation frequencies in intranverse direction to the flow; in their final conclusions they present a number of remarks: If s/d = e/d = 10, then the greatest vibration amplitudes are generated at an angle of approach flow between 6 and 12. If the strips are rounded at the ends, then this is between 8 and 16. This concerns excitation due to vortices in the free boundary layer that is close to the wall. Several frequencies occur during the excitation. Probably the mechanism of excitation is somewhat comparable to what was discussed in Paragraph 4.3 concerning the flow instability that is generated in case of a free boundary layer with a limited length. For angles of approach flow that are greater than 20 there is only a dynamic load as a consequence of the periodic separation of vortices relative to a two-sided separating flow. As was mentioned, unfortunately the data for smaller rod distances are unavailable. Although not explicitly mentioned, for greater distances the data for the single rod may probably be used.

Case IIb: rectangular rods, vibrations in the direction of the flow At a number of trash racks that were caused to vibrate, it was observed that these vibrated in the flow direction. At some strips with an approach flow at right angles, Callander (1987) carried out investigations. From these, Figure A5.32 was distilled. When considering the critical, therefore the maximum permissible, value of Vr, then it appears that for strips with a length/thickness ratio e/d of 30.875 or greater, this is around Vr = 2. For strips with a length of e/d = 5, the critical Vr value is 5, although below that a weak vibration may still be observed. If the strips in the rack are positioned close together, then the effect of the blocking needs to be estimated using Figure A5.28 (i.e. the curve that relates to the rounded strip). The influencing of the Strouhal number due to blocking in case of strips with a length/thickness ratio of e/d > 5, probably is the same as in case of a rounded strip as mentioned in Figure A5.28; at these longer strips, the flow also starts to attach again at the sides. The influence of the obliquity of the approach flow follows from A5.33, taken from Naudascher and Wang (1993). Also the rod distance here has been included as a variable. The reduced flow velocity in this Figure was based again on the effective thickness (Equation A5.16).
In Table A5.1 all Vr' values of Figure A5.33 have been elaborated, resulting in Vr values (based therefore on the thickness of the strip d), to verify whether oblique approach flow or approach flow at right angles is normative for the minimally required natural frequency of the strip in the direction perpendicular to the rack plane. With that, these may be directly compared with the Vr values of the strip that experiences approach flow at right angles.

163

Delft Hydraulics

Figure A5.33: Response in the direction perpendicular to the trash rack plane, of rectangular grid rods, in case of oblique approach flow. Strip distance is varied. From Naudascher and Wang (1993), referring to Nguyen et al (1988). When considering the definition of d' according to Equation A5.16 and the definitions for Vr' and Vr, then: V V Vr = and Vr = (A5.19) fnd fnd
from this, it may then be deduced that: Vr = Vr d d cos + e sin = d d (A5.20)

The elaboration was done for e/d = 10 (the strip cross-section to which also the observations in Figure A5.33 relate) and for e/d = 5. In the latter case it is assumed that the data of Figure A5.33 are valid here too (i.e. that d' is determining for the width of the wake and the magnitude of the vortices). As it may be expected that the strip thickness d has a much smaller influence on the wake (and the corresponding periodicity of the vortices) than the strip length e, this elaboration is justified. Eventually the results only serve to estimate an order of magnitude of Vr. The Vr thus obtained is compared to the minimum value found for a rod that experiences an approach flow at right angles.

164

Delft Hydraulics

From the equation it now appears that in case of approach flow at right angles, the maximum permissible Vr value Vr = 5 is normative relative to the Vr values that apply to oblique approach flow. Application of laboratory data concerning trash racks The data presented above have been obtained from test installations in which vibrations have again and again been investigated with one degree of freedom. The results indicate that the in-flow vibrations are excited in higher frequencies then the vibrations in transverse direction. In case of strips, the stiffness in flow direction naturally is much greater than in transverse direction. In reality however it is desirable to also choose the natural frequencies in transverse direction so high, that in case of crosstalk of the excitation in flow direction no problems are caused. Assuming a structure as shown in Figure A5.22, it will not cause any problems to construct enough coupling rods at the rectangular rack rods. Table A5.1: Elaboration of test results of Figure A5.33 to Vr-values.
s/e Vr' e/d Vr

3 3 1 1 2 2 3 3 2 2 3 3

15 ,, ,, ,, 30 ,, ,, ,, ,, ,, ,, ,, 45 ,, ,, ,, ,, ,,

10.1 ,, 13.2 ,, 4.6 ,, 4.9 ,, 5.3 ,, 6.1 ,, 3.9 ,, 3.9 ,, 5.3 ,,

5 10 5 10 5 10 5 10 5 10 5 10 5 10 5 10 5 10

22.8 35.9 29.8 46.9 15.5 27.0 16.5 28.7 17.8 31.1 20.5 35.8 16.5 30.3 16.51 30.3 17.8 41.2

165

Delft Hydraulics

166

Delft Hydraulics

EXPERIENCES IN PROTOTYPE AND SCALE MODEL INVESTIGATIONS

The examples that are described in this chapter are based as much as possible on experiences in which Delft Hydraulics was involved. As a consequence, these especially relate to gates and related structures. A few examples that have been described in other sources, and that are particularly illustrative of certain phenomena, have been added for the sake of completeness. Examples: 6.1 Gates above a threshold (free-flow (critical) discharge) 6.2 Low-head gates with drowned discharge conditions 6.3 High-head gates and other culvert gates 6.4 Valves 6.5 Seals and leakages 6.6 Overflow gates and flap gates 6.7 Stop logs 6.8 Trash racks 6.9 Dynamic phenomena due to cavitation and partial aeration

6.1 Gates above a threshold (free-flow (critical) discharge)


6.1a TRANSVERSELY POSITIONED WHEEL GATE (prototype experience) GENERAL: This example has been included as it is very illustrative of vibrations that in reality occur unexpectedly. Moreover, very few data on this gate are available. At first it may be expected that a gate with free-flow discharge holds very little danger of vibrations. The vibrations observed would only be discovered in scale models, in case the horizontal freedom of movement (deflection) had been reproduced there as well. DIMENSIONS AND WEIGHTS: Span 28 m. Further data unavailable. WATER LEVELS: Maximum upstream water level 2.7 m above the bottom (the vibrations occurred at small lifting heights). DESIGN: See Figure a. OBSERVATIONS: At this gate, strong horizontal gate vibrations occurred. It is unknown whether this only concerned the deflection of the entire gate or that the plating at the bottom edge was vibrating extra strong. The critical lifting height was 8 to 25 mm. The vibration amplitude was approximately 4 mm and the frequency was 3 to 5 Hz. 167

Delft Hydraulics

CAUSE OF VIBRATIONS: During the vibrations the gap width varied periodically, so that essentially a so-called bath plug vibration was generated. See Paragraph 4.4.2. REMEDIES: Extra stiffness and damping are difficult to realize. It may be best to quickly pull the gate through the critical area of gate openings. REMARKS: A similar situation may also be caused in case of a vertically placed gate that is positioned on a sloping (ascending) bottom or threshold. REFERENCES: Petrikat (1980).

Figure a: Cross-section gate, from Petrikat (1980). 6.1b ADVERSELY CURVED SECTOR GATE (prototype experience) GENERAL: The design of the sector gate includes a number of critical points regarding the situation with flow-through underneath. The shape of the plating is such, that the water load does not pass through the point of rotation. Moreover, its direction varies strongly with the water level. At the bottom edge the first stiffening girder is positioned very low, so that the flow-through underneath often does not remain free from the girder. DIMENSIONS AND WEIGHTS: Span 22.5 m. Further data unavailable.

168

Delft Hydraulics

WATER LEVELS: Maximum 3.4 m above the threshold (the vibrations occurred at small lifting heights of the gate). The downstream water level is lower than the threshold. The flow separates completely at the gate bottom edge. DESIGN: The gate is a sector gate with plating with inverse curvature. For the regulation of the upper water level an overflow flap gate is in place. This is of no significance for the vibrations observed. See Figure a. OBSERVATIONS: Vibrations occurred at a lifting height of 8 mm in the middle of the gate and 25 mm at the outer end. These were horizontal gate vibrations. Frequency 4.1 Hz. Upstream, strong transverse waves were observed, Figure b. Such strongly splashing waves have not been reported in any other source. For an explanation, see Paragraph 4.6. CAUSE OF VIBRATIONS: During the horizontal vibration the gate opening varies periodically, as a result of which bath plug vibrations may be generated. See Paragraph 4.4.2. REMEDIES: A damper has been used that is very effective at small openings, see Figure c. A spring ensures that the damper already touches the bottom before the gate is fully closed. The position of the underside of the damper adapts to the lifting height. REMARKS: The gate vibrations have been investigated in a model with scale 1:2 (see C. Miller: Ein Beitrag zur Bestimmung der Schwingungserregende Krfte an unterstrmten Wehren Mitteilungen Inst. f. Wasserbau, Univ. Stuttgart, Heft 42, 1977). REFERENCES: Petrikat (1980)

169

Delft Hydraulics

Figure a: Vertical cross-section of gate with upper flap (Petrikat, 1980).

Figure b: Transverse waves in the upper water (Petrikat, 1980).

170

Delft Hydraulics

Figure c: Damper structure, from Petrikat (1980). 6.1c SECTOR GATE ON DAM CREST (prototype experience) GENERAL: This example has been included as it concerns a gate that has collapsed due to its vibration behaviour. Although the design seemed to meet all requirements for functioning vibration-free (central sector plating corresponding to point of rotation, sharp bottom edge), in reality the distortion due to the water load was so great that the centre of the sector plating was shifted downward. This vibration behaviour could not have been forecasted with an (over)simplified scale model! DIMENSIONS AND WEIGHTS: Width 9 m, height 12 m and nappe 13 m. Polar inertia moment 5.76 106 kgm2. Mass per unit of gate surface area 340 kg/m2. WATER LEVELS: Upper water 11.7 above the threshold, downstream water level lower than the threshold. DESIGN: See Figure a. OBSERVATIONS: Failure of the gate at small lifting height, due to buckling of the lower radial stay. Figure b shows the probable distortion immediately after the buckling and Figure c shows how the gate subsequently came loose and was washed away. CAUSE OF VIBRATIONS: This is typically a situation in which a bath plug vibration may occur (Paragraph 4.4.2). The authors have also carried out a similar analysis and at a later stage the vibrations were also found in a scale model investigation. 171

Delft Hydraulics

REMEDIES: The gate needs to be so stiff, that even in case of water load, the centre of the curved plating corresponds to the location of the point of rotation. REFERENCES: Ishii et al (1980).

Figure a: Vertical cross-section of gate, from Ishii (1980).

Figure b: Collapse of stays.

172

Delft Hydraulics

Figure c: Tumbling down of gate after collapse. 6.1d BOTTOM EDGES OF SECTOR GATES ARKANSAS RIVER DAMS (prototype experience and scale model investigation) GENERAL: As there are many similar situations at the many dams in the Arkansas River, following the vibration problems that were observed, a search was carried out for a good bottom edge. This relates to sector gates that are positioned on a threshold. DIMENSIONS AND WEIGHTS: As regards more or less representative dimensions, see Figure a. Weight and stiffness of the suspension are unknown. Figure b shows a shape of the bottom edge that was used often. WATER LEVELS: Downstream water level below the threshold, upstream water level coupled to gate position. DESIGN: The centre of the curved plating and point of rotation match. The shape of the initial bottom edge is shown in Figure b. OBSERVATIONS: Very strong vibrations occurred at an opening of a few centimetres. CAUSE OF VIBRATIONS: In case of small gaps, music note rubber seals are very instable. The cause of this is both due to the design and the elasticity: in case of a vibration, the gap varies in width, the discharge through the gap varies and when the discharge slows down, an extra hydraulic head is generated (as with the bath plug vibration, Paragraph 4.4.2) as a result of which the gap tends to close even further. Although the vibration apparently only concerns the rubber, the hydraulic head influences the entire gate, generating a strongly alternating load. REMEDIES: Based on experience a scale model investigation has been carried out, focusing on a number of shapes of the edge, with and without rubber sealing strips. The conclusion is unambiguous: all solutions for a bottom edge with a music note rubber seal are sensitive to vibration. A sharp edge, possibly combined with a sharp stiff rubber strip, does provide a solution. See Figure c.

173

Delft Hydraulics

REMARK: In Chapter 7 it is advised not to use a music note rubber seal for bottom and top sealing of a gate. It is however suitable for a sliding side seal. REFERENCES: Schmidgall (1972) and Petrikat (1980).

Figures a and b: a) Cross-section of gate. b) Detail of bottom edge of the original design, from Schmidgall (1972).

Figure c: Tested gate edges and corresponding sensitivity to vibration. 6.1e GATE WITH EQUAL STIFFNESS AND DIMENSIONS IN BOTH DIRECTIONS (scale model investigations)

174

Delft Hydraulics

GENERAL: This example has been included because it is a design that, according to various criteria, should not be sensitive to vibrations, but nevertheless it is caused to vibrate. The flow pattern is stable. As long as the lifting height is small, the flow separates at the smallest opening and the cornered plating situated upstream plays no role. The instability indicator (Paragraph 4.4.6) indicates that when the gap varies in width, it does not cause an extra hydraulic head in the direction that is amplified by the gap change. There is also no danger of galloping (Paragraph 4.4.5), in which the angle of approach flow is important. This type of gate has been investigated by Ishii and Knisely (1992). This was a scale model investigation in a section model, suspended by wires and springs. These allowed a horizontal and a vertical movement. The authors found that the model was stable when one of the degrees of freedom was eliminated. DIMENSIONS AND WEIGHTS: See Figure a for the cross-section. The gate is approximately 3.5 m high and 3 m wide and has a span of 30 m. WATER LEVELS: During tests no overflow occurred; the upstream water level is limited by the upper side of the gate. At the downstream side the water level is very low and there is free runoff. DESIGN: See Figure a. OBSERVATIONS: In scale model investigations, in which a section was suspended by wires and springs that allowed translatory movements in horizontal and vertical direction, strong vibrations were found as soon as the horizontal and vertical frequencies approximated each other. CAUSE OF VIBRATIONS: Ishii and Knisely state that both vibrations (horizontal and vertical) are coupled by the water. Suppose, first there is a horizontal vibration. In Figure b the forward motion has been indicated by an unbroken line and the retrogressive motion has been indicated by a broken line. Because of the added water mass (during forward acceleration to the left) an extra water pressure will be generated that causes an (added water mass) force to the right and also a force that is directed upward. The result therefore may be that the movement does not remain horizontal, but becomes transversely directed upward. This results in the vibration direction A-A. Likewise, a vibration that is initially vertical (Figure c) will also generate a horizontal force (an acceleration downward causes an overpressure and will generate a deflection to the right, which results in vibration direction B-B). Both cases result in a transverse vibration in the same direction. This becomes one combined movement when both vibrations also have the same resonance frequency. The result is as if the system has one degree of freedom only. In Figure d this has been indicated once again. But the instability indicator (Paragraph 4.4.6) does apply to this system. A bottom-left movement reduces the gap width, as a result of which the hydraulic head increases and the bottom-left movement is amplified. The situation is similar to that of Example 6.1a.

175

Delft Hydraulics

REMEDIES: In the design process, care needs to be taken that stiffness and resonance frequency in the two directions are different. REFERENCES: Ishii and Knisely (1992).

Figure a: Cross-section of the gate, from Ishii and Knisely (1992).

Figures b and c: b) Deflection in case of horizontal vibration. c) Deflection in case of vertical vibration.

176

Delft Hydraulics

Figure d: A system with one degree of freedom in transverse direction.

6.2 Low-head gates with drowned discharge conditions


6.2a VISOR GATE HAGESTEIN (prototype experience) GENERAL: Figure a shows the design of the weirs in the Nederrijn River. The gates consist of a horizontal semicircular arch that is put under strain of tension by water pressure. Despite an extensive scale model investigation (see Part C, Chapters 5 and 6), vibrations were observed in the prototype of local plate fields. These vibrations fell outside the reach of the scale models used in the design phase, as only the elastic properties of cables and bending and torsion formation were brought to scale. DIMENSIONS AND WEIGHTS: Span 48 m, nappe of the visor 26.35 m and gate height 7.55 m. Local plate thickness 12 mm. WATER LEVELS: Downstream water level 4 to 6 m above the threshold, maximum hydraulic head 3.5 m. DESIGN: See Figures a, b and c. Inside the pier and the abutments space has been reserved to facilitate the radial outflow (to keep the load symmetrical). In the central pier a water power turbine and a cylinder gate have been included, the latter in connection with the fine-tuning of the discharge. OBSERVATIONS: When the hydraulic head is greater than 1 m, high-frequency plate vibrations occur, which generate small waves in the upper water. The vibration frequency is between 12 and 20 Hz. It seems that at the location of the vibrations, the local lifting height that is liberated (after the relieve of the rubber under edge) is 5 to 80 cm. The vibrations are audible. As the vibrations are generated at the bottom edge, and are still clearly visible (small waves) at the water surface, these must be rather strong vibrations. Over the years (the weir has been in use

177

Delft Hydraulics

for over 30 years) reinforcements have been mounted as well. And in the proximity of the abutments fatigue phenomena have been observed at the plate fields. CAUSE OF VIBRATIONS: The cause is the shape of the bottom edge. The periodically varying surrounding flow that accompanies the vibration, and that relates to the inertia of the water (added water mass flow), influences the normal discharge in such a way, that it is periodically cut off and accelerated and this again results in the force to maintain the vibration. See Paragraph 4.4.4. REMEDIES: A thinner and sharper edge will strongly reduce the lifting height that is critical. Stiffening of the box beam at the fender profile (Figure c) by propping it to the bottom gate girder has been carried out in 1993. It appears that the effect of this is not sufficient. Up till now it has been tried again and again to avoid the vibrations by keeping the gate outside of the critical area of lifting height during the regulation of the discharge or the upstream water level. In all those cases, the upstream water level was regulated by the second visor gate and the cylinder gate in the central pier. REFERENCES: Schoemaker (1959) (Delft Hydraulics Publication 14), Kolkman (1959), (1980) (Delft Hydraulics Publication 219), Delft Hydraulics Reports M561/700 and Q322-I.

Figure a: Plan of the visor gate.

178

Delft Hydraulics

Figure b: View of the weir from upstream side.

Figures c and d: c) Cross-section of the gate. d) Detail of bottom edge 6.2b STORM SURGE BARRIER KRIMPEN (prototype experience) GENERAL: When constructing the second gate of the storm surge barrier Krimpen it was decided to allow a leakage gap in the closed situation, in order to prevent that in case of alternating load during the changing of the tides, the gate would scrape across the bottom. (The first gate 179

Delft Hydraulics

has a wooden bottom beam). As during that time Delft Hydraulics was also of the opinion that a sharp bottom edge would not give rise to vibrations, the plating has been ended abruptly. DIMENSIONS AND WEIGHTS: The gate has a span of 80 m and is 8 m high. It has been constructed as a horizontal arch bridge. The arch has a height of 15 m. The freely protruding plating at the bottom edge may adopt many vibration modes; no calculations have been carried out concerning possible natural frequencies. The bottom stiffening girder is at 0.465 m from the edge of the plate. The plating is 12 mm thick. The threshold is at N.A.P. 6.5m. WATER LEVELS: Relative to N.A.P., the outer water level may rise up to +4 m. (The gate height is higher, because of the waves). During prototype measurements the water level at the seaward side varied from -0.65 m to +1.20 m, and at the side of the inner water from 0.4 to +1 m. DESIGN: See Figure c for the shape of the bottom edge. As a design magnitude, the gap in closed condition was considered to be approximately 35 mm, but in reality it will be variable due to alternating deflection. OBSERVATIONS: During an initial visit it appeared that there was an audible drone, also audible at a (small) distance from the gate. Measurements (at location A, see Figure b) later confirmed a vibration of 88 Hz. Figure d shows the spectrum. CAUSE OF VIBRATIONS: See Paragraph 4.4.4; this concerns the theory of the varying discharge coefficient. During later investigations concerning this type of edge, it appeared that there are many critical values for the reduced flow velocity Vr (the velocity here is divided by resonance frequency and plate thickness). See Figure A4.18. The bottom plate field has many natural frequencies, because of which there is a greater probability that one of the critical Strouhal numbers is reached. REMEDIES: Lifting the gate a little bit more causes the vibrations to stop (according to results of scale model tests, at a lifting height of approximately 2 times the plate thickness, no vibrations are expected anymore). As water density in case of a storm surge barrier is not a requirement, this regulating measure has been used. Sharpening the edge might also be considered. REMARKS: The first vertical lift gate that was built with a wooden beam at the bottom edge was vibration-free. It might however be expected (see Example 6.2.e). In case of an equal hydraulic head, and in case of a thicker edge, the frequencies of the flow excitation are lower, and would probably fall outside the area of resonance frequencies. REFERENCES: Kolkman (1980) (Delft Hydraulics Publication 219), Delft Hydraulics Report R1305. The general research on sharp edges was published by Jongeling (1987) (also Delft Hydraulics Publication 392) and Delft Hydraulics Report Q190. 180

Delft Hydraulics

Figures a and b: a: Plan view. b: Cross-section of the vertical lift gate.

181

Delft Hydraulics

Figures c and d: c) Detail of the bottom edge of the gate. d) Spectrum of the vibration velocity. 6.2c DISCHARGE SLUICE HARINGVLIET, RIVER GATE (prototype experience) GENERAL: On both sides of a triangular bridge girder, sector gates with double plating have been mounted. At the seaward side the circular plating starts from the bottom, so that alternating wave pressures do not create a force component on the winch. The gates rest on the bridge girders by way of several points of rotation. The gate is suspended on both sides in a winch. The bottom edge of the river gate consists of a rubber fender profile, the side seal is inflatable, so that the gate in non-closed position is free from the wall, as a result of which the system is also extremely low-damped. DIMENSIONS AND WEIGHTS: Gate width 56.5 m. Supporting trusses with a point of rotation centre-to-centre distance 14.8 m. Gate weight 520 tonf. Stiffness of two winches together 115 106 N/m in closed position, and in open position a little lower (in that case there is more oil in the lift cylinder). The gates themselves have been constructed in a very stiff way. WATER LEVELS: The water levels may vary from 2 to +5 m on the seaward side and from 1 to +2 m on the riverside. During the measurements at which vibrations on the riverside were registered, the water level of the river varied from N.A.P. to N.A.P. +0.5 m. The sea water level was 0.5 to 2 m lower. DESIGN: The shape of the vertical cross-section of the gate is normative for the vibrations. Figures b and c show the design of the cross-section.

182

Delft Hydraulics

OBSERVATIONS: The vertical (better: tangential) vibrations were observed during a special measuring campaign with the aim of reaching an understanding of the gate behaviour at all possible operating conditions. Vibrations with a frequency of 2 to 2.5 Hz occurred during gate openings in the area of 1.5 to 2 m and around 4 m. The hydraulic heads used during the measurements are indicated in Figure d, together with the intensity of the vibrations observed. CAUSE OF VIBRATIONS: In case of smaller openings, it may be that the flow that separates at the cornered plating alternatingly attaches again or not near the rubber bottom edge, see Paragraph 4.3. In case of bigger gap openings the flow may not attach, but the point of separation is at the upstream side of the gate, which is always unfavourable, see also Paragraph 4.4.6. REMEDIES: These have not been investigated more closely, as the vibration amplitude remained limited to a few millimetres. REMARKS: The semicircular shape of the fender profile might have given cause for horizontal gate vibrations to occur (see Paragraph 4.4.4), but the gate probably is stiff enough to prevent this. REFERENCES: Kolkman (1980) (Delft Hydraulics Publication 219) and Delft Hydraulics Report M754.

Figure a: Cross-section of the bridge and both gates.

183

Delft Hydraulics

Figures b and c: b) Vertical cross-section of gate at riverside. c) Detail bottom edge.

Figure d: Measured velocities with corresponding gate opening and hydraulic head. 6.2d BUTTERFLY GATES MUIDEN MARINE LOCK (scale model investigation) GENERAL: One of the lock chambers of the renovated ancient shipping lock at Muiden is only used as a discharge sluice these days. In this, butterfly gates have been used, 3.35 m high and 3.75 m wide. Between the two gates there is a small central pier. The structure has been mounted in a steel frame and could be installed as a whole. This design has been tested in a scale model 1:10. In this, the rotation stiffness of the hoisting equipment and the bending stiffness of the plating have been scaled. DIMENSIONS AND WEIGHTS: See Figures a and b. 184

Delft Hydraulics

WATER LEVELS: Outer water maximum N.A.P. +0.35 m, inner water minimum N.A.P. 2 m. DESIGN: In the initial design the plating at the sides was ended abruptly. The plating was supported by horizontal girders centre-to-centre distance 0.65 m. See Figure c. OBSERVATIONS: There were two conditions during which strong vibrations occurred: vibration I, when the gate was completely or nearly closed ( = 0 - 3), and vibration II, at = 50. Both vibrations were audible and were accompanied by small waves on the water surface. The vibration frequency in the model was 150 Hz (prototype translated 50 Hz). Vibration I occurred in case of a hydraulic head of 0.7 m and over; type II started from 0.35 m. Whether smaller hydraulic heads would also cause critical situations to occur has not been verified. CAUSE OF VIBRATIONS: For vibration I: see Paragraph 4.4.4 (theory of the varying discharge coefficient). Due to the surrounding flow related to the added water mass, the leakage discharge is periodically blocked, as a result of which a periodically varying hydraulic head is generated that amplifies the vibration. This type of vibration may occur up to a gap width of 2 to 3 times the thickness of the edge of the plate. Vibration II is especially coupled to an instable flow pattern; after the separation, the flow attaches again, but the point of attachment may fluctuate strongly with the vibration, as a result of which the size of the underpressure zone fluctuates strongly as well. A closing motion causes (in Figure d at the upper side) an extra strong contraction of the flow (influence of the added water mass flow), as a result of which the closing force increases again exactly at that time, and the motion is amplified. This is similar to what has been described in Paragraph 4.4.4: the theory of the varying discharge coefficient. REMEDIES: It has been decided to reinforce the edge of the plating (Figure e). First, thinner ribs have been tried, but these were ineffective. In the central pier seatings have been mounted, which the gate in closed position is pressed against. The abutments have been bevelled off in order to obtain a bigger gap when opening the gate. In the model no vibrations were observed and also in prototype no problems arose. REFERENCES: Kolkman (1980) and Delft Hydraulics Report M1311.

185

Delft Hydraulics

Figures a and b: a) Plan view. frame with two butterfly gates b) Vertical cross-section of one gate.

Figure c: Original design.

Figure d: Flow situation at 50.

186

Delft Hydraulics

Figure e: Applied improvements. 6.2e VERTICAL LIFT GATE LOWER HYDRAULIC HEAD STRUCTURE LITH SHIPPING LOCK (prototype experience) GENERAL: The shipping lock is emptied by pulling up the vertical lift gate slightly. Exactly in this area, considerable vibrations occurred. Initially, these were not very strong or even absent. However, after the wooden bottom beam was replaced and the wheel systems were revised, the vibrations were so strong that they could be felt in the houses of staff members of the shipping lock. The wheels were slightly twisted during the revision and the rubber was worn, as a result of which the side seal was leaking; probably for this reason, the mechanical damping had been reduced relative to the former situation. During a Delft Hydraulics visit to the gate, a simple Askania manual reader was taken along (principle: absolute displacement transducer, see Part C, Paragraph 7.2). With this, the frequency and the amplitude of the vibrations at a point approximately 1 m above the water level could be measured (see picture Figure d). DIMENSIONS AND WEIGHTS: Sluice width 14 m, gate height 9.5 m. Gate mass 90 ton. Gate plating and side seal at the upstream side. WATER LEVELS: Before lifting may begin, the upstream water level is at N.A.P. +4.5 m and the downstream water level in the lock chamber is approximately N.A.P. During the testing, the downstream water level varied from N.A.P. +1.05 m to 1.4 m. In case of a hydraulic head smaller than 1.9 m, the vibrations stopped. DESIGN: See Figures a, b and c. During the revision the bottom beam was replaced, in which the cross-section of Figure c was realized exactly. It is unknown whether this was also the case in the former situation. OBSERVATIONS: During an opening of 0.13 m to 0.35 m considerable vibrations occurred. These were strongest at 0.25 m. The vibration frequency was between 4.8 and 5.2 Hz. Only the first harmonic bending vibration occurred. Probably the vibrations were initially vertical and only at a later stage horizontal. 187

Delft Hydraulics

The vertical amplitude (some periods) was 0.25 mm and the horizontal amplitude was 5 mm. The small waves in the upper water were initially generated parallel to the gate plating, and the transverse or cross waves were only generated at a later stage, similar to those in Figure d. That may be explained using the present state of knowledge about the generation of these waves (Paragraph 4.6). The parallel waves are generated passively and follow the vibration, while the transverse waves are caused by a self-amplifying mechanism. For this, at first an initial transverse wave needs to be present accidentally, before the amplifying mechanism may occur. CAUSE OF VIBRATIONS: See Paragraph 4.4.4 (theory of the varying discharge coefficient). Due to the surrounding flow related to the added water mass, the leakage discharge is periodically blocked, as a result of which a periodically varying hydraulic head is generated that amplifies the vibration. This type of vibration may occur at a rectangular edge up to a gap width of 2 to 3 times the thickness of the edge. In case of a bevelled edge, the critical opening may be smaller. REMEDIES: By mounting pieces of angle steel of 0.4 m long with an intermediate distance of 1 m, see Figure 3, the vibrations could be neutralized. The idea behind this was: by cancelling the two-dimensional situation, two areas are formed, each with a different critical opening. If one situation is critical, the other operates like a damping force. REFERENCES: Kolkman (1980) and Delft Hydraulics Report S50-1.

Figures a, b and c: a) Vertical cross-section of lock gate. b) Horizontal cross-section of wheel system. c) Cross-section of wooden bottom edge. 188

Delft Hydraulics

Blz. 178 Figure d: Vibration measurements and transverse waves upstream of the gate.

Figure e: Remedy: pieces of angle steel (L-beam) of 40 cm wide and 1 m intermediate distance. dimensions in cm 6.2f FLOATING SECTOR GATES STORM SURGE BARRIER ROTTERDAM WATERWAY (scale model investigation) GENERAL: The sector gates are afloat before closing. In case of extra high tides, they are put into place using a traction engine on the shore that moves a cogged rail across the length of the gate. Afterwards, the gates are sunk down. During the sinking down phase (pre-design) there was a critical situation during which a standing wave movement is generated between the abutments with an amplitude that may increase up to a few meters. As a result of this, the gates also move up and down. This was observed during a 1:60 scale model investigation.

189

Delft Hydraulics

DIMENSIONS AND WEIGHTS: Free navigable width between the abutments 360 m. Gate height 22.5 m. Gate thickness at water level: 8 m. WATER LEVELS: Flow and hydraulic head may be in both directions. Seaward side maximum water level during sinking down phase N.A.P. +3.5 m, corresponding minimal inner water level N.A.P. If the hydraulic head is from inside to outside, then the inner water level is approximately N.A.P. +1 m and the hydraulic head is 1 m. DESIGN: See Figures a and b. Width box at cross-section of the water 8 m and at the bottom 14 m. OBSERVATIONS: A. In case of flow from sea to river, a standing transverse wave is generated at the downstream side between the abutments (first harmonic, period 60 sec.). In the most unfavourable situation the (double) amplitude of the wave may increase up to 6 to 8 m. In that case the gates move up and down with an amplitude of 4 to 6 m; all this with a hydraulic head of 3 m. The growth of the standing wave was exponential. B. In case of flow from river to sea a standing transverse wave (second harmonic, period 30 to 35 sec.) is generated at the downstream side, in which the gates each execute a rotating movement around the centre. Here too, the wave amplitude grew exponentially. In case of an (exaggerated) hydraulic head of 2 m, the wave amplitudes could increase up to 3 to 5 m (crest-trough), the gate movement was slightly less than the wave amplitude. C. The gates may start to oscillate in vertical direction with a period of 10 to 15 sec. In that case, the gates move independently. The period corresponds to the natural immersion period. This happens in case of small lifting openings (0 to 0.5 m) and in case of greater openings (greater than 2.5 m). The oscillations did not always occur and strongly depended on the cross-section of the gate. If the shape was favourable for one particular flow direction, then it was often unfavourable for the other flow direction. CAUSE OF VIBRATIONS: As far as oscillations A and B are concerned, these were instable fluid resonances. These occur when the gate (partially) floats on the downstream water. If the downstream water level increases locally, then the discharge increases due to the buoyancy of the gate. That is exactly the right condition for the standing wave to start generating resonant rise. The fact that the first harmonic transverse wave occurs in case of flow from the sea, and the second harmonic occurs in case of flow from the river, is a result of the discharge distribution beneath the gates. From the sea the water especially flows to the abutments, so that a wave is generated with a great amplitude in any event at that location. Flowing from the river the discharge concentrates toward the middle; in the lower reach a wave will be generated in any event that will also have a great amplitude in the middle. Oscillation C relates to instable flow beneath the gate. Either the flow separates at the upstream cornered plating and attaches or not at the bottom edge, or the flow separates at the bottom edge and attaches or not further down the floating body.

190

Delft Hydraulics

REMEDIES: By shifting the bottom edge in downstream direction, it is achieved that the gate floats less on the downstream water. By bevelling off the upstream part of the lower box, the water pressure is still more coupled to the upper water. The problem in the design was the requirement of a stable situation for both flow directions. The eventual design only has critical situations that are outside the area of the expected circumstances. REFERENCES: Bakker et al (1991) (Delft Hydraulics Publication 462), Delft Hydraulics Reports Q958, Q969, Q1140, Q1190, Q1271 and Q1278.

Figures a and b: a) Plan view. b) Cross-section of a gate

Figure c: Standing wave movement in case of flow from sea to river.

191

Delft Hydraulics

6.2g WHEELED GATE DISCHARGE SLUICE VOLKERAK (scale model investigation) GENERAL: This example only relates to vibrations that were found at a nearly closed gate. The vibrations were found in a continuous-elastic model, scale 1:25. It concerns a lifting gate with double plating. The gates were guided by wheels (in the design phase of that moment). In some tests the gate was suspended in horizontal leaf springs to avoid wheel friction (in case of type I vibrations). DIMENSIONS AND WEIGHTS: Span 30 m, gate height 7.5 m and gate thickness 3 m. Weight 120 tonf. See Figures a and b. Stiffness lifting structure per side 2.5 106 N/m. Bending stiffness horizontal 30 to 50 109 Nm2. WATER LEVELS: Downstream water depth 0 and 3.25 m in most tests. Upstream water depth varying up to 7.5 m. Water depth 0 is a situation that does not occur in reality; it means free runoff. DESIGN: See Figures a and b. OBSERVATIONS: Type I vibrations, unloaded upper wheels: (see for shape bottom edge Figure c). Initially vertical vibrations 3 to 3.5 Hz started; at a later stage sometimes changing into strong horizontal vibrations. The vibrations started at h2 = 0 and a hydraulic head of 2.75 m. In case of a situation with h2 = 1.75 m and h1 = 3.25 m also strong vibrations occurred. These situations approximately correspond to a net unloaded condition of the upper wheel, as a result of which the gate is capable of moving freely. Vibrations also occurred in case of a sharp edge, at h2 = 0, and a hydraulic head of 3.5 m. Type II vibrations: in case of h2 = 3.25 and a hydraulic head of 3.25 m the vibrations started and became stronger in case of an increasing hydraulic head. The thickness of the bottom edge was varied (Figure d). At d = 300 mm the vibrations already started at a hydraulic head of 1.25 m. In case of an edge thickness of 0.25 and 75 mm, and with h2 = 3.25 m, no vibrations occurred. For all vibrations in the upper reach small waves were observed, initially moving away from the gate; at a later stage these changed into a kind of small standing transverse waves. The pictures of Figure A4.28 derive from this investigation. CAUSE OF VIBRATIONS: In case of type I the disappearance of the upper wheel load means that there is no stiffness anymore. The gate may rattle. In case of type II: see Paragraph 4.4.4 (theory of the varying discharge coefficient). Due to the fact that the surrounding flow is related to the added water mass, the leakage discharge is periodically blocked, as a result of which a periodically varying hydraulic head is generated that amplifies the vibration. This type of vibration may occur in case of a rectangular edge, up to a gap width of approximately twice the thickness of the edge.

192

Delft Hydraulics

REMEDIES: Ensure that the gate does not experience a long lasting critical condition. In case of a sharp edge it is possible to maintain a leakage gap that is not too big. The gate may also be provided with a seal, provided it has a very sharp shape (see Example 6.1d). Take care that during opening and closing the critical situation passes quickly. Moreover, in the model a little bit of friction on the side was enough to cause the vibration to stop. For the final design no wheels were used, but sliding supports. With these, type I cannot occur. Furthermore, in case of a thickness of the lower part of the seal of 10 mm, it was decided to maintain a gap of at least 30 mm. Also the stiffening girder in Figure c was placed higher up. REFERENCES: Kolkman (1980) and Delft Hydraulics Report M1129-II.

Figure a: Plan view of gate (continuous-elastic model) from above.

Figure b: Cross-section of the gate.

193

Delft Hydraulics

Figures c and d: c) Original bottom edge used. d) The edge with systematically varied thickness of bottom part of seal. 6.2h FLOATING FLAP GATES STORM SURGE BARRIER VENICE (scale model investigation) GENERAL: The storm surge barrier in Venice exists to retain a hydraulic head of maximum 1.5 m. The barrier will be constructed with a number of 20 m wide flap gates that have been sunk to the bottom in a normal way. In case of high tide, air is interposed and next the flap gates arise. The flap gates turn aloft because of the hinges in the bottom and they are then positioned at an angle of 50. The barrier allows 85 to 90% of the waves to pass through. Research has been carried out at Delft Hydraulics on a section model with three flap gates (two whole ones and one half one at the drain walls), and on an overview model. Only the first research is discussed here. DESIGN AND WEIGHTS: Figure a shows a cross-section at the deepest opening (Malomocco). There, 20 flap gates are positioned side by side. WATER LEVELS: In closed position the maximum water level at the outside is +2 m. Hydraulic head is maximum 1.5 m. Wave heights up to 3 m crest-trough. OBSERVATIONS: The flap gates all move simultaneously with the wave period. Next to this, in case of perpendicular incoming regular waves, parasitic movements occur, during which the flap gates move out-of-phase. The period of the parasitic movement is twice the period of the incoming wave. The movements may be different in shape (there are different modes). Seen from above, the patterns observed were as in Figure b. The growth of the movement is purely exponential. In case of still water tests it was observed that the periods during which these movements occurred, exactly correspond to a kind of resonance period in which the 194

Delft Hydraulics

damping is very low (dimensionless damping 5 to 9%). The lowest resonance period occurred at a period of the incoming wave of approximately 11.5 sec., in which the period of the parasitic movement was 23 sec. (The lowest period of a standing wave between the laboratory flume walls is approximately 10 sec., so that the real period relative to the flap gate movement is much longer.) As the modes are excited by waves with the half period comparable to the relative resonance period, the response in Figure c varies with the wave period. CAUSE OF THE PARASITIC FLAP GATE OSCILLATIONS: The cause is exactly the same as that which also causes the transverse waves in case of a wave board and horizontally vibrating gates. See Paragraph 4.6. In theory, it may be demonstrated that if there accidentally is a transverse wave with period 2T, the wave board effect (period T) produces a periodically varying discharge, including amongst others, a component with period 2T. This component is proportional to both the velocity of movement of the wave board, Vgs, and the water level variation, z, of the transverse wave. As a result of this, the oscillation of the transverse wave increases more and more. In theory it may also be explained that because of this, the amplitude of the standing transverse wave grows exponentially. At the sea and lagoon side, the transverse waves are always out-of-phase. This is what causes the parasitic flap gate movements. REMEDIES: As in reality waves are always irregular, the phenomenon causes little hindrance. REFERENCES: Jongeling (1993), Jongeling and Kolkman (1995), Delft Hydraulics Reports Q20, Q744 and Q1033.

Figure a: Cross-section of flap gate.

195

Delft Hydraulics

Figure b: Normal flap gate movement mode O and possible parasitic flap gate modes.

Figure c: Response at various wave conditions.

6.3 High-head gates and other culvert gates


6.3a CULVERT GATES SHIPPING LOCKS KRAMMER / KREEKRAK (scale model investigation) GENERAL: For the filling and emptying culverts, gates have been used that seal completely in closed position. A choice was made for a preserving jar seal in one horizontal plane. Because of a relief installation at the wheels, the gate separates from the seating and may then be pulled along. The consequence of this is that the plating is located at the downstream side and 196

Delft Hydraulics

the girders at the upstream side, which causes flow forces in downstream direction. This is why perforations have been made in the girders. Investigations were carried out (in a 1:15 scale model) with a non-elastic gate, but with an elastic attachment, allowing either horizontal or vertical vibrations to be investigated. For this, the model was suspended from a wall frame using springs. This frame was capable of moving up and down with the gate. Care was taken that the frame was completely outside the area of flowing water. DIMENSIONS AND WEIGHTS: Culvert 3 m high and 7 m wide. Gate weight (including wheels and relief installation) 18 tonf. Horizontal stiffness 108 106 N/m (this was a schematic representation of the prototype situation, concerning the horizontal bending stiffness of the gate). WATER LEVELS: Drowned situation. Maximum hydraulic head 3 m. DESIGN: See Figures a, b and c. OBSERVATIONS: When the gate had just been pulled loose from the seating, very strong vibrations occurred, of which the amplitude increased exponentially. As far as could be observed, the amplitude was limited because the gate touched the seating. The horizontal resonance frequency that normally was approximately 6 to 10 Hz, became 2.5 Hz in the nearly closed position, as the entire volume of water in the neighbouring part of the culvert operated as added water mass. CAUSE OF VIBRATIONS: Following from this investigation, at a later stage the bath plug vibration was developed as a theory (Paragraph 4.4.2). In Paragraph 4.4.2 this example has been cited as a verification of the theory. The Ck represents the spring stiffness divided by the hydrodynamic stiffness. This has been determined as follows. The hydrodynamic stiffness equals twice the static load, divided by the gap. The culvert surface area is 21 m2, but the gross gate surface area is approximately 25 m2. In case of a hydraulic head of 3 m and a gap of 0.015 m, the hydrodynamic (negative) stiffness found is 108 N/m, so that Ck becomes approximately 0.9 to 1. The Cm represents the mass of the gate including the added water mass upstream, divided by the water mass in the culvert downstream. This is not easy to determine. The mass of the gate plays a subordinate role here. It is clear that the water may run off much more easily upstream: the first shaft is positioned in close proximity to the gate itself. In any event, the conclusion must be that Cm is smaller than 1. As a result, the relation Ck = Cm + 1 has been roughly verified. It may be clear that such a verification is far from accurate. REMEDIES: In any event, the gate needs to be pulled loose from the seating to such an extent, that the gap remains outside the critical area. It appeared that for this, a gap of 15 mm would already be big enough. It is however important to take the static deflection of the gate and the elasticity of the rubber into account.

197

Delft Hydraulics

REFERENCES: Kolkman (1980) and Delft Hydraulics Report M865-VI.

Figure a: Vertical cross-section of culvert and gate shafts.

Figure b: Vertical cross-section of main gate.

198

Delft Hydraulics

Figure c: Horizontal cross-section of the prototype gate and of the scale model. 6.3b SECTOR GATE IN FILLING CULVERT HIGH-HEAD LOCK (prototype experience and scale model investigation) GENERAL: The dynamic behaviour of a reversed Tainter gate has been investigated in a 1:10 scale model. As the girders are positioned at the upstream side, relatively important suction forces due to flow may be expected. In order to reduce these, space has been reserved between the horizontal stiffening girders and the plating. This type of gate is used, as air may not be sucked in through the shaft, which is a precondition for shipping locks. When the gate started operating, prototype verification measurements were carried out. These especially focused on cavitation. The design of the gate was represented in great detail in the scale model. DIMENSIONS AND WEIGHTS: The culvert had a height of 2.67 m and was square. Gate mass (a polar inertia moment recalculated to a radius of 3.96 m) was 9300 kg. Stiffness suspension in model 5 106, 10 106 and 18 106 N/m (prototype values). WATER LEVELS: Drowned situation. Hydraulic head up to 25 m. DESIGN: See Figure a concerning the situation in the scale model, and Figure b regarding the gate design. Figure c shows the original design of the upper seal. OBSERVATIONS: In case of the original shape (see Figure c) and small openings (< 5%), considerable tangential vibrations occurred. At the weakest suspension spring a (single) amplitude was reached of 0.1 m, and with a small opening the amplitude was limited because it made contact with the bottom. The dynamic force could increase up to four times the gate weight. On the basis of the measured frequencies in dry condition and in water it could be concluded that the added water mass is approximately 700 kg.

199

Delft Hydraulics

CAUSE OF VIBRATIONS: Because of the protruding upper lip, a bath plug vibration (Paragraph 4.4.2) occurs. When closing the gate, the discharge decreases, and because of the inertia of the water in the downstream part of the culvert, the hydraulic head increases temporarily. This hydraulic head influences the upper lip and amplifies the closing motion of the gate. This example is used to verify the theory of the bath plug vibration. This has been discussed in great detail by Kolkman (1976). REMEDIES: Figure d indicates the shape that is recommended on the basis of the scale model investigation. The size of the protruding lip is minimalised and at the bottom of the gap a protruding strip has been mounted that limits the discharge. In case of a very small lifting height of the gate however, this strip is not yet operational and the leakage discharge is determined by the gap at the rubber seal. In prototype it appeared that in case of a very small lifting height (a few cm), the rubber edge was caused to vibrate. This was a very short-lasting phenomenon during the normal lifting of the gate. Nonetheless, at a later stage it was advised to use a rubber profile that is fastened at two sides (see Figure e). REFERENCES: Kolkman (1976, 1980 and 1984).

Figure a: Longitudinal cross-section of the culvert during scale model investigations.

200

Delft Hydraulics

Figure b: Vertical and horizontal cross-section of the sector gate.

Figures c, d and e: c) Original design of the upper lip (seal). d) Applied design. e) Improved shape.

201

Delft Hydraulics

6.4 Valves
6.4a CYLINDER GATE AMERONGEN WEIR (prototype experience) GENERAL: This is not an example of a vibration measurement, but of a gate that was already put out of service after one week, the cause of which was probably due to vibrations. This is to do with the cylinder gate in the Amerongen Weir that serves the fine-tuning of the discharge. The gate structure consists of a hollow core, that protrudes above the water level, around which a cylinder casing may move up and down. A rubber edge closes the gap between casing and core (wear and tear of the rubber may cause problems). When the cylinder is lowered, it rests on a seating that, besides a radius of curvature, constitutes the edge of the vertical pipe as well. The hollow core, together with the winch of the cylinder gate, has been mounted on the concrete pier by way of a girder structure. DIMENSIONS AND WEIGHTS: The inside diameter of the pipe is 3.2 m. The gate height measures 1.6 m. The gate is horizontally approached by the flow from both sides of the piers. By using guiding elements it has been tried to cause the discharge locally to run off radially. Beyond the gate the flow bends downward, which despite the cone at the underside of the core is probably accompanied by a lot of turbulence. WATER LEVELS: Maximum upper water level N.A.P. +6 m. Downstream water level not fully known, but varying between N.A.P. +2.5 m and N.A.P. +4.0 m. The intake culvert is at the same level as the bottom at N.A.P. +2.4 m and with the ceiling at +4.0 m. DESIGN: See Figures a, b and c. The tailrace culvert is in the direction of the main flow of the river. The intake openings are on both sides of the pier. The hollow core has a cone at the underside to guide the flow. OBSERVATIONS: At the occasion of the mounting of the support beams of core and winch, the concrete was slightly damaged and the steel was hammered in. The horizontal beams were fractured. The shores were intact, but the holes of the fastening bolts had become oval-shaped. The gate could no longer be closed due to the distortion of the vertical sliding beams. Also the wheels of the gate were stuck. The gate itself was not damaged. The vertical sliding beams of the gate (that run from the hollow core to the seating of the gate) had been hammered in at 10 and 30 cm from the underside. The core at the bottom side of the cylinder-shaped part had been hammered in. Just before collapsing, the gate produced a strong banging noise.

202

Delft Hydraulics

CAUSE OF VIBRATIONS: The first thing to be observed is that when the gate is in a certain position, the winch and the gate together are suspended elastically from the support beams. In case of small openings, this irrevocably leads to the danger of bath plug vibrations (Paragraph 4.4.2) with possible amplitudes of the size of the lifting opening. Auxiliary causes of dynamic load may be: When closing the gate, underpressures are generated beneath the core, partially because the discharge in the tailrace pipe needs to be slowed down. If the (local) upstream water level is lower than the upper edge of the cylinder gate, then air may be sucked in. The air will expand. The downstream water rushes through. If the gate is closed, then the air is squeezed out and the water returns with a certain velocity. Next, this is stopped due to the presence of the core, as a result of which shock phenomena occur. During a visit, gushers through the leakage gap between core and cylinder were observed as well. When closing the gate, in case of a very small opening, the negative hydrodynamic stiffness may always become greater than the positive structure stiffness. In that case the closing motion rushes through, and an extra great load is generated. REMEDIES: Very stiff suspension (especially of the core). REFERENCES: None.

Figure a: Longitudinal cross-section of the pier at the location of the cylinder gate and the tailrace culvert.

203

Delft Hydraulics

Figure b and c: b) Horizontal cross-section. c) Cross-section. 6.4b GLOBE VALVE (prototype experience) GENERAL: About this example very few data are available, as Delft Hydraulics was only involved indirectly. The globe valve is positioned at the downstream (almost) end of the high-pressure pipe, upstream of the needle valve that regulates the discharge of the Pelton turbine. The globe valve in closed position fully closes: after closing the valve, the rubber edges are inflated with water pressure. DIMENSIONS AND WEIGHTS: No longer traceable, but the diameter of the pipe was approximately 2 m and the pipe length was a number of kilometres. WATER LEVELS: Hydraulic head 500 m. DESIGN: See Figure a. OBSERVATIONS: After closing the valve, but before pressurizing the sealing rings, strong vibrations were generated. The period corresponded to the period of the standing pressure wave in the pipe. The system to inflate the rubber edges had been damaged due to the great pressure oscillations and could no longer be used.

204

Delft Hydraulics

CAUSE OF VIBRATIONS: This concerns an instable fluid oscillation (in the shape of a standing pressure wave). Instability is generated when an increase in the hydraulic head causes a decrease of the discharge of the valve (in this case, as a consequence of the distortion of the rubber sealing rings). In Paragraph 4.5.1 this is discussed in great detail. The stiffness of the rubber (in relation to the hydraulic head), the gap width and the design of the cross-section of the rubber determine whether vibrations occur or not. If, due to an enlargement of the hydraulic head, the leakage gap is closed, this will occur. REMEDIES: The most important aspect is taking care that the rubber cannot grab the flow. REFERENCES: None.

Figure a: Longitudinal cross-section of the valve. 6.4c HOLLOW-JET VALVES (prototype experience) GENERAL: This example has been taken from other sources (Mercer, 1970) and is also included in the ICOLD bulletin Vibrations of hydraulic equipment of dams (ICOLD, 1995). Hollow-jet valves are often used at outlet structures with a somewhat higher prepressure. The valve is placed at the end of a horizontal pipe. In case the valve is open, the jet is spread out in all directions by a cone. The cone is mounted on the pipe with radial blades. The opening between pipe and cone is closed with a cylinder casing that may move over the pipe and the blades. As there is a free runoff, everything is put under strain of tension; there is relatively little load due to turbulence and at the outflow the water is strongly mixed with air (because of which a small stilling basin may be used). A number of them however have collapsed. DIMENSIONS AND WEIGHTS: Diameter 0.75 to 2.5 m. WATER LEVELS: Pre-pressure up to 140 m.

205

Delft Hydraulics

DESIGN: See Figures a and b. OBSERVATIONS: A large number of gates have collapsed in the past. In this, the internal blades were considerably distorted (see Figure d). CAUSE OF COLLAPSE: The cause of collapse is not univocal and is probably different for each gate. Considering the observed damage of the blades, the collapse probably relates to the sharp upstream edge of the blades. Here vibrations may be generated, and in case of deflection the blade grabs the flow, amplifying the initial deflection even more. REMEDIES: Mercer has developed a criterion to test the strength and stiffness. He has taken a kind of Strouhal number (S = fD/V) as a measure, but in this case the frequency is based on the dry resonance frequency of bending vibrations. Although in theory something might be brought against this, he has succeeded reasonably well in bringing the damage cases together into one system. The dimensionless number is defined as (Q/CTD) *steel/Esteel. In this, Q is the maximum discharge, D is the diameter of the valve, T is the wall thickness of the blades and C is a coefficient that depends on the number of blades in the valve. For four blades C = 2.22, for five blades C = 2.35 and for six blades C = 2.48. All valves for which the Mercer number was smaller than 0.115 remained without damage. REFERENCES: Mercer (1970).

Figures a and b: a) Longitudinal cross-section. b) Vertical cross-section

206

Delft Hydraulics

Figure c: Vibration modes (theory).

Figure d: Picture of the blades of a collapsed gate. 6.4d ANTI-CAVITATION VALVE WITH PLUG SEALING (prototype experience and laboratory tests) GENERAL: The anti-cavitation valve concerned is a kind of bath plug that moves up and down inside a cage with little holes. It must be capable of functioning up to a hydraulic head of 2000 m (200 ato). Instead of a normal discharge lead-through, as with a normal plug or bath plug valve, the water runs off through the little holes in the cage casing in the form of small nappes. In Figure a the water runs off downward through the central pipe. Although here too cavitation will be generated, this especially occurs inside the water, without contacting the wall of the

207

Delft Hydraulics

discharge pipe or the cage casing. Moreover, the cavitation is spread out across the entire water mass, making it more cloud-shaped, without large bubbles being generated. The casing is composed of three layers, in which the little holes have been made in a staggered way. By thus letting the water flow through a labyrinth, each nappe is spread out even better. In case the casing has eroded, it may be replaced. DIMENSIONS, DESIGN AND WEIGHTS: For the shape, see Figures a and c1. Diameter lead-through opening 100 mm. Weights and stiffness are unknown. This type of valve is used in varying sizes. WATER LEVELS: Irrelevant. The installation in which the flap gate was mounted, was entirely filled with water. The flap gates were all positioned at the downstream side of a series of pumps. OBSERVATIONS: In case of very small openings during test runs with steam, strong vibrations occurred of 10 to 30 Hz, in which the frequency corresponded to those of standing compression waves in the system. It took approximately five seconds before the vibration was at full intensity. Figure b shows pressure measurements. The exponential growth of the amplitude is clearly visible. CAUSE OF VIBRATIONS: At first it might be the bath plug vibration or an instable standing wave (see Paragraphs 4.4.2 and 4.5.1). A measure for the occurrence or not of these types of vibrations is the relation between the suspension stiffness of the plug and the (negative) sudden flow stiffness. Although the gate is moved by a very stiff system, in case of very small openings the negative sudden flow stiffness is very great as well. The latter has been deduced for a bath plug with straight seating in Paragraph 3.2.3, Equation A3.36, as being twice the stationary force, divided by the gap width; for a tapered seating this would be smaller. REMEDIES: During an investigation it was found that the shape of the seating is significant. Shapes three and four did not generate any more vibrations. In reality no changes have been made, as in case of good operation, small openings may be avoided. REMARK: The fact that the shape of the seating is still significant might point to the fact that the bath plug also makes a transverse movement. In that case the cause is a mechanism similar to the bath plug vibration, but this time comparable to the situation in Example 6.5c (Stoney Weir gates Sambeek); if at a gap the sealing has been placed too much at the upstream side, then reducing the size of the gap causes an underpressure downstream of the narrowest crosssection, which again influences the gate, causing a tendency for the gap to close even further (see Paragraph 4.4.3 and also Paragraph 6.5, Example e). REFERENCES: Cassidy (1990).

208

Delft Hydraulics

Figure a: Cross-section of the valve.

Figure b: Increase of pressure amplitude with time.

Figure c: Various shapes of the tested seating

209

Delft Hydraulics

6.5 Seals and leakage gaps


6.5a SLUICE HARINGVLIET, SEA GATE (prototype experience) GENERAL: At the sea gate of the discharge sluice Haringvliet, a shoe-shaped profile was used, in which the water pressure amplifies the sealing function. Although extensive scale model investigations took place concerning the vibration behaviour in case of using a hemispherical fender profile (as in the case of the river gate as shown in Example 6.2.c), at a later stage it was however decided to use another profile in which a better sealing might be expected. DIMENSIONS AND WEIGHTS: Figure a shows a cross-section of the entire discharge sluice. The total discharge sluice consists of seventeen openings of a 56.5 m span each. Figure b shows the cross-section of the sea gate and a detail of the bottom edge. Dimensions of the bottom edge and damage are not exactly known. WATER LEVELS: During vibrations the hydraulic head was no greater than 2 m. OBSERVATIONS: After the bottom edge was damaged (probably due to the fact that when lowering the gate, there was a piece of iron obstructing it), at the hydraulic head from the seaward side and in case of a closed gate, very strong vibrations were generated. Based on hearsay, the noise was similar to a slowly running diesel engine. Under these circumstances, while the vibration was still occurring, the diver did not dare to inspect the gate. CAUSE OF VIBRATIONS: This is an example of the bath plug vibration (Paragraph 4.4.2). Although the cause is the periodic distortion of the bottom edge, when the discharge through the leakage gap is quickly pinched off, the entire gate has to endure a heavy load. The shape of the edge is such that the edge grabs the flow. Therefore it is possible that during the vibration the edge periodically touches the bottom, amplifying the vibrations even more. REMEDIES: Preferably another bottom edge. At edges that derive part of their sealing function from the water pressure, vibrations are regularly observed. Damages that cause a leakage gap are never completely avoidable. Here, repairing the distorted bottom edge was sufficient. The support seats were slightly lowered, so that the bottom edge is better compressed, reducing the probability of leakages. REFERENCES: Delft Hydraulics Report M754 part I and IV for general information and the vibration measurements that were carried out. No report available concerning this particular vibration.

210

Delft Hydraulics

Figure a: Cross-section of the complete sluice.

Figures b and c: b) Vertical cross-section of the sea gate c) Detail of the bottom edge

211

Delft Hydraulics

6.5b MUSIC NOTE RUBBER SEAL AND DEFLECTED RUBBER SEAL (prototype experience) Example I: GENERAL: Here, two experiences with seals that are designed in such a way that the water pressure provides the sealing function are mentioned briefly. This first example concerns a seal at a flap gate of a lock chamber. DIMENSIONS AND WEIGHTS: Unknown. Data concerning vibrations were obtained during a foreign visit. WATER LEVELS: Maximum hydraulic head 8 m, downstream water depth 3.5 m. DESIGN: See Figure a. OBSERVATIONS: In case of a hydraulic head between 2 and 6 m, again and again vibrations occurred that disappeared in case of an increasing hydraulic head. According to some, the vibrations must have been the cause of the damage and fractures of the trunnion that were observed. CAUSE OF VIBRATIONS: Damage with leakage has been observed. Therefore a bath plug vibration could immediately be generated (Paragraph 4.4.2). If the rubber vibrates strongly and the leakage discharge is periodically pinched off, then the leakage gap also closes again and again. But in case of (periodically) closing the discharge, the entire gate is heavily loaded as well. (Due to the deceleration and acceleration of the discharge, pressure gradients are generated in the upper and lower water). The fact that the vibration during a situation of greater hydraulic head disappeared again, might be due to the fact that the leakage gap is permanently closed then. Example II: GENERAL: This example concerns the use of a bent rubber strip as a side seal. Just like Example I, using the bent rubber strip may be very dangerous for the situation in which there is a leakage discharge. But in case of using it as a sliding side seal, it is hardly possible that it causes any damage. The fastening of an object in the gap between gate and wall is hardly possible. However, wear and tear will occur over time. Experiences of a manufacturer of low-head gates and flap gates confirm the good experiences (from hearsay).

212

Delft Hydraulics

Figure a: Cross-section heel post with sealing strip.

Figure b: Bent rubber flap as side seal. 6.5c STONEY WEIR GATES SAMBEEK (prototype experience) GENERAL: At the Stoney Weir the upper gate regulates the water level of the Maas River in case of small discharges. After renovation of the gates (around 1975), it turned out to be impossible to lower the upper gate more than 1 m, without heavy vibrations occurring (as a result of which there is also 1 m water overflow). Lowering the gate by more than 1.6 m would have been absolutely irresponsible and this situation was unacceptable for the maintenance of the gate. Therefore it was an obvious step to investigate the cause by looking at the overflowing water. Aeration of the nappe however had no effect. DIMENSIONS AND WEIGHTS: Span between the wheels at the upper gate 17.6 m. At the lower gate this is 18.8 m. Net distance between the piers 17 m. In normal cases, the lower gate remains standing on the bottom. Both gates are only pulled out in case of high tide. See for the dimensions of the vertical cross-section Figure a. Stiffness (inertia moment) per horizontal girder: at the bottom gate 0.03 m4 and at the upper gate 0.018 m4. The top also contributes a little: 0.007 m4. With this, the torsion stiffness is also fixed, but is has not been calculated.

213

Delft Hydraulics

WATER LEVELS: See Figure a. Precondition for the design is that the upper gate is capable of sinking to the bottom. The upper water level in that case remains approximately constant. DESIGN: See Figure a. OBSERVATIONS: See for the circumstances under general. The vibrations were only felt with a stick. Frequency approximately 3 Hz, corresponding to the lowest harmonic vibration. In case of lowering 1.6 m, the vibration amplitude (single) was approximately 2.5 mm. The suspension chain vibrated with an amplitude of a few centimetres. CAUSE OF VIBRATIONS: After it turned out that aeration of the overflow did not help, the gap between the gates was further analyzed. Figure b shows the cross-section in detail. This probably is a situation of a bath plug vibration: if the gap is pinched off, an underpressure is generated beneath the narrowest gap and the gap is sucked to such an extent that it closes even further. Because of this, a horizontal vibration is amplified. REMEDIES: By putting a roof gutter (pipe) on top of the lower gate next tot the upper gate, that is pressed onto the gap in between the gates due to the water pressure, it was possible to fully stop the vibration during a visit to the site. As the pipe was capable of rolling, the upper gate could move normally. For the definitive solution a similar solution was considered; only now the pipe moves around a fixed centre. For the vibration naturally it would also have been good to choose the narrowest gap at the bottom, i.e. by mounting the heavy wooden beam at the upper gate. In that case however, the hydrostatic load on the upper gate would have been considerably amplified; that was unacceptable here. REFERENCES: Kolkman (1980).

214

Delft Hydraulics

Figure a: Cross-section of the Stoney Weir gates.

Figures b and c: b) Detail of gap between upper and lower gate. c) Improved situation.

215

Delft Hydraulics

6.5d GAP WITH RECTANGULAR AND ELLIPTIC EDGES (laboratory tests) GENERAL: In order to test the pressure distribution along a lower gate edge, a 2 cm thick steel plate was wedged between the walls of a 1 m wide laboratory flume. Here considerable horizontal plate vibrations were generated unexpectedly. At first the lower edge was rectangular. At a later stage also an elliptic edge has been tested, in which similar vibrations occurred. DIMENSIONS AND WEIGHTS: Steel plate of 2 cm thickness, 1 m wide and 1 m high. WATER LEVELS: Downstream water level 0.1 to 0.8 m and upstream 0.5 to 1 m. Maximum hydraulic head 0.6 m. Little turbulence of approach flow. DESIGN: See Figures a and b. OBSERVATIONS: Critical and non-critical situations: Edge Straight ,, ,, ,, ,, ,, ,, ,, ,, ,, ,, ,, ,, ,, ,, ,, elliptic ,, ,, ,, ,, ,, opening 13.5 mm ,, ,, ,, ,, ,, 20 mm ,, ,, ,, ,, ,, ,, 10 mm 30 mm 140 mm 55 mm ,, ,, ,, ,, ,, lower water hydraulic head 0.1 m 0.4 m 0.3 m 0.2 m 0.3 m 0.4 m 0.3 m 0.6 m 0.5 m 0.2 m 0.5 m 0.4 m 0.3 m 0.2 m 0.3 m 0.4 m 0.3 m 0.6 m 0.5 m 0.2 m 0.5 m 0.4 m 0.8 m 0.1 m 0.8 m 0.2 m variable variable variable variable variable variable 0.6 m 0.2 m 0.2 m 0.2 m 0.4 m 0.4 m 0.4 m 0.4 m 0.4 m 0.6 m 0.4 m 0.6 m vibration yes yes yes yes yes yes yes yes yes yes yes yes yes no no no yes yes yes yes yes yes remarks

relatively strong

weak weak

CAUSE OF VIBRATIONS:

216

Delft Hydraulics

See Paragraph 4.4.4; the theory of the discharge coefficient that varies due to the added water mass flow. As a result of this, the discharge also periodically varies and strong pressure gradients occur in the upper and lower water that amplify the vibration. Especially those edges are sensitive, in which situations may arise in which the separation of the flow or the attachment or not of the flow is instable. At a later stage further laboratory research concerning this type of vibrations was carried out (Jongeling, 1987 and 1988). REMEDIES: Edge shapes that are sufficiently bevelled hardly give any problems. REMARKS: The small waves in the upper water have a wave length of 2 cm, corresponding to those of a standing wave frequency of 11 Hz. The vibrations were also audible, so that the vibration frequency was just over 20 Hz (corresponding to the lowest natural frequency of the plate). See Paragraph 4.6 for the explanation of subharmonic transverse waves. REFERENCES: Kolkman (1980).

Figures a and b: a) Cross-section of the rectangular bottom edge of the plate. b) Cross-section of the (quarter) elliptic bottom edge of the plate.

Figure c: Transverse waves in the upper water during vibration. 217

Delft Hydraulics

6.5e SECTOR GATE WITH THRESHOLD OFFSET (prototype experience) GENERAL: This concerns a sector gate with overflow. When discharge needs to overflow it, the gate sinks below a threshold. The leakage between the gate and the threshold remains limited. When the gate was lowered more than 1.7 m (relative to the retaining situation without overflow), vibrations of the gate plating occurred. This example has been taken from Neilson and Picket (1980). DIMENSIONS AND WEIGHTS: See Figure a for the dimensions of the cross-section. The gate has a span of 33.5 m. Weight 200 tonf. The gap width at the bottom is 19 mm. The gate plating was reinforced with ribs. Dimensions per panel 0.87 x 2.64 m2. Plate thickness 19 mm. Natural frequency in dry condition 68 to 144 Hz. WATER LEVELS: Vibrations occurred at all possible values of the overflow height. In case of flowthrough beneath (raised gate) no vibrations occurred. DESIGN: See Figures a and b. The concrete threshold was somewhat steeper than the incline of the gate plating (the continuous line in Figure b). OBSERVATIONS: Considerable vibrations of the gate plating, frequency 100 Hz at 1.7 m water column pre-pressure. Further varying from 30 to 110 Hz with a displacement amplitude of 0.005 to 1.5 mm, depending on the place and the pre-pressure. CAUSE OF VIBRATIONS: Bath plug vibrations, Paragraph 4.4.2. In case of horizontal vibrations of the plating, the gap width varies. As the narrowest cross-section, indicated by A (see Figure b), is upstream, periodic reduction or enlargement of the gap width, and therefore of the flowthrough discharge, results in a pressure gradient in the rest of the gap that causes pressures that amplify the vibration. Only when the gate has been sufficiently lowered, the vibrations are generated. Reason for this is that with the lowering a sufficiently large load surface area of the part of the gate that is downstream of the gap is created. REMEDIES: See Figure b, the broken line. In zone B filling material has been applied, so that the gap has the same width across the entire length. REFERENCES: Neilson and Picket (1980).

218

Delft Hydraulics

Figure a: Cross-section of gate and threshold.

Figure b: Detail of the leakage gap at the threshold and the measures taken at a later stage to avoid the vibration (dotted line). 6.5f LEAKAGE GAP WITH AND WITHOUT VIBRATION SENSITIVITY (laboratory tests) GENERAL: During a vibration investigation in a flume, a model was suspended by wires. The model was suspended in such a way that the wires only allowed vertical vibration. Wires in transverse direction to the flume had to prevent the sideways movement of the gate model. At one side of the flume these were stiff wires, on the other side springs were included, these springs served to prestress the stiff wires.

219

Delft Hydraulics

DIMENSIONS AND WEIGHTS: Flume 0.5 m wide and 0.7 m deep. The gate model weighed approximately 5 kgf. The natural frequency in vertical direction varied from 2.5 to 9 Hz. WATER LEVELS: Flume depth 0.7 m. Water levels were varied. DESIGN: Leakage gap on the side walls approximately 1 cm. See for the shape of the horizontal cross-section and of the leakage gap Figures a and b. OBSERVATIONS: Unexpected vibrations occurred that were clearly audible, therefore with a frequency much higher than those of the established vertical vibrations. The vibrations were not measured. CAUSE OF VIBRATIONS: These again are bath plug vibrations, Paragraph 4.4.2. Because of the natural flow contraction, the narrowest cross-section of the gap is at the upstream side. If the gate model vibrates in transverse direction, the discharge varies periodically. When narrowing the gap, the discharge decreases, which results in a pressure decrease downstream of the narrowing. This again influences the part of the model that is downstream of the narrowest cross-section. REMEDIES: See Figure c. The narrowest cross-section now is at the downstream side of the gap and in this case even a positive hydrodynamic damping is generated. REFERENCES: Kolkman (1980).

220

Delft Hydraulics

Figures a and b: a) Horizontal cross-section of the gate model. b) Detail of the gap between gate and flume wall.

Figure c: Shape of the gap that turned out to be vibration-free.

6.6 Overflow gates and flap gates


6.6a GATE WITH UPPER FLAP GATE (PROTOTYPE EXPERIENCE) GENERAL: This example is taken from Ogihara and Ueda (1980). It concerns the upper flap; this serves the regulation of the upper water level. Following observed vibrations relating to the overflowing water, design and distances of nappe splitters at the top have been investigated. DIMENSIONS AND WEIGHTS: Span 40 m. The calculated natural frequency of flap and bottom gate is 3.7 to 4.2 Hz. WATER LEVELS: Water depth upstream 3.5 m and 1 m downstream. DESIGN: For the cross-section see Figure a. OBSERVATIONS: Although nappe splitters were used (type I, see Figure b), vibrations regularly occurred with undulations in the water curtain. This happened mainly in the area in which the top of the flap was lowered 0.8 to 1 m (relative to the retaining position). CAUSE OF VIBRATIONS: See Paragraph 4.4.7. Initial periodic air pressure variations in the enclosed air cushion cause such undulations in the water curtain, that the volume of the air cushion varies. This in 221

Delft Hydraulics

turn causes pressure variations in the air cushion, amplifying the undulations again. Also the air pressure variation may generate flap vibrations, which again generate undulations in the water curtain. REMEDIES: Nappe splitters that break up the nappe sufficiently. The problem with some nappe splitters is that in case of a somewhat greater overflow height, they become submerged. Also, shapes that become narrower toward the top are not very effective. Type II nappe splitters turned out to be effective. Initially the regular distance was 3 m and at a later stage alternatingly 3 and 4.5 m. Both sufficed. REFERENCES: Ogihara and Ueda (1980).

Figure a: Cross-section of gate and upper flap.

222

Delft Hydraulics

Figures b and c: b) Nappe splitters. c) Corresponding vibrations. 6.6b BOTTOM-HINGED GATE PROVINCE SOUTH HOLLAND (scale model investigation) GENERAL: The bottom-hinged gate serves to partition water systems. In case of danger of a dike breach it may be prevented that all the water runs off and that thereby all the dikes that have not breached are damaged due to overpressure in the interstitial water in the dike. The flow may be either in the flow direction or counter flow (as in Figure a). In the model, in case of flow, powerful vibrations occurred in situations with quite a lot of overflow and a somewhat higher downstream water, as a result of which the air below the water curtain was sucked along. In this case the air cushion is eventually replaced by water. The use of large nappe splitters is not possible here, because in the normal situation in which the flap is positioned on the bottom, there should not be any obstacles for ships. Therefore, in case of flow, the top has been staggered in height (to avoid two-dimensional situations). Also teeth have been mounted to achieve a better attachment of the flow in case of counter flow (less probability of vibrations). DIMENSIONS AND WEIGHTS: Span 36 and 50 m. Hoisting equipment at the 36 m flap gate one-sided, at the 50 m flap gate two-sided. In the 1:15 model a 12.5 m section was tested, with a polar inertia moment of 0.2 106 kgm2. The stiffness of the suspension was variable. WATER LEVELS: Two vibration situations occurred: Vibration I: upper water +4.4 m, downstream water +1.3 m. Flap gate position 40 to 55. Vibration II: upper water +4.4 m, downstream water +2.7 m. Flap gate position 40 to 55. 223

Delft Hydraulics

Other water level combinations have not been tested. Maximum vibrations occurred at a flap angle of 50. DESIGN: For the cross-section see Figures a and b. OBSERVATIONS: Very strong vibrations, to such an extent that the measuring system needed to be protected at a later stage. Maximum measured moment in the 12.5 m gate section: 1.25 106 Nm (or 125 103 per running meter). CAUSE OF VIBRATIONS: For this, see Paragraph 4.4.7. The phenomenon is comparable to the vibrations in which an air cushion is locked in below the water curtain (see the previous example); only this time the transfer of volume variation from the cushion to the flap gate movement is much stronger, as the enclosed water is incompressible. REMEDIES: Here too, the nappe splitters are effective. These do need to be high, as it concerns a situation with a broad overflowing nappe. On the other hand, a small opening is sufficient; it concerns the fact that the space below the water curtain is filled with air again. Precisely in case of these broad nappes air cushion vibrations do not occur (see Paragraph 4.4.7 under mechanism A). Eventually a choice was made for two nappe splitters near the abutments. When the flap gate is positioned on the bottom, the shape does not protrude above the rest of the flap gate. See Figure c. REFERENCES: Kolkman (1980) and Delft Hydraulics Report M1300.

Figure a: Cross-section of flap gate.

224

Delft Hydraulics

Figure b: Detail of upper edge.

Figure c: The chosen nappe splitter. In the model no further vibrations occurred.

225

Delft Hydraulics

6.6c NAGDONG WEIR (scale model investigation) GENERAL: The weir is situated in an estuary and serves to maintain a sufficiently high level for navigation and irrigation purposes and to prevent salt intrusion. Normally the gate is closed, but for the low discharge, overflow is possible at two of the ten gates. In that case the gate is lowered; due to the shape of the threshold the bottom leakage is minimal. Initially the design of the top was not specially adapted for overflowing; this was the cheapest solution that was first tested in the scale model. See for the cross-section Figure a. It concerned a model with a 10 m wide gate section on a scale of 1:20. Only rotational vibrations have been investigated. DIMENSIONS AND WEIGHTS: The weir consists of 10 gates with a 47.5 m span each. Nappe of the sector gate 15 m. Weight 142 tonf (without the trusses). WATER LEVELS: Reservoir water level normally +1.5 m, extreme +3 m. The top height of the gate is +2 m maximum. Sea level varying from 0.75 to +0.95 m. Bottom stilling basin 7 m. DESIGN: Initially a sharp upper edge, allowing the overflowing nappe to fall freely onto the upper girder. The horizontal girders have been perforated. OBSERVATIONS: At all possible values of the overflow height, vibrations occurred that could be overcome for the major part by a nappe splitter. Figure b shows the nappe splitter that was initially used. Two situations remained that were critical. I: 1.2 to 2 m overflow and a relatively high downstream water level. In that case the nappe is not aerated anymore. II: At a high upper water level (+3 m) and a low downstream water level (in that case the nappe is aerated). CAUSE OF VIBRATIONS: The cause of type I probably is that the degree of flow contraction varies (comparable to what was discussed in Paragraph 4.4.4) due to the added water mass flow, as a result of which the pressure difference across the upper girder increases and decreases with the movement velocity of the gate. In case of type II it was observed during the investigations that the vibrations are related to dynamic load on the bottom girder due to flow instability. By shielding the girder the vibrations stopped. REMEDIES: By providing the upper side of the gate with a rounded cap, with an irregular toothing (each fifth tooth is bigger) at the end and next by also using nappe splitters, the vibrations could be considerably reduced in amplitude (see Figure c). At an established (dry condition) natural frequency of 4 Hz (prototype) the type I single vibration amplitude (expressed as force in the winch) was reduced from 655 kN per 226

Delft Hydraulics

entire gate to 95 kN. For type II this was from 715 to 145 kN. The suction force due to the flow decreased considerably at this solution; sometimes there even was an upward force of up to 70% of the gate weight. REFERENCES: De Jong and Jongeling (1982) and Delft Hydraulics Report M1767.

Figure a: Cross-section of the sector gate.

Figure b: The original top design with first tested nappe splitter. vertical main girder

227

Delft Hydraulics

Figure c: The top design with nappe splitter applied. 6.6d SEGMENT GATES STORM SURGE BARRIER THAMES RIVER (scale model investigation) GENERAL: Although there are not many data available, this example is included to show how, in case of gates with overflow, an instable flow pattern may cause critical situations. For this, observations of a demonstration in a small scale model during a visit of one of the authors (Kolkman) to the hydraulic laboratory of the Imperial College in London have been used. Dr. D. Hardwick gave his permission to publish about these observations. The gates, consisting of a hollow container, have been fastened to the bottom, thus allowing completely free navigation. In case of expected high tides the gates are lifted, during which the flow pattern varies strongly with the position of the gate and with the water levels. The gates are fastened on both sides with a round disc onto the points of rotation. DIMENSIONS AND WEIGHTS: Radius 12.2 m. Gate width between the piers 64.9 m and between the discs 61.9 m. Weight 1300 tonf. WATER LEVELS: Low tide jump 2.83 m, high tide jump +3.69 m and storm-surge level to be retained +6.90 m. DESIGN: See Figure a. The gap between bottom and gate is not fully closed.

228

Delft Hydraulics

OBSERVATIONS: In a small pilot model all possible instabilities have been investigated and the tensile force in the suspension wire has been measured, Figure b. For the dynamic part of the investigation, the tensile wire has been replaced by an elastic wire. During conditions in which the flow pattern was instable or tended toward complete separation or separation while attachment further on, the gate showed a strong rotational movement in case of a low rotational stiffness. CAUSE OF VIBRATIONS: The size of the area between the point of separation and attachment varies with the gate oscillation. As there is underpressure in this area, the turning moment also varies. The greatest degree of instability occurs at an angle of = 40. At this angle also the instability indicator (Paragraph 4.4.6) may be used: if the gate moves up during the oscillation, the discharge is pinched off. This causes an extra hydraulic head, as a result of which the flow is thrown out and also the size of the underpressure zone increases. This amplifies the upward movement, which is an indication of instability. Also the added water mass flow that generates a surrounding flow opposite to the main flow, cooperates to throw the flow further away (similar to what was discussed in Paragraph 4.4.4: the theory of the varying discharge coefficient). REMEDIES: In his publication, dr. Hardwick indicates that by perforating the gate at the top, the local underpressure may be neutralized, as a result of which the instability no longer occurs either. This has not been used: the real rotation stiffness of the hoisting equipment is big enough to prevent these vibrations. REFERENCES: Hardwick (1977).

Figure a: Cross-section of gate (Clark and Tappin, 1977).

229

Delft Hydraulics

Figure b: The pilot model with wire suspension (Hardwick, 1977).

Figures c and d: c) Flow pattern at various gate positions. d) Corresponding turning moment.

6.7 Stop logs


6.7a EMERGENCY STOP LOG BARRIER WITH CYLINDERS (prototype experience and laboratory tests) GENERAL: Cylinders are not suited as stop logs, if they need to be lowered in flow conditions. The periodic separation of vortices results in a very powerful vibration excitation. If the stop log rests on the previous one or on the bottom, upstream of the point of contact an overpressure is generated at the underside (there is no flow there), while at the upper side there is a strong underpressure; the stop log separates again which results in a jumpy behaviour. Cylinders roll. It was however decided to apply these, after laboratory tests showed that a sufficiently large downward force could be generated by using a fin, so that it could easily hang up till just above the previous stop log. To keep the fin in the same place relative to the approach flow, the rolling movement had to be avoided. Therefore a square sealing ring with wheel guidance has been used (Figure a). DIMENSIONS AND WEIGHTS: Cylinders approximately 19 cm, filled with water. Wall thickness 1.5 cm. Span 8 m.

230

Delft Hydraulics

WATER LEVELS: During the tests the initial flow velocities were approximately 3 m/s. DESIGN: See Figure a. OBSERVATIONS: During the prototype tests the stop logs were dislodged, also when the tubes were coupled two-by-two (Figure c). Also considerable vibrations occurred. CAUSE OF SCALE EFFECTS: The forces that operate on a cylinder strongly depend on the Reynolds number (Paragraph 5.1, Figure A5.1) and this relates to the place where the flow separates. In case of the model values of the Reynolds number, the flow separates earlier than in case of the prototype. Therefore in prototype a greater (upward) suction is generated on top of the cylinder. REMEDIES: A wire has been welded to the cylinder at the place where the flow separates in the model. See Figure b. In a flow tunnel the pressure distribution around the cylinder at higher Reynolds numbers has been verified again. Moreover, between the tubes distance blocks have been mounted, so that the pipes do not touch each other, but a leakage gap remains. During the last prototype test, it turned out to be possible to lower the tubes. REFERENCES: Kolkman (1980) and Delft Hydraulics Report S50-5.

Figures a and b: a) Recommended shape, based on tests in a scale model. b) Applied design.

231

Delft Hydraulics

Figure c: Prototype tests.

232

Delft Hydraulics

6.8 Trash racks


GENERAL: Delft Hydraulics itself has no experience with vibrations of trash racks. Syamalarao (1989) discusses a number of cases, in which damage was caused by vibrations. The summary diagram with references is shown below (Tables I and II).

Table I: Damages concerning which vibration measurements were carried out. REFRENCES USED BY SYAMALARAO: 1: Bugl, H. von and H. Jericha, Fragen der Dimensionierung und Ausbildung der Rechenanlagen von den Turbineneinlufen von Flushkraftwerken, sterreichische Ingenieurzeitchriften, Vol. 11 Nr. 3 und 4, 1968. 2: Fortrey, J.W. and R.F. Tiry, Flow-induced transverse vibrations of trashrack bars, Civil engineering, ASCE, May 1972. 11: Neilson, F.M. and E.B. Picket, Corps of Engineers experiences with flow-induced vibrations, In Naudascher and Rockwell (1980) (see Chapter 8). 12: Schol, G.A and P.A. March, Model testing of trashrack at Hiwassee dam, Proc. ASCE Hydr. Div. Conf. on applying research to hydraulic practice, Jackson Missisippi, Ed. P.E. Smith, Aug. 1982. 13: Vanbellingen, R., A. Lejeune, J. Marschal, M. Poels and M. Salhoul, Vibration of screen at La Plata Taille hydro-storage power station in Belgium, BHRA Fluid Engineering conf. on Flow-induced vibrations in fluid engineering, Reading. UK, Sept. 1982. 15: Liess, C., Waldeck-II Auslaufrechen; Ursache der Schwingungsprobleme und Beurteilung des Neuentwurfs, rep. 14936, J.M. Voith Gmbh, Heidenheim, FRG, Apr. 1984. 16: Schlageter, G., Erfahrungen am Maschinenhauseinlaufrechen des Rheinkraftwerkes Albbruck-Dogern, Wasser, Energie, Luft, Vol. 77 nr. 10, 1985. 18: Crandall, S.H., S. Vigrander and P.A. March, Destructive vibration of trashracks due 233

Delft Hydraulics

21:

to fluid-structure interaction, Journal of Engineering for Industry, ASME, Nov. 1975. Todd, R., B.W. Mefford and T.J. Isbester, Mt Elbert trashrack vibration studies, Proc. ASCE Hydr. Div. Conf. on Applying research to hydraulic practice, Ed. P.E. Smith, Jackson Missisippi, USA, Aug. 1982.

REFERENCE: Syamalarao (1989).

234

Delft Hydraulics

Table II: Damages without further investigation. 235

Delft Hydraulics

6.9 Dynamic phenomena due to cavitation and partial aeration


6.9a EMPTYING CULVERT HIGH-HEAD SHIPPING LOCK (prototype experience) GENERAL: In an emptying culvert of a shipping lock an aeration (by way of a grid) was mounted downstream of the gate to avoid cavitation damage of the gate and the culvert wall. DIMENSIONS AND WEIGHTS: Culvert approximately 2 x 2 m2, culvert length downstream of the gate approximately 65 m. WATER LEVELS: The culvert ceiling was approximately 5 m below the lower water. The hydraulic head of the gate was 25 m. DESIGN: See Figure a. OBSERVATIONS: During a transmittage test of the lock, after the lock chamber had first been partially emptied, the gate was closed. A pressure gauge was mounted in the ceiling. After the gate was closed, a pressure impulse of 110 m water column was measured. At the covering slab of the gate water spurted out of the aeration pipe. CAUSE OF DYNAMIC PRESSURE IMPULSE: In case of flow, there is a lower pressure downstream of the gate compared to the downstream water level. This relates to the contracted nappe that disappears again further down and in which the reduced kinetic energy is partially converted into potential energy. The pressure may also become lower than atmospheric; the aeration then starts working. Although it is being tried, using the aeration grid, to delicately distribute the air to such a point that it flows along with the flow, a lot is left behind at the gate because that is the point of lowest pressure. If during the closing of the gate the gate discharge is reduced, then initially the flow in the culvert is not slowed down enough due to the elasticity of the air cushion. Moreover: as an extra low pressure is generated at the gate during closure, extra suction of air occurs. The water column in the culvert downstream of the gate does slow down, because the pressure at the gate is lower than the downstream water pressure. After a while, the water velocity reverses direction and the column suddenly stops at the gate. The magnitude of the pressure again depends on the amount of air that is left in the culvert at that moment. REMEDY: By closing the culvert more slowly, the underpressure (relative to the downstream water pressure) may decrease to such an extent, that no extra air is sucked in anymore. Due to the reduced underpressure, the water column with a lesser velocity returns to the gate. If no air is sucked in because the average pressure is higher than atmospheric, then due to turbulent pressure fluctuations, the instantaneous pressure may still fall under the vapour pressure and cavitation may be generated. This is always an area of concern when studying high-head gates. 236

Delft Hydraulics

REFERENCE: Kolkman (1980).

Figure a: Longitudinal cross-section of emptying culvert. 6.9b CLOSURE OF CULVERT SLUICE GATE TIEL (annex cylinder gate central pier Weir Driel) GENERAL: Next to the shipping locks in Tiel there is a culvert sluice gate. The upstream operating gate has been implemented with the two gates positioned directly behind each other. The custom was that both gates were moved simultaneously as a way of verifying that both gates indeed function and to prevent that an unused gate would be out of operation for too long. After a few years of functioning, a large pressure impulse occurred, which caused damage. Consequently, several calculations have been carried out. DIMENSIONS AND WEIGHTS: For the intake structure see Figure a. The gates are positioned upstream at a 450 m long culvert. Culvert height 3.65 m. Ceiling at N.A.P. +0.7 m. At the location of the intake structure the culvert is higher (ceiling at N.A.P. +2.25 m). WATER LEVELS: On the day of the pressure impulse the downstream water (the Betuwe reach) was at N.A.P. +3.68 m. (In special circumstances this might even be 20 cm lower). The culvert ceiling at the location of the gates is at N.A.P. +2.25 m. The water level of the Waal River was at N.A.P. +6.30 m on that particular day. DESIGN: See Figure a for the vertical cross-section of the intake structure. OBSERVATIONS: During the closure at one time a heavy noise was heard, after which it turned out that a concrete sealing plate of the emergency gate shaft had been lifted and the surrounding pavement had been damaged. According to eye-witnesses the noise was accompanied by a water fountain several meters high. CAUSE OF THE PRESSURE IMPULSE: 237

Delft Hydraulics

Behind the gates an underpressure that is too great has been generated, as a result of which air was sucked in. See also Example 6.9a. The pressure downstream of the gate may become lower than what corresponds to the Betuwe reach. First of all, locally there are extra high velocities beneath the gate, which coincides with a low pressure. Next, when closing the gate, there is also a slowing down of a long column of water. An additional unfavourable factor here is the fact that two gates, that are positioned behind each other, are being used simultaneously. The contracted jet beneath the upstream gate then widens in case of a gate position of around 50% in two steps; first the jet attaches at the edge of the downstream gate and next the discharge spreads out across the entire crosssection of the culvert. As a result, the losses are smaller and the pressure between the gate is extra low. REMEDIES: The culvert is actually positioned too high up. Taking this as a starting point, it is advisable to close the gate more slowly and to use the downstream gate. Moreover, it might be considered to close the emergency recess, so that no more air may be sucked in. REMARKS: Also at the cylinder gate in the central pier of the Driel Weir (situation fully comparable to Example 6.4a concerning the cylinder gates in the pier of the Amerongen Weir), during a visit it was observed that when closing the gate quickly, there was a sudden spurt between the upper edge of the gate and the centre of the cylinder gate (a Delft Hydraulics staff member was soaked through!). At that moment, the upper edge of the gate protruded above the upper water. Here too, the cause is the same: underpressure downstream of the gate during slowing down of the water column downstream, during the reducing of the discharge. This results in air being sucked in, and the water column downstream spurts through and then returns. When the water column makes contact the gate again, a relatively great overpressure is generated. REFERENCE Culvert Sluice Gate Tiel, Delft Hydraulics Report M1072.

Figure a: Vertical cross-section of the intake side of the culvert with gate house. 238

Delft Hydraulics

6.9c GATES FILLING CULVERT MAASBRACHT (cavitation) (prototype experience) GENERAL: The shipping lock Maasbracht has a hydraulic head of maximum 12.25 m. The sluice is filled through 2 filling culverts, mounted in a compact upper hydraulic head. See Figure a. The design of the culvert is such, that the gates experience an irregular approach flow; the water will also be very turbulent.

Figure a: Vertical and horizontal cross-sections of the upper hydraulic head. DIMENSIONS AND WEIGHTS: The culvert cross-section at the location of the winch is: width 2.5 m, height 3.35 m.

239

Delft Hydraulics

WATER LEVELS: At maximum hydraulic head N.A.P. +32.65 respectively N.A.P. +20.40 m. DESIGN: The intake openings are positioned in the wall of the gate chambers. They deflect downward and next run horizontally (in the direction of the upper reach) with the bottom height, corresponding to the lock chamber bottom. The gate has been mounted in this horizontal part, as well as a shaft accommodating an emergency gate. Next, both culverts deflect in such a way, that both outlet openings end up being opposite of each other. The jets from each of the outlet openings collide with splitter beams and subsequently with each other. The discharge enters the lock chamber through the stilling basin. OBSERVATIONS: The gates function reasonably well, but it is noisy, the wheels and the rail tracks experience a lot of wear and tear and the fouling of algae in the proximity of the gates has eroded. The latter is not a serious issue, but altogether these factors are an indication of the fact that cavitation occurs.

Figure b: Limit values i of Thoma number , for weak and strong cavitation. CAUSE OF DAMAGE AND NOISE: It appears that the altitude of the culvert is such, that cavitation at the gates may be expected. If Figure b is taken as a reference, elaborated following Brandao and Menezes (1969), and based on research by Sograh at Grenoble, it appears that halfway the height of 240

Delft Hydraulics

the culvert the pre-pressure relative to the downstream water is 1.7 m. When taking into account the fact that in most critical flow conditions the lock chamber is already partially filled (say 2 m), but that in case of a remaining hydraulic head of 10 m there is a local underpressure directly downstream of the gate of 5 m (these factors naturally necessitate a more extensive consideration than may be accomplished here), and assuming that the vapour pressure is 9.7 relative to the atmospheric pressure, then it may be found that relative to the vapour pressure there is an overpressure of: (1.7 + 2 5 + 9.7) = 8.4 m The velocity in the contracted jet is coupled to the hydraulic head, reduced by the culvert resistance, but increased with the local underpressure downstream of the gate (where the nappe is maximally contracted). Therefore, in this case the velocity height in the contraction is taken to be 10 to 12. This results in:

(p

pdamp ) / g u / 2g
2 0

8.4 8.7 = 0.84 0.7 10 12

This value is so low because, due to Figure b, it has been concluded that there is rather strong cavitation. Through the emergency gate shaft however some air will be sucked in, which will reduce the cavitation erosion. On the other hand, air in a culvert by itself also generates instable phenomena. REMEDIES: On the basis of the existing design it seems best to regulate the filling for a while by using a small gate opening, and only pulling the gate higher up when the lock chamber is already partially filled. In case of a small gate opening the cavitation is still situated in the free boundary layer between jet and surface eddy. In this, the flow remains stable. The velocities of the jet however are high (approximately 16 m/s), but still below the value at which (provided there is cavitation) the concrete starts to erode (according to Kenn and Garrod, 1981, 30 m/s, but according to Delft Hydraulics experiences, 25 m/s). REFERENCE: Investigation filling system Delft Hydraulics Report M627; the experience with cavitation has not been reported.

241

Delft Hydraulics

242

Delft Hydraulics

RECOMMENDATIONS FOR THE PREVENTION AND COMBAT OF VIBRATIONS 7.1 General

There are no panacea for the design and construction of vibration-free structures in flowing fluid. It is however possible, by using the knowledge presented in the chapters above, to prevent design faults. Naturally, any design of a gate or a trash rack will include many compromises due to the various requirements for each. Below are a few remarks on this: The requirements of large-scale structures certainly increases the danger of vibration, because cost-effectiveness goes hand in hand with a relatively reduced stiffness and lower natural frequencies. The extra damping that might be desirable to prevent vibrations from occurring, conflicts with the requirement for gates to be capable of opening and closing smoothly. The best location for a bottom edge, as far as vibrations are concerned, gives rise to a sealing around the gate that is not in the same plane and this results in a complex geometry of the sealing and in clearance problems. Also the stiffness of the rubber in corners is often relatively great, as a result of which the compression of the sealing at other places may end up being insufficient. The requirement that in case of engine failure, the gate should be capable of closing due to the water pressure, results in extra risk of vibration. But also the requirement of being capable of functioning vibration-free, when focusing completely on safety, possibly results in a structure that is too heavy and unmanageable. Using scale model investigations however, it has been possible in many cases to achieve a vibration-free design by applying detailed design changes for structures about which there was a lot of insecurity concerning their vibration behaviour. However, even if the design is not optimal with regard to the possible dynamic behaviour, this does not always mean that vibrations will occur. Perhaps accidentally there is a lot of damping or friction. As a result however, it may be that for structures that appear to be the same, one does and the other does not generate any vibration. The replacement of riveted structures by a similar structure that has been welded, may sometimes result in unexpected vibrations as a consequence of the lower damping. Vibrations that are observed in reality sometimes only occur under special conditions. If those conditions are sufficiently understood on the basis of measurements or experience, then good operational management might possibly prevent damage due to vibration, and the structure does not need to be changed. Over the past few years however, there has been a trend toward remote control operation, in which one has to be sure that no unwanted vibrations occur. If vibrations are observed in a structure, in case of a correct analysis of the cause, the remedy may often be accomplished using simple methods. Along with the examples of Chapter 6 the following may be mentioned: 6.6c, the weir gate in Japan, where another nappe splitter was used for aeration, 6.5c concerning the Stoney Weir gates at Sambeek, where a rolling sealing was used and furthermore 6.6f, where a small change of the shape of the leakage gap sufficed to stop a vibration. 243

Delft Hydraulics

In some cases vibrations may be acceptable. The biggest danger of vibrations often only concerns the large number of load alternations, which may result in fatigue of the structure material. If the critical situation occurs very little or if the amplitudes are small, then vibrations are probably not normative for the life span of the structure. However, vibration noise may also be a nuisance, and it attracts so much attention that it is associated with a badly designed structure. If the structure is still in its design phase, the strategy is as follows: 1. 2. First of all, care needs to be taken that self-exciting vibrations of the structure or fluid oscillations are avoided, as these vibrations may be very strong. This especially concerns the design and the location of the point of rotation, wheels and so on. If there is no self-excitation, then the remaining excitation may still be very strong. This especially applies to round and elliptic rods and cylinders. Angular shapes are safer on this point, although there too, self-excitation is possible. Shapes at which the flow separates and attaches again further on are also unfavourable. Finally, it needs to be verified what the frequencies are of possible excitation sources of the chosen design. The stiffness of the structure needs to be adapted on that basis. Thickening of rods has a twofold effect: the stiffness is greater and the excitation frequency decreases (at a given Strouhal number, based on the rod thickness). The natural frequency of the structure, taking into account the added water mass, always needs to be above the excitation frequency. Extra damping is always favourable for safety, but this is difficult to realize. Extra dampers are especially used when vibrations have already been observed and are rarely included in the initial design. When using guy ropes in case of high chimneys, towers and bridges sometimes dampers are mounted.

3.

4.

The following paragraphs will discuss what conclusions regarding the design may already be drawn immediately, when using the knowledge of the chapters above.

7.2 Overall design 7.2.1 Gates


Design first of all involves the avoidance of movement-induced excitation (selfexcitation). For gates, a good tool for this is the instability indicator as introduced in Paragraph 4.4.6. In reality this means that if the water at small openings generates a closing force, there is a probability of vibration. For the sake of safety it needs to be assumed that in case of closure the discharge does not decrease, so that closure always temporarily results in a greater hydraulic head. Figure 7.1 shows a number of examples, in which the design with closing force has caused problems.

244

Delft Hydraulics

Figure A7.1: Examples of gates with closing force. At smaller openings the shape of the remaining gap is very important for the occurrence or not of the closing force. The location of the greatest flow contraction is best situated downstream, the closing force is then neutralized and even reverses its sign (see as an example Figure 7.2).

Figure A7.2: Gap shape with no vibration. Many vibrations observed in reality were vibrations at small gaps; whenever possible these need to be avoided. Naturally, opening and closure of a gate or valve always causes a small opening for a short while. It is important that the gate does not remain in the critical situation for long. At the side of the gates, gaps may be avoided by using sliding sealing strips. Freezing in may be an argument not to use these. In Paragraph 7.3 seals are discussed in more detail. If side gaps cannot be avoided, then it may also be best that the sealing edge is situated at the downstream side of the gap (Figure 7.2). In case of wheel gates in tapered recesses side gaps are always caused when the gate is lifted. Moreover, for these cases no examples of vibrations are known; probably the stiffness of the gate in transverse direction is big enough. Another important point concerning the avoidance of vibrations is the prevention of instable flow. This concerns the shape of the underside of the gate, where a not well-defined separation of flow is caused, see Figure 7.3. This also applies to gates that overflow at the top. 245

Delft Hydraulics

Figure A7.3: Examples of gate edges with instable separation of flow. A sharp or thin edge is always preferable, but it is not always possible with regard to the sealing wanted. A narrow edge means that the flow, at the location where the velocities are high, only has a small load surface area, so that the dynamic excitation forces are limited as well. Moreover, as the critical opening width is coupled to the edge thickness, a narrow edge only offers a small area of lifting heights where the situation may become critical. In that case there is only a small probability that the gate remains standing in exactly that position for a longer period of time. Critical for possible vibrations (see Paragraph 4.4.4) is an opening that is smaller than 2 to 3 times the plate thickness, and even more so in case of rounded edges.

Figure A7.4: Examples of alternating flow separation and attachment (unfavourable situation). Also the possibility of flow that separates, and attaches again further down at the gate edge, may generate instabilities, see Figure 7.4. Therefore a sharp edge that has been reinforced at the downstream side should not have a stiffening girder within the area of 45. If the bottom edge itself has a certain thickness, then it should preferably be shaped such that the flow separation is without uncertainty situated downstream, see Figure 7.5. This also appears to apply to floating doors and gates. 246

Delft Hydraulics

Figure A7.5: Relatively favourable shapes of gate edges. Next to the design of the gate, other factors that relate to the design also play a role, such as creating a free runoff downstream of the gate. If this concerns a culvert gate, then there needs to be ample aeration, preferably in combination with a sloping bottom. For this reason it may be preferable to place the gate or valve downstream of a culvert (Figure A7.6).

Figure A7.6: A gate with free runoff. Good aeration may also be important for stabilizing the flow, in case of a drowned outflow of the gate. In case cavitation may be expected, it is often recommended not to use recesses. In that case a sector gate is the obvious choice. Cavitation, moreover, is an aspect that has hardly been dealt with in this report. In case of a structure with a hydraulic head of 10 m or more, the possibility of cavitation needs to be recognized (see Paragraph 6.9, Example c). If wheel gates are used, then care must be taken that the wheels remain constantly loaded across the entire trajectory of gate positions. This too may be influenced by the design 247

Delft Hydraulics

and the right positioning of the wheels. Also the rattling of the gates that are lifted into the shaft should preferably not be permitted.

7.2.2 Rods and trash racks


Right from the start, some shapes are unfavourable, because the surrounding flow is instable or because galloping (see Paragraph 5.4.1) may have been generated. These shapes are: circular cylinders (though often used); rectangular beams or I-beams with 0.25 < L/B < 2 that are positioned at right angles to the flow (L is a measure in flow direction and B is a measure perpendicular to the flow direction); shapes that are similar to the wing of an airplane, therefore streamline-shaped with a rounded shape at the front. In case of an approach flow at right angles there are no problems, but when flow threatens to separate at the front because of transverse approach flow, then a profile like that will experience a very heavy dynamic load. A number of examples are shown in Figure A7.7. If strong excitation is expected at circular cylinders, the following may be considered: shrouds (perforated casing), as a result of which the velocity gradients between the flow and the wake are reduced, preventing the vortex trail to be generated; flags, preferably stiff (separates the vortex trail in two half vortex trails that are symmetrical and therefore do not result in excitation in transverse direction). The flag must be capable of rotating, otherwise galloping may be caused again in case of transverse approach flow; spoilers (small fins that interfere with the two-dimensional character of the vortex trail); single or double spiral (same effect as spoilers).

For a number of these measures (shroud and spiral) the flow resistance of the rod may increase considerably.

Figure A7.7: Examples of interference elements in case of cylindrical rods. 248

Delft Hydraulics

It may also be attempted to prevent a purely two-dimensional vortex trail at cylinders by: a cross-section that varies across the length; local interference elements; torsion of the profile.

At trash racks that are always constructed with strips and rods it is preferable to ensure that the approach flow is more or less at right angles, without local flow concentrations. This also means that the grid should not be clogged up locally. Short or rather long rod profiles are preferable for trash racks. Special care must be taken to ensure that the flow either separates immediately and does not attach again or that the flow attaches again anyway. In the area between those points, galloping may be generated. Thang and Naudascher (1991) recommend using rectangular profiles with the length e (in flow direction) greater than six times the thickness of the strip and not to use bevelling of the nose of the rods. This also means that the trash rack needs to be cleaned on a regular basis. In situations in which the flow may occur in both directions, it needs to be taken into account that the flow velocities for both flow directions may be very different, even when the discharge is the same. In case of a narrowing culvert the flow remains more or less equally divided across the cross-section; in case of a widening profile separation of the flow occurs very quickly, the velocity profile is irregular and there is a strong turbulence. As far as known, in the Netherlands no vibration problems have occurred at trash racks.

7.3 Gate edges and seals


A sealing edge must meet a number of requirements at the same time: when the gate is closed it needs to prevent leakage; when the gate moves there should not be too much friction; sliding of the sealing strip should not cause too much wear and tear; it must be capable of handling the force of the hydraulic head and in some cases, in closed position, the edge must also be capable of supporting the weight of the gate. Moreover, the edge needs to be wearresistant against sand in the water, and the rubber needs to be sunlight-proof. With regard to gate vibrations it is important to distinguish between the sealing strips that are used at the underside, at the side and, in case of culvert gates, also at the upper side. a Seals at the underside of a gate: Here, sharp gate edges are preferable, see Example 6.1d for possible solutions for the rubber sealing strip. Sealing strips that have an extra sealing effect due to water pressure (as an example the music note profile) are unfavourable with regard to vibrations, as long as they do not seal fully. This also follows from the use of the instability indicator of Paragraph 4.4.6. Especially when, due to damage of the gate locally, in case of a closed position of the gate, there is leakage, very strong vibrations may occur for a protracted period of time. This has been observed a number of times. As with the visor gates at the Hagestein Weir (Example 6.2a), it is unfavourable to use a hemispherical fender profile; in case of gap flow this causes an instable flow because

249

Delft Hydraulics

the point of separation of the flow is not clearly fixed. This appears to generate horizontal vibrations in the plating.

Figure A7.8: Examples of sliding sealing strips. b Side seals: Here, sliding strips with a music note profile or flat bent strips that have an extra sealing effect when the pressure difference across the edge is greater, are fully in order. In case of sliding side seals there is hardly any risk of damage. It needs to be prevented as much as possible that the side seal leaks, first of all because this again concerns a small leakage gap that may be critical, and secondly because dirt may have gathered, which in case of further closure would damage the seal or edge, which in turn would cause another leakage gap. Wear and tear does not need to be of great concern in this sliding direction, provided there is a sufficiently smooth surface and the right kind of rubber is used. Sometimes sliding strips are used with a vulcanized sliding strip (Teflon or nylon with graphite). Seals at the upper side of a gate: There always is a tendency to choose a rubber strip as the upper sealing, mounted on the gate, that seals on a vertical plane relative to the flow-through shaft. Thus a sealing is achieved regardless whether the gate is positioned a few centimetres higher or lower. The rubber then slides a short distance along the vertical plane. This however is not satisfactory, because the sliding direction perpendicular to the strip causes the rubber to crease. The best solution here is a bulb profile of a rather tough kind of rubber that, after opening of the gate, separates after a short distance. See also what has been proposed in Example 6.3b. The profile has two ears, so that it may be fastened on two sides. The distance across which the profile needs to slide is very limited. As soon as a certain gap width has been reached, there is no hydraulic head above the rubber strip anymore, because the metal adjusting strip that protrudes from the concrete limits the discharge. For this too, rubber profiles have been developed with composite materials to obtain better properties. The softer rubber serves for compression and sealing, and the sliding strip serves to reduce the friction.

7.4 Stiffness
There are two reasons to focus on stiffness. 250 At flap gates and valves, the critical gate opening, multiplied with the stiffness of the suspension of the flap gate relative to the flow force, is a measure for the occurrence

Delft Hydraulics

or not of so-called bath plug vibrations. This will be further discussed below. Other types of self-excitation are coupled to the stiffness as well. This is usually expressed in a minimum value of the reduced frequency (a kind of Strouhal number, but in this case based on the resonance frequency). The greater the stiffness, the smaller the area of gap width that is critical. The resonance frequency must be higher than the dominant flow excitation.

As far as self-exciting vibrations are concerned, it always holds that: the stiffer, the better. Often a Strouhal number or critical stiffness may be indicated to avoid self-exciting vibrations. One exception is situations in which the excitation is concentrated on specific frequencies. In those cases it may be the case that by increasing the stiffness, the frequency actually becomes critical. However, even then it may not be safe to keep the resonance frequency below the critical frequency. An example of a situation with a number of discrete frequencies is the overflowing nappe, in which the overflow height needs to contain an exact number of undulations. It may happen that the overflowing nappe vibrates, in which especially the enclosed air cushion operates elastically, without the gate vibration itself having any influence on this. In that case the nappe is an external excitation source for the gate. The stiffness of the gate or the gate suspension may also be an additional influence. Excitation that is generated by instable flow, such as in situations with separating flow that attaches again downstream, often includes a number of discrete frequencies as well. The limited length of the free boundary layer must again contain an exact number of wave lengths, before vibrations may be generated. The critical frequencies in this case may be expressed by a number of discrete values of the Strouhal number. An example of this is Figure A4.3, together with Figure A4.4 in Paragraph 4.3. To illustrate the importance of stiffness, two cases that relate to self-excitation (from Chapter 4) are quantified in more detail in the following examples, a and b. Next, a third case is discussed, taken from other sources, in which a stiffness criterion applies as well. The quantitative elaboration only serves to get a sense of the order of magnitude. A. BATH PLUG VIBRATION The relation found in Paragraph 4.4.2 was that, in order to prevent vibrations, the stiffness and mass factor need to have the following relation: Ck > 1 + C m (A4.26)

This relation in reality means that only small gate openings are critical. Besides the factor (Cm + 1), so when Ck = 1, it appears that the critical opening may be calculated by dividing twice the static load by the spring stiffness, of which the bath plug has been suspended (see Equation A4.24). In case the Cm is greater than zero, this critical opening becomes greater proportional to Cm + 1. Roughly, Cm is the mass of the bath plug (plus added water mass upstream of the bath plug, and following Equation A3.36 this is 4R3) divided by the water mass present in the pipe. The longer the pipe, the less significant this factor is.

251

Delft Hydraulics

Calculation example: Tube with diameter 1 m, length 10 m, mass plug 500 kg, spring stiffness fastening 106 N/m. Diameter plug 1.02 m. Hydraulic head 10 m. Upstream water mass (with R = 0.51 m) (Equation A3.32): 4R3 = 530 kg Water mass in the pipe (with D = 1 m): (D/4)L = 8000 kg Static load (D = 1,02 m): 1000 9.81 10 (1.02/4) = 63 000 N 1 + Cm = 1 + (500+530)/8000 = 1.13 Critical opening = 1,13 x (2 x 63000)/1 000 000 = 0.14 m From zero to 0.14 m opening there is danger of vibration. For a normal procedure in case of relatively fast opening and closure it seems just about acceptable, but in this particular area the gate may certainly not be used as a regulating valve. If the calculation example concerned the gate of the Kreekrak sluices (Example 5.3a), then the 0.14 m would relate to the minimum required stroke of the relief installation (on top of which there is the outward spring of the rubber edge) and in that case this size would be too great. B. VERTICAL GATE VIBRATIONS This calculation example takes the vibration mechanism as described in Paragraph 4.4.3 as a starting point. For these vertical gate vibrations it holds that the degree of self-excitation decreases with an increasing vibration frequency, see Figure A4.14. This relates to the inertia of the water flowing through beneath. In case of higher frequencies the discharge no longer follows the periodically varying discharge capacity of the opening, as a result of which fact the element that causes selfexcitation falls away. The so-called turnover frequency, in which the (negative) damping changes into a (negative) flow stiffness, has been calculated in Appendix I. In Equation A4.41 an analytical expression has been presented for the gate in open water:
S= f 1 = 2 g H 2 mCi

(A4.41)

m = discharge coefficient of the gate opening, Ci = the coefficient for flow inertia (see further Equation A4.33 together with Figure A3.13), f = vibration frequency, = gate opening and H = hydraulic head. Calculation example: Lets take a rather low value for the flow inertia and for the discharge coefficient: Ci = 6 (in reality this is between 6 and 10), m = 0.5. The part on the right now becomes 0.05. In case of a water velocity of 8 m/s (well over 3 m hydraulic head) we find that:

f * = 0.05*8 = 0.5

252

Delft Hydraulics

In case of a resonance frequency of 5 Hz (the gate vibrates vertically, always in the resonance frequency) the critical opening is 0.1 m. If the opening is smaller than 0.1 m, this type of vibration may occur. If the natural frequency is higher, then the critical gate opening is lower. If higher values had been introduced for Ci and m, then even lower values for the critical openings would have been found. Whether the vibration indeed occurs and how strong the degree of self-excitation is, again fully depends on the shape and thickness of the bottom edge.
(a) C. HOLLOW-JET GATE

The hollow-jet gate is used in high-head structures. It concerns a gate that is placed at the downstream side of a tube. See Chapter 6, Example 6.4c. Although the design meets most of the criteria for a vibration-free design, a number of gates have collapsed, probably due to dynamic behaviour. Mercer (1970) has developed a stability criterion for this gate, that is somewhat analogous to that for the bath plug. He indicates that the minimum required plate thickness is:

t = plate thickness t= Q steel 0.115 C D Esteel

For the unfavourable situation with four blades inside the valve, it holds that C = 2.22 (for five blades this is 2.35 and for six blades this is 2.48). When taking as a calculation example for the diameter D = 1 m, Q = 30 m3/s, = 7850 kg/m3 and E = 2 1011 N/m2, then we find for the minimum plate thickness of the ribs 23.3 mm. A point that needs special attention in case of a thick gate or stop log, is the fact that if there are equal resonance frequencies in case of both horizontal and vertical vibrations, then even a design that would normally not cause any vibrations, would still become instable. A last point that also relates to the concept of stiffness, is what was already mentioned earlier, i.e. that in case of wheel gates it needs to be prevented that one of the wheels is relieved to such an extent that it starts to rattle. In that case, the stiffness disappears and it has been observed (in scale model investigations, Example 6.2j, vibration type II) that vibrations do indeed occur in that case.

7.5 Damping and friction


Damping always has a favourable effect with regard to reducing the danger of vibration. In case of self-excitation (or negative damping) the mechanical damping needs to be greater than the negative damping. If that is not the case, then damping slows down the generation of vibration, but the vibration amplitude that is eventually reached, hardly becomes any smaller.

253

Delft Hydraulics

Friction is not equivalent to damping. If the friction force is not overcome, then friction causes a new fixed point. In general however, this fixed point will have a relatively great stiffness, so that it may possibly result in a greater security. There is another critical point concerning which one needs to be on guard for friction damping. Relative to a linear damper, friction has a decreasing damping effect in case of increasing vibration amplitude. This may be observed in the vibrations that are freely damping out in Figure A2.3. In case of friction damping the gradient for all amplitudes remains the same, but in case of linear damping the gradient of the enveloping curve increases at a greater amplitude. At greater amplitudes the damping effect of friction therefore decreases. This means that if a vibration with great amplitude is generated due to an external cause, it may maintain itself or even increase, while up to that point, the structure had been functioning vibration-free due to friction damping. In some cases such as the bath plug vibration at gates, the damping required may be calculated (if the stiffness factor Ck is smaller than the desired 1 + Cm). These calculations have not been carried out. It does appear however, that the negative damping is never greater than what follows from Equation A4.31. In Paragraph 5.4.1, Figure A5.17, it is indicated how a linear damping may prevent the vibrations of a square rod. The diagram shows that the higher the natural frequency, the lower the required damping. The problem in using a damper is finding a fixed point. If this is not present, then the material damping will have to supply the damping force. This however usually is so small, that its influence may be ignored. At a gate, a damper could be mounted in a hydraulic lifting system, but this is not easy. If observed vibrations only occur at small lifting heights, there are possibilities of using a damper that is parallel to an elastic fender. This functions as soon as the fender both makes contact with the bottom and with the gate or door. This remedy has been used successfully, see Example 6.1b. It is recommended to have this kind of remedies readily available, but a good gate design will not primarily include the use of a damper.

7.6 Aeration to prevent vibrations and cavitation


Aeration of gates (at the downstream side) is often used to create a situation with free discharge outflow. This may be very favourable for the following reasons. The flow becomes more stable and a possible hydraulic jump will then be at some distance from the gate. If a sloping bottom has been used in the tailrace culvert, it may even prevent that a hydraulic jump is generated in the culvert. The pressure increases due to aeration and therefore the discharge decreases as well. Both are favourable with regard to cavitation. Downstream turbulence no longer plays a role as excitation source for the gate in case of free runoff. All downstream gate girders may now be positioned in air, as a result of which the stationary and dynamic load on the winch is reduced as well. This means girders of greater dimensions may also be used. In case of free runoff, a choice is always made for a gate with smooth plating upstream and girders at the downstream side.

254

Delft Hydraulics

Also, for those cases in which, due to aeration of a culvert, no free runoff is generated at the gate, a number of arguments for the aeration remains valid. In case of the expected cavitation, aeration is recommended, but in that case only in a finely distributed form. Because of this, the water becomes compressible, as a result of which impulse phenomena that may occur in case of cavitation, are strongly damped. There are however risks connected with aeration, i.e. when its effect is that aeration only occurs locally, and the culvert further down is again completely filled with water. In case of closure of a culvert gate it is possible that, due to the fact that air may expand, the water downstream spurts through, while at the gate it has already been stopped. The air pocket generated will later be filled again, but this may be accompanied by a heavy shock (Example 6.9a). In case of shipping locks, aeration of the filling culvert is hardly acceptable for reasons of turbulence, foaming, and likewise at a culvert with branches, as the discharge distribution is strongly influenced by air. Air gathers in places where the pressure is at its lowest, so that the flow is strongly influenced. If air escapes, then this often happens in a thrustlike way with relatively large quantities at the same time. The air with a certain overpressure in the culvert is often released in an explosive way due to compression.

255

Delft Hydraulics

256

Delft Hydraulics

REFERENCES 8.1 Delft Hydraulics Reports (in Dutch)3

WL M561

part A

part B

part C

Vizierschuiven Stuw Hagestein; modelonderzoek dynamisch gedrag: schaalproeven, 1962, ir P.A. Kolkman. (Visor Gates Hagestein, scale model investigation dynamic behaviour) Vizierschuiven Stuw Hagestein; modelonderzoek dynamisch gedrag: onderrandproeven, 1962, ir P.A. Kolkman. (Visor Gates Hagestein, scale model investigation dynamic behaviour; under edge tests) Vizierschuiven Stuw Hagestein; ontwerp elastisch gelijkvormig model, 1961, ir. P.A. Kolkman (Visor Gates Hagestein, design of elastic similarity model)

WL M627

Drielingschutsluis in het Julianakanaal bij Maastricht, 1962, ir A.P.H. van Baardewijk, ir W. Tromp. (Triplet shipping lock Julinan Canal; scale model investigation)

WL M667A Zeesluis Terneuzen; rapport modelonderzoek rioolschuiven, 1967, Ir M.A. Geleedst. (Sea lock Terneuzen, scale model investigation) WL M754 part I Buitenmeting spuisluis Haringvliet; trillen van de schuiven door stroom. 1977, ir J.D. van den Bunt. (In situ measurements Haringvliet sluice; flow induced vibrations of Gates) Buitenmeting spuisluis Haringvliet; evaluatierapport. 1981, ir J.D. van den Bunt. (In situ measurements Haringvliet sluice; evaluation)

part IV

WL M833

Stuwklep in de Haelense beek; rapport modelonderzoek. 1963, ir M. Geleedst. (Flap gate Haelense beek; scale model investigation) part VI Zoutbestrijding Kreekraksluizen,onderzoek rioolschuiven en woelbak bij zoutbestrijding volgens systeem Duinkerken. 1972, ir E.B. Zegers. (Measures against salt water Kreekrak shipping locks; scale model investigation of gates and stilling basin)

WL M865

Reports can only be studied by others after Delft Hydraulics has received approval for this from the client involved.

257

Delft Hydraulics

WL M1072

Nota betreffende de metingen aan het westelijk spuiriool van de schutsluis te Tiel, 1970, ir E.B. Zegers. (In situ measurements discharge culvert of shipping lock Tiel) part I Inlaatsluis Volkerak; rapport vooronderzoek. 1972, ir H.A. Nuhoff. (Inlet sluice Volkerak, pre-investigation) Inlaatsluis Volkerak; elastisch gelijkvormig model; rapport modelonderzoek, 1973, ir H.A. Nuhoff. (Inlet sluice Volkerak, elastic similarity model investigation)

WL M1129

part II

WL M1272

Schuiven van de stroomsluis in de Brouwersdam, onderzoek krachten en trillingen, 1983, ir R.J. de Jong. (Gates of Brouwerdam sluice, scale model investigation of forces and vibrations) Klepstuwen in boezemkeringen; krachten en trillingen; rapport modelonderzoek, 1975, ir R.J. van der Wal. (Flap Gates in canal systems; scale model investigation forces and vibrations) Grote Zeesluis te Muiden, rapport modelonderzoek, 1975, ir P. van Groen. (Sealock Muiden; scale model investigation butterfly gate) Toegevoegde watermassa en instabiele trillingen van schuiven met een verticale bewegingsmogelijkheid, 1977, ir A. Vrijer. (Scale model investigation added mass and unstable vertical vibrations of gates) part I Stormvloedkering Oosterschelde; krachten en afvoercofficinten bij roosterschuiven; onderzoek bij stijve modellen, 1978, ir R.J. van der Wal. (Storm surge barrier Eastern Scheldt; scale model investigation forces and discharge coefficients of grid gate)

WL M1300

WL M1311 WL M1322

WL M1327

WL M1338

Stormvloedkering Oosterschelde; Trillingsgedrag van roosterschuiven; rapport Modelonderzoek, 1979, ir R.J. van der Wal. (Storm surge barrier Eastern Scheldt; elastic similarity model investigation of vibrations grid gate) part II Nota schuifonderzoek stormvloedkering Oosterschelde hefschuiven, 1976, ir C.R.M Oudshoorn (Storm surge barrier Eastern Scheldt; vertical lift gate scale model investigation) Hefschuifonderzoek svk Oosterschelde; tweede nota stromingsexcitatie, 1976, ir C.R.M. Oudshoorn. (Storm surge barrier Eastern Scheldt; vertical lift gate scale model investigation; flow excitation)

WL M1377

part III

258

Delft Hydraulics

part IV

Nota hefschuifonderzoek stormvloedkering Oosterschelde; gecombineerde stromings- en golfexcitatie, 1976, ir C.R.M. Oudshoorn. (Storm surge barrier Eastern Scheldt; vertical lift gate scale model investigation; combined flow and wave excitation)

WL M1424

Stormvloedkering Oosterschelde; krachten en trillingen bij de hefschuiven in de pijlerdam; vooronderzoek met sectiemodel, 1978, ir R.J. de Jong. (Storm surge barrier Eastern Scheldt; vertical lift gate scale model investigation; section model investigation forces and vibrations) Stormvloedkering Oosterschelde. Waterspiegelfluctuaties tussen de pijlers, hulpmiddelen daartegen en de invloed op afvoercofficinten; rapport modelonderzoek, 1978, ir L. Haas. (Storm surge barrier Eastern Scheldt; scale model investigation water level fluctuations between piers) Stabiliteitsgedrag van schuiven met diverse onderrandvormen bij een verticale bewegingsmogelijkheid, 1979, ir J. Uwland. (Scale model investigation of under edges of vertically movable gates) Stormvloedkering Oosterschelde; pijleroplossing. Vooronderzoek met behulp van een stijf sectiemodel naar stroom- en golfbelastingen op dorpelbalken, bovenbalken en plaatliggerschuiven, 1981, ir T.H.G. Jongeling. (Storm surge barrier Eastern Scheldt; investigation with a scale model section of flow and wave forces on upper beams and open girder gates) Bepaling cofficinten van de verticale kracht op de schuifonderrand, 1979, ir J. Uwland. (Scale model investigation vertical forces on gate under edge)

WL M1436

WL M1490

WL M1494

WL M1497 & M 1498

WL M1561

Stormvloedkering Oosterschelde; pijleroplossing; onderzoek naar trillingsgedrag van plaatliggerschuiven en balken met behulp van een elastisch gelijkvormig model, 1981, ir H.G. Jongeling & ir H.W.R. Perdijk. (Storm surge barrier Eastern Scheldt; investigation with an elastic similarity model of flow and wave induced forces and vibrations on upper beams, sill beams and open girder gates) Stormvloedkering Oosterschelde; onderzoek naar toegevoegde watermassa's plaatliggerschuiven met behulp van een elektrisch analogon, 1982, ir C. Deelen. (Storm surge barrier Eastern Scheldt; investigation using an electric analogon of the added mass of gates)

WL M1582

259

Delft Hydraulics

WL M1648

part I

Stormvloedkering Oosterschelde pijleroplossing. Onderzoek met behulp van een elastisch gelijkvormig model naar het responsiegedrag van de bovenbalken bij golven,1981, ir H.W.R. Perdijk & ir T.H.G. Jongeling. (Storm surge barrier Eastern Scheldt; investigation with an elastic similarity model of wave induced forces and responses on upper beams and open girder gates) idem, maar nu met betrekking tot het gedrag van de plaatliggerschuiven bij golfklapbelastingen, 1981, ir H.W.R. Perdijk & ir T.H.G. Jongeling. (id.; investigation with an elastic similarity model of wave shock induced responses on upper beams and open girder gates) idem, berekening van optredende krachten in de aanslagen van de vakwerkliggerschuiven; berekening van de waterspiegelbewegingen in de eindkokers en in de sponningsruimten; berekening van botskrachten bij bewegen van de schuiven in lengterichting, 1984, ir T.H.G. Jongeling. (id.; various computations of water fluctuations and collision forces of gates in length direction)

part II

part III

WL M1688

Stormvloedkering Oosterschelde; stroomkrachten en trillingen bij het plaatsen van de dorpelbalk in de kering. Vooronderzoek met een geschematiseerd massaveersysteem; rapport modelonderzoek, 1980, ir H.W.R. Perdijk. (Storm surge barrier Eastern Scheldt; scale model investigation of flow induced forces and movements during positioning of sill beams (building stage)) Lozingsmiddel Zoommeer; verificatie en aanpassing ontwerp, bepaling afvoerkarakteristieken en rekenmodel spuisluis met vrije waterspiegel, 1982, ir G. Heijdra. (Discharge sluice Zoommeer; check and adaptation of design; scale model investigations and computations) Westerschelde oeververbinding; krachten op en verplaatsingen van de tunnelelementen ten gevolge van stroomdruk, rapport modelonderzoek, 1981, ir W.D. Eysink. (Western Scheldt, immersed tunnel project; scale model investigation flow induced forces and movements of tunnel elements) Nakdong Estuary barrage and reclamation project; gate vibrations, flow forces and wave loads. Report model investigation and desk study, 1982, ir T.H.G. Jongeling. Kifil Shinafya project; Kufa barrage. Radial gates; response on hydrodynamic Streamforces, 1984, ir T.H.G. Jongeling. Kifil Shinafya project, Kufa barrage; flap gates, hoist load and splitter Configuration, 1985, ir J. Uwland.

WL M1711

WL M1739

WL M1767

WL M2068 WL M2087

260

Delft Hydraulics

WL M2101

Kifil Shinafya project, Kufa barrage; radial gates with I-beams; respons on hydrodynamic stream forces. Report on model investigations, 1985, ir T.H.G. Jongeling. Nuovi interventi per la salva guardia di Venezia, modele fisico di una schiera di paratoie, Sudio B.6.5, 1993, ir T.H.G. Jongeling. Onderzoek in een hydraulisch model naar het ontstaan van in-flow trillingen in schuifranden, rapport modelonderzoek, 1986, ir T.H.G. Jongeling. (Research of in-flow vibrations of under edges of gates)

WL Q20

WL Q190 (M1906)

WL Q282

Trillingen in de stroomrichting van rioolschuiven bij varirende spleetgrootte; rapport bureaustudie, 1986, ir J. Uwland. (Research of in-flow vibrations of culvert gates) part I Vizierschuiven in de Rijn te Driel; orinterende trillingsmeting aan de noordelijke vizierstuw; rapport buitenmeting, 1986, ir T.H.G. Jongeling. (In situ measurements of gate vibrations; Driel weir (visor gates)) Spuisluizen Volkerak, spuisluizen afsluitdijk Den Oever; trillingsmetingen aan hefschuiven, 1988, ir T.H.G. Jongeling. (In situ measurements of gate vibrations; sluice Volkerak & sluice Den Oever)

WL Q322

part II

WL Q491

Een eenvoudig rekenschema voor de berekening van toegevoegde watermassa bij schuiven; presentatie rekenprogramma, 1987, dr ir P.A. Kolkman. (Simple calculation scheme for added water mass) Venice barrier; study on the influence of the inclination angle and the gate side shape on gate response, 1988, ir T.H.G. Jongeling. part I Maeslantkering; vooronderzoek naar het responsiegedrag van de sectordeuren met behulp van een sectiemodel, 1989, ir T.H.G. Jongeling. (Maeslant barrier; pre-investigation to the dynamic behaviour of the sector gates using a model section)

WL Q744 WL Q958

WL Q969

Maeslantkering; onderzoek naar het responsiegedrag van de sectordeuren in een overzichtsmodel, 1989, ir T.H.G. Jongeling. (Maeslant barrier; investigation to the dynamic behaviour of the respons of the sector gates using an overall scale model) Venice barrier; study on gate response, forces in the gate supports and leakage discharge, 1990, ir T.H.G. Jongeling.

WL Q1033

261

Delft Hydraulics

WL Q1140

Maeslantkering. Vervolgonderzoek naar het responsiegedrag van de sectordeuren in een overzichtsmodel, rapport modelonderzoek, 1990, ir T.H.G. Jongeling & ir J.J.A. van Huijstee. (Maeslant barrier; continuation of investigation to the dynamic behaviour of the respons of the sector gates using an overall scale model) Maeslantkering. Onderzoek met behulp van een sectiemodel: optimalisatie vormgeving sectordeur; drukmetingen op drempel en sectordeur, rapport modelonderzoek, 1990, ir T.H.G. Jongeling. (Maeslant barrier; optimisation of the under edge of the gates and pressure measurements in sill and gates, using a section model) Rioolzuiveringsinstallatie Velsen; goottrillingen. Rapport bureaustudie, 1990, ir R.J. de Jong. (Flume oscillations at the sewage water plurification system Velsen) Maeslantkering; analyse van het responsiegedrag van de sectordeuren met behulp van een rekenmodel (met gebruikmaking van drukmetingen in sectiemodel); 1991, dr ir P.A. Kolkman. (Maeslant barrier; response computations using pressure measurements of a section model) Maeslantkering; additioneel onderzoek voor de sectordeuren in een overzichtsmodel, 1991, ir T.H.G. Jongeling. (Maeslant barrier; additional investigation dynamic behaviour of gates in an overall model) Trillingen brugpijler Oosterschelde; rapport studie en berekening, 1962, ir P.A. Kolkman. (Computations of vibrations of bridge piers Eastern Scheldt bridge) Stuwklep bij de Winkelmolenbrug in de Neerbeek; rapport trillingsonderzoek (prototype), 1963, ir M.A. Geleedst. (In situ investigation of vibrations of weir gate Neerbeek) Elastisch gelijkvormig model roosterschuiven; rapport voorbereiding Onderzoek, 1977, ir H. Depeweg. (Preparation of elastic similarity model for the grid gate of the Eastern Scheldt barrier) part 2 Invloed getal van Reynolds en relatieve ruwheid op stroombeeld om cylinders, 1977, ir H.C.N. Breusers. (Influence of Reynolds number and wall roughness on the flow pattern around cylinders) Invloed turbulente aanstroming en getal van Reynolds op de stroming rond stompe voorwerpen, 1977, ir H.C.N. Breusers. (Influence of turbulence and Reynolds number on the flow pattern around bluff bodies)

WL Q1190

WL Q1199

WL Q1271

WL Q1278

WL R166

WL R251

WL R1068

WL R1186

part 3

262

Delft Hydraulics

WL R1280

Stormvloedkering Oosterschelde; bewegingsgedrag schuiven onder invloed van het beweegsysteem en de wrijving op de glijdopleggingen. Wiskundig model met Coulombse wrijving, 1979, ir T.H.G. Jongeling. (Storm surge barrier Eastern Scheldt; computations of the influence of the (vertical) gate operating system and friction in bearing slides (for horizontal forces)) part I Invloed richtingsgevoeligheid wrijvingskracht op het bewegingsgedrag van de schuif; aanvullende berekeningen met wiskundig model massaveersysteem met Coulombse wrijving, Notitie onderzoek, 1980, ir T.H.G. Jongeling. (Storm surge barrier Eastern Scheldt; additional computations of the influence of the (vertical) gate operating system and friction in bearing slides (for horizontal forces)) Dynamische verschijnselen bij verticaal bewegen en belasten van de vakwerkschuiven; berekeningen met behulp van een geschematiseerd massaveersysteem model met Coulombse wrijving, 1982, ir T.H.G. Jongeling. (Storm surge barrier Eastern Scheldt; additional computations of the influence of the (vertical) gate operating system and friction in bearing slides (for horizontal forces) in case of open truss girder gates)

WL R1280

part II

WL R1304

Trillingsmetingen aan de hefschuif noord van de stormvloedkering te Krimpen a/d IJssel; rapport buitenmetingen en bureaustudie, 1980, ir T.H.G. Jongeling. (In situ vibration measurements of the storm surge barrier gate Krimpen a/d IJssel) Doorlaatsluis Brouwersdam; trillingsmetingen schuiven (rapport buitenmetingen), 1980, ir C. Deelen. (In situ measurements vibrations of the Bouwerdam sluice Gates) Stormvloedkering Oosterschelde; turbulentiemetingen in de monding van de Oosterschelde; rapport uitvoering en verwerking prototypemetingen, 1980, ir C. Deelen & ing.W.J.Vos (RWS). (Storm surge barrier Eastern Scheldt; turbulence measurements in the Eastern Scheldt) Duwvaartsluizen in de Philipsdam; maximale krachten op schuiven (bureaustudie), 1981, ir C.R.M. Oudshoorn. Philipsdam shipping locks; study on maximum forces on culvert gates) Spuisluis, trillingsmetingen; rapport buitenmetingen, 1979, ing.W.Klinkenberg & ir A.C.M. Vermeer. (In situ measurements of gate vibrations of a sluice gate)

WL R1347

WL R1437

WL R1506

WL R1510

263

Delft Hydraulics

WL R1594

Stormvloedkering Oosterschelde; bepaling meebewegende watermassa met behulp van een elektrisch analogon ten behoeve van de trilplaatproblematiek, 1988, ir C. Deelen. (Storm surge barrier Eastern Scheldt; added mass assessment applying an electrical analogon) Stormvloedkering Oosterschelde; turbulentiemetingen in de Hammen. Notitie verwerking prototypemetingen, 1981, anoniem. (Storm surge barrier Eastern Scheldt; turbulence measurements in the Eastern Scheldt (Hammen); data processing) Bewerking spectra van de turbulentiemetingen in de monding van de Oosterschelde (R1437 en R1699), 1983, F. van Stralen. (Storm surge barrier Eastern Scheldt; turbulence measurements in the Eastern Scheldt (Hammen); spectra) Stormvloedkering Oosterschelde; responsie groutpijp; rapport bureaustudie, 1985, ir T.H.G. Jongeling. (Storm surge barrier Eastern Scheldt; study on response of grouting pipe) part 1 Investigations on rapidly varying forces on gates; prototype measurements at Lith and possibilities for model, 1958, ir P.A. Kolkman & ir R.W. Hart Supplement to the IAHR Report Studies on the interaction between turbulent pressure fluctuations and movements of a structure, 1959 R.W. Hart. Nota onderzoek 1967 betreffende het dynamisch gedrag van Sectorschuiven, ir P.A. Kolkman. (General study on dynamic behaviour of sector gates) Rapport onderzoek naar oorzaak en gedrag van trillingen in overstortende of verticaal vallende straal, 1968, ir. N. Schoemakers. (Cause and swinging behaviour of falling water curtain) Nota onderzoek aan ronde schotbalken op prototypeschaal, 1968, ir P.A. Kolkman & P. Bosland. (In situ investigation behaviour of circular stop logs) Cavitatie bij rioolschuiven; cavitatie-inceptie getallen en hierbij optredende schaaleffecten, 1967 & 1968, ir C. Jonker & ir C. Bos. (Caviation at culvert Gates; cavitation inception and scale effects)

WL R1699

WL R1920

WL R2055

WL S50

part 2

part 3

part 4

part 5

WL S120

parts I IV

264

Delft Hydraulics

part V WL S230

Velocity and turbulence measurements behind a gate in a wind tunnel, 1968, ir K.W. Pilarczyk.

Drukfluctuaties in turbulente stroming; rapport literatuuronderzoek, 1972, ir. H.C.N. Breusers. (Pressure fluctuations in turbulent flow) Cavitatie en waterkwaliteit, 1979, dr ir D. Oldenziel. (Cavitation and water quality)

WL S232 parts 1 15 WL W254

Trillingen van sluisdeuren, 1976, dr ir J.C.W. Berkhoff. (Vibrations of lock gates)

8.2 Other reference material


Abelev, A.S. (1959): "Investigations of the total pulsating hydrodynamic load acting on bottom outlet sliding gates and its scale modelling", 8th IAHR congress, Montreal, paper 10A. Allersma, E. (1959): "The virtual mass of a submerged sluice gate", 8th IAHR congress, Montreal, paper 23A. Aso Delft Hydraulics publication 18. Aschenbach, E. en E. Heinecke (1981): "On vortex induced shedding from smooth and rough cylinders in the range of Reynolds numbers 6x103 to 5x106", Journal of Fluid Mech., 109, 239-51. Bakker, A.D., T.H.G. Jongeling, P.A. Kolkman en Yan Shi Wu (1991): "Self-excited oscillations of a floating gate related to the gate discharge characteristics", XXIV IAHR congress, Madrid. Also Delft Hydraulics publication 462. Binnie, A.M. (1972): "The stability of a falling sheet of water", Proc. Royal Soc. London series A326 jan. '72, 149-63. Bishop, R.E.D. en A.Y. Hassan (1964): "The lift and drag forces on a circular cylinder oscillating in flowing fluid" Proc. Royal Soc., London, 277 (1368). Blevins, R.D. (1977 en 1990): "Flow-induced Vibration", Van Nostrand Reinhold Co, N. York. Brandao de Menezes, V., D. Pinto de Silves en A. Pinto de Magahaes (1969): "L'cluse de Carrapatello sur le fleuve Douro", 22th PIANC congress, Paris. Inland Navigation, Subject 3. Callander S.J. (1987): "Flow-induced vibrations of rectangular bars (in German)", Pd.D. Thesis Univ. of Karlsruhe, Inst. for Hydromechanics.

265

Delft Hydraulics

Callander S.J. (1988): "Streamwise oscillations of circular-sectioned trashrack bars" Sonderforschungsberiech 210, Univ. Karlsruhe rapp. SFB 210/e/43. Cassidy, J.J. (1990): "Fluid mechanics and design of hydraulic structures", J. of Hydr. Eng. ASCE Vol 116 nr. 8 pp. 961-977. Cassidy, J.J. (1996): "ICOLD bulletin about Vibrations of Hydraulic Equipment for Dams Review and recommendations" Number 102-1996. Clark, P.J. en R.G. Tappin (1977): "Final design of Thames barrier gate structures" in Conf. on Thames barrier design, Inst. of Civ. Engrs, London. Crausse, E. (1939): "Sur un phnomne d'oscillation du plan d'eau provoqu par l'coulement autour d'obstacles en forme de pile de pont" comptes rendus de sances de l'Academie des Sciences, p. 209. Ethembababoglu, S. (1978): "Some characteristics of unstable flow past slots", ASCE Journ. Hydr. Div. Vol. 104, HY5, p. 649. Delaney, N.K. en N.E. Sorenson, (1953): "Low speed drag of cylinders of various shapes", NACA Techn. Note 3038. DenHartog, J.P. (1956): "Mechanical Vibrations", McGraw Hill. Dirkzwager, M. en J.E. Prins (1959): "Tangential forces exerted on segment gates, in partially raised positions by discharge and waves", 8th iahr CONGRESS, Montreal. Also Delft Hydraulics publication 20. DiSilvio, G. (1969): "Self-controlled vibration of a cylinder in a fluid stream", J. Mech. Div. ASCE 95, EM2, paper 6498. Feng, C.C. (1968): "The measurement of vortex induced effects in flow past stationary and oscillating circular and D-section cylinders", Ma. Sc. Thesis, Univ. of Brittish Columbia, Canada. Fung, Y.C. (1960): "Fluctuating lift and drag on a cylinder in a flow at supercritical Reynolds numbers" Inst. of Aerospace sciences, 28th IAS meeting New York, paper 60-6. Griffin, O.M. (1980): "Cold water pipe design for problems caused by vortex-excited oscillations", Naval Research Lab. Washington D.C., Ref. 4157. Hardwick, J.D. (1977): "Hydraulic model studies of the rising sector gate conducted at Imperial College", in Conf. on Thames barrier design, Inst. of Civ. Engrs, London. Harrisson, A.J.M. (1967): "Boundary layer displacement thickness on flat plates", Proc. ASCE J. Hydr. Div., HY4, paper 5339. Hart, R.W. en J.E. Prins (1959):; "Studies on the interaction between turbulent pressure fluctuations and movements of a structure", 8th IAHR congress, Montreal, paper 27A. Also Delft Hydraulics publication 19. 266

Delft Hydraulics

Den Hartog, J.P. (1956): "Mechanical Vibrations", McGraw-Hill Book cy, N. York. Haszpra, O. (1979): "Modelling hydroelastic vibrations", Pittman, Londen. Hooft, J.P. (1972): "Hydrodynamic aspects of semi-submersible platforms" Dr-thesis, Delft University of Technology, Dept. Naval Arch. Hooft, J.P. (1982): "Advanced dynamics of marine structures" publ. Wiley New York. ICOLD, Committee Hydraulics for Dams (1996): "Vibrations of Hydraulic equipment for dams", Bulletin of Subcommittee 2: J.J. Cassidy (chairman), M. Hino, P.A. Kolkman, H. Makarechian en K. Ogihara. Number 102-1996. Ishii, N., K. Imaichi en A. Hirose (1980): "Dynamic instability of tainter gates", see: Naudascher en Rockwell 1980. Ishii, N. (1992): "A design criterion for dynamic stability of tainter gates", J. Fluids and Structures vol. 6, pp. 67-84. Ishii, N. en C.W. Knisely (1992): "Flow-induced vibration of shell-type long-span gates", Journal of Fluids and Structures Vol. 6, Nr 6. Jones, G.W., J.J. Cincotta en W. Walker (1969): "Aerodynamic forces on a stationary and oscillating circular cylinder at high Reynolds number", NASA Rep. TR-300. Jong, R.J.de en J.W.G. van Nunen (1979): "Excitation and vibration of a grid gate", zie Naudascher en Rockwell 1980. Also Delft Hydraulics publication 220. Jong, R.J.de en T.H.G. Jongeling, (1982): "Fluid elastic response study of the Nakdong barrage gates", Int. conf. on Flow-induced Vibrations in fluid-engineering. Reading (Eng.). Also Delft Hydraulics publication 285. Jongeling, T.H.G. (1987): "In-flow vibrations of gate edges": BHRA conf. on Flow-induced Vibrations, Bowness-on-Windmere England. Also Delft Hydraulics publication 392. Jongeling, T.H.G. (1988): "Flow-induced self-excited in-flow vibrations of gate plates", J. Fluids and Structures, Vol. 2, Nr. 6, Also Delft Hydraulics publication 420. Jongeling, T.H.G. (1993): "Wave-induced resonance of a flap-gate barrier" in Structural Dynamics-Eurodyn'93 conf., Trondheim. Ed. Balkema Rotterdam. Jongeling, T.H.G. en P.A. Kolkman (1995): "Subharmonic standing waves leading to lowfrequency resonance of a submersible flap-gate barrier", 6th Intern. Symp. on Flowinduced Vibration. Imperial College, London. Kanne, S., E. Naudascher en Yinan Wang (1991): "On the mechanism of self-excited, vertical vibration of underflow gates" Fifth Int. Conf. on Flow-induced Vibration, Brighton, England, organ. by Inst. of Mech. Engrs.

267

Delft Hydraulics

Kenn, M. en A.D. Garod (1981): "Cavitation dammage and the Tarbela tunnel collapse of 1974", Inst. Civ. Engrs. Part I, nr. 70. King, R. en M.J. Prosser (1974): "Criteria for flow-induced Oscillations of a cantilevered cylinder in water", in Naudascher 1972. Kolkman, P.A.(1959): "Vibration tests in a model of a weir with elastic similarity on Froude scale", 8th IAHR congress, Montreal, paper 29A. Also Delft Hydraulics publication 15. Kolkman, P.A. (1972): "Instability of a vertical water-curtain closing an air-chamber" in Naudascher 1972. Kolkman, P.A. (1976): "Flow-induced gate vibrations; prevention of self-excitation, computation of dynamic gate behaviour and the use of models" Dissertation of the Delft University of Technology. Also Delft Hydraulics publication 164. Kolkman, P.A. (1977): "Self-excited gate vibrations" 17th congr. IAHR, Baden-Baden, invited lecture of section c.c. Also Delft Hydraulics publication 186. Kolkman, P.A. en A. Vrijer (1977): "Gate edge suction as a cause of self-exciting vertical vibrations", 17th IAHR-congress, Baden-Baden paper C49. Also Delft Hydraulics publication 188. Kolkman, P.A. (1980): "Development of vibration-free gate design; learning from experience and theory" see: Naudascher en Rockwell 1980. Also Delft Hydraulics publication 219. Kolkman, P.A. (1984): "Vibrations of hydraulic structures" and "Gate vibrations", Chapters I and II in Developments in Hydraulic Engineering vol. 2, Ed. P. Novak, Elsevier publ. Kolkman, P.A. (1988): "A simple scheme for calculating the added mass of hydraulic gates" J. Fluids and structures Vol. 2, Nr. 4. Also Delft Hydraulics publication 439. Lamb, H. (1932): "Hydrodynamics", 6th ed. Cambridge Univ. Press, Cambridge, UK. Levin, L. (1957): "tude hydraulique des grilles de prise d'eau", 7th IAHR congress Lisbon, paper C11. Liebl, A. (1973): "High pressure sluice gates", 11th Congress on Large Dams, Madrid, Vol. II, Question 41, R42. Martin, W.W., E. Naudascher en M. Padmanabhan (1975): "Fluid-dynamic excitation involving flow-instability", Proc. ASCE, J. Hydr. Div., HY6 paper 11361. Meier-Windhorst, A. (1939): "Flutter of cylinders in steady liquid flow (in German)", Mitt. Hydraulik Inst. Techn. Hochschule Mnchen, nr. 9. Mercer, A.G. (1970): "Vane failures of hollow-cone valves, "IAHR-symp. Hydraulic Machinery Stockholm. 268

Delft Hydraulics

Morkovin M.V. (1964): "Flow around circular cylinders -a kaleidoscope of challenging fluid phenomena", Symp. on fully separated flow, Proc. of Eng. Div. Conf. ASME, Philadelphia. Ed. A.G. Hansen New York. Naudascher, E. (1964): "Hydrodynamic and hydro-elastic type of load at high-head gates" (orig. in German) in "Der Stahlbau", Nr. 7 en 9. Naudascher, E. (1972) (ed.):"Flow-induced structural vibrations", IAHR/IUTAM symp. Karlsruhe 1972, Springer Verlag Berlin 1974. Naudascher, E. en D. Rockwell (editors) (1980): "Practical Experiences with Flow-Induced Vibrations" Publ. Springer after IAHR/IUTAM-symp. in Karsruhe Germ. 1979. Naudascher, E. en N.D. Thang (1986): "Self-excited vibrations of underflow gates", J. Hydraulic Research (ed. IAHR), Vol. 24, Nr.5. Naudascher, E. (1987): "Flow-induced streamwise vibrations of structures", Journal of fluids and Structures, Vol. 1, p. 265. Naudascher, E. (1991): "Hydrodynamic forces" in the series Hydraulic structures design manual edited by IAHR, publ. Balkema, Rotterdam. Naudascher, E. en Y. Wang (1993): "Flow-induced vibrations of prismatic bodies and grids of prisms", Journal of Fluids an Structures, Vol. 7, p.341. Naudascher, E. (1994): "Flow-induced vibrations"in the series Hydraulic structures design manual edited by IAHR, publ. Balkema. Neilson, F.M. en E.B. Picket (1980): "Corps of Engineers experiences with flow-induced vibrations". In Naudascher en Rockwell 1980. Nguyen, D.T., B.C.S. Rao en E. Naudascher (1988): "Field and model studies on trashrack vibrations", Proc. Int. Conf. on Model Prototype Correlation of Hydraulic Structures, ASCE, Colorado Springs, Colorado, USA. Novak, M. (1969): "Aeroelastic galloping of prismatic bodies", Proc. ASCE, J. Mech. Div., EM1, paper 6394. Ogihara, K. en S. Ueda (1980): "Flap gate oscillation", zie Naudascher en Rockwell 1980. Ogihara, K., H. Nakagawa en Y. Ueda (1991): "Self-excited vibration of long span shell roller gate by three-dimensional experimental model", 24th IAHR Congress Madrid session D. Oldenziel, D.M. en J. Tijema (1976): "Cavitation on valves in correlation to liquid properties", IAHR symposium "Problems of hydraulic machine- hydraulic structure interaction", Leningrad. Also Delft Hydraulics publication 173.

269

Delft Hydraulics

Padoussis, M.P. (1980): "Flow-induced vibrations in nuclear reactors and heat exchangers and state of knowledge" zie Naudascher en Rockwell 1980. Partensky, H.W. en I. Sar Khloeung (1971): "Overflow nappe oscillations without aeration" (orig. in French) 12th IAHR-congress, Paris; Seminar paper S-6. Patel, M.H. (1989): "Dynamics of offshore structures"publ. Butterworth London etc. Pettigrew, M.J., Y. Sylvestre en A.O. Campagna (1978): "Vibration analysis of heat exchanger and steam generator designs", Nuclear engineering and design, Vol. 48, p. 97. Petrikat, K. (1980): "Seal vibration", zie Naudascher en Rockwell 1980. Rao, B.C.S., N.D. Thang en E. Naudascher (1987): "Vibration of trash racks in flow with different incidence angles" BHRA Intern. Conf. on Flow-induced Vibrations, Bowness-on-Windmere, UK. Rhee, C. van (1984): "Trillingsgedrag van een klepstuw bij hoge afvoer" Afstudeerrapport in de vakgroep Vloeistofmechanica aan de Technische Hogeschool Afd. der Civiele Techniek. Richter, A. en E. Naudascher (1976): "Fluctuating forces on a rigid circular cylinder in confined flow", Journal of Fluid Mechanics, Vol. 78, p.561. Sarpkaya, T. (1960): "Added mass of lenses and parallel plates" Journ. of Eng. Mech. Div. Proc. ASCE, June 1960. Sarpkaya, T. (1976): "In-line and transverse forces on smooth and sand-roughened cylinders in oscillatory flow at high Reynolds numbers", Naval post-graduate School, Montery, Calif. Rep. NPS-69 SL, 76062. Sarpkaya, T. (1978): "Fluid forces on oscillating cylinders", Proc. ASCE, J. Waterway, Port and Ocean Div. WW3, paper 13941. Sarpkaya T. (1982): "Flow-induced vibration of roughened cylinders", BHRA Int. Conf. on Flow-induced Vibrations in Fluid Engineering, Reading, paper D1. Saveira, J.G. en C.M. Ramos (1983): "Hydrodynamic loads and vibrations of trash rack elements" 20th IAHR-congress Moscow. Schmidgall, T. (1972): "Spillway gate vibrations on Arkansas river dams", Proc ASCE, J. Hydr. Div., HY1 paper 8676. Schoemaker, H.J. (1959): "Dynamics of a sluice gate and movable weir subject to forces exerted by waves and vibrations" 8th IAHR congress, Montreal paper 30A. Also Delft Hydraulics publication 14.

270

Delft Hydraulics

Schoemaker, H.J. (1971): "Virtuele massa bij golfklappen en daarop volgende trillingen in een constructie", in Manuscripten van H.J. Schoemaker in de periode 1946-1971, publ. Delft Hydraulics. Schwarz, H.I. (1964): "Nappe oscillation", Proc. ASCE, J. Hydr. Div. HY6 paper 4138. Scruton, C. (1963): "On the wind-excited oscillations of stacks, towers and masts", First conf. on Wind effects on buildings and structures. Nat. Phys. Lab., Teddington UK, Aero rep. 305. Streeter, V.L. en E.B. Wylie (1967): "Hydraulic transients", McGraw-Hill book cy, New York. Syamalarao, B.C., N.D. Thang en E. Naudascher (1987): "Vibration of trashracks in flow with different incidence angles", Int. Conf. on Flow-induced Vibrations, Bowness-onWindmere, England org. BHRA. Syamalarao, B.C. (1989): "A review of trashrack failures and related investigations", Water Power and Dam Construction, Vol. 41, p. 28. Thang, N.D. (1990): "Gate vibrations due to unstable flow separation", ASCE Journal Hydraulic Eng., Vol. 116, Nr.3. Toebes, G. (1969): "The unsteady flow and wake near an oscillating cylinder", transactions ASME, J. Basic Eng. pp. 493-505. Treiber, B. (1972): "Theoretical study of nappe oscillations", in Naudascher 1972. University of Washington (1952): "The role of vortex shedding in the aerodynamic excitation of suspension bridges" Univ. of Washingtom Engineering, Experimental station, Bull. 116, part III, appendix III. Vrijer, A. (1980): "Stability of vertically movable gates", zie Naudascher en Rockwell 1980. Also Delft Hydraulics publication 222. Walshe, D.E.J. (1967): "The aerodynamic investigation for the proposed 850-ft Chimney stack for the Drax Power Station", Nat. Phys. Lab., Aero rep. 1227. Wendel, K. (1950): "Hydrodynamische Massen und hydrodynamische Massentragheitsmomente", Jahrbuch der Schiffsbautechnischer Gesellschaft, Vol. 44, blz. 207-255. Westergaard, H.M. (1933): "Water pressures on dams during earthquakes", Trans. ASCE, Vol. 98, paper 1835 p.418. Weaver, D.S. (1980): "Flow-induced vibrations in valves operating at small openings", zie Naudascher en Rockwell 1980. Wootton, L.R., M.H. Warren en D.H. Cooper (1972): "Some aspeects of the oscillations of full scale piles", zie Naudascher 1972.

271

Delft Hydraulics

272

Delft Hydraulics

APPENDIX

THE FREQUENCY AT WHICH, IN CASE OF A VERTICALLY VIBRATING GATE, THE FLOW-INDUCED DAMPING CHANGES INTO A FLOW STIFFNESS EFFECT

273

Delft Hydraulics

APPENDIX THE FREQUENCY AT WHICH, IN CASE OF A VERTICALLY VIBRATING GATE, THE FLOW-INDUCED DAMPING CHANGES INTO A FLOW STIFFNESS EFFECT

In case of a vertically vibrating gate, the vertical forces are determining for both the flow damping and the flow stiffness. It is assumed that the flow force that operates vertically on the gate is proportional to the hydraulic head of the gate. Whether the forces operate upward or downward depends on the design. The issue represented in the title of this appendix is now translated into the question: at what frequency is the extra hydraulic head of the gate that is generated by a gate vibration, in-phase or out-of-phase with the gate vibration? Out-of-phase means that there is flow damping; in-phase means there is flow stiffness. Whether this damping and stiffness will be positive or negative is not discussed here. This concerns the (periodic) extra hydraulic head that is generated by a vibration with a very small amplitude.

Figure 1: Gate in culvert.

Figure 2: 274

Delft Hydraulics

Gate in a situation with a free water surface. The gate may be a culvert gate, as in Figure 1, or a gate in free flowing water as in Figure 2. In the latter, an equivalent tube is introduced that is representative of the flow inertia. The tube length equals Ci (see Paragraph 4.4.3, Equation A4.33) and the tube height then becomes . In Paragraph 4.4.3 it has been deduced (see also Figure A4.14) that in case of a very low-frequency vibrating gate, flow damping occurs that strongly decreases at higher frequencies. It appears that the flow damping is negative when a suction force influences the gate (this suction force is proportional to the local hydraulic head of the gate). Suppose the gate has the following discharge relation: q = m 2 g H s (1)

In this, q = discharge, = opening beneath the gate, Hs = hydraulic head of the gate and m = discharge coefficient of the gate (assumed to be constant). The flow resistance in the culvert, relative to the gate resistance of a pinched-off gate, is negligible. For the non-vibrating gate: q0 = m 0 2 g H 0 (2)

The index 0 is introduced to indicate that this is a situation in which the gate does not vibrate. Now, assuming that in case of a vibrating gate and a fluctuating discharge, the inertia of the culvert does play a role, but its resistance does not, then the hydraulic head of the gate equals the external hydraulic head, H0, subtracted by the hydraulic head required to accelerate or decelerate the water in the culvert. Therefore, at a culvert height D:
H s = H 0 L dV pipe L dq = H 0 g dt gD dt

(3)

From Equation 3 it immediately follows that the dynamic part of the hydraulic head fully depends on the dq/dt. (b) The limit case = 0 First, the dynamic part of the hydraulic head of the gate is calculated for the limit case = 0 ( = angular frequency of the vibration of the gate). The discharge variation as a consequence of vibration in this case is only determined by the gap variation; the hydraulic head remains constant, because the interference on the hydraulic head due to the (small) dq/dt term, relative to the hydraulic head itself, is very small. The complete calculation, presented below in this appendix, in which this simplification is not being used, eventually results in = 0, the same value for the flow damping as deduced below. From Equation 1, combined with the assumption that the flow inertia, coupled to dq/dt does not influence the gate discharge, follows:
dq dy =m 2 g H 0 dt dt (4)

275

Delft Hydraulics

As long as the vibration is small, relative to the gap height underneath the gate, all variables may be linerarised. All magnitudes that vary with time may be considered as a superposition of the value in the non-vibrating situation (indicated with index 0) and a small fluctuation around it (indicated with an accent, provided that ' is replaced by the vibration displacement y). Now, the dynamic part of the hydraulic head of the gate Hs' is calculated from the dq/dt found, using Equation 3. This hydraulic head, H0, is again expressed in the initial discharge q0, using Equation 2. This results in: H s = Lq dy dy mL 2 g H 0 = 0 dt gD gD dt (5)

In the maximum case = 0 the hydraulic head is therefore out-of-phase with the vibration. This results in a damping, as also the dynamic component of the vertical force on the gate is now proportional to the vibration velocity of the gate. (c) The limit case = If the gate vibrates with a high frequency, then the discharge tends to remain constant as a consequence of the flow inertia in the tube. From Equation 1 it may be deduced that if the discharge remains constant, and the gate is widened by a small percentage (a%), the hydraulic head must be reduced by 2a% in order to compensate for the effect of the gap change. In case of a small value of this may be presented as:
(1 2 ) = 1

(6)

(squared, and next ignoring 2, shows that this is correct). It also follows that: (1+ )(1 ) = 1 Therefore it may now be stated that: dH gate H 0 or: = 2 y

H s =

2H 0

(7)

As the hydraulic head (and consequently also the vertical force) is in-phase due to vibration, and considering the proportionality with the displacement, this results in a spring stiffness. (d) The full calculation for the entire area From Equations 3 and 1 it follows that:
( H s = ) q2 L dq = H 0 2 2 2 gm gD dt

(8)

276

Delft Hydraulics

H0, the external hydraulic head, does not vary with time. Here again it is introduced that the periodic discharge variation is indicated by q' and the gap variation by y. The coefficient with time only depends on the fluctuating part. This applies to , Hs and also to q. Equation 8 now changes into:
2 2q0 dq 2q0 dy L d 2 q = 2 gm 2 02 dt 2 gm 2 03 dt gd dt 2

(9)

As all equations are linearised (this happened during the transition from Equation 8 to Equation 9) and again the calculation is limited to very small amplitudes, it holds that if the gate moves harmonically, the other fluctuating magnitudes vary harmonically around the mean position (which matches the value without the vibration) as well. For the fluctuating magnitudes: y = Yeit it q = qe and also: eit = H H gate s (12) (10) (11)

As y is a forced movement, Y may be chosen as a real number. The other magnitudes may have shifted in-phase and therefore have a complex value. Equation 9 now transforms into:
2 2q0 2q0 L 2 i =0 i q Y q 2 gm 2 02 2 gm 2 03 gD

(13)

or, after dividing by and multiplying with i: = q


2 2q0 / 2 gm 2 03 Y 2q0 / 2 gm 2 02 + i L / gD

(14)

Because it may be deduced from Equation 3 that: = H s and therefore: = i H s L dq gD dt (15)

L
gD

(16)

it may be deduced from Equations 16 and 14 that:

277

Delft Hydraulics

= i H s

2 2 Lq0 / 2 g 2 m 2 03 D Y 2q0 / 2 gm 2 02 + i L / gD

(17)

After multiplying the numerator and denominator with the added complex of the denominator, the following equation is obtained:
2 2 2 3 2 2 = Y (2 Lq0 / 2 g m 0 D) *( L / gD + i 2q0 / 2 gm 0 ) Y H s (2q0 / 2 gm 2 02 ) 2 + ( L / gD) 2

(18)

Lets consider the imaginary and the real terms separately. The imaginary term is representative of the situation that is out-of-phase with the vibration and therefore results in the flow damping. For = 0, the imaginary term is compared with the maximum value found in Equation 5. The real part of the hydraulic head of the gate results in the flow stiffness; the maximum value for = is compared with what is found in Equation 7. The imaginary term may be written as:
) = iY I (H s (2 Lq0 / 2 gm 2 02 ) 2 *( Lq0 / g 0 D) 2 2 2 + (2q0 / 2 gm 2 02 ) 2 * 1 ( / ) /(2 / 2 ) L gD q gm { } 0 0 (19)

and this again may be written as:


) = I (H s Lq0 dy 1 gD dt 1 + {( L / gD) /(2q / 2 gm 2 2 )}2 0 0 (20)

It is easy to see that for = 0, Equation 20 changes into Equation 5. The real part of Equation 18 may be elaborated in a similar way:
) = Y R (H s
2 ( L / gD) 2 (2q0 / 2 gm 2 03 ) 2 ( L / gD) 2 * 1 + {(2q0 / 2 gm 2 02 ) /( L / gD)}

(21)

Using Equation 2, this may be written as:

) = 2H 0 Y R (H s

1 + {(2q0 / gm 2 02 ) /( L / gD)}

(22)

For high values, the real part therefore changes into the maximum value found in Equation 7. Calculation of the transition frequency at which the flow damping decreases and the flow stiffness increases. The transition frequency may be defined, in which both the real and the imaginary part are bisected relative to their limiting values at = and at = 0. Both for the real and the

278

Delft Hydraulics

imaginary part of the hydraulic head of the gate (Equation 22, respectively Equation 20) this frequency is:

L 2 gm2 02
2 gDq0 Using Equation 2, this may be written as:

=1

(23)

Lm
D 2 g H

=1

(24)

This results in the angular frequency, at which the stiffness and the damping of the culvert gate in Figure 1 are bisected, relative to their limiting values. If this relation is based on vertical gate vibrations in a situation with a free water surface, then:
D =

(25)

The tube length is defined as: L = Ci For the turnover frequency it is then found that: S= f 0 1 = 2 g H 0 2 mCi (26)

These results fully match what was presented in Figure A4.14 for the negative damping in case of vertical gate vibrations.

279

Delft Hydraulics

280

Delft Hydraulics

INDEX BY SUBJECT (PART A)

Added damping due to wave radiation (Paragraph 3.4.1) due to flow (Paragraphs 3.4.2 3.4.4) Added stiffness due to immersion (Paragraph 3.4.1) due to flow (Paragraphs 3.3.2 3.3.3) Added water mass general considerations (Paragraph 3.1) polar (added) mass inertia moment (Paragraph 3.2.4) calculation methods (Paragraph 3.2.1) values (Paragraphs 3.2.2 3.2.4 and 3.2.7) influence culvert ceiling and culvert length (Paragraph 3.2.5) wall influence (Paragraph 3.2.3) Aeration (Paragraphs 6.9 and 7.6) Amplification of vibrations due to fluid oscillations (Paragraph 1.3 and 5.5) Bath plug vibration (Paragraphs 4.4.1 and 4.4.2) Blockage (influence on Strouhal number) (Paragraph 5.7.2) Causes of vibrations turbulence and separation of vortices (Paragraphs 1.3, 4.4.2 and 5.2) flow instability (Paragraphs 1.3, 4.3 and 5.3) gate vibrations (general) (Paragraph 4.1) self-excitation (general) (Paragraph 1.3) self-excitation of gates (Paragraph 4.4) self-excitation of flow-surrounded objects (Paragraph 5.4) fluid oscillations (Paragraphs 1.3 and 4.5) Cavitation (Paragraphs 1.5 and 6.9c) Classification of vibrations (Paragraphs 1.3 1.4) Communicating vessels (Paragraph 2.2.3) Correlation (decreasing correlation, relative to distance) (Paragraph 5.3.7) Couplings between vibration modes (Paragraph 3.6) discharge variations coupled to gate movements (Paragraph 4.5.1) discharge coefficient variation with gate movements (Paragraph 4.4.4) Culvert gates (including flap gates and valves in pipes) (Paragraphs 4.4.2 and 6.3 6.4)

281

Delft Hydraulics

Damping and friction Coulomb friction (Paragraphs 2.2.5 and 7.5) influence damping in case of harmonic excitation (Paragraph 2.2.4) influence damping in case of noise excitation (Paragraph 2.2.8) damping as a remedy in case of vibrations (Paragraph 7.5) negative damping: see self-excitation added damping due to wave radiation (Paragraph 3.4.1) added damping due to flow (Paragraphs 3.4.2 3.4.4) Degrees of freedom (couplings) (Paragraph 3.6) Design of gates (Paragraph 7.2.1) of rods (paragraph 7.2.2)

Discharge coefficient related to (and cause of) gate vibration (Paragraph 4.4.4) Discharge sluices (Paragraph 5.6) Discharge variations as a cause of unstable fluid oscillations (Paragraphs 1.3, 4.5.1 and 5.6) Excitation frequency: see Strouhal number Excitation sources for vibrations turbulence and separation of vortices (Paragraphs 1.3, 4.2 and 5.2) flow instability (Paragraphs 1.3, 4.3 and 5.3) gate vibrations (general) (Paragraph 4.1) self-excitation (general) (Paragraph 1.3) self-excitation of gates (Paragraph 4.4) self-excitation of flow-surrounded objects (Paragraph 5.4) self-excitation in case of fluid oscillations (Paragraphs 1.3 and 4.5) Experiences in prototype (Chapter 6) Experiences in scale models (Chapter 6) Flap (or flap gate) vibrations flap gates, gates and valves in culverts and pipes (Paragraphs 4.4.2 and 6.3 6.4) overflow (Paragraph 4.4.7) pressure relief valve (Paragraph 4.4.2) Flow inertia (Paragraph 3.5) Flow-surrounded objects (Paragraphs 5.1 5.5) Flow turbulence (excitation due to) (Paragraphs 1.3, 4.2 and 5.1 5.2) Fluid oscillations resonance (Paragraphs 1.3, 2.3 and 5.5) unstable resonance (Paragraphs 1.3, 4.5 and 5.6) 282

Delft Hydraulics

amplification of vibrations due to fluid oscillations (Paragraphs 1.3 and 5.5)

Fluid oscillator general properties (Paragraph 2.3.1) situations of occurrence ((Paragraph 2.3.2) communicating vessels (Paragraph 2.3.3) standing wave (Paragraph 2.3.4) causes of fluid oscillations (rocking) (Paragraph 2.3.5) Flutter (Paragraph 5.4.3) Frequency domain (Paragraph 2.2.4) Friction (Coulomb) (Paragraphs 2.2.5 and 7.5) Galloping in case of gates (Paragraph 4.4.5) in case of rods (Paragraphs 5.4.1 and 5.7.2) Gap shape (Paragraphs 6.5c, 6.5f and 7.2.1) Gates in culverts and pipes (including valves and flap gates) (Paragraphs 4.4.2 and 6.3 6.4) Gates in free flowing fluid (Paragraphs 4.4.3 4.4.4, 4.4.7, 6.1 6.2 and 6.6) Grid (vibrations) (Paragraph 5.7) Harmonic flow excitation (Paragraph 2.2.7) Immersion stiffness (Paragraph 3.3.1) In-flow vibrations (in the main flow direction, parallel to the wall or bottom) in case of gates (Paragraph 4.4.4) in case of rods (Paragraph 5.4.2) Instability indicator (Paragraphs 4.4.6 and 4.5) Mass inertia moment of the fluid (Paragraph 3.2.4) Mass-spring-damper system dry (Paragraph 2.2.1 2.2.5) wet (Paragraph 2.2.6 2.2.8) multiple degrees of freedom (Paragraph 2.2.9) combined fluid / mechanical system (Paragraph 2.4) Nappe splitter (Paragraphs 4.4.7 and 6.6.a 6.6.c) Natural frequency (Paragraph 2.2.2) Noise excitation due to flow (Paragraph 2.2.8) 283

Delft Hydraulics

Overflow (vibrations) (Paragraphs (Paragraphs 4.4.7 and 6.6) Polar (added) mass inertia moment (Paragraph 3.2.4) Pressure relief valve (Paragraph 4.4.2) Pumps (Paragraph 5.6) Radiation of waves due to vibrations (Paragraphs 3.1.1 and 3.4.1) Reduced (flow) velocity (Paragraphs 2.2.6, 5.3.8 - 5.3.10, 5.4.1, 5.7.2) Regulating system (paragraph 5.6) Separation of vortices and turbulence (Paragraphs 1.3, 4.2 and 5.2) Remedies (Chapter 7) Resonance in case of fluid oscillations (Paragraphs 1.3 and 5.5) Response in case of harmonic flow excitation (Paragraph 2.2.7) in case of noise excitation due to flow (Paragraph 2.2.8) in case of various loads (Paragraph 2.2.5) Reynolds dependency of Strouhal number (Paragraphs 5.1 and 5.3.4 5.3.6) Rods angular rods (Paragraphs 3.2.2 and and 5.2) circular or bevelled (Paragraphs 5.1 and 5.3 5.5)

Scale rules; see Part C Sealing structures (Paragraphs 6.5 and 7.3) Self-excitation (or negative damping) general (Paragraphs 1.3 and 2.2.10) in case of gates (Paragraph 4.4) in case of flow-surrounded objects and rods (Paragraph 5.4) Shroud (Paragraph 7.2.2) Specification of concepts (Paragraph 1.2) Spiral (Paragraph 7.2.2) Spoiler (Paragraphs 6.2.e, 6.6.a c and 7.2.2) Standing wave (Paragraph 2.3.4)

284

Delft Hydraulics

Stiffness -

influence on vibration behaviour (Paragraph 7.4) added stiffness due to immersion (Paragraph 3.3.1) added stiffness due to flow (Paragraphs 3.3.2 3.3.3) sudden hydrodynamic stiffness (Paragraph 3.3.3)

Stop logs (Paragraph 6.7) Strouhal number definition (Paragraphs 1.3 and 4.2) values for angular rods (Paragraph 5.2) values for circular and rounded rods (Paragraphs 5.1, 5.3.4 and 5.3.6) values for grid (trash rack) rods (Paragraphs 5.7.1 5.7.2) influence blocking (Paragraph 5.7.2) Reynolds dependency (Paragraphs 5.1, 5.3.4 and 5.3.6) Sub-harmonic transverse waves (Paragraph 4.6) Sudden stiffness (Paragraph 3.3.3) Terminology (Paragraph 1.2) Time domain (Paragraph 2.2.5) Transverse waves (Paragraphs 4.5, 4.6 and 6.1b) Trash rack: see grid Turbines (Paragraph 5.6) Turbulence and separation of vortices (Paragraphs 1.3, 4.2 and 5.2) Unstable resonance in case of fluid oscillations (Paragraphs 1.3, 4.5 and 5.6) Valves (Paragraph 6.4) Vibrations bath plug vibration (Paragraph 4.4.2) vibration of sealing strips (Paragraphs 6.5 and 7.3) vibration of flap gates with overflow (Paragraphs 4.4.7 and 6.6) vibration of rods and grids (trash racks) (Paragraphs 3.2.2, 5.2 5.5, 5.7, 6.6 and 6.8) vibration of gates, flap gates and valves in culverts and pipes (Paragraphs 6.3 6.4) vibration of gates in free flow conditions (Paragraphs 4.43 4.4.4, 6.1 6.2 and 6.6) Wall influence in case of added water mass (Paragraph 3.2.3) Waves 285

Delft Hydraulics

sub-harmonic transverse waves, induced by vibrations (Paragraph 4.6) wave radiation (Paragraphs 3.1.1 and 3.4.1) standing wave (Paragraph 2.3.4)

286

Delft Hydraulics

287

Delft Hydraulics

Biographies of the authors


Dr P.A. Kolkman was born in 1932 and graduated from the Delft University of Technology in 1956. He was awarded a doctorate in 1976 on the subject of vibrations in closure mechanisms. He spent his entire career at Delft Hydraulics. In addition, he was a scientific researcher at the Delft University of Technology from 1968 to 1990. Dr Kolkman has been involved in the design of almost all major Dutch navigation locks, river dams and tidal barriers and has also acted as an advisor to large hydraulic engineering projects abroad. He has also been responsible for various leading publications. In particular, he has acquired an international reputation for his work in the area of combating vibration in movable gates and hydraulic valve mechanisms. He is the co-author of a number of books published in English on saline-fresh water separation systems in sea locks and drainage relationships in man-made structures, among other subjects. At this time he is the chief editor of the International Association of Hydraulic Research (IAHR) journal. T.H.G. Jongeling was born in 1947 and completed his studies in 1976 in Applied Mechanics at the Delft University of Technology. He has worked at the HBG and IBBC-TNO and has been with Delft Hydraulics since 1977. Mr. Jongeling specialises in the dynamics of hydraulic structures and has carried out research and consultancy projects in the field of vibration and wave impact, particularly in connection with large sea defence projects in the Netherlands and overseas. He has also contributed to the design of numerous large and small hydraulic engineering projects, as well as writing a number of works in the field of structural vibration. Mr. Jongeling is now also active in the area of irrigation systems. Dienst Weg- en Waterbouwkunde (civil engineering unit of Rijkswaterstaat) Bouwdienst (construction office of Rijkswaterstaat) Delft Hydraulics 1996

288

You might also like