computational fluid dynamics investigation and ... - dspace cover page

339
Computational fluid dynamics investigation and optimisation of marine waterjet propulsion unit inlet design Author: Seil, Gregory Juergen Publication Date: 1997 DOI: https://doi.org/10.26190/unsworks/5538 License: https://creativecommons.org/licenses/by-nc-nd/3.0/au/ Link to license to see what you are allowed to do with this resource. Downloaded from http://hdl.handle.net/1959.4/57490 in https:// unsworks.unsw.edu.au on 2022-05-31

Upload: khangminh22

Post on 08-Mar-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

Computational fluid dynamics investigation and optimisationof marine waterjet propulsion unit inlet design

Author:Seil, Gregory Juergen

Publication Date:1997

DOI:https://doi.org/10.26190/unsworks/5538

License:https://creativecommons.org/licenses/by-nc-nd/3.0/au/Link to license to see what you are allowed to do with this resource.

Downloaded from http://hdl.handle.net/1959.4/57490 in https://unsworks.unsw.edu.au on 2022-05-31

Computational Fluid Dynamics Investigation and Optimisation of Marine Waterjet Propulsion

Unit Inlet Design

by

Gregory Juergen Seil BE (Hons. I)

A thesis submitted to fulfil the requirements for admission to the degree of Doctor of Philosophy

of The University ofNew South Wales

Department of Mechanical and Manufacturing Engineering

The University of New South Wales Sydney, Australia.

December 1997

Abstract The primary objective of the work presented in this thesis was to use computational

fluid dynamics (CFD) to investigate and optimise the design of flush-type marine

waterjet propulsion unit inlets. The CFD methodology used was based on the solution of

the Reynolds-averaged Navier Stokes equations with two-equation RNG k-E turbulence

modelling, on single-block body-fitted-coordinate structured grids.

The accuracy of a commercially-available CFD code (Fluent) was validated against

three different experimental data sets, in order to assess the suitability of CFD as a

reliable tool for the investigation of waterjet inlet design. The validation cases

examined, corresponded to flow in a 90° bend, S-Duct and waterjet inlet. It was

consistently found that the RNG k-E turbulence model gave more accurate flow

predictions than the Standard k-E turbulence model, yielding results in good agreement

with experimental data.

The effect of inlet velocity ratio on the static pressure distribution within the waterjet

inlet and the flow at the duct exit was examined. Due to the variation in possible

boundary layer thicknesses with different hull forms, the effect of hull boundary layer

thickness was investigated and found to have a significant effect on the flow within the

waterjet inlet, by virtue of ingested momentum and energy fluxes.

A generic parametric waterjet inlet geometry was defined and its design hyperspace

investigated, in order to correlate the flow within the waterjet inlet with its underlying

geometry. The parametric geometry was then optimised for maximum static pressure on

the surface of the waterjet inlet. This resulted in a sharp, raised-lip profile. The design

hyperspace investigations and the waterjet inlet optimisation were made with a thick

boundary layer upstream of the inlet, in the absence of a surrounding hull form, at a

vessel cruise condition. It was found that careful attention must be given to the design of

the inlet lip profile and operating conditions, in order to avoid cavitation. The lip must

be designed in such a way as to direct the flow symmetrically over it and so maximise

static pressure on its surface.

i

Acknowledgements I would firstly like to express my deepest sense of gratitude and appreciation to my wife

Elizabeth and my parents for their continual support and encouragement, without which

this work would not have been possible.

I would also like to thank my supervisors Prof. C. A. J. Fletcher and Assoc. Prof. L. J.

Doctors for initiating this interesting research project, which I have enjoyed studying.

Furthermore I would like to thank both gentlemen for their supervision, guidance,

motivation and encouragement during the course of my research. I would also like to

express my appreciation to DrS. Di for his co-supervision.

I thank the Commonwealth Government and the Australian Maritime Engineering

Cooperative Research Centre (AMECRC) for providing both the direct and indirect

financial support that has made this research possible.

I am indebted to Mr J. Roberts for providing me with experimental data so that I could

validate Fluent against an actual waterjet inlet flow. I would also like to thank Dr G.

Walker and Prof. M. Davis for their active enthusiasm and interest in this project as part

of the AMECRC propulsion program.

Finally I would like to thank Mr N. A. Armstrong for sharing his knowledge on the

development of waterjet-propelled high-speed catamarans and Mr B. W. Matthews for

his encouragement.

ii

Table of Contents Abstract..................................................................................................................... 1

Acknowledgements................................................................................................. ii

Table of Contents.................................................................................................... 111

Nomenclature ........................................................................................................... vii

1 Introduction......................................................................................................... 1 1.1 Alternative Waterjet Concepts .... ............... ............. ... ... .... ... .. ........ .......... ... . .. 3 1.2 Historical Development of W aterjet Propulsion Systems . . . .. . . . . . . . . . . . .. . . . . . . . . . . . . 4 1.3 Technical Overview........................................................................................ 4

1.3.1 Inlet.................................................................................................... 5 1.3.2 Pump.................................................................................................. 6 1.3.3 Nozzle................................................................................................ 7 1.3.4 Steering and Reversing Gear.............................................................. 8

1.4 Issues Associated with Waterjet-Propelled Vessels ....................................... 11 1.4.1 Waterjet-Hull Interaction................................................................... 11 1.4.2 Ship Design Optimisation.................................................................. 12

1.5 Research Issues and Objectives ...................................................................... 14 1.6 Overview of Thesis......................................................................................... 16

2 Parametric Model of Waterjet Performance ........................................... 18 2.1 Waterjet Thrust ............................................................................................... 19

2.1.1 Definition ofWaterjet Control-Volume for Thrust Analysis ............. 20 2.1.2 Thrust Relationships .. . . . . . . . .. . . . . . . . . . . . . . . .. . . .. . .. . . . . . . . . . . . . . . . . .. .. . . . . . . . . . .. . . . . . . . . . 21 2.1.3 Waterjet Thrust Model ....................................................................... 25

2.2 Propulsive Efficiency ......... .. ... ................ .. ............. ......... ... ............ ........... ..... 27 2.2.1 Control-Volume Definition for Efficiency Analysis.......................... 27 2.2.2 Development of a Model for Propulsive Efficiency . . . . . . . . . . .. . . . .. . . . . . . . . . 28

2.3 Parametric Study of Waterjet Efficiency........................................................ 31 2.3.1 Parametric Analysis with Boundary Layer Ingestion ........................ 33 2.3.2 Points of Maximum Efficiency .......................................................... 36

2.4 Closure ............................................................................................................ 37

3 Computational Fluid Dynamics Modelling .............................................. 38 3.1 Governing Flow Equations ............................................................................. 41

3 .1.1 N a vier Stokes Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . 41 3.1.2 Reynolds-averaged Navier Stokes Equations .................................... 42

3.2 Turbulence Modelling .................................................................................... 43 3.2.1 Standard k-E Turbulence Model.. ....................................................... 46 3.2.2 RNG k-E Turbulence Model .............................................................. 46

iii

3.2.2 Limitations of k-E Turbulence Modelling .......................................... 48 3.3 Modelling of the Near-Wall Flow .................................................................. 51

3.3.1 Boundary Layer Structure .................................................................. 51 3.3.2 The Wall Function Approach ............................................................. 53 3.3.3 Standard (Equilibrium) Wall Function .............................................. 54 3.3.4 Two-Layer-BasedNonequilibrium Wall Function ............................ 56

3.4 Finite-Volume Discretisation of the Governing Equations............................ 58 3.4.1 Finite-Volume Formulation ............................................................... 58

3.5 Convective Differencing ................................................................................. 61 3.5.1 Power-Law Scheme ........................................................................... 63 3.5.2 Higher-Order Convective Differencing ............................................. 64 3.5.3 Numerical Diffusion .......................................................................... 65

3.6 Solution of the Discretised Equations ............................................................ 65 3.6.1 Solution Methodology ........................................................................ 66 3.6.2 Determination of Convergence .......................................................... 69 3.6.3 Iterative Solution of the Discretised Equations .................................. 70

3.7 Closure ............................................................................................................ 73

4 Generic Geometry and Grid Generation .................................................. 76 4.1 Generic Flush-type Waterjet Inlet Geometry .................................................. 80

4.1.1 Geometric Simplifications .......... ..... ........ .... ......... ......... .. .. ... .. ......... .. 80 4.1.2 Parameterisation of the Generic Geometry........................................ 81 4.1.3 Representation of Geometric Features............................................... 84

4.2 Mesh Topology............................................................................................... 85 4.3 Boundary Mesh............................................................................................... 86

4.3.1 Transfinite Interpolation ..................................................................... 86 4.3.2 Stretching Function ............................................................................ 87 4.3.3 Smoothing of the Surface Mesh ......................................................... 88

4.4 Interior Mesh ......... .................................... ...... .............. .. . ... ..... ...................... 89 4.4.1 Transfinite Interpolation..................................................................... 89 4.4.2 Smoothing of the Interior Mesh......................................................... 90

4.5 Quality of the Generic Waterjet Inlet Grid..................................................... 91 4.6 Grid Generation for Bends and S-Ducts ......................................................... 94

4.6.1 Grid Topology.................................................................................... 94 4.6.2 Meshing Procedure ............................................................................ 95 4.6.3 Grid Quality ....................................................................................... 95

4.7 Closure ............................................................................................................ 96

5 Experimental Validation . .. ......................... ...... ......... .. .................................... 98 5.1 Generic Flow Behaviour ................................................................................. 100

5.1.1 Flow in Bends .................................................................................... 101 5.1.2 Flow inS-Shaped Ducts ..................................................................... 102 5.1.3 Flow in Flush-Type Waterjet Inlets ................................................... 104

5.2 Flow in a 90° Bend ......................................................................................... 107 5.2.1 Experimental Configuration ............................................................... 107 5.2.2 Computational Modelling of Experimental Configuration ................ 107

iv

5.2.3 Computational Simulation ................................................................. 110 5.2.4 Experimental and Theoretical Comparisons ...................................... 111

5.3 Flow in an S-Duct. .......................................................................................... l24 5.3.1 Experimental Configuration ............................................................... 126 5.3.2 Computational Modelling of Experimental Configuration ................ 127 5.3.3 Computational Simulation ................................................................. 128 5.3.4 Experimental and Theoretical Comparisons ...................................... 129

5.4 Flow in a W aterjet Inlet .................................................................................. 141 5 .4.1 Experimental Configuration ............................................................... 141 5.4.2 Computational Modelling of Experimental Configuration ................ 143 5.4.3 Computational Simulation ................................................................. 145 5.4.4 Experimental and Theoretical Comparisons ...................................... 147

5.5 Discussion of Results ..................................................................................... 168 5.5.1 Boundary Conditions ......................................................................... 168 5.5.2 Grid Size and Quality ......................................................................... 169 5.5 .3 Turbulence Modelling ........................................................................ 170 5.5.4 Near-Wall Modelling ......................................................................... 172

5.6 Closure ............................................................................................................ 174

6 Boundary Layer Investigations .................................................................... 177 6.1 Assessment of Hydrodynamic Performance ................................................... 178'

6.1.1 Cavitation ........................................................................................... 179 6.1.2 Inlet Total Pressure Losses ................................................................. 181 6.1.3 Flow Distortion at the Duct Exit.. ...................................................... 182 6.1.4 Internal Volume ofWaterjet Inlet.. .................................................... 187 6.1.5 Vertical Forces acting on the Waterjet Inlet.. ..................................... 187 6.1.6 Dimensions of the Inlet Streamtube ................................................... 188

6.2 Computational Modelling and Simulation ..................................................... 188 6.2.1 Waterjet Inlet Geometry ..................................................................... 189 6.2.2 Computational Modelling of Flow Domain ....................................... 190 6.2.3 Computational Simulation ................................................................. 194

6.3 Results ............................................................................................................ 196 6.4 Discussion of Results ..................................................................................... 210 6.5 Closure ............................................................................................................ 215

7 Design Subspace Investigations .................................................................... 218 7.1 Investigation Methodology ............................................................................. 220 7.2 Computational Simulation .............................................................................. 222 7.3 Results ............................................................................................................ 225

7.3.1 Variation of Lip Radius and Inlet Inclination .................................... 225 7.3.2 Variation of Lip Profile ...................................................................... 237 7.3.3 Variation of Lip Radius with Internal Diffusion ................................ 246

7.4 Discussion of Results ..................................................................................... 259 7 .4.1 Hydrodynamics of Inlet Lip ............................................................... 259 7 .4.2 Vertical Forces acting on the Waterjet Inlet.. ..................................... 262

7.5 Closure ............................................................................................................ 266

v

8 Optimisation of Waterjet Inlet Design ....................................................... 268 8.1 Optimisation Methodology ............................................................................. 270

8.1.1 Overview of Optimisation Procedure ................................................ 273 8.1.2 Optimisation Algorithm ..................................................................... 276

8.2 Computational Simulation and Optimisation ................................................. 279 8.3 Results ............................................................................................................ 281 8.4 Discussion of Results ..................................................................................... 292

8.4.1 Correlation between Flow and Geometry .......................................... 292 8.4.2 Convergence Behaviour of the Optimisation Algorithm ................... 298

8.5 Closure ............................................................................................................ 299

9 Conclusions and Recommendations ............................................................ 302 9.1 Conclusions .................................................................................................... 302

9.1.1 Parametric Model ............................................................................... 302 9 .1.2 CFD Modelling .................................................................................. 303 9.1.3 Design Subspace Investigation and Optimisation Methodologies ..... 304 9.1.4 Effect of Upstream Boundary Layer .................................................. 305 9 .1.5 W aterjet Inlet Design ......................................................................... 306

9.2 Recommendations .......................................................................................... 309 9.2.1 CFD Modelling .................................................................................. 309 9.2.2 Waterjet Inlet Design ......................................................................... 310

References .................................................................................................................. 312

Appendix A - Design Subspace Data .............................................................. .323 A.1 Tabulated Results for Design Subspace 1 ..................................................... 323 A.2 Tabulated Results for Design Subspace 2 ..................................................... 324 A.3 Tabulated Results for Design Subspace 3 ..................................................... 324 A.4 Lip Flow and Minimum Lip Static Pressure ................................................. .325 A.5 Lip Flow and Vertical Forces ........................................................................ 325 A.6 Study of Vertical Forces on Waterjet Inlet .................................................... 326

A.6.1 Design Subspace 1 ............................................................................ 326 A.6.2 Design Subspace 2 ............................................................................ 327

Appendix B- Optimisation Data ....................................................................... 329 B.1 Tabulated Results for Optimisation Algorithm ............................................. 329 B.2 Results for Optimum Lip Radius and Lip Profile Inclination ........................ 331

vi

Chapter 1 Introduction

Marine waterjet propulsion is a form of marine propulsion. Water is first drawn from

around the marine vehicle into the inlet. A pump is then used to add energy to the water,

increase its momentum and produce a thrust by expulsion of a jet of water. Fig. 1.1

shows a typical waterjet propulsion unit, with a flush-type inlet, used on displacement,

semi-displacement and planing hulls. Fig 1.2 shows a waterjet propulsion unit, with a

ram-type inlet, typically used on hydrofoil craft.

Pump Shaft

Duct/Diffusor Nozzle

Inlet Ramp

Outflow

Inlet Lip

Inflow

Fig. 1.1 Waterjet propulsion unit with flush-type inlet

The components of the waterjet propulsion unit are shown in both Fig. 1.1 and Fig. 1.2.

Water enters the waterjet propulsion unit through the inlet opening and flows through

the inlet ducting to the pump. The pump adds energy to the flow, by increasing the total

pressure of the water. This is usually accomplished via a rotating impeller driven by the

pump shaft. Stator vanes (in the pump unit or nozzle) eliminate the rotational

component of the flow downstream of the impeller. The momentum of the water is

increased as the flow is accelerated in the converging nozzle, before being ejected

rearward from the vessel. Steering of the vessel is accomplished by deflection of the jet

of water exiting the nozzle using steering and reversing gear. The jet is deflected either

sideward for steering to port or starboard, or forward for stopping or reversing. Not all

waterjet propulsion units are equipped with steering and reversing gear. Some waterjet

1

units are used exclusively as "booster" units designed only to produce a forward thrust.

Impeller Blades

Outflow

Fig. 1.2 Waterjet propulsion unit with ram-type (pitot) inlet

According to Allison (1993), Brandau (1967), Griffith-Jones (1994) and Roy (1994), the

use of waterjet propulsion offers the following advantages over conventional open

propeller arrangements:

1) Reduced draft depending on hull-type

2) Excellent manoeuvrability

3) Minimal appendage drag and elimination of appendages such as shafting, struts,

skegs and rudders

4) More uniform engine and transmission loading profile provided that there is no air

ingestion into the inlet

5) Essentially constant torque over the range of ship speed at a given power

6) A properly designed pump unit produces a rotation-free flow, thus improving the

propulsive efficiency

7) No transmission reversing gear is required, although in some cases transmission

reversing gear may be installed to allow a back-flushing of the inlet for clearing

debris from the inlet

8) Unrivalled stopping ability and reduced stopping distances

9) Minimal damage susceptibility to floating debris

1 0) Reduced inboard noise

2

11) Reduced vibration

12) The absence of an externally rotating blade provides a safety benefit.

Other advantages which may be of significance in military applications include:

1) Greatly reduced underwater noise

2) A reduction in the turbulence of the wake downstream of the vessel

3) Reduced magnetic signature.

There are, however, some notable disadvantages inherent in the selection of waterjet

propulsion as a means of ship propulsion. These include:

1) Substantially higher initial cost in terms of engineering, purchase and installation

2) Higher overall fuel consumption, as compared with equivalent propeller installations,

due to reduced efficiency at off-design conditions (such as low speed operation in the

case of a high-speed vessel)

3) The potential for inlet plugging due to the build-up of weeds and other debris on the

inlet grill

4) Corrosion and biological growth within the inlet causes increased pressure loss within

the inlet and hence leads to a degradation of performance

5) Integration of the propulsion unit with the hull form is more complex than with the

equivalent propeller installation on non-hydrofoil applications

6) The ingestion of air into the waterjet when certain hull types, such as planing hulls

and surface effect ships, operate in a seaway

7) Impeller access, when compared with conventional propeller designs is poor, making

inspection and repair, or removal of debris difficult.

1.1 Alternative Waterjet Concepts

The field of waterjet propulsion is not limited to the types of waterjet propulsion units

shown in Fig. 1.1 and Fig. 1.2. Perhaps the most futuristic waterjet concept is that of

magnetohydrodynarnic (MHD) sea water propulsion, where water is electrolysed by an

electric field acted upon by large magnetic fields. The interaction of the electric current

produced by ion movement in the MHD duct and the magnetic field, forces water

through the duct and creates a thrust. Detailed discussions on MHD sea water

3

propulsion can be found in Owen (1962), Doragh (1963), Swallom et al (1991) and

Doss and Geyer (1993). Motora et al (1991) discussed the development of the world's

first MHD-powered ship, the Y amato-1.

Gany (1993) proposed a "bubbly" waterjet, a ramjet-type propulsion unit where thrust is

generated by forming a high-speed two phase exhaust jet flow of water and air, created

by the addition of compressed air into the water flow. Allison (1993) cited references to

water piston propulsors. Roy (1994) cited references to other forms of waterjet

propulsion such as "pulse jet" propulsion. It is clear that there exist a large number of

potential waterjet concepts that can be used for marine propulsion.

1.2 Historical Development of Waterjet Propulsion Systems

There are several papers in the literature discussing the historical development of

waterjet propulsion. Roy's (1994) paper is an excellent and comprehensive overview of

the history of waterjet development and future development possibilities. Allison ( 1993)

briefly reviewed some of the historical developments of waterjet propulsion and Youngs

( 1994) provided an overview of the development of jet boats.

It is interesting to note that the idea of waterjet propulsion dates back to 1661 when

English inventors Toogood and Hayes were granted a patent by King Charles II, 19

years before Hooke suggested the screw propeller (see Roy (1994)). In the latter part of

the twentieth century there has been a rapid increase in the number of waterjet

manufacturers and the number of waterjets produced. There has also been a trend of

increasingly more powerful waterjet propulsion units.

1.3 Technical Overview

In this section the basic components of conventional waterjet propulsion systems which

include the waterjet inlet, the pump, nozzle and steering and reversing gear are

discussed in greater detail.

4

1.3.1 Inlet

The waterjet inlet and associated ducting from the inlet opening to the waterjet pump

shall (in the remainder of this thesis) be termed the "waterjet inlet". The use of the term

"inlet", in the context of waterjet propulsion, will be restricted to that section of the

waterjet inlet extending from the inlet opening to the throat of the waterjet inlet. The

function of the waterjet inlet is to supply the waterjet pump with water. There are

essentially two types of inlets as shown in Fig. 1.1 and Fig. 1.2. These are the flush-type

inlet commonly used on monohulls and catamarans and the ram-type (or pitot-type)

typically used on hydrofoil craft. The type of inlet and ducting arrangement to be used

on a particular vessel is really a pragmatic choice of the engineer.

At high Reynolds number, the flow in the waterjet inlet (for a given thickness of

upstream boundary layer) is determined primarily by the inlet velocity ratio (NR). The

inlet velocity ratio is defined here as

(1.1)

where U0 is the volumetrically-averaged velocity at the duct exit of the waterjet inlet and

Us is the ship speed. The volumetrically-averaged velocity (U0 ) at the duct exit is

therefore

u =_g_ o A

0

(1.2)

where Q is the volumetric flow-rate through the waterjet inlet and Ao is the cross­

sectional area of the duct exit.

The flow entering the inlet stagnates at the inlet lip. Depending on operating conditions,

cavitation and flow separation may occur at the inlet lip. At high NR cavitation may

occur on the inlet lip above the stagnation line, while at low NR cavitation may occur

below the stagnation line. Flow separation may also occur on the inlet ramp if the

adverse pressure gradient associated with diffusion of the flow into the inlet becomes

too large. The efficiency of the inlet duct is also dependent on the length of ducting,

number and type of transitions as well as the roughness of the duct walls. The impeller

shaft passing through the inlet, the presence of the duct bend, internal boundary layer

growth and total pressure losses resulting from flow inside the waterjet inlet create a

5

distorted total pressure field at the pump inlet. This distorted total pressure field can

reduce the operating efficiency of the pump and assist in exciting pump vibration.

1.3.2 Pump

The waterjet pump consists of one or more impeller-stator pairs. Hence, the waterjet

pump may be of a single-stage or multi-stage construction. The pump imparts an

increase in total pressure to the fluid flowing through it. This increase in total pressure is

usually manifested as increased fluid static pressure.

Pumps may be of either the axial, centrifugal or mixed-flow type. It is evident from

Trillo (1994) that most waterjet pumps are of the mixed-flow type. According to

Verbeek ( 1992), pump type is associated with the dimensionless parameter specific

speed (Ns) defined as

(1.3)

where n is the impeller speed (rad/s), Q the volumetric flow rate through the pump

(m3/s), g the gravitational constant and Hp the pump head in metres. Pumps of low

specific speed produce higher head, for a given volumetric flow rate and impeller speed,

than pumps of higher specific speed. Pumps of low specific speed are of the centrifugal­

flow type, whereas pumps of high specific speed are of the axial-flow type. With

increasing specific speed, there is a transition from a radial-flow to an axial-flow type.

Another important quantity in pump analysis is that of suction specific speed (Nss),

given in dimensionless form by Verbeek ( 1992) as

(1.4)

where Hs is the net positive suction head (NPSH), given by

- I ( I u2 ) Hs -- Patm +po +2P o -pgH-pv pg

(1.5)

In Eqn 1.5, p is the fluid density, Patm is the atmospheric pressure, Po the static pressure

at the pump inlet (neglecting the hydrostatic component of pressure), U0 the velocity at

the pump inlet in rn/s, H the elevation of the pump shaft centreline above sea level and

6

Pv the vapour pressure of water. All head quantities are expressed in metres. NPSH is

essentially the total manometric head available at the pump inlet above the vapour

pressure of the fluid as noted by Verbeek (1992). Suction specific speed is an extremely

useful parameter in relating the operating conditions of the pump, such as impeller

speed, volumetric flow-rate and NPSH, to the likelihood of impeller cavitation. From

Eqn 1.4 it is clear that an increase in impeller speed, volumetric flow-rate or a decrease

in NPSH, will result in a higher value of suction specific speed and hence a greater

possibility of impeller cavitation.

Note that an inducer can be used to increase the suction specific speed limit. Allison

(1993) listed the following characteristics as being desirable for a waterjet pump:

1) High hydraulic efficiency at high flow

2) Minimum outside diameter and weight for a given nozzle size

3) Cavitation free operation down to maximum pump speed and low inlet head

4) The capability of sustained operation with some cavitation without noticeable erosion

of blades, stator vanes and nozzle

5) High impeller speed enabling the use of a lighter gearbox

6) Tolerance to flow distortion at the pump inlet

7) Superior mechanical design of bearings, pump lubrication system, shaft seals and

other associated components

8) The use of lightweight corrosion-resistant materials for pump housing and parts

The above characteristics favour pumps of relatively high specific speed and high

suction specific speed.

1.3.3 Nozzle

The waterjet nozzle converts the increase in total pressure (mostly in the form of static

pressure) imparted by the pump into increased fluid momentum. This occurs as the high

pressure water at the nozzle entrance is accelerated in the converging nozzle to a higher

velocity at the nozzle exit, exiting the nozzle at ambient pressure. It is evident from

Trillo ( 1994) that nozzles are usually machined from stainless steel castings with

integral stator vanes. Nozzles generally have very high efficiency and since they carry

fluid of high total pressure they must have high efficiency in order to minimise fluid

7

power losses.

1.3.4 Steering and Reversing Gear

Steering of a waterjet-propelled vessel is typically achieved by deflection of the jet of

water leaving the propulsion unit nozzle. The jet may be deflected by a steering bucket,

steering nozzle, or a duct type deflector. Fig. 1.3 below shows an example of

commercially-available steering and reversing gear.

' "

Fig. 1.3 Typical waterjet steering and reversing unit (Picture from Lips Jets B.V.)

By deflecting the jet of water leaving the nozzle, large steering forces can be produced,

giving the vessel excellent manoeuvrability. The sideward steering force (achieved by

jet deflection) of each steerable waterjet is

Fs = mUJ sin8 (1.6)

where Fs is the side force produced by the waterjet, m the mass flow-rate through the

waterjet, UJ the jet velocity of the water exiting the nozzle relative to the waterjet and 8

the angle of deflection relative to the centreline of the waterjet. Eqn 1.6 is used widely

in the literature (see for example Allison (1993)). For small values of 8, Eqn 1.6 may be

approximated by

(1.7)

The stopping or reversing of a waterjet-propelled vessel is usually achieved by using a

reversing bucket that deflects the jet of water either down and forward, or sidewards and

forward. By partial lowering of the reversing bucket, it is not only possible to control the

8

amount of forward or reverse thrust available, but also to create zero net thrust. The

prime mover can therefore run the waterjet pump at constant revolutions per minute

(RPM), while the reversing bucket is used to control the vessel speed. Fig. 1.4 shows the

mass fluxes associated with reversing.

Steering Nozzle

Reversing Bucket

Fig. 1.4 Operation of reversing gear

The reversing force produced by the waterjet reversing gear can be calculated as

FR = (rn- rnl )UJ cosa. (1.8)

where FR is the reversing force produced by the waterjet bucket, rn and U1 are defined as

for Eqn 1.6, rn1 is the mass flow-rate by-passing the reversing bucket and a. is the angle

of the jet of water emerging from the reversing bucket (measured in a clockwise

direction from the horizontal). Due to the limitations of the geometry of the reversing

gear, the astern bollard pull of the waterjet will be less than that of the ahead bollard

pull. The application of the reversing bucket also generates a vertical force (Fy) given by

Fy = (rn- rn1)UJ sina. (1.9)

which will create a bow-down pitching moment on the vessel.

If vessel thrust (T) is approximated by

T = rn(UJ- u.) (1.10)

where Us is the vessel speed, the resulting force of the vessel as a result of application of

the reversing gear is then

(1.11)

From Eqn 1.11 it is evident that very large reversing thrust can be obtained as the

9

reversing thrust is the sum of the force on the reversing bucket and the momentum drag

associated with ingestion of fluid into the waterjet inlet. Since the momentum drag is a

function of vessel speed, it is obvious that the waterjet-propelled vessel will exhibit

excellent stopping characteristics, especially at high-speed.

The zero thrust position of the waterjet bucket under operating conditions may be

determined by setting Eqn 1.11 to zero, resulting in

_1(:rh1 - :rhUs I UJ J ot=cos :rh- :rhl

(1.12)

The mass flow rate deflected by the reversing gear is a direct function of its design and

therefore affects the dynamics of reversing bucket operation according to the above

equations.

Voulon and Wesselink ( 1995) presented an excellent paper discussing the technical

issues associated with manoeuvrability of waterjet-propelled passenger ferries. The use

of the waterjet steering and reversing bucket can be used to execute a variety of vessel

motion such as turning, yawing and transverse motion. When only stem-mounted

waterjets are installed, the vessel's length-to-beam ratio is of prime importance to

manoeuvrability. For large length-to-beam ratios, the possible steering angles for

transverse motion become small, thus resulting in low side forces. The use of bow­

thrusters or jet pumps can have a marked influence on the manoeuvrability of vessels

with large length-to-beam ratios. Wave-piercing catamarans with low length-to-beam

ratios therefore possess greater manoeuvrability than monohulls, according to Voulon

and Wesselink (1995) and Torneman (1994).

On nearly all commercially-available steerable waterjet units, the steering and reversing

gear is fitted to the waterjet propulsion unit, usually on the pump or nozzle body. This

type of mounting is acceptable for small and medium sized waterjets. Allison and Dai

(1994) described a new concept for mounting steering and reversing gear on the vessel

hull instead of the pump body. This concept can be used for very large waterjet units,

when it becomes impractical to mount these components on the pump body.

10

1.4 Issues Associated with Waterjet-Propelled Vessels

Technical issues that must be taken into account when designing the waterjet propulsion

system for a vessel are discussed in this section. These issues must be appropriately

addressed if the waterjet-propelled ship is to function successfully.

1.4.1 Waterjet-Hull Interaction

For waterjet-propelled ships, the waterjet inlet duct is an integral part of the hull form and

hydrodynamically it is impossible to separate the two for the purpose of analysis, as the

hull form affects the flow entering the waterjet while the waterjet affects the flow around

the hull form. This mutual interaction is called waterjet-hull interaction in the literature.

W aterjet -hull interaction will change the following hydrodynamic phenomena:

1) Hull boundary layer growth

2) Potential flow around the hull

3) Wavemaking

4) Sinkage and trim effects

relative to the bare hull flow, with consequent effects on thrust deduction and propulsive

efficiency.

Dai et al (1995) stated that the drag can be either augmented or reduced relative to the

bare hull resistance when there is an inlet with impeller operating. This situation is

further complicated by changes in hull pressure distribution due to the interaction

between the inlet and hull which can effectively change the trim of the vessel. Dai et al

(1993) also noted that the inlet flow can also change the boundary layer characteristics

near the stem. There is some flow acceleration ahead of the inlet and the formation of a

"new" boundary layer aft of the stagnation line on the inlet lip. This modification of the

boundary layer ahead and aft of the inlet can have a significant effect on the drag of the

vessel relative to the bare hull resistance. The flow into the inlet may also affect the

boundary layer on the stem adjacent to the inlet area.

The hydrodynamic problem posed is therefore very complex. At present experimental

model testing in towing tanks offers the best means of a systematic investigation into

waterjet-hull interaction effects. The proceedings of the 1987 International Towing Tank

11

Conference (ITTC) offered guidelines for the evaluation of data from self-propulsion

model tests using waterjets. Dyne & Lindell ( 1994) discussed the evolution of wateijet­

propelled self-propulsion tests at the SSP A towing tank in Sweden. Steen & Minsaas

( 1995) discussed the testing of waterjet propulsion system models and self-propulsion tests

at MARINTEK in Norway. It is therefore clear that a significant number of waterjet­

propelled self-propulsion model tests have been carried out worldwide.

In the literature there has been some systematic work in the area of waterjet-hull

interaction for waterjets with flush-type inlets, most notably by van Terwisga (1996). Van

Terwisga used a combination of experimental and computational work to investigate the

mechanisms influencing waterjet-hull interaction. For planing craft, van Terwisga (1996)

found a large peak in the thrust deduction fraction at low Froude numbers. This was

caused by the clearing of the transom stem which affects the nozzle pressure and hence

thrust. The resistance increment of the hull due to the jet action was found to be dominated

by pressure drag. For Froude numbers below unity, this pressure drag was governed by

transom sinkage, with the trim angle becoming more important at higher Froude numbers.

Other works have included papers by Alexander et al (1994), Alexander and van Terwisga

(1993) and Coop and Bowen (1993). Hoshino and Baba (1984) investigated the effect on

thrust deduction of using a waterjet with twin propellers on a self-propelled semi­

displacement craft model. There remains, however, a great deal of work which can be

undertaken in this field, both experimentally and computationally. The interested reader is

referred to van Terwisga (1996) for a comprehensive literature review on the subject of

wateijet -hull interaction.

1.4.2 Ship Design Optimisation

The design of the waterjet propulsion system, i.e. the vessel prime mover, wateijet pump

and inlet ducting, poses another challenge. It is clear from the literature that the wateijet

propulsion system cannot be optimised independently of the total vessel design, but must

be selected on the basis of a total vessel design optimisation, in order to obtain the most

cost-effective design.

12

Venturini ( 1980) noted that an optimal vessel and propulsor arrangement does not

generally coincide with the minimum propulsor power for a given cruising speed.

According to Allison and Goubault ( 1995), there exists a need to optimise the vessel and

waterjet as an integrated entity through a whole ship design synthesis process, rather than

fit available waterjet propulsors into a ship design. Gallin et al ( 1995) for example,

concluded that designing the waterjet propulsor to match the prime mover resulted in a

smaller life-cycle cost than matching a prime mover to an existing waterjet design.

The design of hydrofoil craft provides interesting examples of the need to optimise the

total vessel design rather than individual sub-systems. Schultz (1974a, 1974b) discussed

issues associated with the design of the Boeing Jetfoil, a gas-turbine-powered, waterjet­

propelled high-speed passenger ferry, that entered service in the early 1970s. In order to

design the Jetfoil, Boeing conducted extensive parametric analyses looking at issues

such as suitable hydrofoil routes, hydrofoil configuration and size, vehicle performance

and cost and the number and distribution of passengers. Other factors examined

included fares, block-speed, percent of market carried, load factor, hours of utilisation,

required fleet size, profit and return on investment.

A fundamental design consideration for Boeing was to keep the vehicle weight fractions

(structure, propulsion and machinery) as low as was safely possible to achieve the best

payload/performance for the installed power, in order to obtain the most efficient

design. Life-cycle costs such as construction, maintenance and support were also

analysed.

Batte and Davis ( 1967) discussed the methods developed at Boeing during the 1960s for

waterjet propulsion system selection for hydrofoil craft. For waterjet systems, Boeing

consistently found that a higher propulsive efficiency would result in lower overall

system performance because of the effects of parasitic drag and waterjet weight

fractions. Propulsive efficiency therefore proved to be a misleading criterion in

determining the best waterjet system. More applicable criteria were required, such as

range, payload and the efficient use of power. Boeing found that early design analyses

demonstrated performance optimisation of individual subsystem components resulted in

13

relatively poor overall craft performance, thus creating a need for the examination of a

large number of overall craft performance parameters and various sub-system

compromises.

1.5 Research Issues and Objectives

The focus of this research project was the investigation and optimisation of the design of

generic flush-type waterjet inlets using computational fluid dynamics (CFD). Ideally the

total waterjet inlet/hull configuration should be optimised as part of a total vessel design

optimisation process, taking into account the effects of waterjet -hull interaction, as

discussed in Section 1.4. The systematic hydrodynamic optimisation of a waterjet inlet

itself, in the absence of hull form, presents a sufficiently complex problem that must be

addressed prior to any waterjet inlet/hull optimisation. In addition, investigations related

to the design and optimisation of the waterjet inlet allows a large amount of data to be

gathered on how aspects of the inlet design affect inlet hydrodynamic performance. This

is clearly useful in correlating the hydrodynamic performance with the underlying inlet

geometry.

In recent years, there have been an increasing number of publications appearing in the

literature presenting CFD results for flows in waterjet inlets and discussing the use of

CFD as a tool for waterjet inlet design and analysis. Among these publications are the

works of Ff/Srde et al (1991 ), Dai et al (1995), Seil et al (1995) and Seil et al (1997).

Other researchers have concentrated on comparing CFD results with experimental data

obtained from the testing of model waterjet inlets in wind tunnels. Griffith-Jones (1994)

and Roberts (1998) have carried out such studies. The CFD results of Griffith-Jones

( 1994) were not in good agreement with the experimental data, possibly due to

inadequate near-wall flow resolution resulting from the grid used.

Allison ( 1996) discussed the cost savings that were achieved by using CFD as a design

tool for the hydrodynamic design of the waterjet inlet for the US Marine Corps

Advanced Amphibious Assault Vehicle. Costs savings were attributed to a reduction in

the experimental testing required for the development of the propulsor. Indeed, the use

of CFD provides a necessary foundation for practical, cost-effective waterjet inlet

14

optimisation work as Seil et al ( 1997) noted.

The waterjet inlet is likely to spend most of its operational time in a "cruise" (low IVR)

condition and so will tend to be designed by the manufacturer with a bias toward this

condition. The waterjet inlet must also offer adequate hydrodynamic performance at off­

design conditions such as when manoeuvring in a harbour or docking (high IVR). Under

these off-design conditions, the NPSH at the pump inlet is low thus limiting the power

input into the pump, the flow-rate through the pump and hence the thrust produced. In

addition, if severe cavitation and flow separation exist in the inlet, the thrust available

for manoeuvring may also drop dramatically thus degrading manoeuvring performance.

This is of special concern for a large vessel attempting to manoeuvre in a crowded

harbour (or dock) while being acted upon by currents, wind and the vessel's own inertia.

As a consequence of the importance of obtaining adequate hydrodynamic performance

from the waterjet inlet in off-design conditions, the hydrodynamic optimisation of a

waterjet inlet must be treated as a multi-point optimisation problem where the resulting

optimised design expresses a compromise between the cruise and manoeuvring

conditions.

There are many other questions and issues that arise in conjunction with the design of

the inlet and its optimisation. These include, but are not limited to the following:

1) The key parameters that are used to evaluate the hydrodynamic performance of the

waterjet inlet

2) The aspects of the inlet geometry which have the greatest effect on hydrodynamic

performance

3) The effect of the upstream hull boundary layer thickness on the flow at the duct exit,

the flow in the vicinity of the inlet lip and its effect on an optimised design

4) The optimality of current commercially-available inlet designs

5) The accuracy of CFD flow computation

6) The ease with which existing CFD technology can be applied to the optimisation of

waterjet inlet design

As a consequence of the above issues, the following research objectives were set and

15

form the basis of the work presented in this thesis. These objectives were:

1) To validate that the accuracy of a commercially-available CFD code (in this case

Fluent) based on the solution of the Reynolds-averaged Navier Stokes equations with

two-equation k-E turbulence modelling, is sufficient to allow it to be used as a

suitable tool for predicting the flow in waterjet inlets.

2) To develop a mathematical model to describe waterjet propulsion system

performance.

3) To develop a mathematical model to describe the hydrodynamic performance of the

waterjet inlet for use with an optimisation methodology.

4) To develop a parametrically-defined generic flush-type waterjet inlet geometry that

can be systematically varied (by varying the defining geometric parameters) in order

to produce different waterjet inlet designs. The systematic variation of the waterjet

inlet geometry is a necessary prerequisite for parametric design subspace

investigation and waterjet inlet optimisation.

5) To investigate the use of CFD and a formal optimisation methodology in order to

optimise the waterjet inlet (in the absence of hull-form) with respect to cavitation

performance and the elimination of flow separation, for a vessel cruise condition.

This corresponds to a single point optimisation and allows the optimised geometries

to be compared with existing commercially-available waterjet inlet designs.

1.6 Overview of Thesis

A theoretical model for the determination of waterjet propulsion unit thrust and

efficiency is presented in Chapter 2, together with a parametric study of the waterjet

efficiency model. The parametric study examines the effect of waterjet component

efficiencies and upstream boundary layer thickness on overall waterjet efficiency and

the point of optimum waterjet efficiency.

The CFD methodology used as the basis of the flow computations presented in this

thesis is presented in Chapter 3. The computational techniques discussed are based on

the solution of the RANS equations with two-equation k-E turbulence modelling, using a

co-located finite volume formulation for discretising the governing equations. Iterative

solution techniques used to solve the discretised governing equations are also discussed.

16

A description of the parametrically-defined generic flush-type waterjet inlet geometry,

used for the purposes of waterjet inlet design and optimisation investigations, is

presented in Chapter 4. The mesh generation techniques used for meshing this geometry

and creating single-block body-fitted-coordinate structured grids for use with Fluent are

also presented. Chapter 4 also describes the mesh generation techniques used for

meshing the 90° bend and S-Duct experimental geometries discussed in Chapter 5.

In Chapter 5 the accuracy of Fluent is validated against three experimental data sets.

These include turbulent flow in the 90° bend of Enayet et al (1982), the 45°-45° S-Duct

of Bansod and Bradshaw ( 1972) and the flush-type waterjet inlet wind tunnel model of

Roberts (1998).

A set of parameters used to assess the hydrodynamic performance of the waterjet inlet is

presented in Chapter 6. These parameters are used in Chapter 6 and subsequent chapters

to assess and compare the hydrodynamic performance of different waterjet inlet designs.

A systematic investigation of the effect of upstream boundary layer thickness on the

flow within the waterjet inlet is then undertaken.

In Chapter 7 three two-parameter design subspaces of the parametric design hyperspace

of the author's parametric waterjet inlet design are investigated for an IVR of 0.6. In

other words, three two-parameter combinations of the generic geometry are varied with

the other parameters held constant. The large amount of hydrodynamic data obtained as

a result of these investigations allows a clear link between the hydrodynamics of the

waterjet inlet flow and the underlying geometry to be established.

In Chapter 8 the author's generic parametrically-defined waterjet inlet is optimised for

an IVR of 0.6, in order to maximise the static pressure on the surface of the waterjet

inlet. The initial and optimum geometries are compared and the hydrodynamics of their

respective flows are correlated with the underlying geometry. This further enhances

knowledge of waterjet inlet design. The conclusions of this research are summarised in

Chapter 9 which also contains a list of recommendations relating to waterjet inlet design

and further CFD research directions in this field.

17

Chapter 2 Parametric Model ofWaterjet Performance

Physically realistic and accurate mathematical models for waterjet thrust and efficiency

are crucial in determining the performance of a waterjet propulsion unit. It is also

necessary to know to what extent changes in the parameters which influence waterjet

propulsion system performance would affect overall system performance. These issues

form the basis of the work contained in this chapter. In this chapter a model for waterjet

thrust and efficiency, that can be used to estimate the performance of a waterjet

propulsion unit from experimental or CFD analysis, shall be presented.

Of the waterjet literature surveyed, van Terwisga (1996) and Coop and Bowen (1993)

provided the most detailed and physically realistic definition of thrust. In fact van

Terwisga (1996) derived a logical and comprehensive model of waterjet thrust and

efficiency and provided a good literature review of the subject. Svensson (1993) gave a

simple and yet effective expression for waterjet thrust applicable to waterjet model

testing in a cavitation tunnel. Etter et al (1982) provided a detailed analysis of waterjet

thrust based on momentum analysis, but demarcated the definition of waterjet and hull

differently from other authors. Dyne and Lindell (1994) used momentum analysis of the

frictional wake behind the vessel to derive their expression for waterjet thrust. Authors

who presented models for waterjet efficiency in the literature have included Allison

(1993), Brandau (1967) and Etter et al (1982). The ITTC 87 suggested a model based

primarily on the work of Etter et al (1980) and Haglund et al (1982).

In Section 2.1, the concept of waterjet propulsion system thrust is defined. Subsequently

a mathematical model for waterjet thrust is developed from a control volume analysis of

the conservation of mass and momentum for flow through the waterjet. The terminology

used is based on that in the aeronautical literature.

A mathematical model for waterjet propulsion system energy efficiency, derived using a

18

control volume approach for the conservation of energy for flow through the waterjet, is

presented in Section 2.2. In addition, formulae for calculating the jet velocity ratio

(JVR) corresponding to maximum waterjet efficiency (for a given set of flow parameters

and component efficiencies) is derived.

A parametric study investigating the effect on overall waterjet propulsion unit efficiency

of variation in the efficiency of sub-system components, is presented in Section 2.3. The

effect on waterjet performance of ingesting fluid from the hull boundary layer is also

examined. Realistic efficiency values for the various propulsion unit sub-system

components, reflecting the current state-of-the-art, are used in the parametric study

presented. The point of optimum operational efficiency, corresponding to the efficiency

of sub-system components and the thickness of ingested boundary layer is also

examined.

2.1 Waterjet Thrust

An accurate determination of waterjet propulsion unit thrust is essential for the

determination not only of the thrust capability of the waterjet propulsor itself, but also of

propulsive efficiency and the effect, both beneficial and adverse, of waterjet-hull

interaction.

Although waterjet propulsion is a hydrodynamic problem, it is in fact analogous to

subsonic aircraft inlet design and hence the aeronautical literature provides a good basis

for obtaining information on flow behaviour and thrust. Information on thrust definition

and subsonic inlet flows (in the aeronautical literature) can be found in Seddon and

Goldsmith (1985). It is surprising that so little reference has been made in the waterjet

literature to aeronautical publications.

The term waterjet shall be defined as the installed waterjet propulsion unit, from the

inlet duct welded to the bottom plating of the hull, to the nozzle exit plane. Any surface

in contact with water and not contained within the waterjet shall be deemed to be the

hull. In the following sections, equations for waterjet thrust are derived from the

equations of conservation of mass and momentum for a finite control-volume.

19

The analysis contained in this section is used to determine correct expressions for

propulsor thrust based on an analysis of a finite control-volume. This type of analysis is

the typical analysis approach used for determining the performance of jet engines and

waterjet propulsors. Initially, detailed expressions for propulsor thrust shall be

determined, followed by simplifications for ease of interpretation of computational or

experimental results.

Expressions for thrust are derived from Newton's Second Law applied to a finite

control-volume. Conceptually, forces due to pressure and shear acting over the surface

of the control volume, in addition to body forces acting on the fluid within the control­

volume, effect a net change in momentum of fluid entering and leaving the control­

volume. In a conventional jet engine or waterjet propulsor, it is only those solid control

volume surfaces in contact with the fluid that can transmit force to the aircraft or ship.

Magnetohydrodynamic propulsors work on a different principle where the thrust is

transmitted to the ship through the magnet structure, the thrust itself being generated

from the interaction of the magnetic field with ions in the propulsor duct.

2.1.1 Definition ofWaterjet Control-Volume for Thrust Analysis

A control-volume (ABCEFGA) encompassing the waterjet unit, impeller shaft housing,

impeller and a streamtube of fluid external to it, is shown in Fig. 2.1. Area A5 represents

all surfaces inside the waterjet unit exposed to water. This includes the waterjet inlet

duct, impeller shaft housing/fairing, impeller, nozzle and stator vanes. It must be noted

that these surfaces are the only means by which force can be transmitted to the vessel.

Area A2 represents a streamtube surface through which no transport of mass occurs, thus

dividing the flow entering the waterjet unit from the flow by-passing it. The location of

Area At is arbitrarily chosen. Van Terwisga ( 1996) specified the location of AB at a

distance of 10% of the physical intake length DG. Since area A3 represents the portion

of the inlet streamtube in contact with the hull, the location of AB will affect the shear

force acting on A3 and velocity profile over At. For reasons of simplicity, it is therefore

beneficial to locate AB as close to the inlet as possible, but as far upstream as is

necessary to avoid the major effects of the pressure field caused by the inlet ramp. This

may be greater than 10% of the length DG, depending on operating conditions. Point C

20

is representative of a series of points defining the stagnation line along the inlet lip. The

location of the stagnation line and the geometry of areas A1 and Az are dependent on the

geometry of the waterjet inlet, the IVR and the external flow conditions.

The flow is discharged through the nozzle, area A6. Area ~ is an imaginary surface

bounded by the waterjet inlet and representing the location where the flow is first fully

bounded by the waterjet inlet. In the forward part of the inlet ~ will lie in the plane of

the inlet opening, whereas near the inlet lip, A4 will rise above the plane of the inlet

opening to the location of the stagnation point C. Point D may be referred to as the

"trailing edge" of the lip profile. Area A7 encompasses the inlet duct area below ~ and

is therefore bounded on top by the inlet streamtube. This area is dependent upon the

IVR and the external flow conditions and being part of the waterjet ducting, it provides

a means of transmission of force to the hull.

A G

~ v2 -----L ~--------l---

AI A2

Fig. 2.1 Control-volume for thrust analysis

2.1.2 Thrust Relationships

Some of the terminology used in the following thrust analysis shall be based on

aeronautical terminology. In order to develop appropriate formulae for waterjet

propulsion unit thrust, it is essential to begin with the relevant conservation equations of

mass and momentum for a finite control-volume.

Conservation of mass:

!P (U · n)dA = o jAcv

(2.1)

21

Conservation of linear momentum:

,( pU(U · fi)dA = -JpdA- J:rdA + f pgV jAcv Acv Acv Jv1+V2

(2.2)

where p is the density of water, D the velocity vector, fi the unit normal vector of the

differential control-volume surface dA (pointing in a direction normal and outward from

the control-volume), p the static pressure, 1 the shear stress vector and V is volume.

Subscript CV denotes control-volume. The first two terms on the right of Eqn 2.2

represent the forces acting on the surface of the control-volume due to pressure and

shear force, respectively. The third term on the right of the equation is a body-force term

resulting from the weight of the water inside the control-volume.

For the control-volume shown in Fig. 2.1, conservation of mass (Eqn 2.1) yields

f p(U · fi)dA = f p(U · fi)dA A6 At

Defining the following quantities:

1) Momentum flux ( M1

) through control volume surface i

2) Surface pressure force ( FPI) acting on fluid over surface i

3) Surface shear force ( F"tl) acting on fluid over surface i

4) Gravitational body-force (Fv1

) acting through volume i

Eqn 2.2 may therefore be expressed as

1=1 1=5 1=2

In the waterjet literature, the thrust ('T) acting on the waterjet is often defined as

T = - (M6 - M I) . fi6 fi6

(2.3)

(2.4)

(2.5)

Van Terwisga (1993) and Svensson (1991, 1993) correctly called the expression for T given by Eqn 2.5, the gross thrust. We shall use the term "thrust" here as being the

propulsive force acting on the waterjet propulsion unit itself, rather than the ship per se.

On some commercially-available waterjet propulsion units, the majority of the force is

taken by the shaft and transmitted to a large axial bearing outside the waterjet. Hence the

22

force in the direction of the negative of the nozzle area vector provides a convenient

direction for describing the net propulsive force acting on the waterjet propulsion unit,

as this vector is generally parallel to the shaft centreline. Gross thrust is not, in fact, the

actual thrust generated for the waterjet under consideration. Seddon and Goldsmith

(1985) discussed the various thrust definitions used in the aeronautical literature. The

practical thrust definition in the aeronautical literature is termed the net standard thrust

( T Ns ) • A revised definition of net standard thrust, accounting for body forces (weight of

entrained water) acting inside the control-volume and thus relevant to waterjet

applications, is

(2.6)

It must be noted that in the aeronautical literature FP1 is taken as being equal to ambient

pressure. Since the thrust analysis presented here is applicable to a waterjet unit installed

as part of a hull system, FP1 is non-zero in value due to hydrostatic and hydrodynamic

forces and therefore must be included in the definition of net standard thrust.

The actual expression for the force (thrust) transmitted to area A5, called the intrinsic

thrust, is defined by

Tlntrlnslc = -(Fps + F'ts). n6 n6 = -(M6- M4- (Fp4 + Fp6)- FYI). n6n6 (2.7)

Intrinsic thrust may also be written as

TlntrmSIC = -(M6- Ml- (Fp6 + Fpl + Fp2 + Fp3)- (F't2 + F't3)- FYI- Fy2). n6 n6 (2.8)

The difference in thrust between net standard thrust and the intrinsic thrust, is termed

the "Pre-entry Drag" CDPre) and may be evaluated as

Eqn 2.9 states that the Pre-entry Drag is primarily a function of the geometry of the inlet

streamtube and may be either positive or negative depending on whether the streamtube

pre-diffuses or contracts prior to entry into the propulsor. It is clear that the

determination of net standard thrust, intrinsic thrust and pre-entry drag is by no means

simple, requiring extensive CFD analysis or experimental investigation.

23

While Eqn 2.7 specifies the actual thrust transmitted to the hull via the propulsor, for the

control-volume under consideration, this is not the total thrust force acting on the

propulsor. There is also a force generated by the pressure distribution over area A7. The

total thrust (r Net ) generated by the propulsor is then

(2.10)

with an inlet drag ( Diniet) defined as

(2.11)

Neglecting body forces, Seddon and Goldsmith (1985) stated (for subsonic inlets) that

the drag of the inlet is composed of three components. These are the pre-entry drag,

frictional drag and the pressure drag on the cowl. In an inviscid flow condition, the pre­

entry drag is balanced by suction on the cowl lip. In practice, because the boundary layer

on the cowl displaces the potential flow and there are skin frictional effects, the suction

of the cowl lip is insufficient to balance the pre-entry drag and so there is a net drag. In

the case of the waterjet, area A7 is analogous to the inlet cowl of a subsonic aircraft. It

can thus be expected, in the case of inviscid flow, that inlet drag will be zero. The work

of van Terwisga ( 1996) supports this conclusion. In this case Eqn 2.11 becomes

(2.12)

Substitution of Eqn 2.12 into Eqn 2.10 and using the definition of pre-entry drag from

Eqn 2.9 gives

(T Net ) Invtsctd = {T Intnns•c + f) Pre ) lnv•sc•d = T NS (2.13)

It is therefore clear that in the case of in viscid flow, the net thrust produced by the

waterjet is simply equal to the net standard thrust. In the case of viscous flow, the net

thrust becomes

TNet = {TNet )lnvJscJd + {Ft2 + Ft3 - Ft7) · ff6 °6 = {TNS )corrected - Ft7 · ff6 ff6 {2.14)

where

(2.15)

is the corrected net standard thrust, that is the net standard thrust that would be obtained

if AB were moved close to the ramp tangency point. Eqn 2.15 explicitly deals with the

drag contribution of the inlet stream-tube outside of the waterjet. Eqn 2.13 provides a

close approximation to thrust if the pressure losses on A7, A2 and A3 are small.

24

If lip pressure losses due to fluid viscosity (or changes due to cavitation) become

significant, the deviation between net thrust and corrected net standard thrust can be

accounted for by the introduction of a lip loss thrust deduction factor ( 1-q_J, defined by

(2.16)

where tL is the lip loss thrust deduction fraction. The significance of this parameter is

that it gives an indication as to the magnitude of the lack of pressure recovery on the

inlet lip outside of the inlet streamtube. Lip loss thrust deduction factor defined by Eqn

2.16 is similar in form to the jet thrust deduction factor as proposed by van Terwisga

( 1996), but differs in its definition and significance.

Since many authors use gross thrust as their definition of thrust, without taking into

account pre-entry drag or the force contribution of area A7, the effects of inlet drag

inevitably appear in the thrust deduction factor (1-t) defined by

(1- t) = IR.BH 1/ltl (2.17)

where R BH is the bare-hull resistance vector and T the gross thrust vector defined by

Eqn 2.5. With the nozzle and A1 of the inlet streamtube at ambient pressures and, low

lip losses, gross thrust will provide a close approximation to the waterjet thrust.

2.1.3 Waterjet Thrust Model

The equation for the net thrust of a waterjet (Eqn 2.14) was derived in Section 2.1.2. In

order to develop a model for waterjet thrust, it shall be assumed that the direction of the

area vectors representing the nozzle throat area (A6) and the control-volume inlet area

(A1) are parallel to each other and parallel to the horizontal direction. This

simplification eliminates the body force terms in the expression for net thrust.

A simplified expression for net thrust can be derived using the definition of net standard

thrust (Eqn 2.6), corrected net standard thrust (Eqn 2.15), net thrust (Eqn 2.14) and lip

loss thrust deduction factor (Eqn 2.16). The net thrust (T) of the waterjet may now be

written as

(2.18)

where UJ is the average jet velocity through the nozzle throat (with the assumption that

25

the vena contracta of the flow is at the nozzle exit plane) determined on a

volumetrically-averaged basis and :rh the mass flow-rate through the waterjet. The skin­

friction force caused by the flow of water on the solid surface A3 is accounted for by the

term incorporating the skin-friction coefficient (Cr) in Eqn 2.18. Pressure losses at the

lip reduce the lip suction and hence the thrust. They are therefore accounted for through

the lip loss thrust deduction factor. The nozzle momentum flux coefficient (Cmn)

accounts for the change in nozzle momentum due to flow non-uniformity over the

nozzle throat and may be defined as

cmn = . 1 Jp(U·n6 )

2dA

(mU) A6

(2.19)

where dA is the differential area and n6 is the unit direction vector of differential area

of the nozzle. The upstream momentum flux coefficient (Cm) accounts for the reduction

in ingested momentum resulting from the ingestion of fluid from an upstream boundary

layer and may be defined as

(2.20)

The negative sign in Eqn 2.20 is required to make Cm a positive quantity, since n1 is

negative. An average static pressure coefficient ( CP) can be defined where

(2.21)

In Eqn 2.21, Prer is the reference static pressure. An average skin-friction coefficient

(2.22)

may be defined to account for the viscous force on A3.

Eqn 2.18 may then be rewritten as

T= (1-tL)(m(CmnUJ -CmUs)+tpu;(CrA3 -CPA 1)) (2.23)

The pressure coefficient term can be written in terms of mass flow-rate so that Eqn 2.23

becomes

26

(2.24)

where CA and C8 are the ratios of A3 and A1 to the streamtube cross sectional area of A 1

under free-stream conditions, respectively.

2.2 Propulsive Efficiency

The development of a theoretical model of waterjet efficiency is outlined below. This

model can be used in conjunction with CFD or experimental results to quantify the

changes in waterjet efficiency resulting from changes in parameters that affect

performance.

With the increasing use of waterjet propulsion as the preferred means of providing high­

speed marine propulsion, there is a need to clarify waterjet -hull interaction and

efficiency in a systematic manner. A good mathematical model of waterjet efficiency

should therefore be straightforward to use and should relate the performance of the

waterjet to the performance of individual components of the waterjet propulsion unit

(such as the inlet, pump and nozzle) and the flow conditions upstream of the inlet. This

must be done in a physically correct and consistent manner. The model presented in this

thesis meets all these requirements and can be readily applied to the interpretation of

experimental or CFD results.

2.2.1 Control-Volume Definition for Efficiency Analysis

In order to properly analyse the waterjet propulsion unit, it is necessary to define a

suitable control-volume encompassing the flow through the waterjet. The defined

control-volume (ABCDEFGHA) is shown in Fig. 2.2. This control-volume encompasses

the inlet streamtube, waterjet inlet, impeller and other surfaces in contact with water

contained in the waterjet ducting. Station 1 represents an upstream cross-section of the

inlet streamtube, Station 2 the pump inlet, Station 3 the pump exit and Station 4 the

nozzle throat.

27

Station 3 - Pump Outlet -

Station 2 - Pump Inlet

......::;-- --c D

A

Station 4 - Nozzle Throat

Fig. 2.2 Control-volume for efficiency analysis

2.2.2 Development of a Model for Propulsive Efficiency

The efficiency of a waterjet is often defined in the literature as

(2.25)

where 1lw1 represents the waterjet system efficiency, TNet the net waterjet thrust vector,

D. a vector representing the ship's velocity, W the power output of the vessel prime

mover driving the waterjet pump and TIT the efficiency of power transmission from

prime mover to waterjet pump. Waterjet system efficiency may be also written as

TJ WJ = TJ Rot TJ Pump TJ Duct TJ I (2.26)

where TJPump is the efficiency of the waterjet pump at the required operating point with

spatially uniform inflow conditions, TJRot is the rotative efficiency of the waterjet pump

as installed in the waterjet, TJouct the efficiency of the total waterjet ducting and 111 the

ideal waterjet efficiency derived from momentum considerations as

TII = 2U. I (V. + V) (2.27)

The propulsive efficiency of the total waterjet-hull system, on the other hand, can be

written as

28

(2.28)

In Eqn 2.28, R8H is the bare-hull resistance vector and (1-t) is a thrust deduction factor

linking the waterjet thrust and the bare-hull resistance. It is clear that the value of t will

depend on how the operation of the waterjet modifies the flow around the vessel hull.

For the derivation of a mathematical model for waterjet efficiency, it will be assumed

for simplicity that the vectors for thrust, resistance and ship's velocity are parallel in

direction. Using the result of Eqn 2.24, thrust (T) may be calculated as

(2.29)

A mass-averaged total pressure (P) can be defined as

P= ~L (p+pgH+tpU 2 )dm (2.30)

where p represents static pressure, g the gravitational constant and H is the elevation

above a given reference point. The energy flux (E) may be then expressed as

E= mP p

(2.31)

which is similar in essence to that provided by Haglund et al (1982). The inlet efficiency

can be defined as

(2.32)

and the nozzle efficiency as

(2.33)

Ingestion of fluid from an upstream hull boundary layer will result in the total energy of

the flow passing through these cross-sections being different from the case of uniform

flow. The energy at Station 1 may be expressed as

E -I . C uz AB- 2m e s

where ce is the inlet energy flux coefficient defined by

(2.34)

(2.35)

The negative sign appearing in Eqn 2.35 is required to make Ce a positive quantity, as

29

i'i 1 is negative. A nozzle energy flux coefficient (Cen) can be similarly defined. Waterjet

efficiency can now be determined as

TUS (2.36)

or in other words,

(2.37)

Defining J..L=U/Us as the jet velocity ratio (JVR), waterjet efficiency can be written as

In terms of JVR, ideal efficiency (Eqn 2.27) becomes

Tli = 2/ (1 + J..L)

(2.38)

(2.39)

Waterjet duct efficiency can be determined from Eqn 2.26 by substitution of Eqn 2.38

and Eqn 2.39. The result is

(2.40)

It can be seen from Eqn 2.40 that the duct efficiency of the waterjet is a non-linear

function of JVR and correspondingly of IVR. Duct efficiency is also dependent on the

flow behaviour upstream of the waterjet inlet. Inlet efficiency and nozzle efficiency

themselves are not fixed in value but vary as functions of IVR and JVR. Seddon and

Goldsmith ( 1985) discussed the variation of inlet efficiency with IVR for subsonic

aircraft intakes. To assign a fixed value of duct efficiency over a range of IVR of interest

is clearly erroneous. Similarly, since inlet efficiency will change with IVR, it may be

expected that the pump rotative efficiency will change as a result of the different flow

conditions at the duct exit of the waterjet inlet, therefore affecting propulsive efficiency.

The optimum value of JVR at which waterjet efficiency is maximised can be found be

setting the derivative of Eqn 2.38, with respect to JVR, equal to zero. The resulting

expression is

30

cmncen 2 2Cen I - I -f..l ---(Cm -zCACr +zCsCp)f..l+CmnCellin =0

llN llN (2.41)

The optimum value of JVR is therefore

(2.42)

where A, B and C are the coefficients of the polynomial described by Eqn 2.41 such that

(2.43)

2.3 Parametric Study of W aterjet Efficiency

The development of a model for waterjet thrust and efficiency was presented in Section

2.1 and Section 2.2, respectively. This model incorporates all of the relevant physics

associated with the performance of the waterjet propulsion unit. In order to understand

how changes in model parameters affect the performance of the waterjet, it is necessary

to conduct a parametric study in order to gauge the relative importance of these

parameters.

The parametric study is presented in two parts. The effect of the variation in component

efficiency and upstream boundary layer thickness on the efficiency of the waterjet is first

examined in Section 2.3.1. The optimum value of JVR at which waterjet efficiency is

maximised (for variations in inlet and nozzle efficiency), with different amounts of

boundary layer ingestion will be examined in Section 2.3.2.

For the purposes of investigating the effect of the ingestion of fluid from the hull

boundary layer, it has been assumed for simplicity that the cross-section of the inlet

streamtube is rectangular, thereby allowing the boundary layer to be treated as two­

dimensional. In reality the cross-section of the inlet streamtube is more of an

ellipticaVtrapezoidal shape as Griffith-Jones (1994) and Roberts (1998) showed. In this

case evaluation of the inlet momentum flux coefficient (Cm) and the inlet energy flux

coefficient (Ce) would require an integration of the boundary layer velocity and pressure

profile over the cross-sectional area of the inlet streamtube. The assumption of a two­

dimensional profile would be more realistic when wide flush-type inlets are used as

noted by Purnell (1976).

31

With the simplifying assumption of a rectangular inlet streamtube cross-section and a

two-dimensional boundary layer profile, Cm and Ce may be evaluated as

1 Jh 2 n (h)~ Crn =- (y I 8) 11 dy = - -h o n+2 8

(2.44)

1 Jh ; n (h)~ ce =- (y/8) 11dy=--h o n+3 8

(2.45)

when the depth of the inlet streamtube (h) is less than that of the boundary layer

thickness (8). The power law exponent (n) is a function of the Reynolds number of the

flow. For the flows under investigation here, n is taken as 9 according to Steen and

Minsaas ( 1995). When the depth of the inlet stream tube is greater than the boundary

layer thickness, the following expressions for Cm and Ce are used

(2.46)

(2.47)

The dimensionless ratio h/8 provides a measure as to the amount of boundary layer fluid

ingested. Values of h/8 less than unity correspond to the ingestion of fluid exclusively

within the boundary layer. In this case the waterjet may be considered to be operating in

a "thick" boundary layer. On the other hand, when h/8 is greater than unity, in particular

h/8>>1, a "thin" boundary layer is being ingested. The extreme case of hl8=oo

corresponds to no boundary layer ingestion. In reality, this is of course impossible, but

the values of h/8 may become large depending on how thin the boundary layer is relative

to the depth of the inlet streamtube.

In all of the figures presented, the lower value of JVR is limited to unity, as it is

impossible for the waterjet to produce thrust at JVR values less than or equal to unity. It

should be noted that the parametric results presented are also applicable to waterjets

with ram-type inlets. Values for the model parameters used in the parametric analysis

are shown in Table 2.1.

32

Figure Tlln llN TlPump TlRot (1-tL) Cm Ce Cmn Cen 2.3 Tlln 0.98 0.9 1.0 1.0 1.0 1.0 1.0 1.0 2.4 0.83 llN 0.9 1.0 1.0 1.0 1.0 1.0 1.0 2.5 0.83 0.98 0.9 TlRot 1.0 1.0 1.0 1.0 1.0 2.6 0.95 0.98 0.9 1.0 1.0 Cm Ce 1.0 1.0 2.7 'Tlln llN 0.9 1.0 1.0 1.0 1.0 1.0 1.0 2.8 Tlln 0.98 0.9 1.0 1.0 Cm Ce 1.0 1.0

Table 2.1 Values used in parametric analyses

The fixed values of inlet efficiency (0.83) and nozzle efficiency (0.98) are taken from

Verbeek (1992) and represent realistic values for current large waterjet installations. The

pump efficiency of 0.9 is taken from Allison (1993). The author believes that the value

of inlet efficiency quoted by Verbeek implicitly includes the effect of the ingestion of a

thin boundary layer rather than separating this effect viaCe (c.f. Chapter 6).

2.3.1 Parametric Analysis with Boundary Layer Ingestion

Fig. 2.3 shows the variation of waterjet efficiency with inlet efficiency over a range of

JVR values. It can be seen from the figure that there is a general trend of decreasing

waterjet efficiency with decreasing inlet efficiency. At larger values of JVR, after the

point of maximum waterjet efficiency, the sensitivity of the waterjet efficiency to

changes in inlet efficiency becomes less. The optimum value of JVR at which peak

efficiency is reached increases as inlet efficiency decreases. It is interesting to note that

in the case of "no boundary layer ingestion", the peak efficiency of the waterjet for an

inlet efficiency of 83% is close to 61%.

The variation of waterjet efficiency with nozzle efficiency over a range of JVR is shown

in Fig. 2.4. As would be expected, the general trend shows decreasing waterjet

efficiency with decreasing nozzle efficiency. Although it is not apparent from the above

figures, each percentage point of lost nozzle efficiency translates into a larger decrease

in overall waterjet efficiency than the equivalent loss of inlet efficiency. This is to be

expected since the nozzle carries water with a higher head and so there is greater scope

for losing energy in the nozzle for each percentage point of lost efficiency. It is therefore

crucial to keep viscous losses in the nozzle to a minimum. This does not of course

undermine the importance of keeping waterjet inlet losses to a minimum and ensuring as

33

uniform a flow as is possible at the exit of the waterjet inlet in order to maximise the

rotative efficiency of the pump (llRot).

08

07

06

05

~ 04 ~

03

02

0 I

00 0 0 "' N

TIIn -1000 --0950 ---0900 ---···0850 --0800

-+-0750 ----- 0 700

0

"' JVR

8 N

Fig. 2.3 Variation of waterJet efficiency with waterJet mlet efficiency

~ ~

07

06

05

04

03 I 000 --0975 --- 0950

02 ••••• ·0 925 --0900

0 I -l--0875 ----- 0 850

00 0 "' 0 "' 8 "' 0 0 N "' r- N "' - - N N N

JVR

Fig. 2.4 Variation of waterjet efficiency w1th nozzle efficiency

The effect of reduced pump performance on waterjet efficiency, as a result of operating

the pump in a non-uniform flow field is shown in Fig. 2.5. It can be seen that small

decreases in pump rotative efficiency translate to significant decreases in waterjet

efficiency. This is not surprising as the pump rotative efficiency term (llRot) appearing in

Eqn 2.26, multiplies all other terms in the equation. In physical terms, a non-uniform

flow at the pump inlet can result in a number of possible mechanisms that may reduce

llRot . These may include variations in the local angle of attack of the impeller blades

relative to the flow, the possibility of impeller blade stall and impeller blade cavitation.

In Fig. 2.3 it has been assumed that the rotative efficiency of the pump remains constant

with JVR. In reality, it must be acknowledged that as JVR changes, the changed flow

conditions in the inlet and hence at the duct exit, are likely to change the rotative

efficiency of the pump. The same statement can be made for the assumptions underlying

the construction of Fig. 2.4.

The effect of ingesting hull boundary layer fluid, on the efficiency of the waterjet is

shown in Fig. 2.6. It can be seen from Fig. 2.6 that as h/8 decreases (greater ingestion of

boundary layer fluid), the waterjet efficiency increases. The increase in efficiency with

34

decreasing h/8 becomes less pronounced as JVR increases, however. For some values of

h/8, the waterjet efficiency exceeds unity! While initially this may appear to be

physically impossible, it is a consequence of the definition of the system (control

volume) under consideration, rather than a violation of physical reality. The definition of

waterjet efficiency presented herein treats propulsor/hull interaction effects, such as a

hull boundary layer or pressure field, upstream of the waterjet inlet (expressed through

momentum and energy flux coefficients) as contributions to waterjet efficiency, rather

than as separate interaction factors. Van Terwisga ( 1996), for example, treated

propulsor/hull interaction effects as contributions to overall vessel propulsive efficiency

rather than as contributions to waterjet efficiency. Hence the values of waterjet

efficiency predicted by his mathematical model are always less than unity.

07

06

05

04 ~

!=' 0 3

02

0 I

oo~~~~~~~~~~~~~y

8

Fig. 2.5 Variation of waterjet efficiency with pump efficiency

14

12

I 0

0 8 T ..t'T"fO+o.&._o_' ~ ~ .. -.,...:-~~~

!=' 06 1

I I

04 I

02

h/8 ---o 25 --o so - - - 0 75 - - - - - - I 00 --150 --+-300

----- NoBL

JVR

Fig. 2.6 Variation of waterjet efficiency with boundary layer ingestion

The physical explanation for the increase in waterjet efficiency with the ingestion of

fluid from a hull boundary layer is as follows. As a vessel moves through the water a

boundary layer grows on the hull due to the viscosity of the water. Energy is therefore

expended in the shear layer that the hull boundary layer represents, resulting in the flow

in the boundary layer being retarded relative to the "free-stream" flow. Water must flow

into the inlet in order to feed the waterjet pump. If this water is fed from the "free­

stream" condition, there is a greater momentum drag associated with the inlet flow. If on

the other hand, water is drawn from the hull boundary layer, a smaller inlet momentum

drag results. The consequence of this is therefore an increase in thrust and an increase in

35

waterjet efficiency. The ingestion of hull boundary layer fluid in essence "recovers"

energy lost in growing the boundary layer.

It has been assumed, in the above discussion, that the ingestion of hull boundary layer

fluid does not affect the rotative efficiency of the pump in such a way that the potential

for increased efficiency (by virtue of boundary layer ingestion) is counteracted by a

decrease in the rotative efficiency of the pump. From Svensson (1994), it appears that

the effect of boundary layer ingestion is beneficial as the propulsive efficiencies

appearing in Figure 3 of Svensson's paper are significantly larger (in the upper vessel

speed ranges) than for the case of negligible boundary layer ingestion, even allowing for

thrust deduction.

17.-----------------------~

16

I 5

~ 14 ~

> ...., I 3

12

II

8 V'l a.. 0

0 a.. 0

'I") 00 0

llin

TJN -1ooo --0975

--- 0950

----- -o 925 --0900

- -0875 ----- 0 850

0 'I") 0 00 ,.... ,.... 0 0 0

Fig. 2.7 Optimum JVR (neghgible upstream boundary layer thickness)

2.3.2 Points of Maximum Efficiency

llin

Fig. 2.8 Optimum JVR (upstream boundary layer)

The optimum value of JVR at which waterjet efficiency is maximised when fluid from a

boundary layer of negligible thickness is ingested is shown in Fig. 2. 7. It can be seen

that as the efficiency of the inlet decreases, the optimum value of JVR increases.

Similarly as the efficiency of the nozzle decreases, the optimum value of the JVR also

increases.

Fig. 2.8 shows the variation of optimum JVR (for a fixed nozzle efficiency) with inlet

efficiency for different values of hi~. It is evident that the optimum value of JVR

decreases with increased ingestion of hull boundary layer fluid, but increases as the

efficiency of the inlet decreases. The latter trend reflects that of Fig. 2. 7. For larger

36

values of inlet efficiency there may not be an optimum JVR, but rather a continuous

decrease in efficiency with increasing JVR. Fig. 2.6, for example, shows a continuous

decrease in waterjet efficiency with increasing JVR over the range shown for h/0=0.6. If

a waterjet is to operate at its peak efficiency, then the selection of the JVR must take

into account the depth of the inlet streamtube relative to the thickness of the upstream

hull boundary layer.

2.4 Closure

Physically realistic, accurate and simple models for waterjet thrust and efficiency are

crucial to determining waterjet performance and for providing a framework for the

proper interpretation of the results of experimental or CFD analysis of the waterjet

propulsion system. Therefore, a mathematical model for the thrust and efficiency of a

waterjet propulsion unit was derived from control-volume analyses of the conservation

of mass, momentum and energy for flow through a waterjet propulsion unit. A

parametric study of the mathematical model for waterjet efficiency has been undertaken

in Section 2.3 in order to investigate the parametric model.

The following conclusions can be drawn from the work presented in this Chapter:

1) Flow separation (spillage), cavitation and boundary layer growth on the inlet lip

create drag, by virtue of reduced lip suction. In the parametric model, a lip loss thrust

deduction fraction tL is introduced to account for this effect.

2) Theoretically, the ingestion of fluid from an upstream hull boundary layer results in

increased waterjet efficiency. This is primarily a result of increased propulsor thrust

resulting from a lower momentum drag associated with the flow into the inlet. This

assumes that the ingestion of boundary layer fluid does not degrade pump

performance and so counteract the beneficial effect of ingesting fluid from an

upstream boundary layer.

3) There is a reduction in the optimum jet velocity ratio (JVR) as the upstream boundary

layer increases.

4) The pump rotative efficiency has a significant effect on overall waterjet efficiency for

a given JVR. This indicates the importance of ensuring that the quality of the flow

delivered to the waterjet pump does not have an adverse effect on pump performance.

37

Chapter 3 Computational Fluid Dynamics Modelling

In the literature there have been several approaches to the computational simulation of

waterjet inlets. These have ranged from potential flow methods, such as the method of

Hess and Smith (1964) used by Kashiwadani (1997), to the solution of the Reynolds­

averaged Navier Stokes (RANS) equations with turbulence modelling which Seil et al

( 1995) used. Some workers have used an interesting combination of computational

techniques. F!1Srde et al ( 1991 ), for example, used a combination of the time-dependent

form of the Euler equations with thin-layer RANS equations to calculate the flow in both

ram and flush-type waterjet inlets.

Several authors such as Kashiwadani (1997), have used potential flow methods as

preliminary analysis tools to examine the pressure distribution over the surface of the

waterjet inlet in the immediate vicinity of the inlet. This approach allows the likelihood of

cavitation and flow separation to be determined. Kashiwadani's results were in reasonable

agreement with experimental data possibly due to the thin nature of the boundary layers on

the inlet ramp resulting from his scoop-type inlet. Van Terwisga ( 1996) compared

potential flow predictions with experimental data for flow within a waterjet inlet model.

He found that the use of the potential flow methodology gave relatively large inaccuracies

in the predicted static pressure distribution over the inlet lip and ramp. The discrepancies

can be primarily attributed to the neglect of boundary layer growth within the inlet. Large

differences were evident at the inlet lip as a result of an incorrectly calculated stagnation

line resulting from the neglect of boundary layer effects.

Dai et al ( 1995) and Steen and Minsaas ( 1995) have used potential flow when seeking to

"optimise" an inlet design by examining a large number of prospective designs. These two

sets of authors, however, complement their potential flow analyses with RANS

computations for designs which appear to be promising. This was done in order to provide

38

more accurate computations accounting for the effects of viscosity on flow behaviour.

From the results of van Terwisga ( 1996), it is evident that the assumptions underlying the

use of potential flow computations are clearly violated for conventional flush-type waterjet

inlets, which have relatively shallow inlets and large internal boundary layer growth.

Furthermore, the assumptions of potential flow theory become irrelevant in the presence of

the large boundary layer thicknesses developed on high-speed catamaran ferries. For large

flush-type waterjet inlet installations, upstream boundary layer thicknesses must be

accounted for in the design of these inlets. Therefore computational techniques based on

the solution of the RANS equations offer the only realistic means of solving the flow in

the vicinity of the inlet and inside the waterjet inlet itself.

fu the literature there has been some work towards the computational simulation of the

complete waterjet-hull geometry. Yang et al (1995) used a boundary element method to

calculate the flow around a mathematical hull form in order to provide boundary

conditions for a RANS computation of the flow in a waterjet inlet. Viscous effects

associated with the hull boundary layer development were neglected and this is a clear

limitation of their methodology. Widmark and Gustafsson (1997) of SSPA, used that

organisation's SHIPFLOW code to provide boundary conditions for a RANS simulation

of flow in a waterjet inlet. SHIPFLOW used a combination of potential flow, boundary

layer integral and RANS methodologies to calculate the flow around the vessel hull. No

experimental validation of the methodology was presented in the paper, so the accuracy of

the total methodology could not be determined.

The optimisation of the hydrodynamic design of waterjet inlets is an applied problem,

where the emphasis is on design investigation and optimisation. As a consequence, the

choice was made to use commercially-available CFD software. This allows attention to

be focused on the practicalities of waterjet inlet design investigation and mesh

generation, with CFD used as an analysis tool. This is analogous to the way in which

modem industry uses finite element structural analysis as an analysis tool in the design

of machine parts.

Since interest is directed at the design of the waterjet inlet itself, in the absence of hull

39

form geometry, a simplified flow domain external to the inlet is chosen. This makes the

meshing of the waterjet inlet and flow domain external to the inlet amenable to the

single-block structured grid topology described in Chapter 4. As a result, a structured

mesh can be used to mesh the complete flow domain and consequently a structured CFD

solver can be used to simulate the flow internal and external to the waterjet inlet. This is

a primary reason for the choice of the commercially-available CFD software Fluent. The

good international reputation of this software also favours its use.

An overview of the CFD methodology and techniques that form the foundation of the

numerical work presented in this thesis will be given in this chapter and is applicable to

obtaining flow solutions on body-fitted-coordinate (BFC) structured meshes. The RANS

equations are derived from the isothermal and incompressible form of the continuity and

Navier Stokes equations in Section 3.1. In Section 3.2, the Reynolds stresses that appear

in the RANS equations are modelled using two-equation turbulence models based on

the hypothesis of Boussinesq (1877), where mean velocity gradients are related to the

Reynolds-stresses via an eddy-viscosity formulation. Two, two-equation turbulence

models based on the transport of turbulent kinetic energy (k) and turbulent dissipation

(E) are described. These are the Standard k-E model of Launder and Spalding (1974) and

the RNG k-E model of Yakhot and Orszag (1986) which was extended by Yakhot et al

(1992) to account for the effects of irrotational strain on the turbulence. The relative

merits and limitations of these turbulence models are then discussed. In Section 3.3 the

structure of the turbulent boundary layer is presented and the reasons why wall functions

are used to model the near-wall behaviour of the flow are discussed. Details of the wall

functions used in this work are subsequently given.

The finite-volume discretisation of the governing transport equations is discussed in

Section 3.4. In order to complete the discretisation process, it is necessary to employ

suitable convective differencing schemes to relate the values of the transported quantity

<j> at the control-volume faces to their surrounding finite-volume cell-centre values. This

forms the subject of Section 3.5. The nonlinearity of the governing transport equations

and the present state of computer hardware necessitate the solution of the discretised

equations by iterative means. The manipulation of the discretised equations into a form

40

suitable for use with iterative solution techniques is discussed in Section 3.6. The CFD

methodology used by the author is summarised in Section 3.7. The Dirichlet and

Neumann boundary conditions used in the computational simulations presented in this

thesis will be discussed in later chapters in the context of the description of those

simulations.

3.1. Governing Flow Equations

The flows studied in this thesis may be considered to be steady, isothermal and

incompressible. As a consequence of this, the continuity and Navier Stokes equations

provide a full description of the flow behaviour.

3.1.1 Navier Stokes Equations

The continuity equation and Navier Stokes equations express Eulerian conservation

relationships for mass and momentum respectively. According to Fletcher (1987b,

p.10), the coordinate invariant form of the Navier Stokes equation is

_i_(pu) + (u · V)u = pf- Vp-2

V(J.!V · u) +2V · (J.!D) at 3

(3.1)

In Eqn 3.1 V represents the gradient operator, u the instantaneous fluid velocity, Jl the

molecular viscosity of the fluid, p the fluid density, p the pressure, f the volume force

per unit mass and D is the deformation (rate of strain tensor) whose components can be

written in Cartesian tensor notation as

D,,=~(~: +~:J (3.2)

Using Cartesian tensor notation, the steady-state, incompressible form of the continuity

and N a vier Stokes equations may be written as

a -a (u)=O x.

(3.3)

a a a = p-(uJu.) = _ _£_+2-(J.!D)+pf.

axJ ax. axJ (3.4)

Eqn 3.3 is the continuity equation and Eqn 3.4 is the Navier Stokes equations,

respectively (see Wilcox (1993)). In the above equations, U1 is the instantaneous velocity

and f. the body force per unit mass in the ith coordinate direction (x1), respectively.

41

Following the conventions of tensor notation, repetition of the same index implies a

summation over the range of that index.

As the Reynolds number of a flow increases, the ability of the fluid's viscosity to

suppress instabilities within the flow (and so maintain laminar flow) decreases.

Eventually, at a sufficiently high Reynolds number, the flow becomes unstable and is

then characterised by random fluctuations of the flow quantities. Turbulence is therefore

a random, dissipative, time-dependent phenomenon displaying a large range of excited

length and time-scales. These range in size from the integral scales of the turbulence

down to the Kolmogorov microscale as was discussed by Tennekes and Lumley ( 1972).

The smallest scales of the turbulent fluctuations are many orders of magnitude smaller

than the largest scales present in the flow. Inherent in turbulent flow is a cascading

process whereby, as the turbulence decays, its kinetic energy transfers from larger eddies

to smaller eddies, this being achieved mainly by the process of vortex stretching as was

noted by Wilcox (1993). The apparent fluid stresses that are developed within the

turbulent flow are several orders of magnitude larger than in a corresponding laminar

flow. Since turbulence is a continuum phenomenon, the three-dimensional Navier

Stokes equations contain all of the physics necessary to describe that flow.

3.1.2 Reynolds-averaged Navier Stokes Equations

Turbulence is a random phenomenon which must be handled using a statistical

approach. Therefore, the time averages of the continuity and Navier Stokes equations

are sought. The instantaneous velocity at a point in the flow (u1) can be considered as

being composed of a mean (average) component (U1) and a fluctuating component ( u; ),

such that

The mean velocity is defined by

where T is a characteristic time-scale. The time average of the mean velocity t+T

- 1 J ul = lim - ul dt = u I T-7~ T

(3.5)

(3.6)

(3.7)

is simply the mean velocity. In Eqn 3.7 the overbar denotes an average value. The time-

42

average of the fluctuating component of the instantaneous velocity t+T

u; =lim_.!_ Jcul- UJdt = ul- ul = o T--t~ T

(3.8)

is zero. A limit ofT of oo in the above definitions is appropriate for stationary turbulence

(a turbulent flow that on average does not vary with time). For flows containing very

slow variations of the mean velocity with time, an alternative definition

t+T

UJ = ~ f U 1dt , T1 <<T<<T2 (3.9)

is required. fu Eqn 3.9, T1 is the time scale characteristic of the random velocity

fluctuations associated with the turbulence and T 2 the time scale associated with the

variation of the slowly varying mean flow behaviour. It has been assumed here that T2 is

several orders of magnitude greater than T 1• A more detailed exposition of Reynolds­

averaging (time averaging) was given by Wilcox (1993).

Reynolds-averaging of the steady, incompressible form of the continuity and Navier

Stokes equations results in the following equations in Cartesian tensor form

(3.10)

(3.11)

The correlated quantity -pu:u; appearing in Eqn 3.11 is the Reynolds stress tensor

denoted by 't1J • The quantity S is a "source" term and accounts for such effects as body

forces. The Reynolds stress tensor is symmetric and there are therefore six Reynolds

stresses that must be determined in order to close Eqn 3.11. The determination of the

Reynolds stresses in the RANS equations forms the fundamental problem of turbulence

modelling.

3.2 Turbulence Modelling

A brief overview of the theory behind turbulence modelling and the two-equation

turbulence models available in Fluent and used in this research, shall be presented in

this section. Fluent also allows the option of using a Reynolds Stress Model (RSM)

turbulence model which should allow more accurate solutions to be obtained for

43

anisotropic flows. Fluent ( 1996) noted that despite this advantage, more time is required

to obtain converged solutions as there are a larger number of equations being solved and

the model is computationally less robust than the k-E models available. In view of the

fact that a large number of runs would be required for the design space investigation and

optimisation-related work presented in later chapters (where minimal execution time

and robustness of calculation are essential) it was decided to make exclusive use of k-E

models.

The hypothesis of Boussinesq (1877), referred to by Wilcox (1993), can be written for

incompressible flow as

-,-, (au. auJ J 2 k~ -pu.uJ = J..lt -a +-a --p u.J XJ X1 3

(3.12)

can be used to relate the mean strain rate tensor (Dij) to the Reynolds stresses via an

isotropic eddy/turbulent-viscosity (J.11) formulation. In Eqn 3.12, 01J is the Kronnecker

delta and the quantity k is the specific turbulent kinetic energy, commonly called the

turbulent kinetic energy and is defined by

1-- -- 1-k = -(u'u' + v'v' + w'w') = -u'u' 2 2 I I

Dimensional arguments dictate that eddy viscosity is given by

J..l 1 = const · pk112 f

(3.13)

(3.14)

where f is the turbulent length-scale, the determination of which is not unique as was

noted by Wilcox (1993).

The steady form of the transport equation for turbulent kinetic energy can be written as

U ak au. a ( ak 1 , , , -,-,J p -='t ---pE+- J..L---puuu -pu J ax IJ ax ax ax 2 I I J J

J J J J (3.15)

The left-hand-side of Eqn 3.15 represents the convection of turbulent energy by the

mean flow. The terms on the right-hand side of Eqn 3.15 represent the production,

dissipation, molecular diffusion, turbulent transport and the pressure diffusion of

turbulent kinetic energy, respectively. The production term represents the rate at which

kinetic energy is transferred from the mean flow to the turbulence. It is also the rate at

which the mean flow "works" against the Reynolds stresses. The dissipation is the rate

44

at which turbulent kinetic energy is converted into heat. Molecular diffusion represents

the diffusion of turbulence energy as a consequence of natural molecular transport

processes. The turbulent transport term represents the rate at which turbulence energy is

transported through the fluid by turbulent velocity fluctuations. The pressure-diffusion

results from correlated velocity and pressure fluctuations.

The turbulent transport and pressure-diffusion terms are approximated together by a

gradient -diffusion representation

1 ~ ak -pu'u'u' + p'u' = --1

-2 IIJ J crax

k J

(3.16)

as stated by Wilcox (1993). Replacement of the turbulent transport and pressure

diffusion terms in Eqn 3.16 with a gradient-diffusion approximation yields the following

generic equation for k

(3.17)

The turbulent dissipation is defined by the correlation

au: au: E=V--

axJ axJ (3.18)

If R is a turbulent length-scale of effective eddy size, then the velocity scale of the

corresponding eddy is k112• Its turbulent dissipation should scale by dimensional

reasoning as

(3.19)

(see White (1991)), which when substituted into Eqn 3.14 allows the eddy-viscosity to

be directly related to k and E via

(3.20)

Therefore, in order to determine the eddy-viscosity so as to close the RANS equations, it

is necessary to solve two additional transport equations, one for the turbulent kinetic

energy (k) and the other for the turbulent dissipation (E).

The exact equation for E is actually a transport equation for the dissipating eddies. What

45

is actually sought is an equation for the transfer of energy from the energy-containing

eddies, since the length-scale required is that of the energy-containing, Reynolds-stress­

bearing eddies (Wilcox (1993)). As a result, the transport equation for turbulent

dissipation is taken as an empirical equation for the rate of energy transfer from the

large eddies (equal to the rate of dissipation of turbulent energy in the small eddies). The

e equation is therefore modelled in an analogous way to the k equation (Wilcox (1993)).

3.2.1 Standard k-e Turbulence Model

Fluent (1996) noted that the Standard k-e model of Launder and Spalding (1974) has

been a popular two-equation turbulence model for over 20 years and derives its

acceptance from its robustness, computational economy and reasonable accuracy. The

Standard k-e turbulence model may be written as

p ae +pU ae =_j_(J..LT ae )+ CEIJ..lTf.(aul + auJ)aul- pCE2£2

at J ax J ax J (j E ax J k ax J ax 1 ax J k

where the eddy viscosity is determined by

k2 II =pC -t"! J.l f.

(3.21)

(3.22)

(3.23)

The empirically determined closure coefficients of the transport equations for k and e

are

cJ.l = o.o9, eEl = 1.45, CE2 = 1.90, crk =to, crE = 1.3 (3.24)

The constants for the Standard k-e model were determined by calibrating the model

against experimental data (Launder and Spalding (1974)). A brief history of the

development of the k -e turbulence model was provided by Wilcox ( 1993).

3.2.2 RNG k-e Turbulence Model

The Renormalisation Group (RNG) k-e turbulence model used by Fluent was originally

developed by Yakhot and Orszag (1986) and extended by Yakhot et al (1992) to account

for the effect of irrotational strain on the production of e. Only the high Reynolds

number form of the model shall be presented, as this is the only form of the model used

46

to obtain the CFD results presented in this thesis. The RNG method is essentially a scale

elimination technique that can be applied to both the Navier Stokes equations and scalar

transport equations. Removal of successively larger scales leads to differential transport

equations. Constants in the model are derived explicitly from theory.

The RNG k-E turbulence model used in Fluent may be summarised as

where the closure coefficients for the transport equations of k and E are

c,1 =1.42, c.2=l.68, a=l.39

The turbulent kinetic energy production term in Eqn 3.25 is

Gk = JlrS2 = 2J.1rS,JSIJ

where S,J is the mean rate of strain tensor

and S is its modulus defined by

s =~S,1S,1

(3.25)

(3.26)

(3.27)

(3.28)

(3.29)

(3.30)

The structure of Eqn 3.25 and Eqn 3.26 are similar in form to that of the Standard k-E

model, but differs due to the addition of a rate of strain term (R) in the transport

equation for E and different closure coefficients. It is this rate of strain term that

sensitises the production (or destruction) of E to the effect of irrotational strains in the

flow and hence improves the model's predictive accuracy. The rate of strain term (R) is

given by

II au' au' r- I I R=2-S --p ljax ax

I j

and is expressed in the model equations as

47

(3.31)

(3.32)

where ll = S k/c is the ratio of the turbulent to mean strain rate time-scales, llo""4.38 and

B=0.012. For the high Reynolds number form of the model, the isotropic eddy-viscosity

is calculated by

(3.33)

where the closure coefficient CJ.1=0.0845. This is close to the value of CJ.1=0.09 of the

Standard k-£ model. The inverse of the Prandtl number (a.), appearing in the turbulent

transport equations is obtained from the effective viscosity by the solution of

a- 1.3929 o 6321 a+ 2.3929 o 3679 = Jlo

a 0 -1.3929 a 0 + 2.3929 Jl 1

(3.34)

where ao= 1.

3.2.3 Limitations of k-£ Turbulence Modelling

A well known deficiency of the Standard k -£ model is the tendency of this model to

over-predict the eddy-viscosity under adverse pressure gradients. This acts to suppress

flow separation and leads to an over-prediction of the skin friction coefficient for wall­

bounded flows. Rodi and Scheuerer ( 1986) demonstrated that the limitations of the

Standard k-c model under adverse pressure gradient may be attributed to an over­

prediction of the turbulent length-scale by the c equation.

Rodi and Scheuerer ( 1986) noted that the results of Bradshaw ( 1967) suggested that the

turbulent length-scale is essentially independent of pressure gradient over a wide range.

Thus there is a need to modify the c equation so that either the generation of c is

enhanced or the destruction of c is reduced under adverse pressure gradient. Rodi and

Scheuerer ( 1986) adopted the former approach based on augmenting the production of c

using the model of Hanjalic and Launder ( 1980). Hanjalic and Launder ( 1980) modified

the c equation in order to promote higher rates of dissipation for irrotational strains than

for rotational strains, thus suggesting that energy transfer rates across the turbulence

spectrum is promoted by irrotational deformations.

The Boussinesq hypothesis represents another limitation of the Standard k-c model and

48

indeed other two-equation models as well. The notion that the Reynolds stresses can be

related to mean flow gradients via an isotropic eddy-viscosity becomes inaccurate when

the Reynolds-stresses are unrelated to mean flow gradients. Wilcox (1993) noted that

examples of such flows include:

1) Flows subjected to sudden changes in mean strain rate

2) Flows over curved surfaces

3) Flows with boundary layer separation

4) Flows in ducts with secondary motion

5) Secondary flows in the comers of ducts of rectangular cross-section driven by

anisotropy in the turbulence

6) Three-dimensional flows and flows with significant boundary layer cross-flow

7) Flows subjected to rotation

Flows over curved surfaces, flows subjected to rotation, secondary flows (and swirl) and

flows with boundary layer separation are themselves examples of flows subjected to

streamline curvature and what Bradshaw (1973) termed as "extra rates of strain". These

"extra rates of strain" modify the structure of the turbulence to a much greater extent

than the extra production terms that appear in the Reynolds stress equations would

suggest. The phenomenon is related to the higher-order structural parameter of the

turbulence, which causes changes in the Reynolds stresses and therefore modifies the

structure of the turbulence. These processes are unrelated to the mean strain rates of the

flow.

Lakshminarayana ( 1986) reviewed the subject of turbulence modelling for complex

flows. Lakshminarayana ( 1986) stated the following reasons for the limitations of k-E

models:

1) The assumption of isotropic eddy-viscosity

2) Inadequate modelling of the pressure strain terms in the turbulent transport equations

3) The calibration of model constants from two-dimensional flows

4) Gradient-induced diffusion as the only diffusive mechanism

Lakshminarayana (1986) and the references cited therein suggested that a constant value

49

of C~ is inadequate for the prediction of complex shear layers. In order to improve the

predictive accuracy for three-dimensional flows, the k-E model may be modified to

include a vectorial representation of C~ via an Algebraic Reynolds Stress Model

(ARSM). For example, Leschziner and Rodi (1981) used this approach and were able to

obtain closer agreement with experimental data for annular and twin parallel jet flow.

Alternatively, the k-E model should be coupled with either an ARSM or Reynolds Stress

Model when adequate prediction of the mean and turbulence flow field is sought.

There have been numerous attempts to modify two-equation turbulence models to give a

better prediction of flows subjected to streamline curvature. Wilcox and Chambers

( 1977) for example, proposed a curvature correction to the turbulent kinetic energy

equation for their k-ro model and obtained improved predictive accuracy for flow over a

curved surface. They argued that the equation for k should be thought of as an equation

for v'2 for curved flow over a surface. Launder et al (1977) modified the dissipation

term in the E equation to account for the effect of streamline curvature. The constant C£2

was modified using a turbulent Richardson number as the significant curvature

parameter. This yielded an improvement in predictive accuracy.

From the references cited, it is clear that there is a diverse range of modifications to two­

equation models to account for adverse pressure gradient and streamline curvature.

These essentially involve the modification of C~ , or the turbulent length-scale obtained

by modification of the production or dissipation terms in the E equation.

The RNG k-E turbulence model offers improved accuracy over the Standard k-E model

due to the rate of strain term (Eqn 3.32) in theE equation. This effectively sensitises the

production (or dissipation) of E to the effect of irrotational strains. Therefore, in regions

of large irrotational strain (Tl>Tlo) R becomes negative and E increases. This leads to

smaller turbulent length-scales and reduced eddy-viscosity. Hence, the tendency of the

flow to separate will be enhanced. For flows over curved surfaces, the reduction in

eddy-viscosity provides a better modelling of the collapse of turbulence phenomena (see

Gillis and Johnston (1983)) for flow over convex surfaces.

50

3.3 Modelling of the Near-Wall Flow

The viscosity of the fluid, irrespective of its magnitude, results in the "no-slip"

condition at a solid boundary (wall). This requires that the velocity of the fluid at a solid

surface be equal to the velocity of the surface, as noted by Tennekes and Lumley (1972).

As a consequence of the no-slip condition, fluid in its flow over a solid boundary is

subject to significant shearing in the streamwise direction. A boundary layer evolves

over the surface of the solid boundary. Solid boundaries therefore act as sources of

turbulence and energy loss.

3.3.1 Boundary Layer Structure

The no-slip (zero velocity) condition at the wall is the essence of the formation of the

turbulent (high flow Reynolds numbers) boundary layer. The difference between the

free-stream velocity and the zero velocity at the wall results in sharp velocity gradients

and hence shearing in the fluid. The consequence of this being the formation and growth

of a boundary layer adjacent to the wall.

The turbulent boundary layer may be described by several layers in accordance with the

behaviour of the turbulence in these "layers". This is ultimately a direct consequence of

the turbulent length-scales involved, since there exists a close analogy between the

spatial structure of a turbulent boundary layer and the spectral structure of turbulence

(Tennekes and Lumley (1972)). The velocity profile of a turbulent boundary layer

consists of an inner layer where viscous effects are predominant, an outer layer (wake

region) where turbulent shear dominates, and an overlap layer providing an asymptotic

matching between the inner and outer layers (see for example White (1991) and Wilcox

(1993)).

The behaviour of the inner layer, or "wall region" is directly affected by the conditions

at the wall and hence by the wall shear stress, the wall roughness, fluid viscosity and the

distance (y) from the wall, but not on free-stream properties as both Hinze (1975) and

White (1991) noted. The functional form of the velocity in the wall region may be

determined from dimensional analysis as

51

where ut is the "friction velocity" defined by

and y + is the dimensionless sublayer-scaled distance from the wall defined by

y+ = pyU-r

J.1

The behaviour in the outer layer follows a "velocity-defect law" of the form

Urer- u = f(x_) ut 8

(3.35)

(3.36)

(3.37)

(3.38)

(Hinze (1975)) where 8 is the boundary layer thickness and Uref the velocity at the edge

of the boundary layer. From Eqn 3.35 and Eqn 3.38 it can be seen that the inner layer

corresponds to small turbulent length-scales of order y, the distance from the wall, down

to the Kolmogorov microscale of dissipative eddies. The outer layer corresponds to

large scale turbulent behaviour in the boundary layer with length-scales of the order of

8. It is possible to find a range of distances (y) from the surface where y+>>1 and y/8<<1

simultaneously. This range covering the transition from small to large-scale turbulence

dynamics corresponds to the overlap layer. This range is also called the logarithmic

region or the inertial sublayer, in an analogous manner to the inertial subrange of the

spectral structure of turbulence as Tennekes and Lumley ( 1972) explained. It is here that

asymptotic matching of the velocity profiles of the inner and outer layers occurs.

Three subregions may be identified in the inner layer. These are the viscous sublayer

(y+<5) , the transition or "buffer" layer (5<y+~30) and a fully turbulent region (y+>30),

(see Hinze (1975)). In the near-wall region, the integral scale of the turbulence is

proportional to the distance from the wall and must be of order y. Sufficiently close to

the wall the length scale of the turbulence becomes less than the Kolmogorov

microscale and hence Reynolds stresses become negligible and viscous effects

dominate. This is the viscous sublayer. As distance from the wall increases, the integral

scale becomes larger than the Kolmogorov microscale and Reynolds stresses

progressively become larger in magnitude until viscous stresses become negligible in

comparison to Reynolds stresses in the fully turbulent region. This region of transition

52

from viscous to Reynolds stresses is the buffer layer. The buffer layer is a region of

vigorous turbulence dynamics, because the turbulent energy production reaches a

maximum where the Reynolds stresses become equal to the viscous stress (y+:::::12) (see

Tennekes and Lumley (1972)).

In the viscous sublayer, velocity scales in a linear manner directly with distance from the

wall as

(3.39)

as noted by White (1991). In the fully turbulent region, the velocity can be described by

a logarithmic velocity distribution

u 1 ( +) -=-In y +B u't K

(3.40)

In Eqn 3.40, K is the von Karman constant and B is an empirical constant. Coles and

Hirst ( 1968) proposed the following values of the two constants

K=0.41 , B=5.0. (3.41)

Note that the above discussion has assumed flow over a smooth surface. In the buffer

layer the velocity profile is neither linear nor logarithmic but rather a smooth transition

from linear to logarithmic. Spalding (1961) proposed the following single composite

profile

(3.42)

for the wall region.

3.3.2 The Wall Function Approach

In order to utilise a two-equation turbulence model for computational simulation of a

wall-bounded flow, it is necessary to specify boundary conditions for velocity, turbulent

kinetic energy (k) and turbulent dissipation (c) at the wall, since the presence of the wall

is the primary source of turbulence formation. Since the Standard k-c model is of a high

Reynolds number form, it fails to predict a satisfactory constant B in the law of the wall

as Wilcox ( 1993) discussed. As a consequence, simply applying the no-slip condition at

the wall and integrating through the viscous sublayer will yield an unsatisfactory result.

53

Apart from the issue of the behaviour of two-equation turbulence models in the near­

wall region, there are also issues of computational requirements. As the Reynolds

number of the flow increases, the range of turbulent length-scales also increases and

hence there will be a larger discrepancy in size between the integral scale of eddies and

the Kolmogorov microscale of dissipative eddies. It can thus be expected that as the

Reynolds number of the flow increases, the inner layer of the boundary layer will

decrease in thickness. As a consequence of this fact, if an integration of the governing

equations through the sublayer is sought, a finer grid would be required in order to

adequately resolve the flow down through the viscous sublayer. This increases the

computational requirements, both in terms of memory and execution time.

The above-mentioned problems can be overcome via the use of "wall functions" as a

means of providing suitable boundary conditions for the wall adjacent cells. Launder

and Spalding (1974) noted that the use of wall functions would not only reduce the

computational requirements of the simulation, but also allow the introduction of

additional empirical information in special cases, for example when the wall is rough.

3.3.3 Standard (Equilibrium) Wall Function

Wilcox (1993) and Cebeci and Smith (1974) showed that when pressure gradients are

small or negligible compared to other terms in the momentum equations, the momentum

equations in the inner layer simplify to

(3.43)

where U is the velocity tangent to the surface and y is the perpendicular distance from

the wall. Integration ofEqn 3.43 yields

(3.44)

In other words the sum of the molecular shear stress and the Reynolds stress is constant

in the inner layer, therefore shear stress is constant. Since shear stress is constant in the

inner layer, outside the buffer layer (y+>30), molecular shear stresses become negligible

in comparison to turbulent shear stress, hence the Reynolds stress is z-pU~ and the

mean velocity gradient is given by Uj(Ky) (Tennekes and Lumley (1972)). It follows

54

that the production of turbulent kinetic energy in the inner layer is given by

Gk = -ulvl au= U~ ay Ky

(3.45)

In an equilibrium boundary layer, the rates of turbulence production and dissipation are

in balance, therefore

u3 E=Gk = ~ (3.46)

Applying the Boussinesq hypothesis for the flow in the buffer layer, the Reynolds stress

can be related to the mean velocity gradient by

I I au -pu v = Jlt ay (3.47)

The turbulent kinetic energy may be obtained by substituting Eqn 3.46 into Eqn 3.23

and solving fork, with the mean velocity gradient in Eqn 3.47 being evaluated from the

law of the wall (Eqn 3.40). The resulting expression is

k= u~ ,JC;

Using the result ofEqn 3.48, y+ can be written as

and the following expression for the law of the wall may then be evaluated as

~CI'4kl/2 = ..!_ln(Ey p(C~2kp)It2) 'tw / p 11 p 1( p Jl

(3.48)

(3.49)

(3.50)

where kp is the turbulent kinetic energy at the wall-adjacent cell centres, at a

perpendicular distance of yp from the wall. The production of turbulent kinetic energy

then becomes

and turbulent dissipation can be rewritten (using the result ofEqn 3.48) as

c3'4k312 11 p

Ep =--=---K)'p

(3.51)

(3.52)

Fluent solves the momentum equations throughout the computational domain, including

the wall adjacent cells, with the no-slip condition specified at the wall. The near-wall

55

values of turbulent kinetic energy are determined by solving the complete transport

equation for k in the near-wall control-volumes subject to the boundary condition of a

zero normal gradient assumed for k at the wall. The production term for k in the

transport equation for k is obtained from Eqn 3 .51.

When the Standard (equilibrium) wall function option is specified, the wall shear stress

is determined from Eqn 3.50. Once the wall shear stress has been determined, the

production of k and the turbulent dissipation can then be determined using Eqn 3.51 and

Eqn 3.52, respectively. It is interesting to note that the transport equation forE, unlike

the momentum equations and the transport equation for k, is not solved at the near-wall

cells, but rather the solution fore is obtained via the use ofEqn 3.52.

It has been assumed in the above discussion, that the cell-centre location of the wall­

adjacent cells have values of y+ sufficiently large such that the predominant shear

stresses present in the flow are turbulent and that Eqn 3.50 will provide a reasonably

accurate approximation to the velocity profile. In the viscous sublayer, where the

stresses are viscous and in the lower part of the buffer layer where viscous stresses are

predominant (y+s;12), a different form of wall function must be used. In many near-wall

modelling approaches using wall functions, the buffer layer is eliminated by linking the

linear velocity profile in the viscous sublayer to the logarithmic profile in the inertial

sublayer. This causes an abrupt change from purely viscous stress to purely turbulent

stress at y + = 11 as Tennekes and Lumley ( 1972) discussed. Following this approach,

when y +> 11.225 Fluent uses the logarithmic velocity profile of Eqn 3.50 to calculate

wall shear stress. When y+<11.225, the laminar stress-strain relationship ofEqn 3.39 is

used (see Fluent (1996)). The abrupt change from a viscous to a turbulent shear stress

relationship, resulting from the changeover of velocity profile at y + = 11.225 (as a

consequence of the elimination of the buffer layer in the modelling), is clearly a source

of modelling inaccuracy.

3.3.4 Two-Layer-Based Nonequilibrium Wall Function

For complex flows involving separation, reattachment and impingement, where the

mean flow and turbulence are subject to severe pressure gradients and change rapidly,

56

the equilibrium assumptions underlying the development of the Standard wall function

break down. The boundary layer can no longer be considered as being an equilibrium

boundary layer, indeed it is now a nonequilibrium boundary layer. Fluent offers the

option of selecting a Two-Layer-Based Nonequilibrium (TLBN) wall function which

can partly account for nonequilibrium effects neglected by the Standard wall function

(see Fluent (1996)). The discussion of the TLBN wall function presented below is taken

exclusively from Fluent ( 1996).

In the TLBN wall function, the wall function of Launder and Spalding (1974) is

sensitised to the effects of pressure gradient and is modified to the following form

(3.53)

where

(3.54)

and Yv is the physical thickness of the viscous sublayer determined by

* JlYv (3.55)

where y: = 11225. Unfortunately Fluent (1996) does not give the value of K *. For the

cells adjacent to the wall, the following profile assumptions are made for the turbulence

quantities

{

0, 'i =

I 'i W'

k = {(y /Y v t kp, Y < Y v

kp, Y>Yv {

2j.lkj(py 2 ), Y < Yv

e = kat2 j(Cey), y > Yv (3.56)

where Ce = KC~314 , 't1 is the turbulent shear stress and Yv is the dimensional thickness of

the viscous sublayer defined by Eqn 3.55. Using the profiles for 't1 , k and E specified in

Eqn 3.56, the cell-averaged production of k can be approximated by its depth-average

- 1 fYn dU 1 't2

( y ) Gk = -Jc 'tt ::h, dy = -- u: 112 In _n

Y 0 0 v y Ky 0 pC ~ k p Y v

(3.57)

The over-bar on Gk denotes a cell-averaged quantity and Yn is the height of the cell

(yn=2yp). Similarly, the cell-averaged dissipation rate is

57

_ 1 JYn 1 ( 2fJ. k~2 (Y"J] E=- cdy=- -+-In- kp y n 0 y n py v cl y v

(3.58)

The solution procedure for determining velocity, k and E in the wall-adjacent cells is

identical to the case of the Standard wall function, but in the case of the TLBN wall

function, wall shear stress ('tw) is determined by Eqn 3.53. The value of 'tw obtained

from Eqn 3.53 is subsequently used in the calculation of the production term of the

transport equation fork using Eqn 3.57. The turbulent dissipation in the wall-adjacent

cells is then calculated using Eqn 3.58.

3.4 Finite-Volume Discretisation of the Governing Equations

The equations governing fluid flow describe a continuum phenomena. Exact solution of

the governing equations would yield a solution at an infinite number of points in the

flow field. Whereas simple two-dimensional partial differential equations may be solved

analytically, the complexity of the governing equations described in this thesis are only

amenable to numerical solution using powerful computers. In order to solve the

governing equations numerically, it is necessary to replace these partial differential

equations with a system of algebraic equations so that a computational solution can be

obtained as noted by Fletcher (1987a).

A brief overview of the finite-volume method of discretising the governing equations

shall be presented in this section. Much of the discussion is based on Patankar (1980),

Fletcher (1987a) and Fluent (1996). The simple case of the application of the finite­

volume formulation to three-dimensional, rectangular grids is used for illustrative

purposes. The extension of the methodology to general curvilinear coordinate systems is

not discussed, but the reader is referred to Naser (1990) and the references cited therein.

3.4.1 Finite-Volume Formulation

Fletcher (1987a) noted that the finite-volume (or control-volume) formulation belongs

to the class of numerical techniques collectively known as the "method of weighted

residuals" (MWR). The MWR may be stated as

JJfvwm (x 1 ,x2 ,x3 )Rdx 1dx 2dx 3 = 0 (3.59)

58

where R is the residual expressing the difference between the exact solution and an

approximate solution and W m is the weighting function. It is therefore required that the

integral of the weighted residual over the computational domain be zero. R is a

continuous function of x1, x2, x3 and t (for time-dependent flows) (see Fletcher (1987a),

p.99-100). In the finite-volume formulation, the computational domain is divided into a

finite number of control-volumes or computational cells ~ V such that

W = 1 inside ~V m

W = 0 outside ~V m

(3.60)

as shown in Fig 3.1. The centre of the control volume is denoted by point P, which is

surrounded by six other points corresponding to the centres of surrounding control

volumes (points E,W,N,S,T,B).

ww w

ss BB

Fig. 3.1 Definition of control volume used in finite volume method

Fig. 3.2 shows how Fluent defines the discrete control volumes using a non-staggered

Cell-Centre Point

(i-1,j-1,k-1) (i,j-1 ,k-1)

Fig. 3.2 Relationship between grid points and cell-centre location

59

grid storage scheme. It is clear that the node points of the computational grid do not

correspond to the control-volume centres, but are used to determine the control-volume

centres and define the cell boundaries. The same control-volume is employed for

integration of all of the conservation equations. All variables are stored at the centre of

the control-volume and therefore Fluent employs a co-located finite-volume

formulation.

The weighted residual inside the control-volume becomes

JffvRdx 1dx 2 dx 3 =0 (3.61)

where R is now the governing equation being discretised, since the governing equations

describe conservation relationships which must reduce R to zero if the correct solution

is obtained. Gauss's theorem (see for example Dubbel (1990))

JJfvv. F(f)dV =JL (F. n)ds (3.62)

can be used to convert the volume integral of a governing equation into a corresponding

surface integral over the six faces of the control-volume as follows

JII V · FdV =JFedA- JFwdA + JF0 dA- JF5dA + JF1dA- JFbdA (3.63) V A 0 Aw A 0 As A1 Ab

Eqn 3.63 may be alternatively expressed as

Ie - Iw +I" -Is + I 1 - Ib = S

where

represents the flux of F across the ith control volume face and

S = Sp .1x 1.1x 2 .1x 3

(3.64)

(3.65)

(3.66)

is the linearised source term obtained by integration of the volumetric source (sp) over

the control volume (Naser (1990)).

The fluxes across each control volume face (Eqn 3.65) are composed of a convective

component and a diffusive component. For the continuity equation, the cell face fluxes

are purely convective (mass fluxes). For the momentum equations, the convective fluxes

are momentum fluxes and the diffusion fluxes are due to viscous stresses. For the

continuity equation, the mass fluxes across the control-volume face are given by

60

II = J (pU)dA = (pUA)I A,

(3.67)

for the control-volume shown in Fig. 3.1. Similarly, the convective fluxes across the

control-volume faces, for a transported quantity<!>, may be written as

II = J (pU<!>)dA = (pU<!>A)I A,

(3.68)

The diffusion fluxes contain both normal-derivative and cross-derivative contributions

as discussed by Naser (1990).

It is clear from the above discussion that the values of the transported variable at the

control-volume faces are sought. Since the finite-volume methodology used by Fluent

solves the value of the transport variables at the control-volume cell-centres, it is

therefore necessary to perform a suitable interpolation in order to express the value of

the transport variables at the control volume faces in terms of the cell-centre values.

This is the underlying motivation behind the development of convective differencing

schemes which are the focus of the next section.

Patankar ( 1980) wrote the finite-volume discretised governing equations in the

following generic form

ap<l>p =I anb<l>nb + b nb

(3.69)

where the value of the dependent variable at point P (<!>P) is a weighted average of the

neighbouring cell-centre values of q,. In Eqn 3.69 the subscript nb denotes neighbouring

points. The coefficients in the discretised equations (a) multiply the cell-centre point

values. The constant, b, is the volume integral of the source term. It is interesting to note

that the coefficients in the discretised equations are nonlinear due to the nonlinearity of

the governing transport equations being solved. As a consequence of this and the large

number of equations to be solved, iterative solution techniques are used to solve the

discretised equations.

3.5 Convective Differencing

As mentioned in the previous section, the finite-volume methodology results in

expressions for the discretised governing equations in terms of cell face values, which

61

must be related to cell-centred values via appropriate local solution interpolation.

Patankar ( 1980) noted two guiding principles that must be adhered to in a finite-volume

discretisation formulation. These are the principles of physically realistic solution

behaviour (or boundedness) and of integral conservation (conservativeness).

Roundedness implies that in the absence of sources, the values of the transported

quantity <1> in the interior of the computational domain, should be restricted to lie within

the limits specified by the boundary conditions as noted by Naser (1990).

Conservativeness implies that overall conservation of transported quantities applies over

any number of control-volumes taken together. This implies an overall integral balance

between fluxes and internal sources. With an inappropriate choice of convective

differencing scheme for a given size of grid, the above guidelines may be violated

resulting in poor or physically unrealistic solutions.

The Peclet number, defined by

Pe= pUL r

(3.70)

where L is a characteristic length and r the diffusion coefficient, describes the relative

strengths between convection and diffusion in a flow (see Patankar (1980)). In the

context of CFD, a cell Peclet number is defined as

Pe= pU(~) r

(3.71)

where the characteristic length is now replaced by the distance between adjacent grid

points (~). For large cell Peclet number where pU(~)>>r the flow is dominated by

convective transport, whereas for very small cell Peclet number r>>pU(Ax) the

transport is largely diffusive. The extreme cases are IPel=oo and Pe=O corresponding to

pure convection and pure diffusion, respectively.

The behaviour of the solution between two cell-centre locations, with Peclet number,

can be understood by examining the solution of a steady one-dimensional convection­

diffusion equation in the absence of sources,

j__CpU<!>) = ~cr d<l>) dx dx dx

(3.72)

over an interval L with boundary conditions

62

x=O <1>=<1> 0

x=L <j>=<j>L

as discussed by Patankar (1980). The exact solution to the Eqn 3.72 is

<I>- <l>o exp((Pe)x I L) -1 '----~=-::....C....'--'----'---

<1> L - <1> 0 exp(Pe)

(3.73)

(3.74)

Patankar showed that for the case of pure diffusion a piecewise linear (centred

difference) approximation provides an exact solution of the variation of <1> between 0 and

L. For convection dominated flows, where Pe>>l, the flow is dominated by the

upstream behaviour, so that the value of <1> at the control-volume face could simply be

set as the value of <1> at the upstream node. This forms the basis of the upwinding scheme

(see Patankar (1980) and the references cited therein). Therefore for small Pe (1Pe1~2) a

centred-difference approximation provides an adequate interpolation of face values,

whereas for large Pe>>l an upwind difference scheme provides a good approximation.

For a given application, the cell Peclet number can be reduced by grid refinement, but

this increases the computational requirements.

3.5.1 Power-Law Scheme

In order to provide a convective differencing scheme that is applicable over a wide

range of Peclet numbers, a suitable scheme could use the exact solution of the steady

one-dimensional convection-diffusion equation, or an approximation thereof. Finite­

volume discretisation of the steady one-dimensional convection-diffusion equation

yields

aP<I>P = aE<I>E + aw<l>w

where a E = De - Fe I 2

aw = Dw +Fw 12

aP = De +Fe I 2 + Dw - Fw I 2

F = pu , D = r I (Ax)

(3.75)

according to Patankar (1980). Substitution of the exact solution to the one-dimensional

convection diffusion equation into Eqn 3.75 gives

63

Fe aE = ___ .:::._ __ exp(Fe I De - 1)

Fw exp(Fw I Dw) aw =

exp(F w I D w ) - 1 (3.76)

ap = aE + aw +(Fe - Fw)

Eqn 3.76 is the "exponential" convective differencing scheme. Since the exponential

scheme involves the computation of exponentials, it is not used but rather an

approximation to the exponential solution, known as the "Power-Law" convective

differencing scheme is used. The basis of the power-law scheme is the representation of

the variation of the coefficient aE (or likewise aw) with Peclet number for the

exponential scheme with an accurate Power-Law profile. The power-law expression for

aE can be written as

Pe<-10 aE /De =-Pe

-10 ~ Pe < 0 a E I De = (1 + 0.1 Pe) 5 - Pe

0~ Pe ~ 10 aE I De= (1- 0.1Pe) 5 (3.77)

Pe> 10

(see Patankar (1980)). Fluent (1996) noted that the evaluation of the above coefficients

using the Power-Law scheme, is computationally economical and provides a robust

scheme of formal accuracy between first and second order.

3.5.2 Higher-Order Convective Differencing

Higher-order convective differencing schemes, such as the Quadratic Upstream

Interpolation for Convective Kinematics (QUICK) scheme of Leonard (1979) offer the

possibility of second-order (or higher) accuracy and reduced numerical diffusion.

Considering the nomenclature depicted in Fig. 3.3, the one-dimensional form of the

Flow Direction w p f E • • I •

«<>w cjlp $r <Ill;

Fig. 3.3 Nomenclature for higher-order convective differencing

64

higher-order convective differencing schemes used by Fluent can be written in the

following generic form

<l>r =8(.1,xE<I>P +.1.xP<I>E]/(.1,xp +.1.XE]

+ (1- 8)[(.1.xw + 2.1.xp) <I>P - .1.xp<l>w] I [.1.xw + .1.xP] (3.78)

where <l>r is the value of <1> sought at the desired cell face f and <l>c, <l>w, <!>E represent the

centre cell-centre, upstream cell-centre and downstream cell-centre values of <!>,

respectively.

Variation of the parameter 8 leads to different discretisation schemes. For example the

Second-Order Upwind scheme corresponds to 8=0, whereas the QUICK scheme

corresponds to 8=0.75. Bounding of the higher-order schemes used by Fluent is

necessary to avoid numerical instabilities. Fluent ( 1996) noted that the use of the

Second-Order Upwind or QUICK scheme will produce undershoots or overshoots at

discontinuities. On a uniform grid the QUICK scheme becomes

(3.79)

3.5.3 Numerical Diffusion

Additional numerical diffusion (for a given mesh size) will occur when the flow is

oblique to the gridlines and there is a non-zero gradient of the dependent variable in a

direction normal to the flow. Numerical diffusion enhances the actual diffusion in the

flow and will lead to a "smearing" of the solution. Numerical diffusion is most

pronounced when the flow is aligned at 45° to the gridlines and for convection

dominated (high Pe) flows. It is reduced to zero when the flow is aligned at 0° or 90° to

the gridlines. Higher-order convective differencing schemes such as QUICK are less

susceptible to numerical diffusion than the Power-Law scheme, as more points are used

in the interpolation and hence more flow information is available in the interpolation.

3.6 Solution of the Discretised Equations

Once the governing equations have been discretised as algebraic equations, they can be

solved numerically using a computer. Due to the elliptic and nonlinear nature of the

governing equations in steady-state form, the discretised equations can be assembled

65

into the following matrix form (Fletcher (1987a))

A(V)V=B (3.80)

In this case V is the vector of unknown cell-centre values and A is a matrix of

discretised equation coefficient values which is itself a function of V. B is composed

of algebraic coefficients associated with discretisation of the governing equations,

source terms and boundary condition information.

In order to solve the system of equations denoted by Eqn 3.80, iterative techniques must

be used, due to the dependency of A on the solution vector V . It is therefore possible to

set up an outer iteration where the coefficients in A are "frozen" while the system is

solved for V. Once V is obtained the components of A are updated and the procedure

repeated until convergence. Solution of the system A V = B can only be realistically

achieved using iterative solution techniques, due to the enormous computational

requirements that would otherwise be necessary (in terms of computer memory and

speed) if a direct Gaussian elimination were to be attempted.

3.6.1 Solution Methodology

In Fluent, the discretised governing equations are solved sequentially, with the

momentum and continuity equations being solved prior to solution of the transport

equations for turbulent kinetic energy (k) and turbulent dissipation(£).

In order to solve the momentum equations and the continuity equation, it is necessary to

solve for all the velocity components and the pressure. Since the incompressible form of

the continuity equation does not contain any explicit pressure-related terms, it is

necessary to construct an equation for updating the pressure. Fluent uses the Semi­

Implicit Method for Pressure-Linked Equations (SIMPLE) scheme of Patankar and

Spalding ( 1972) to accomplish this. The SIMPLE scheme used in Fluent has been

formulated for use in a co-located finite-volume formulation (rather than on a staggered

grid configuration as presented in the above references). The SIMPLE formulation used

by Fluent overcomes the pressure-velocity decoupling problem (see Patankar (1980))

that was prevalent in early attempts at producing co-located finite-volume formulations

of this scheme.

66

For illustrative purposes, the SIMPLE scheme has been described below for the three­

dimensional continuity and momentum equations discretised on a rectangular Cartesian

grid. The discussion is taken largely from Patankar ( 1980) and Fluent ( 1996). If a

"guessed" pressure field (p*) is substituted into the momentum equations an

approximated velocity field (U*, v· ,W*) can be determined. The U momentum equation,

for example, becomes

apu; = L:anbu:b -(p:Ae -p:Aw)+S (3.81) nb

The approximated velocity field does not in general satisfy continuity except when a

converged solution to the discretised equations is obtained. The actual velocity and

pressure fields can be related to the approximated velocity and pressure fields by

up= u; + u;·

vp = v; + v;·

Wp =W; +W;*

p = p* + p**

where u**, v**, w** are the velocity corrections and p** is the pressure correction.

(3.82)

(3.83)

Since the actual velocity and the pressure satisfy the momentum equations, an equation

linking the velocity correction with the pressure correction can be obtained by

substituting Eqn 3.82 and Eqn 3.83 into the momentum equations and subtracting the

equations for the velocity approximations (for example Eqn. 3.81) from the momentum

equations. This results in a momentum balance for the corrections. For example, the

equation for the u** velocity correction is therefore

U** ~ U** ( **A **A ) ap P = .£..Janb nb- Pe e -pw w (3.84) nb

If the term L anb u:~ is omitted from Eqn 3.84, the velocity correction can be directly nb

linked to the pressure correction resulting in the following equation for the velocity

correction

U** =--1 ( **A - **A ) p Pe e Pw w

aP (3.85)

How can such an omission be made? At convergence the omitted term will be zero as

67

the velocity corrections of the neighbouring points will be zero. Therefore this omission

can be made. A pressure correction equation can now be formed from the continuity

equation by substitution of Eqn 3.82, Eqn 3.83 and Eqn 3.85 into the continuity equation

resulting in the following generic form of the pressure correction equation

app;• = :L/nbP:~ + b (3.86) nb

Patankar ( 1980) noted that the coefficient b in this case acts as a "mass source" term

which the pressure corrections must reduce to zero at convergence.

The overall solution methodology used by Fluent for the solution of isothermal and

incompressible turbulent flow can be summarised as follows:

1) Estimate (guess) the pressure field.

2) Solve the discretised momentum equations (Eqn 3.81) in order to obtain the velocity

approximations in the sequence u·' v· 'w·

3) Solve the pressure correction equation (Eqn 3.86) in order to obtain the pressure

. ** correctiOn p .

4) Calculate the pressure using Eqn 3.83

5) Calculate the velocity corrections u**, v**, w** from Eqn 3.85 and subsequently the

velocities from Eqn 3.82.

6) Solve the transport equation for turbulent kinetic energy (k).

7) Solve the transport equation for turbulent dissipation (E).

8) Regard the corrected pressure as the new pressure estimate p, return to step 2 and

repeat the above steps until a converged solution is obtained. It must be noted that the

updated velocity field is treated as the new approximated velocity field when solving

Eqn 3.81.

In practice, the pressure correction and momentum equations are underrelaxed in order

to successfully converge the calculations. In the case of the pressure calculation, the use

of underrelaxation leads to

• ** P = P +app (3.87)

where ap is the underrelaxation factor for pressure correction. In the case of the

momentum equations for the velocity approximations (Eqn 3.81), underrelaxation is

68

applied to the results of the iterative solver so that

(3.88)

The result (U*) n+l of an iteration is therefore a linear combination of the current

solution (U~as) n+l obtained by the Line Gauss-Seidel (LGS) solver and the result of the

previous solution (U*)n at iteration n, where A is the underrelaxation factor.

3.6.2 Determination of Convergence

The calculation of equation residuals allows the convergence behaviour of the

computational simulation to be determined. Equation residuals provide a measure of

how closely each finite-difference equation is balanced given the current state of the

solution. The generic form of the finite-volume discretised governing equations was

stated previously as

ap<l>p = L anb<l>nb +b. nb

(3.89)

Fluent computes the global residual of the particular discretised transport equations by

summing the residual of the discretised equation for each control-volume over the

computational domain. This results in the following expression for the global residual

N

R = L L anb<l>nb + b- ap<l>p (3.90) •=I nb

where N is the total number of cell-centres in the computational domain. According to

Fluent (1996), the global residuals expressed by Eqn 3.90 are dimensional having units

of kg units of <1> per second and can be normalised to make them non-dimensional. The

non-dimensionalised form of the global residual is therefore

N

L L anb<l>nb + b- ap<l>p R = •=I nb (3.91)

The global residual for the pressure is actually the imbalance in the pressure correction

equation (and hence an imbalance in mass conservation) and is calculated by

N

R= LIFw -Fe +Fs -Fn +Fb -F11 (3.92) •=I

(see Fluent (1996)), where F1 in this case is the mass flux through face i. The pressure

69

residual is normalised by dividing it by the pressure residual at the second iteration of

the calculation, hence

- R R=-"

R2 (3.93)

Computation of the discretised equations terminates when the sum of the normalised

residuals is reduced below a specified "termination criteria".

3.6.3 Iterative Solution of the Discretised Governing Equations

Fluent uses Line-Gauss-Seidel (LOS) iteration as the fundamental iterative solution

technique for its iterative solver. The LOS technique applies Gauss-Seidel iteration to

"lines" of constant curvilinear coordinate direction, in order to update the values at the

cell-centres along these lines. It achieves this by "sweeping" along the line in the

direction of increasing index of the cell-centres. Each line (of constant curvilinear

coordinate direction and increasing coordinate index) in the domain is solved

sequentially, with the solver moving in the "marching" direction along the domain. This

is shown in Fig.3.4.

Sweep t Direction

Flow Direction

Solved values from current iteration

Marching Direction

Values from previous iteration

Line being solved

Fig. 3.4 Line Gauss Seidel iteration

The use of the LOS technique results in a tri-diagonal matrix for the particular line

under consideration. This can be solved by a direct method of solution such as the

Thomas Algorithm. Fletcher (1987a) described the Thomas Algorithm which uses a

simplified form of Gaussian elimination. Generally, the sweep direction should be

chosen to be in a direction normal to the primary flow direction. This allows boundary

70

condition information to be propagated into the domain which can then be marched

along the domain as discussed by Fluent (1996).

LGS iteration is good at reducing local errors (short wavelength errors) in the solution

but is rather poor at reducing global errors (long wavelength error) as noted by Fletcher

(1987a). As a consequence, the Additive Correction Multigrid (ACM) technique of

Hutchinson and Raith by ( 1986) is used by Fluent to reduce global error in the solution

so as to accelerate convergence of the calculation. Fluent ( 1996) recommended that

ACM be used for iterative calculation of the pressure correction equation, because the

mass balance inherent in this equation depends on the reduction of long-wavelength

error. Since the momentum and turbulent transport equations tend to be dominated by

localised conditions, LGS iteration will generally perform in a satisfactory manner in

reducing the associated local errors.

The long-wavelength solution errors that retard the convergence of LGS on a fine grid

become small-wavelength errors on a coarse grid and hence the discretised equations

become amenable to practical solution by LGS iteration. It is this relationship between

error wavelength and grid size that provides the underlying motivation behind the

·development of multigrid techniques. The ACM procedure works by computing

corrections (<j>**) to the current solution field (<j)) within successively coarser blocks of

control-volumes, as shown in Fig. 3.5. In the figure, Level 1 is the finest grid and the

- Level3Gnd

- Level2Gnd

····· Levell Gnd

Fig. 3.5 Example of multigrid grouping of cells with multigrid level

71

higher levels represent successive conglomerations of cells to form coarser grids. The

corrections obtained (based on integral conservation over the cells of the coarser grid)

on these multiple coarse-grid levels are used to "correct" the fine grid solution and

thereby accelerate convergence. The resulting corrected fine grid solution, (q>*+<l>**) then

obeys global conservation on the coarser grid levels.

Since a finite-volume formulation is used, the equations to be solved on each multigrid

level are obtained simply by summing together equations on the original fine grid level.

The summation used to construct the coarse grid equations can be written in generic

form as

~~~(aP <1> =aE <1> +aw <1> +aN <1> +a5 <1> .£..J .£..J .£..J 0 J k l,j,k o J k t+),j,k 1 J k t-),j,k 0 J k l,j+),k o J k l,j-),k

I J k (3.94)

where a1,J,k are the coefficients of the neighbouring points relative to point P of cell i,j,k

on the fine grid. The values of <I>1,J,k are assumed to consist of the current solution in each

cell and a correction, <I>;~,K of the I,J,K th cell of the coarse grid level. The corrected

solution at cell i,j,k on the fine grid is therefore

(3.95)

The logic controlling the movement between different grid levels is shown in Fig. 3.6.

Solve for cp** on Level n Gnd (n~8)

R,>l3R,1

I Solve for cp** on Level 2 Gnd J R,«XR,

I Solve for cp* on Level I (fine) Gnd I R,<OCR0

or 1 >max. IteratiOns

Fig. 3.6 Control of movement between multigrid levels

72

When the rate of reduction of the residuals is insufficient on the current grid level, the

algorithm moves to the next coarser grid level. In order to move to the next coarser grid

level

R R, p<--

R,_I (3.96)

where~ is the residual reduction rate parameter, R1 the absolute sum of residuals on the

current grid level after the ith iteration of the line solver and R1.1 is the residual from the

previous iteration.

A move is made from a coarser grid level to a finer grid level when either the equations

are sufficiently converged, or a maximum number of iterations of the solver at that level

has been executed. The convergence or termination criteria

(3.97)

states that the residual after the ith iteration of the solver on the current level (R1) has

fallen to some fraction (a) of the residual (R0 ) at the beginning of iteration at the current

level.

3.7 Closure

The use of computational techniques based on the solution of the full Reynolds­

averaged Navier Stokes equations provides the only practical means by which

computational simulation of waterjet inlets can be undertaken. This is primarily due to

the necessity to account for viscosity and turbulence in the flow. The decision was made

to use the commercially-available CFD software, Fluent, for the purposes of flow

analysis. The use of existing proven CFD software allows attention to be focused on

grid generation and waterjet inlet design and optimisation-related investigations, rather

than on programming a CFD solver.

An overview of the derivation of the Reynolds-averaged Navier Stokes equations by

Reynolds-averaging of the Navier Stokes equations was presented in this chapter.

Adequate closure of the RANS equations can be effected by relating the Reynolds

stresses inherent in these time-averaged equations, to mean velocity gradients via an

eddy-viscosity formulation according to the hypothesis of Boussinesq (1877). The

73

Standard k-E turbulence model of Launder and Spalding (1974) and the RNG k-E model

of Y akhot and Orszag ( 1986) which was extended by Y akhot et al ( 1992) to account for

the effect of irrotational strain on the turbulence, were described. Both models are used

in the context of work presented in this thesis. The RNG k-E turbulence model contains

an "extra rate of strain" term in the E equation which improves flow prediction for flows

subject to rapid strain and streamline curvature and so represents an improvement over

the Standard k -E model.

The use of wall functions for modelling the near-wall region is a robust and

computationally efficient approach to the modelling of the flow behaviour adjacent to

the wall. The Standard wall function, which assumes an equilibrium boundary layer was

presented together with the Two-Layer-Based Nonequilibrium wall function. This latter

wall function is sensitised to the effect of pressure gradient on boundary layer behaviour

and is meant to represent an improvement over the Standard wall function for flows

experiencing large pressure gradients.

An overview of the co-located finite-volume method, used by Fluent, for discretisation

of the governing equations was presented. The finite-volume method enforces integral

conservation of the transported quantities on a finite control-volume and results in a

balance between fluxes across the faces of the control-volume (cell) and source terms.

Convective differencing schemes such as the Power-Law scheme, outlined in Patankar

(1980), are used to relate the values of the transported quantities at the cell faces (which

appear in the discretised equations) to the values at the cell-centres. Higher-order

convective-differencing schemes such as the QUICK scheme of Leonard ( 1979) offer

greater calculation accuracy and reduced numerical diffusion, but require the use of

limiters in order to retain numerical stability during the course of calculation.

Due to the nonlinearity of the discretised equations and computational requirements

necessary to solve them, the discretised equations are solved by iterative means. Since

the flows analysed in this work are incompressible, the SIMPLE scheme of Patankar and

Spalding (1972) was used to provide a suitable coupling between velocity and pressure

in the solution process. A Line Gauss Seidel iterative solver is used by Fluent to solve

74

the discretised equations and is good at reducing short-wavelength error in the solution,

but is poor at reducing global error. Multigrid acceleration, such as the Additive

Correction Multi grid scheme of Hutchinson and Raithby ( 1986), can be used to reduce

this global error, particularly for the pressure correction equation. The termination

criterion used by the iterative solver for termination of the calculation, is that the sum of

the normalised residuals fall below a specified value. At this point the calculations may

be considered to be converged.

75

Chapter 4 Generic Geometry and Grid Generation

In order to investigate the aspects of the geometry which affect the hydrodynamic

performance of a waterjet inlet, as well as optimise the waterjet inlet for selected

criteria, it is necessary to be able to vary the geometry of the waterjet inlet in a

systematic and logical manner. This can be achieved by a parametric description of a

generic waterjet inlet geometry as noted by Seil et al (1997). A parametric description of

the generic waterjet inlet geometry must therefore be established. Suitable grids must

then be obtained for the discretisation of the continuum of space representing the flow

domain of interest, as grid generation is an integral part of the total CFD analysis cycle.

Conventional flush-type inlets, with their shallow inlet angles and long curved inlet

ramps present a challenging grid generation problem, which will be discussed later in

this chapter.

It was found that a single-block body-fitted-coordinate (BFC) structured grid could be

used to mesh both a waterjet inlet geometry and a simple flow domain representing a

volume of water beneath the waterjet inlet, as shown in Fig. 4.1. This volume of water

external to the inlet was arbitrarily chosen to be semi-ellipsoidal in shape and lying

underneath a flat semi-elliptical surface surrounding the inlet opening. A single-block

BFC structured grid topology allowed the commercially-available CFD software, Fluent,

to be used. Structured grid solvers such as Fluent have traditionally meant greater

computational efficiency than the use of unstructured grid solvers. The mesh topology

used as the basis for meshing the waterjet inlet and external domain is possibly the most

efficient regarding the demands on computer memory, as it is a pure single-block

topology.

For most practical problems of interest, mesh generation takes the largest proportion of

time in the CFD analysis cycle. When the focus is on analysing a large number of

different geometric configurations (as is the case for waterjet inlet geometry parametric

76

design space investigation and optimisation) the mesh generation component of the

overall CFD analysis cycle must be reduced to a minimum. The process must therefore

be automated as much as is possible in order to save time. Therefore, it was decided to

develop a grid generator specifically for the purpose of producing a suitable grid for the

flow domain considered here.

y

~X z

Flat Solid Surface

Waterjet Inlet

Volume of Water External to Waterjet Inlet

Fig. 4.1 Flow domain of interest

The grid generator developed, Inlet3D, was written in Fortran 77 and produced an

output file compatible with both Fluent and PreBFC (one of Fluent's preprocessors).

Grid generation could thus be carried out as an interactive process, with the grid output

of Inlet3D imported into PreBFC for grid smoothing and visualisation. Improvements to

the mesh could be made by repeating the process of mesh generation, smoothing and

visualisation until the desired grid was obtained. The grid generation cycle time required

to obtain a suitable grid for a new geometric configuration was dramatically reduced to a

matter of minutes. A flow chart of the grid generation cycle described above is shown in

Fig. 4.2.

77

Input File • Geometric data • Grid Input data

No

Inlet3D • Grid generation

PreBFC/Fluent Compatible

Grid File

PreBFC • Visualisation

Yes

PreBFC • Elliptic smoothing

No

PreBFC/Fluent Compatible Grid

File

Yes

Fluent

Output File

1---~· Volume •Surface area

Fig. 4.2 Grid generation cycle for meshing waterjet inlet

78

A clear disadvantage of using the above methodology is that the quality of the resulting

meshes are heavily dependent on the user's experience and understanding. In addition,

extensive grid investigations must be initially undertaken in order to determine the

appropriate number of grid lines in each coordinate direction and their distribution.

Solution-adaptive meshing techniques, described by Thompson et al ( 1985), overcome

these limitations and allow a more optimum grid to be developed during the course of

solution. Solution adaptive meshing is not however, available in Fluent and so the above

limitations apply to the grids generated.

In view of the above discussion, this chapter is devoted to discussing the parametric

description of the author's generic flush-type waterjet inlet geometry. This generic

geometry forms the basis of the design and optimisation-related investigations that are

the focus of the work presented in this thesis. The various grid generation techniques

used to mesh the different geometries presented in later chapters are also discussed.

In Section 4.1 the parametric definition of the author's generic flush-type waterjet inlet

geometry is presented. The grid topology used as the basis of the meshing of the

waterjet inlet and a simplified flow domain external to it is discussed in Section 4.2. The

grid topology establishes the link between physical space (x,y,z) and computational

space (~,T).~) for the flow domain being meshed. The choice of topology used in

computational space will inevitably affect the resultant grid in physical space.

&

In order to generate the computational mesh, it is first necessary to generate the

boundary (surface) mesh that encloses the volume of the flow domain of interest. This is

the subject of Section 4.3. Once the domain boundaries have been meshed, the interior

grid can then be generated. This is discussed in Section 4.4. The relative merits and

limitations of the grid obtained, as a direct consequence of the grid topology used, are

discussed in Section 4.5 within the context of the role played by the grid in obtaining

accurate CFD solutions.

In Section 4.6 a brief discussion of the meshing procedure and mesh topology used for

the meshing of the individual geometries for the first two experimental validation cases

79

of Chapter 5 are presented. The discussion presented is brief because the meshing of

these two geometries is a relatively straightforward process. A summary of the contents

of this chapter will be given in Section 4. 7.

4.1 Generic Flush-type Waterjet Inlet Geometry

The geometric simplifications made in order to mesh the generic waterjet inlet

geometry, using a single-block structured mesh are discussed in Section 4.1.1. The

parameterisation of the geometry is discussed in Section 4.1.2. The simplicity and

effectiveness of using Bezier curves as a means of defining aspects of the waterjet inlet

geometry will be discussed in Section 4.1.3.

4.1.1 Geometric Simplifications

There are practical limits to the amount of geometric modelling required to adequately

describe the waterjet inlet for the purposes of CFD computation. Furthermore, geometric

modelling limitations result directly from the grid topology used to mesh the waterjet

inlet. In addition to modelling the shape of the waterjet inlet itself, it would be desirable

to model such aspects of the geometry as:

1) The impeller shaft or impeller shaft housing/fairing

2) The inlet grille (if fitted)

3) The tubing on which the inspection cover is fitted.

For the purposes of good flow design, the impeller shaft should always be housed in

some type of fairing. The impeller shaft (or housing/fairing) passing through the

waterjet inlet duct, will clearly influence the flow inside the waterjet inlet and therefore

should be represented either geometrically, or its effect on flow behaviour modelled.

The grid topology used for the work presented in this thesis makes geometric

representation of the impeller shaft or impeller shaft housing/fairing impractical and so

the effect of this geometric feature on the flow must be modelled. For the purposes of

investigating and optimising the geometric shape of the waterjet inlet, this issue will not

be given further consideration.

The representation of the inlet grille would require a large number of gridlines in order

80

to represent the individual grille bars. The smaller the pitch of the grille bars, the greater

the number of gridlines required for representation of the grille. It must be noted that not

all waterjets have grilles fitted, as there is a tendency for grilles to 'clog up' with debris.

This feature will not be represented in the computational mesh for two main reasons.

Firstly, the mesh topology makes representation of an inlet grille bar impossible and

secondly, the emphasis of this work is on the geometric shape of the waterjet inlet,

rather than on any effects that can be attributed to an inlet grille. No significant

influence on the flow is expected from the tubing associated with the inspection cover.

Therefore, this geometric feature is omitted from further consideration.

4.1.2 Parameterisation of the Generic Geometry

After careful examination of the shape of several flush-type waterjet inlets used on high­

speed catamaran vessels, as described in Trillo (1994), Seil et al (1997) found that the

generic shape of the waterjet inlet could be approximated by eight parameters for a

given inlet throat diameter (D). This assumes, of course, that the following aspects of

the geometry are fixed:

1) Cross-sectional shape of the inlet throat

2) Generic profile of the inlet lip

3) Profile of the inlet ramp

4) Profile of the inlet opening

5) Maximum width of the inlet opening

6) A set variation of the cross-sectional area of the waterjet inlet from the inlet throat to

the duct exit

7) An effective deadrise of the inlet opening

Seil et al ( 1997) used the following eight geometric parameters to describe a generic

waterjet inlet geometry, subject to the above-mentioned geometric assumptions:

1) a - The angle of inclination of the inlet to the horizontal plane

2) RL - Radius of the inlet lip

3) H - Height of the pump centreline above the inlet opening

4) R0 - Radius of curvature of the centreline of the duct bend

5) LH - Length of the horizontal duct section downstream of the bend

81

6) AJ AT - Ratio of the waterjet inlet duct exit area to the inlet throat area

7) HL - Height of the inlet lip centreline above the inlet opening plane

8) y- Angle of inclination of the raised lip profile

This parametric waterjet inlet geometry is shown in Fig. 4.3.

All of the above geometric parameters with the exception of parameters one, six and

eight may be non-dimensionalised by the diameter of the inlet throat (D), thus allowing

the waterjet inlet geometry to be scaled with the diameter of the inlet throat. This

approach has the following advantages:

1) The inlet shape can be geometrically scaled in size allowing an inlet to be easily

designed for a given application

2) If parameters describing the hydrodynamic performance of the waterjet inlet are

presented in a non-dimensional form with the inlet throat diameter taken as a

characteristic length, then these results can be scaled with the size of the waterjet

inlet. This assumes that the inlet is operating at the same IVR and with similar

upstream boundary conditions.

For the generic waterjet inlets used for the purposes of parametric design space

investigation and optimisation, the aspects of the geometry which were fixed are

discussed below in more detail.

Parametric Definition of Lip Profile

_____.L.-------l t H

~~ I Fig. 4.3 Parameterisation of the generic flush-type waterjet inlet geometry

82

The variation of the cross-sectional area of the waterjet inlet with distance along the

duct centreline, from the inlet throat to the duct exit, can be described by a third-order

polynomial of the form

(4.1)

where A is cross-sectional area, AT the cross-sectional area at the inlet throat, Ao the

cross-sectional area at the duct exit and u the non-dimension length along the centreline

from the inlet throat to the duct exit, defined on the unit interval by

u=S/STot O~u~1 (4.2)

In Eqn 4.2, S represents the arc length along the duct centreline measured from the inlet

throat and STot is the total centreline arc length between the inlet throat and the duct exit.

The coefficients of the cubic polynomial (a1) must be chosen so that the following

constraints on the variation of A with u are satisfied

At u=O, A=AT A>O for O~u~1 (4.3)

For area ratios greater than unity, a linear variation of cross-sectional area along the inlet

duct centreline between the circular inlet throat and duct exit has been assumed. In this

case Eqn 4.1 is reduced to

(4.4)

The choice of polynomial is worthy of further study in order to find an optimum

streamwise cross-sectional area distribution. The linear variation selected for this work

is unlikely to be the optimum profile.

The raised lip profile of the parametric geometry shown in Fig. 4.3, was chosen as this

type of profile is commonly used in waterjet inlets currently installed in many of today' s

large high-speed catamarans as is evident from a reading of Trillo ( 1994 ). This profile is

economical to fabricate from aluminium by virtue of its simplicity. Aluminium is the

metal which is widely used for the construction of the large high-speed catamarans

produced by Australia, due to its relatively light weight in comparison with steel. The

profile of the inlet ramp is taken as a circular arc, again following the generic design of

inlets used on large high-speed catamarans. There is, however, no reason why other

ramp profiles cannot be used. Ffl)rde et al ( 1991) for example, used Bezier curves to

describe the ramp profile of their waterjet inlets.

83

The profile of the inlet opening was obtained by using two Bezier curves, one for the

forward part of the inlet opening, near the inlet ramp and one for the aft section of the

inlet opening, near the inlet lip. The forward section of the inlet opening was taken to be

rectangular in shape, as flat ramp surfaces are used on conventional inlets due to ease of

fabrication. The aft section of the inlet opening was taken to be semi-elliptical in shape,

again following the generic design of current large inlets. This choice was arbitrary and

alternative profiles could have been used instead. The maximum width of the inlet

opening was taken to be 70/6. For the generic geometry described, no effective deadrise

angle was included. Thus the inlet opening lies flat on the horizontal plane. It is

recognised here that many vessels will have an effective deadrise angle on their hull so

that the plane of the inlet opening will be inclined at an angle to the horizontal in such

cases.

4.1.3 Representation of Geometric Features

The use of Bezier curves to represent geometric features of the inlet allows excellent

geometric flexibility and simplicity of formulation. Other possibilities are the use of

cubic curves, Spline curves or B-Spline curves. F~~Srde et al (1991) made extensive use of

Bezier curves and surfaces in defining their waterjet inlet geometry. Bezier curves are

formulated as

n

r(u) = L ~BI,n(u) U E [0,1] (4.5) 1=0

according to Mortenson (1985). In Eqn 4.5, ~ represent the n+1 vertices of a

characteristic polygon and are called "control points" as they directly control the shape

of the Bezier curve. The Bernstein polynomials (B1,n) provide blending (interpolation)

and are defined by

B =nC U 1 (1-u)n-l 1,n 1

(4.6)

where ncl is the binomial coefficient

nc = n! 1 i!(n-i)!

(4.7)

Two Bezier curves are used to define the profile of the inlet opening. By specifying

multiple coincident points at the vertex of the characteristic polygon for the Bezier

curve, the curve is pulled toward that vertex. This effect increases with an increasing

84

number of points coincident with the specified vertex. As a result of this interesting

feature of Bezier curves, it is possible to change the inlet profile from a semi-elliptic

shape to a rectangular shape simply by increasing the number of coincident points at

vertex p2 • Inlet3D allows the order of the Bezier curves representing the inlet opening

profile curve to be specified, thereby allowing different inlet opening shapes to be

generated. Bezier curves could be used to represent the lip and ramp profiles of the

generic waterjet inlet geometry, but were not used for the investigations presented

herein.

4.2 Mesh Topology

The mesh topology for the waterjet inlet flow domain is shown in Fig. 4.4. The ~. 11, ~

directions correspond to gridlines meshing the waterjet inlet in the circumferential,

radial and streamwise directions, respectively. The boundaries of the computational grid

are detailed below. The grid plane at ~=0 represents the waterjet inlet duct exit, whereas

the plane at ~=1 is the curved boundary surface of the external domain which delineates

the boundary of the volume of water examined in the vicinity of the inlet. The grid

G

~X z

Grid in Physical Space

A s Grid in Computational Space

Fig. 4.4 Topology of meshed flow domain

85

boundary at 11=0 is a cylindrical boundary of zero radius lying on the vertical

centreplane (plane of flow symmetry) of the grid. The grid boundary at 11= 1 represents

the surface of the waterjet inlet and the flat horizontal surface of the external flow

domain. The vertical centreplane of the waterjet inlet and the external flow domain are

represented by grid planes at ~=0 and ~= 1, with the grid plane at ~=0 upstream of the

surface at ~= 1.

4.3 Boundary Mesh

The grid generation techniques used by Inlet3D for the meshing of the boundary

surfaces of the grid are discussed in this section. The boundaries of the computational

domain are meshed using two-directional transfinite interpolation. Depending on the

surface being meshed, a two-dimensional Poisson equation solver (elliptic partial

differential equation solver) is additionally used to effect smoothing of the resultant

surface mesh.

4.3.1 Transfinite Interpolation

The two-directional transfinite interpolation algorithm can be written as the sum of

projectors

F1 = P~r F2 = P

11(r-F1)

r(~, 11) = F1 + F2

(4.8)

where ~ and 11 (in this case), represent any two curvilinear coordinate directions. The

projectors P~ and P11

are idempotent linear operators (see Gordon and Hall (1973))

allowing unidirectional interpolation in each computational coordinate direction. The

order of the unidirectional interpolation is immaterial as noted by Thompson et al

(1985). The projectors used in the grid generator described in this work are of the form

(4.9)

according to Thompson et al ( 1985), where ~ and 11 again represent arbitrary coordinate

directions. The functions <P and '¥ in Eqn 4.9 are blending functions responsible for

providing blending (interpolation) of points and first derivatives in the ~ direction,

86

respectively. It is of course possible to use projectors that interpolate higher-order

derivatives, as discussed by Eriksson ( 1982), Eriksson ( 1985), Smith ( 1982) and

Fletcher ( 1987b ). This allows greater control of the resultant interior mesh. Both linear

and Hermite cubic interpolation are used to effect blending. Linear blending functions

result in the following form for <I> and '¥

<l>z = u

'¥2 = 0

whereas the Hermite cubic interpolant takes the form

<1> 1 =2u3 -3u2 +1

'¥1 = u 3 - 2u 2 + u

<1>2 = 1- <I> I

'¥z=U3-u

(4.10)

(4.11)

The blending function parameter (u) is related to the general computational coordinate

(~1) through a control function of the form

(4.12)

For the waterjet inlet boundary mesh, the following form of the control function was

used

(4.13)

when there is little stretching of gridlines in the ~~ direction. For large stretching in the ~~

direction, h takes the form of the stretching function used, in this case the stretching

function of Vinokur (1983). The control function h(~1) is used as an approximate means

of relating the computational coordinate to the arc length along the gridlines in the ~~

direction. The reader is referred to Smith ( 1982) for a more detailed discussion of the

relationship between the computational coordinate, gridline arc length and the blending

function parameter (u). The grid planes at ~=0 and ~=1 are initially generated by

transfinite interpolation.

4.3.2 Stretching Function

In the context of the grid generation procedures presented in this chapter, there is a need

for a general two-sided stretching function allowing arbitrary grid point clustering to be

specified independently at each end of the interval under consideration. The primary

reason for such a requirement is often the need to resolve different length scales in the

87

flow. This is particularly pertinent in the case of applying grid point clustering in the T\

(radial) direction inside of the waterjet inlet, where the resolution of the near-wall flow

requires the resolution of much smaller length-scales than the flow outside the boundary

layer. The two-sided stretching function of Vinokur (1983) is implemented in Inlet3D.

The interested reader is referred to Vinokur' s paper for a detailed discussion of the

mathematical formulation of this stretching function.

4.3.3 Smoothing of the Surface Mesh

The vert-ical centreplane of the waterjet inlet and external domain (comprising the ~=0,

~=1 and T\=0 boundaries) is initially meshed by using transfinite interpolation and then

subsequently smoothed by iteration of a two dimensional Poisson equation solver. The

two-dimensional Poisson solver solves the following set of equations

a2ll a2ll a2s a2s ox2 + oy2 = P(T\,s) , ox2 + oy2 = Q(T\,s) (4.14)

The source terms on the right-hand side of the equations (P and Q) are "control

functions" used to control the gridline spacing in the interior of the grid. Solution of Eqn

4.14 requires a solution forT\, sin the physical domain. The actual solution of Eqn 4.14

is carried out in the computational domain (T\,s). A transformation from the physical

domain to the computational domain is therefore necessary. This yields the following set

of equations

(4.15)

where gu, g22 and g12 are the components of the covariant metric tensor defined by

The control functions P and Q are chosen as

P=- ~·~11- r;,·r~~ 1~/ l~f

r~ · r~~ r~ · r;,11

Q=- 1~12 - 1~12

88

(4.16)

(4.17)

in order to ensure that the grid is orthogonal at the boundary of the domain, as outlined

in Thompson et al ( 1985). The subscripts appearing in Eqn 4.17 denote partial

differentiation with respect to the subscript. Using this choice of P and Q, an iterative

solution methodology is used to smooth the centreplane grid. The control functions are

initially calculated on the boundary curves from the centreplane grid. The boundary

values of P and Q are then interpolated into the domain using a simple linear

interpolation,

P(Tt.s) = (1- u)P(T\,0) + uP(Tt,l)

Q(T\.s) = (1- v)Q(O,s) + vQ(l,s)

The parametric variables u and v ofEqn 4.18 are

(4.18)

(4.19)

Eqn 4.15 is then solved for the spatial coordinates x and y of the centreplane grid. The

values of P and Q on the domain boundaries are recalculated and the above process is

repeated for the specified number of iterations, or until a convergence criteria is

satisfied. Sorenson ( 1982) used this approach for meshing a two-dimensional

augmentor-wing configuration. Eqn 4.15 is then solved using Gauss-Seidel iteration

with successive over relaxation (SOR).

4.4 Interior Mesh

The interior mesh is initially generated from the boundary mesh using three-directional

transfinite interpolation and subsequently smoothed by several iterations of a Poisson

equation solver which provides elliptic smoothing of the grid.

4.4.1 Transfinite Interpolation

The three directional transfinite interpolant mapping the relationship between the grid in

computational space and the grid in physical space can be written as the sum of

projectors

F1 =P~;r

F2 = PTJ (r- F1 )

F3 = i\ (r- F1 - F2 )

r(~.T\.s)=FI +Fz +F3

89

(4.20)

and is in essence an extension of the case of two-directional transfinite interpolation of

Eqn 4.8. Hermite cubic interpolation is used for all unidirectional interpolations as it is a

simple and effective means of providing blending. The parametric variable (u) used in

the blending functions is related to the computational coordinate (~') through a control

function

u=h(~') (4.21)

For the interior mesh of the flow domain, h is of the same form as Eqn 4.13 when there

is little stretching of gridlines in the ~· direction, whereas for a large stretching in the ~· ,

h takes the form of the stretching function used, in this case the stretching function of

Vinokur (1983).

4.4.2 Smoothing of the Interior Mesh

The smoothing of the grid is carried out in PreBFC, a preprocessor for use with Fluent.

The interior grid generated by transfinite interpolation provides a starting solution for

PreBFC's three-dimensional Poisson equation solver, which is iterated several times

using SOR in order to smooth the grid. The three-dimensional Poisson solver solves the

following equations

a2~ a2~ a2~ ax2 + ()y2 + az2 = P(~. 11.~)

d211 d211 d211 dX 2 + dy2 + dZ2 = Q(~. 11'~) (4.22)

a2~ a2~ a2~ ax2 + ()y2 + ()z2 = R(~, 11.~)

The actual solution of Eqn 4.22 is carried out in the computational domain (~.11.~). The

spatial variation of the control functions P, Q and R must be determined throughout the

interior grid. In order to achieve this, the control functions are determined on the

boundary from the spatial coordinates of the boundary mesh. For a detailed discussion

of the calculation of these control functions on the boundary surfaces, the reader is

referred to Thompson et al ( 1985) and Thomas ( 1982). The control functions

determined on the boundary surface meshes are subsequently interpolated into the

interior using transfinite interpolation with linear blending functions. For example, the

interpolation of the control function R into the interior of the flow domain

90

(4.23)

This is achieved by interpolating the boundary values on the four faces on which ~

varies into the interior of the flow domain, as outlined by Thompson et al ( 1985).

Similar expressions can be derived for the variation of P and Q throughout the interior

of the domain.

4.5 Quality of the Generic W aterjet Inlet Grid

Due to the finite size of finite-volume cells, discretisation errors arise from the finite

difference representation of derivatives in the governing equations. These discretisation

errors represent the difference between the exact solution to the set of governing partial

differential equations to be solved and their finite difference representation. It is

therefore necessary to solve the set of governing partial differential equations on a

sufficiently fine grid in order to reduce these discretisation errors to an acceptable level.

Ideally, the mesh should be of sufficient density to reduce the discretisation error to such

a degree that the CFD solutions so obtained may be considered to be "grid independent"

or "grid convergent". The accuracy of the turbulence model used can then be assessed

independently of the numerical errors associated with the discretisation of the governing

equations as Wilcox (1993) discussed.

In order to demonstrate grid convergence, it is necessary to refine the grid so as to

decrease discretisation error. One common way of doing this is to double the number of

grid points in each computational coordinate direction. If this is computationally

impractical, then the grid can be halved and the results compared. Richardson

extrapolation described by Roache (1976), can be used to determine discretisation error

using the computational results obtained on two different grids. In reality, the ability to

obtain grid-independent solutions for a series of simulations may be beyond the

91

capability of the hardware available. Therefore, a compromise must be made between

accuracy, memory requirements and solution time, all of which influence the size of grid

used.

Qualities of a grid considered desirable (in order to facilitate accurate solution of the

governing equations and good calculation convergence behaviour) include good cell

orthogonality, cell aspect ratios close to unity, uniform cell size, grid smoothness and

alignment of the grid with streamlines (see Fluent (1996)). Practical resolution of

boundary layers and other regions of large flow gradients inevitably requires a stretched

grid. This will degrade solution accuracy due to the introduction of diffusive and

dispersive terms in the truncation error for the finite-difference representation of

derivatives in the governing equations as noted by both Fletcher ( 1987b) and Thompson

et al (1985). Therefore, the growth of the grid in any curvilinear coordinate direction

should be minimised. In regions where the solution does not vary appreciably, the rate

of cell size growth is of less concern as Thompson et al ( 1985) noted. Large cell aspect

ratios resulting from the stretching of grid lines may result in problematic calculation

convergence behaviour when the local flow vector has a component across the smallest

dimension of the cell. This may also result in a degradation of calculation accuracy.

Convergence difficulty may also be experienced when there is excessive cell skewness

throughout the domain. Excessive cell skewness near wall boundaries can cause the

accuracy of the boundary condition treatment to deteriorate.

Smoothness of gridlines (curve smoothness along the curves defining the gridlines) is

required for accurate application of convective differencing schemes in the numerical

algorithm, although higher-order convective differencing schemes are less susceptible to

lack of gridline smoothness (see Fluent (1996)). In order to reduce numerical diffusion,

it is desirable to align the gridlines with the general flow direction. Patankar ( 1980)

noted that numerical diffusion can be reduced by the use of higher-order convective

differencing in the discretisation process.

Excessive cell skewness does occur for some cells in the forward lower inlet region and

in parts of the external domain close to the inlet, (close to the surface and near the side

92

of the inlet). The amount of cell skewness increases as the angle of inclination of the

inlet to the horizontal decreases. The primary cause of this cell skewness is the angle of

intersection of grid lines in the ~ and ~ directions on the inlet surface. This is an

unfortunate consequence of the topology used and may result in a degradation of

solution accuracy for boundary layer flow. Near the inlet lip, where high flow gradients

are encountered (due to stagnation and lip flow) the grid exhibits good orthogonality, so

accuracy in this region should not be degraded by cell skewness. Good cell

orthogonality is also exhibited in the region of the inlet ramp, toward the inlet

centreplane (~=0, ~=1 and 11=0) so accuracy should not be degraded in this region. This

region is of key concern in determining the possibility of flow separation as flow enters

the inlet. Excellent cell orthogonality is present in the boundary and interior mesh in the

upper part of the inlet, inclined duct section, bend and horizontal duct section. Fig. 4.6

illustrates aspects of the mesh discussed above.

Good Cell Orthogonality in Lip Region

Good Cell Orthogonality Near Cent rep lane

Good Cell Orthogonality

Large Cell Skewness

Complete Waterjet Inlet Geometry Shown

Fig. 4.5 Examination of grid quality (Complete geometry shown)

The grid is highly stretched in the 11 direction due to the need to resolve boundary layer

flow and there is therefore considerable variation in cell size within the domain. The

93

growth in cell size is, however, smooth and gradual and so the increase in the truncation

error of the discretised equations should be acceptable. Large cell aspect ratios occur

close to the wall boundaries as a consequence of grid stretching in the 11 direction and

may cause convergence difficulties. The use of Additive Correction Multigrid (ACM)

(discussed in Chapter 3), helps to overcome these difficulties by reducing global

solution errors.

4.6 Grid Generation for Bends and S-Ducts

The first two validation cases of Chapter 5 are a 90° bend and an S-Duct, both of

constant circular cross-sectional area. In comparison to the grid generation techniques

described in previous sections, the meshing of these geometries is a relatively

straightforward process.

4.6.1 Grid Topology

Pure single-block BFC structured meshes are used for both geometries, as shown in Fig.

4.6. The ~. 11. ~ directions correspond to gridlines meshing the geometries in the

G

F y

~ z

B

Grid in Computational Space

c

Grid in Physical Space

Fig. 4.6 Mesh topology for duct of circular cross-section

94

circumferential, radial and streamwise directions, respectively. From Fig. 4.6 it is

evident that an 0-grid topology is used to mesh the cross-section of these ducts.

4.6.2 Meshing Procedure

The basic idea behind the generation of these grids is to generate an 0-type grid at the

cross-section corresponding to the inlet of the flow domain under investigation and then

sweep this cross-sectional mesh (aligned normal to the tangent to the centreline) along

the centreline of the respective duct in order to generate the complete mesh. Using this

approach, the grid point locations relative to the axis can be written as

x = xc + fs (D I 2) cos9cos~

Y = Yc +fs(D/2)cos9sin~

z = fs (D I 2) sin 9

(4.24)

where D is the diameter of the duct, ~ is the angle of inclination of the cross-section to

the horizontal and 9 is the angle of the radial gridline to the vertical plane of symmetry

of the duct. Subscript c denotes centreline location and fs is the value of the stretching

function (in this case the stretching function of Vinokur (1983)) applied in the 11 (radial)

direction.

The above-mentioned approach is much simpler than specifying the mesh on the domain

boundaries, either by transfinite interpolation or other means and then using transfinite

interpolation to generate the interior mesh. Simple programs were written to accomplish

the meshing of both geometries. The above methodology may be termed "direct grid­

point specification".

4.6.3 Grid Quality

The grids produced by "direct grid-point specification" possess excellent orthogonality.

Since the primary flow direction will essentially be aligned with the l; curvilinear

coordinate, additional numerical diffusion caused by flow/grid alignment effects should

be minimised. There will be some additional numerical diffusion associated with

secondary flow behaviour in the cross-stream direction. The stretching of the grid in the

radial direction is necessary in order to resolve the growth of boundary layers on the

duct walls and will increase the truncation error in the discretised equations. The large

95

aspect ratios of the cells near the wall may create convergence difficulties, but as

discussed in Section 4.5, ACM can be used to overcome these problems.

4.7 Closure

In this chapter, a description was given of the generic parametrically-defined flush-type

waterjet inlet geometry used as the basis of the design and optimisation-related

investigations presented in later chapters. The process of mesh generation takes a

significant proportion of the CFD analysis cycle time and therefore must be reduced to a

minimum when a large number of different configurations are to be analysed. It was

therefore decided to develop a grid generator specifically for the purpose of minimising

the time dedicated to grid generation.

A Fortran 77 program, lnlet3D, was written specifically for meshing the generic waterjet

inlet geometry. lnlet3D uses a single-block body-fitted-coordinate (BFC) structured grid

to mesh the geometry of the waterjet inlet and a semi-ellipsoidal external flow domain

below the inlet. lnlet3D uses a combination of transfinite interpolation and elliptic

partial differential equation solution techniques (based on the iteration of a Poisson

equation solver) in order to mesh all flow domain boundaries. The interior of the flow

domain is generated initially by transfinite interpolation and then smoothed by iteration

of a Poisson equation solver in the Fluent preprocessor, PreBFC. The stretching function

of Vinokur ( 1983) is used to effect grid point clustering where required.

The merits and limitations of the grid produced by lnlet3D were discussed. In particular,

the generic grid produced possesses good grid orthogonality in the upper part of the

waterjet inlet and in the inlet lip region. The cell skewness in the lower-forward-side

part of the inlet increases as the angle of inclination of the inlet to the horizontal

decreases. This growth in cell skewness is a direct result of the grid topology used and

the geometry being meshed. It is also a fundamental reason why obtaining suitable

meshes for flush-type waterjet inlet geometries is difficult when BFC meshes are used.

The grid is also highly stretched in the radial direction in order to resolve the wall

boundary layers and this results in large cell aspect ratio variations.

96

The grids used in the first two experimental validation cases presented in Chapter 5 are

generated using a single-block BFC structured grid topology. Owing to the simplicity of

these geometries it is possible to generate these grids directly by specifying their grid

point locations utilising explicit trigonometric relationships. Again, the stretching

function of Vinokur (1983) is used to effect grid-point clustering. The grids produced

possess excellent orthogonality by virtue of the topology used, but large cell aspect

ratios and significant stretching in the radial direction result from the need to resolve

wall boundary layers.

97

Chapter 5 Experimental Validation

A necessary component of any CFD study is an assessment of the accuracy of the

computational solution against relevant experimental data. By "benchmarking" the

computational solution against experimental data, aspects related to the computational

simulation such as the grid size, boundary conditions, or the turbulence model, can be

adjusted in order to improve the agreement with the experimental data, within the

limitations of the modelling used.

Good agreement between theoretical computations and experimental results lends

credibility to the efficacy of the CFD methodology used, therefore justifying the use of

CFD as a reliable analysis tool for the particular application under consideration. This is

of particular importance to design and optimisation-related work.

In this chapter, computational solutions obtained using Fluent are compared with

experimental data in order to assess the accuracy of the CFD modelling methodology

that was outlined in Chapter 3 and used in subsequent chapters of this thesis. Of

particular relevance to the work presented in later chapters are the grid sizes necessary

to reduce discretisation error of the discretised governing equations to an acceptable

level, the accuracy of the two-equation k-e turbulence modelling used and the near-wall

modelling (via wall functions) used to "bridge" the viscosity-affected near-wall region.

Boundary conditions are also of great importance in ensuring accurate solution, as they

reflect the realism of the boundary data, but vary according to the problem analysed.

The flow in a flush-type watetjet inlet is subject to significant streamline curvature as a

result of the different aspects of the waterjet inlet geometry, such as the inlet ramp, lip

and the bend in the waterjet inlet. At the inlet lip there is a line of flow stagnation and

then large pressure gradients as flow accelerates into the inlet or outside of the inlet.

98

Since the work presented in this thesis is based upon the CFD simulation of flush-type

waterjet inlets, it is necessary for the purposes of credibility and completeness, to

validate the accuracy of CFD computation using Fluent against an actual flush-type

waterjet inlet.

In addition to the flow in a flush-type waterjet inlet, experimental data sets for the flow

in S-shaped inlet ducts and pipe (or duct) bends offer rigorous test cases for validation

of a CFD program, as the essential physics of the flow are similar to that in a waterjet

inlet. This is due to the streamline curvature, adverse pressure gradients and secondary

flow behaviour inherent in such flows. Using experimental data for flow in a bend or an

S-shaped duct as a validation case, offers the additional advantage that this type of flow

is essentially equivalent to the flow in a ram-type waterjet inlet. This, therefore, allows

the accuracy of such waterjet inlet flows to be assessed as an additional outcome of the

validations presented herein. No further discussion of this issue is, however, presented.

Three different validation studies were conducted. These were the 90° bend of Enayet et

al (1982), an S-Duct of Bansod and Bradshaw (1972) and a perspex model of a flush­

type waterjet inlet tested in the wind tunnel at The University of Tasmania by Roberts

(1998). The model waterjet inlet of Roberts (1998) is based upon generic designs used

in the marine industry for the propulsion of large high-speed catamaran vessels and is

therefore a realistic, industrially-relevant example of a commercially-available waterjet

inlet design.

The generic flow behaviour for flows in bends, S-Ducts and flush-type waterjet inlets is

discussed in Section 5.1 for reasons of completeness and for the benefit of the reader. In

Section 5.2 a validation study of the flow through the 90° bend of Enayet et al (1982)

is presented. Validation studies are also presented for an S-Duct of Bansod and

Bradshaw (1972) and the model waterjet inlet of Roberts (1998), in Section 5.3 and

Section 5.4 respectively. The accuracies of the computational solutions obtained for

each validation case will be discussed in the relevant section. Section 5.5 provides a

general discussion of the results obtained in previous sections, relating the accuracy of

the computational results obtained to the strengths and limitations of the underlying

99

CFD modelling methodology. The conclusions of the experimental validation studies

are presented in Section 5.6.

5.1 Generic Flow Behaviour

An understanding of the generic physical behaviour of a particular flow provides a

necessary foundation for an intelligent interpretation of results obtained from a CFD

simulation of that flow. The generic behaviour of flows in bends, S-Ducts and flush-type

wateljet inlets is presented in this section, since these are the flows which the validation

studies presented in later sections of this chapter are based upon.

Flow development in bends, S-Ducts and flush-type waterjet inlets involves a complex

interaction of pressure gradients, streamline curvature and secondary flow effects. It is

interesting to note that the development of secondary flow within a curved duct is

essentially an inviscid phenomena, given an initial upstream boundary layer, as Bansod

and Bradshaw ( 1972) noted.

An inviscid-flow theory relating the effect of upstream vorticity on the downstream

secondary flow in a cascade of airfoils was developed by Squire and Winter (1951 ).

Using this theory, Squire and Winter related the strength of the axial vorticity of the

flow leaving a cascade to the angle of deflection of the flow by the cascade and the

upstream velocity gradients. The Squire and Winter relationship is

~ =-2eau ~ dZ (5.1)

In Eqn 5.1 ~~ is the streamwise vorticity at the duct exit, 8 the angle of deflection of the

streamwise flow by the duct, U the streamwise velocity at the entrance to the duct and z

the cross-stream direction (which is evident from the geometric definitions of the flow

domains presented later in this chapter). It is clear from Eqn 5.1 that the greater the

cross-stream vorticity ((}U/(}z) upstream of the duct and the larger the deflection of the

flow is, the stronger will be the strength of the streamwise vortex developed in the duct

and hence the secondary flow. The upstream vorticity is equivalent to the boundary layer

thickness in this case.

100

5.1.1 Flow in Bends

In the literature there is a significant number of publications dealing with the subject of

flow in pipe and duct bends of both circular and rectangular cross-section. Papers such

as those by Ward-Smith (1963), Rowe (1970), Enayet et aJ (1982) and Taylor et aJ

(1982a) are examples of such publications. A characteristic which distinguishes flows in

duct bends from those in straight ducts is the generation of stream wise vorticity, or

"secondary motion" within the duct. This results in a pressure loss, the spatial

redistribution of streamwise velocity and increased heat transfer at the duct wall, as

noted by Taylor et al (1982a).

As the flow enters the duct bend cross-stream static pressure gradients develop between

the outside of the bend and the inside of the bend, as a result of the streamline curvature

present. A larger static pressure develops at the outside of the bend than at the inside.

The adjustment of streamwise flow in the bend creates ~ cross-flow in the core region

from the inside of the bend to the outside. The larger static pressure at the outside of the

bend creates an adverse pressure gradient that thickens the boundary layer as a result of

flow deceleration. The boundary layer at the inside of the bend experiences a beneficial

pressure gradient and flow acceleration, thus thinning it. The cross-stream static

pressure gradient causes a flow of boundary layer fluid from the outside of the bend

toward the inside. This creates a secondary flow which convects boundary layer fluid to

the inside of the bend. This results in a build-up of low momentum fluid at the inside of

the bend and the formation of a pair of counter-rotating vortices which further assists in

convecting boundary layer fluid to the inside of the bend. The reversal of the streamwise

pressure gradient over the second half of the bend also assists in the thickening of the

boundary layer at the inside of the bend. Ward-Smith (1963) showed that the effect of

the duct bend is to influence the static pressure distribution both upstream and

downstream of the bend.

As Taylor et aJ (1982a) explained, factors which affect the development of the flow in

the pipe bend and the nature of the secondary flow include: the radius of the pipe bend,

the distribution of the cross-stream vorticity upstream of the bend (the boundary layer

profile) and the flow Reynolds number. Bends of greater curvature experience larger

101

pressure gradients and hence stronger secondary flows. The larger pressure gradients

present cause a greater flow acceleration at the inside of the bend and a greater flow

retardation at the outside of the bend. This has the effect of displacing the core flow

toward the inside of the bend. Larger centrifugal forces are applied to the in viscid core

flow with increasing Reynolds number, due to higher bulk fluid velocities. This will

affect the location of the "inviscid" core flow and hence the overall flow behaviour in

the bend.

5.1.2 Flow in S-Shaped Ducts

Much experimental and CFD work has been presented in the aeronautical literature

dealing with the subject of flow in S-shaped ducts. The S-shaped duct is of particular

relevance to the design of gas turbine inlet ducting, where the engine compressor inlet is

offset from the air inlet. Little and Trimboli ( 1982) for example, presented the results of

an experimental investigation of S-Duct diffusers for high-speed prop-fans. Kitchen and

Sedlock (1983) discussed the development and challenges faced in the design of

diffusers for advanced tactical aircraft.

S-Ducts may be diffusing or non-diffusing depending on the variation of duct cross­

sectional area along the duct centreline. Diffusing S-ducts not only have centreline

curvature but also a cross-sectional area increase. As Wellborn et al (1994) noted, the

adverse pressure gradient caused by the increasing cross-sectional area of the duct can

lead to flow separation in the duct.

There are numerous references in the literature presenting experimental data for the flow

inS-Ducts. Among these are the velocity and turbulence measurements (obtained from

experiments) for flow in a diffusing S-Duct by Whitelaw and Yu (1993). Wellborn et al

(1994) provided an extensive set of pressure, velocity and flow visualisation data for

flow in a diffusing S-Duct of circular cross-section. Guo and Seddon (1982), Guo and

Seddon ( 1983a) and Guo and Seddon ( 1983b) investigated the effect of mass flow-rate

and flow incidence on the swirl in S-Ducts. They were able to link the resultant swirl at

the exit of the duct with flow separation effects at the duct inlet caused by flow

incidence and proposed ways of reducing swirl by the use of spoilers.

102

A description of the flow behaviour in S-Ducts of circular cross-section is presented

below and is taken largely from Bansod and Bradshaw (1972) and Wellborn et al

(1994). It must be noted that the flow over the first bend of an S-Duct is essentially the

same as that for an equivalent bend of the same angle. As discussed in Section 5.1.1, the

cross-stream pressure gradients that develop over the first bend create a movement of

the low momentum fluid in the boundary layer toward the inside of the bend.

As the flow enters the second bend, the cross-stream pressure gradient reverses, by

virtue of the adjustment of the streamwise flow in the bend. This results in a greater

static pressure at the outside of the second bend than at the inside. The reversal of the

sign of the cross-stream pressure gradient tends to make the boundary layer fluid

migrate circumferentially to the inside of the duct bend. The boundary layer fluid near

the outside of the duct bend experiences a negligible circumferential component. The

circumferential movement of boundary layer fluid developed over the first bend requires

time to reverse. The combination of this secondary flow behaviour and the presence of a

favourable stream wise pressure gradient on the outside of the bend (after the middle of

the second bend), results in the formation of a pair of strong counter-rotating vortices.

These vortices convect low momentum boundary layer fluid to the inside of the duct

bend. Toward the duct exit, the result is a characteristic region of low total pressure and

velocity at the outside of the second bend.

It is interesting to note that although the mechanisms by which these vortices are formed

in bends and S-Ducts involve a combination of cross-flow, convergence and streamwise

acceleration, the process is basically an inviscid one given an initial boundary layer, as

Harloff et al (1993) showed from the solution of the Euler equations for flow in an S­

Duct with an initial upstream boundary layer velocity distribution.

As Wellborn et al (1994) pointed out, the convection of low momentum fluid toward the

centre of the duct by stream wise vortices results in a decrease in both the uniformity and

magnitude of the total pressure at the exit of an S-Duct. This is also applicable to the

flow in bends.

103

5.1.3 Flow in Flush-Type Waterjet Inlets

There have been few publications presenting experimental flow results for flush-type

waterjet inlets available in the literature. Much of the experimental work undertaken has

been by waterjet manufacturers who are reluctant to release commercially sensitive

information to the public domain. Publications appearing in the literature include the

works of Okamoto et al (1993), Griffith-Jones (1994) and Roberts (1998). The design of

condenser scoops (engine cooling inlets) is in many respects similar to that of flush-type

waterjet inlets. Information on the performance of condenser scoop designs can be

found in Hewins and Reilly (1940) and English (1974).

In Section 1.3 .1, aspects of the flow through a flush-type waterjet inlet were briefly

discussed. The flow through a flush-type waterjet can be understood by considering the

underlying geometry. The presence of the inlet ramp causes an acceleration of the flow

into the inlet, with a reduction of static pressure on the ramp surface. Stagnation of the

flow at the inlet lip causes a high static pressure along the stagnation line with

subsequent acceleration of the flow into the inlet (above the stagnation line), or out of

the inlet (below the stagnation line). This causes rapid changes in static pressure in the

vicinity of the inlet lip.

It is interesting to examine how the bend will affect flow behaviour in the inlet. As

discussed in Section 5.1.1, it may be expected that the outside of the bend will

experience an increase in static pressure that will extend both upstream and downstream

of the bend. An adverse pressure gradient experienced by the flow upstream of the bend

undoubtedly affects the boundary layer development. Similarly, the pressure distribution

associated with the inside of the bend will generate a favourable pressure gradient

upstream of the bend. The secondary flow behaviour at the duct exit is dependent, not

only upon the geometry of the bend, but also upon the vorticity upstream of the bend.

The generic centreline static pressure distribution (including the effect of the bend) for a

flush-type waterjet inlet can be seen from the results of Okamoto et al (1993). It is

interesting to note that as the inlet velocity ratio (IVR) increases, there is a general

decrease in static pressure coefficient. It must be further noted that the values of static

104

pressure coefficient presented in the paper are indicative of hull boundary layer

ingestion (due to the decreased dynamic pressure upstream of the inlet) as static pressure

recovery would have been greater had there been no boundary layer ingestion. The

upstream hull boundary layer characteristics will therefore clearly affect the flow in the

waterjet inlet including the location of the lip stagnation line, the total pressure recovery

in the inlet, static pressure distribution over the surface, the velocity, static pressure and

secondary flow behaviour at the duct exit.

Griffith-Jones (1994) used a wind-tunnel to investigate a complete waterjet model

(including fan and nozzle) and the corresponding flush-type waterjet inlet (without

impeller shaft). He found that the presence of the fan shaft and fan had little effect on

the bare duct flow. Consequently it is acceptable to omit the modelling of the pump

shaft and the pump prewhirl when either experimentally or computationally modelling

the flow in the waterjet inlet. An experimental modelling of the complete waterjet is,

however, necessary to determine the effect of flow non-uniformity (at the duct exit) on

pump performance.

The development of the secondary flow within the waterjet inlet is of particular

relevance to the quality of the flow at the duct exit and is therefore discussed below. The

discussion presented below is constructed from the results of Griffith-Jones (1994),

Roberts (1998) and the author's own CFD investigations. The secondary flow present at

the duct exit is qualitatively similar to that in an S-Duct, but the mechanisms responsible

for its formation are in fact different. From the results of Griffith-Jones (1994), it is

evident that the diffusing flow into the inlet creates an upward secondary flow pattern in

both the core flow and the boundary layer flow on the side-walls of the inlet. In addition,

flow diffusion and the physical length of the inlet ramp cause a thickening of the

boundary layer on the ramp surface. Depending on the geometry of the waterjet inlet and

the IVR, the flow may or may not separate on the upper-ramp surface. The formation of

additional secondary flow, as stream wise vortices in the inlet comers (between the ramp

surface and the inlet side-walls) may also occur (depending on the geometry), thus

affecting the development of the boundary layer to some extent. In the absence of

detailed experimental data it is difficult to quantify this effect and its importance, which

105

is probably minor. The behaviour of the secondary flow in the inlet side-wall boundary

layers is equivalent to that in the first bend of an S-Duct. In the wateijet inlet, the

secondary flow in the core is toward the ramp (convex) surface, whereas over the first

bend of the S-Duct it is toward the outer radius of the bend and so the two flows differ

in this regard.

As the flow enters the waterjet inlet duct bend, the secondary flow in the core continues

to be directed toward the top of the duct, in this case from the inside of the bend toward

the outside of the bend. The larger static pressures at the outside of the bend act to direct

the secondary flow in the boundary layer toward the inside of the bend. The boundary

layer fluid near the outside of the bend experiences a negligible circumferential

component. The circumferential movement of boundary layer fluid developed upstream

of the bend requires time to reverse. The combination of this secondary flow behaviour

and the presence of a favourable streamwise pressure gradient at the outside of the bend,

after the middle of the duct bend, results in the formation of a pair of counter-rotating

vortices that convect low momentum boundary layer fluid to the outside of the bend.

This assists in the development in the low total pressure and velocity region at the

outside of the bend.

As mentioned above, the larger static pressures at the outside of the bend act to direct

the secondary flow in the boundary layer toward the inside of the bend. This is certainly

the case for the secondary flow in the boundary layer at the side of the bend, which is

now directed toward the inside of the bend. With the secondary flow in the core directed

toward the outside of the bend and the secondary flow in the boundary layer directed

toward the inside of the bend, a second pair of streamwise vortices appears on the side

of the duct. This vortex acts to convect boundary layer fluid toward the lower part of the

duct.

The secondary flow development over the waterjet inlet bend is thus analogous to

secondary flow development in the second bend of an S-Duct with similar generic

secondary flow behaviour at the duct exit. As discussed above, the mechanisms for the

formation of upstream vorticity are different for the waterjet inlet and an S-Duct. The

106

actual secondary flow behaviour in the waterjet inlet will depend on several factors.

These include the geometry of the waterjet inlet, IVR, upstream boundary layer

thickness and the presence or absence of flow separation in the waterjet inlet.

5.2 Flow in a 90° Bend

In this section the accuracy of the CFD techniques described in Chapter 3 are assessed

and benchmarked against experimental data for flow in the goo bend of Enayet et al

(lg82). Enayet et al (lg82) investigated the laminar and turbulent flow of water in a goo

bend of circular cross-section using Laser-Doppler velocimetry. Their study was focused

on obtaining mean velocity and turbulence profiles at various cross-sections in the bend

and downstream of the bend, as well as wall static pressure measurements. The study of

flow in a goo bend is an interesting and rigorous CFD validation case because of the

combination of streamline curvature, adverse pressure gradient and secondary flow

development. A turbulence model must adequately model all of these features in order

to produce an accurate solution.

5.2.1 Experimental Configuration

The experimental geometry shown in Fig. 5.1 , consists of a goo bend of circular cross­

section with an internal diameter of 48 mm and a centreline radius of curvature of 134

mm. The bend is machined from two halves of a perspex block fixed in the horizontal

plane and fitted with upstream and downstream tangents that are 240 mm and 480 mm

long, respectively. The experimental geometry is fitted in a special flow rig through

which water was pumped at a Reynolds number of 43000 (based on the 48 mm duct

diameter and a volumetrically-averaged velocity in the duct of o.g2 mls).

5.2.2 Computational Modelling of Experimental Configuration

The experimental configuration is modelled from the start of the upstream tangent to six

duct diameters downstream of the bend. The topology and techniques used to mesh the

modelled flow domain are discussed in Section 4.6 and shown in Fig. 4.6.

Since the experimental data presented by Enayet et al ( }g82) showed that the flow is

symmetrical about the plane of geometric symmetry, only half of the experimental

107

configuration is meshed as shown in Fig. 5.2. Since Fluent solves an elliptic form of the

Reynolds-averaged Navier Stokes (RANS) equations, Dirichlet or Neumann boundary

conditions must be specified on all boundaries of the flow domain. Table 5.1 lists the

grid planes bounding the flow domain and the Fluent boundary condition cell type (see

Fluent ( 1996) for a detailed description of Fluent boundary condition cell type

definitions and applications) applied to each boundary.

48±01 bore 76

Sectaon through bend

0"

Fig. 5.1 Experimental geometry of Enayet et al ( 1982)

Global conservation of mass implies that the mass flow into the upstream tangent must

be equal to the mass flow out of the flow domain at the exit of the downstream tangent.

Therefore Boundary 1 is set as a Dirichlet boundary condition with the inlet velocity

specified on it. Sufficiently far downstream of the bend, it may be assumed that the

stream wise gradients of the transported quantities ( <l>) are much smaller than in the

vicinity of the bend and may be set to zero as a reasonable approximation.

108

y

~ z

Fig. 5.2 41x41x81 surface grid bounding the modelled flow domain

Grid Plane in Computational Space Boundary Plane ~rnm ~max Tlmm Tlmax ~min lmax Fluent

1 ~=0.0 0.0 1.0 0.0 1.0 0.0 0.0 Inlet

2 ~=1.0 0.0 1.0 0.0 1.0 1.0 1.0 Outlet

3 1'}=0.0 0.0 1.0 0.0 0.0 0.0 1.0 Axis 4 11=1.0 0.0 1.0 1.0 1.0 0.0 1.0 Wall

5 ~=0.0 0.0 0.0 0.0 1.0 0.0 1.0 Symmetry

6 ~=1.0 1.0 1.0 0.0 1.0 0.0 1.0 Symmetry

Table 5.1 Relationship between boundary conditions and mesh topology

109

Thus, a Neumann boundary condition of the form

(5.2)

was set for Boundary 2. Boundary 5 and Boundary 6 lie on the plane of flow symmetry

and are therefore set as mixed Dirichlet/Neumann boundary conditions such that

velocities normal to the plane of flow symmetry and normal gradients of transported

quantities are set to zero. In other words

U·n=O cv~)·n =O

(5.3)

where U is the velocity vector, ~ a transported quantity and n a vector normal to the

symmetry plane. Boundary 3 which lies on the centreline of the flow domain, is a

Dirichlet boundary condition obtained by averaging the values of ~ from the cell-centres

of neighbouring cells according to

- 1 ""' ~P = N L.J ~nb nb

(5.4)

In Eqn 5.4, N is the number of neighbouring cell-centre locations surrounding point P.

Boundary 4 represents the wall bounding the flow domain. It is here that the no-slip

condition is applied in conjunction with the near-wall modelling (using wall functions)

as discussed in Section 3.3.

5.2.3 Computational Simulation

A plug-flow velocity profile of 0.92 rnls is specified on Boundary 1 at the entrance to

the modelled flow domain. A water density of 1000 kg/m3 was specified and the

molecular viscosity adjusted to give the appropriate Reynolds number.

All computations were executed on Hewlett Packard K210 workstation servers. In fact,

all computations used to obtain the results presented in this thesis were executed on the

above-mentioned hardware. Calculations are deemed to have converged when the sum

of the normalised residuals of the transport equations and the pressure correction

equation are reduced below lxl0-3 (see Section 3.6.2) which is Fluent's default

convergence criteria.

110

Computations were initially executed using the RNG k-e turbulence model and the

Standard wa11 function on a number of successively finer grids, with progressive

refinement in the circumferential (I), radial (J) and streamwise (K) directions. This was

done in order to determine a suitable grid size necessary to reduce discretisation error to

levels sufficiently low that the solution may be considered to be "grid-independent". For

all intents and purposes, it was found that a reasonably grid-independent solution could

be obtained on a 41 x41 x81 grid.

With discretisation errors reduced to acceptable levels, the effect of turbulence and near­

wall modelling on the solution was examined, using both the RNG k-e turbulence model

and the Standard k-£ turbulence model, with the Standard wall function and the Two­

Layer-Based Nonequilibrium (TLBN) wall function. The results of the validation study

are presented in the next section.

5.2.4 Experimental and Theoretical Comparisons

In this section, the accuracy of the computational results is assessed against

experimental data obtained from the measurements of Enayet et al ( 1982). The predicted

flow behaviour will be discussed in relation to the underlying physical flow behaviour.

The computed distribution of streamwise velocity and turbulence intensity at 0.58

diameters upstream of the bend is compared with the corresponding experimental data

in Fig. 5.3. The data has been plotted against distance from the wall, along the

centreplane of the duct (y), non-dimensionalised by the duct diameter (D). In all relevant

figures presented in this section, the direction of increasing y is taken to be from the

inside of the bend to the outside. The computational results presented in Fig. 5.3 are

obtained on a 4lx41x81 grid using the RNG k-e turbulence model with the Standard

wall function. From Fig. 5.3a it can be seen that the core velocities are within a relative

error of 7% of the measured values. Since Enayet et al (1982) quoted systematic and

random errors in their laser-Doppler velocimetry (LDV) measurements of U!Urer of

2.5% and ±1.5% respectively, it is possible that the actual error between the measured

and predicted velocities may in fact be less.

111

1.20

1.00

0.80

'i! ~ 0.60 ::J

0.40

0.20

• Expenment -CFD 7% Error bars shown

q - N ~ ~ ~ ~ ~ ~ ~ 0 0 d d d d d d d d d ~

y/D

a) Boundary layer velocity distribution

0.08

O.o7

0.06

0.05 'i! ~0.04 -=

0.03

0.02

0.01

• Experiment -CFD

0.00 -t'-'-'-"'t'-'-'-'t'-'-'.Lf-'-1-JL.I..f-'-LLL.j-'-LJ.J.fllJ.J.fll.L.I.fJ-'.L.I.fJ-'u.J..j o-N~~~~r-:~~q dddddddodd-

y/D

b) Boundary layer turbulence distribution

Fig. 5.3 Boundary layer velocity and turbulence data 0.58 diameters upstream of bend

The computed turbulence intensity, calculated as a root mean square (rms) value

(V(2k/3)) based on the computed distribution of k, agrees well with the measured

streamwise intensity. This suggests that there is little anisotropy in the normal velocity

fluctuations. Therefore, the upstream turbulence intensity has been well modelled.

The inaccuracy in the modelling of the boundary layer velocity profile upstream of the

bend will affect the flow behaviour downstream and this must be considered in the

interpretation of results. The modelling of the flow upstream of the bend is deemed to be

acceptable by the author, despite the apparent inaccuracy in the calculation of the

velocity profile, due to the lack of sufficient information on near-wall velocity and

turbulence data in the paper of Enayet et al (1982) for y/D<0.05.

Before evaluating the accuracy of various turbulence and near-wall modelling

combinations, a number of simulations were run using the RNG k-e turbulence model in

combination with the Standard wall function in order to determine the required grid size

necessary to reduce discretisation error to an acceptable level. The results of the

investigation are shown in Fig. 5.4, which compares the measured and computed

streamwise velocity across the centreplane of the duct at the 8=75° plane in the bend

(Fig. 5.4a) and one diameter downstream of the bend (Fig. 5.4b).

112

~ ::::>

1.25

1.00

0.75

0.50

0.25

• Experiment -21 x21 x81 -31 x31 x81 ······41 x41 x81 --31 X 31 X 121 - - 41 X 41 X 161 · - - - 51 X 51 X 81

0.00 ll-'-'..l.f-ll'-'+'-'.J.J.fJ-I..I..Lf-'.!.i.Lf-l..llJ..j-J.J..L.1.1f.!-LLLI-'-'-'-'+'-'-'-.

0 Nl"''"'<tlt")\Of"-000\C! 0000000000

y/D

a) &=75° plane

1.25

1.00

... 0.75

ti ::::>

0.50

0.25

Experiment --21 x21 x81 --31x31x81 ······41x41x81 --31 X 31 X 121 - - 41 X 41 X 161

- - 51 X 51 X 81

0 N 1"'1 "'<t lt") \0 f"- 00 0\ 0. 0 . o· o· o· · · · · · · 0 0 0 0 0 0

y/D

b) One diameter downstream of bend

Fig. 5.4 Effect of grid size on computed streamwise velocity

Comparison of the results for the 31x31x81 grid with those of the 31x31x121 grid and

comparison of the 41 x41 x81 grid with the 41 x41 x 161 grid reveal that refinement of the

grid in the streamwise direction appears to have no effect on the solution. The solution

instead changes with successive refinement of the gird in the circumferential and radial

directions. The initial 2lx2Ix81 grid is clearly too coarse. The 41x41x81 grid may be

considered to have reduced discretisation error to an acceptable level and thus may be

taken as yielding a "grid independent" solution. While this is not exactly true, a further

refinement to a 51x51x81 grid results in essentially no change in the streamwise

velocity one diameter downstream of the bend. Only a slight change in the streamwise

velocity in the 8=75° plane, in the range Qs;y/Ds;O.l6, is noticeable.

With an appropriate grid size selected, simulations were run with both the RNG k-e and

the Standard k-e turbulence models using both the TLBN wall function and the Standard

wall function. The results are compared with the experimental data in Fig. 5.5.

Examination of the results in Fig. 5.5a, reveals that the streamwise velocity is relatively

well predicted for y/D>0.36, but deteriorates for y/D<0.36. The use of the Standard wall

function yields qualitatively better results with either turbulence model than the use of

the TLBN wall function. Overall, the use of the RNG k-e turbulence model provides

results that may be judged to be in closest agreement with the experimental data.

113

Toward the outside of the bend, all turbulence and near-wall modelling options yield

essentially identical results for the mean velocity profile.

1.25

1.00

- 075 e

2 ::>

0.50

0.25

• Experiment

--RNG k-e, TLBN Wall Fn. • • • • • • RNG k-e, Std. Wall Fn.

--Std. k-e, TLBN Wall Fn.

Std. k-e, Std. Wall Fn. 0. 00 .-w-'-1-'-'--'-'+'-'-'-'-t'-'-'-'+.u..w...I-'-'-'-'-P-'-'-'+'-'-'-'+'-'--'-'+'-.......

0 N M ~ ~ ~ ~ 00 ~ ~ d d d d d d d d d d

y/D

a) 9=75° plane

e ::> ....... ::>

1.25

1.00

0.75

0.50

0.25

• Experiment --RNG k-e, TLBN Wall Fn . •••••• RNG k-e, Std. Wall Fn.

-- Std. k-e, TLBN Wall Fn.

0 N M ~ ~ ~ ~ 00 ~ ~ d d d d d d d d d d

y/D

b) One diameter downstream of bend

Fig. 5.5 Effect of turbulence and near-wall modelling on computed streamwise velocity

Examination of the results in Fig. 5.5b clearly shows that the RNG k-E turbulence model

provides more accurate solution prediction than the Standard k-E turbulence model with

either wall function. Both wall functions appear to give identical results. All turbulence

and near-wall modelling combinations examined here give essentially identical results

for y/D>0.6, but under-predict the core velocity. For y/D<0.5, the accuracy of the

predictions deteriorates, but good results are obtained when the RNG k-E turbulence

model is used.

In Fig. 5.6 and Fig. 5.7 computed distributions of streamwise velocity distribution over

the duct cross-section are compared with experimental measurements in the 75° plane of

the bend and at one diameter downstream of the bend, respectively. The computed

results shown are obtained with the RNG k-E turbulence model and the Standard k-E

turbulence model, both using the Standard wall function. Also shown in the figures is

the secondary flow behaviour calculated with the corresponding turbulence model.

114

b) Normalised Velocity RNGk-£

- =0.20mls

d) Secondary flow RNGk-£

a) Normalised Velocity Experiment

c) Normalised Velocity Standard k-E

- =0.20m/s

e) Secondary flow Standard k-£

Fig. 5.6 Comparison of measured and predicted flow behaviour at the 8=75°plane

115

b) Normalised Velocity RNGk-e

- =020rnls

...

d) Secondary flow RNGk-e

a) Normalised Velocity Experiment

c) Normalised Velocity Standard k-e

- =0.20rnls

e) Secondary flow Standard k-e

Fig. 5.7 Comparison of measured and predicted flow behaviour one diameter downstream of bend

116

The results obtained for the streamwise velocity distribution over the duct cross-section

using the RNG k-E turbulence model are in closer agreement with the experimental data

than those of the Standard k-E model. The difference in the computational results for the

different turbulence models can be understood by examining the secondary flow

behaviour at both locations. Both reveal the existence of a streamwise vortex located

toward the inside of the bend (lower duct in the figure) near the duct centreplane. This

results from the interaction of the boundary layer and the transverse pressure gradient

associated with the bend. Both turbulence models predict a reduction in the intensity of

this streamwise vortex and reduced secondary flow one diameter downstream of the

bend (Fig. 5.7) due to the absence of the transverse pressure gradient at this location.

The stronger secondary flows predicted by the RNG k-E turbulence model at both

locations results in a greater convection of boundary layer fluid toward the centreplane

of the duct and hence a greater accumulation and up-welling of low-momentum fluid

from the inside of the bend toward the duct centreline. In fact, the RNG k-E turbulence

model tends to over-predict this secondary flow, whereas the Standard k-E model under­

predicts the strength of the secondary flow.

As an interesting study of how the predicted turbulence behaviour follows the measured

trend of streamwise velocity fluctuations, the rms velocity fluctuations .V(2k/3) obtained

from the results of k along the centreplane of the duct are compared with the measured

streamwise velocity fluctuations in Fig. 5.8. If the correct distribution of k can be

calculated across the duct, then differences between the calculated rms turbulence

intensity and the streamwise intensity of the turbulence will indicate anisotropy in the

turbulence.

It can be seen from Fig. 5.8a and Fig. 5.8b that the distribution of calculated rms

turbulent fluctuations follows the trend of the measurements. Quantitative agreement

between computation and experiment appears to be relatively poor. In Fig. 5.8b

agreement between the computation and experiment appears to improve with relatively

good agreement obtained in the range 0.3<y/D<0.6 and y/D>0.9.

Two conclusions may be drawn from the results which indicate an uncertainty in the

117

interpretation of the results presented in Fig. 5.8:

1) The different turbulence and near-wall modelling combinations yield different

distributions of k (and hence the rms turbulence intensity) leading to uncertainty

about the accuracy of the calculation of k. Without a definitive experimental

distribution of k to compare the calculated results against, it is impossible to draw a

conclusion as to which of the calculated distributions is the most accurate, hence it

is therefore not possible to gauge the level of anisotropy in the flow.

~ ::s

'i!

0.10 • • Experiment

0.09 • -- RNG k-£, TLBN Wall Fn.

0.08 • • • • • ·- • · RNG k-e, Std. Wall Fn. -- Std. k-£, TLBN Wall Fn.

0.07 • • • Std. k-e, Std. Wall Fn. 0.06 • 0.05

0.04

0.03

0.02 • 0.01 • • • 0.00

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 y/D

a) Effect of turbulence and near-wall modelling- 8=75° plane

0.10 •

0.09 •

0.08

0.07

0.06

1.0

e o.o5 "::s

0.04

0.03

0.02

0.01

• Experiment

-- RNG k-£, TLBN Wall Fn. • • • - • · RNG k-£, Std. Wall Fn. -- Std. k-e, TLBN Wall Fn.

Std. k-£, Std. Wall Fn. 0.00 -f-'-i..L-'-I--'-'-'--'-+-L..l-L...l....\-'-'-....L..J....+-'-1....L...l-f-'-..J....1..-'-+.L-L.J'-'-f....l-L..L..1..+-'-.L-L.J'-I--'-'-'-..J....!

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 y/D

b) Effect of turbulence and near-wall - One diameter downstream of bend

Fig. 5.8 Comparison between measured and computed turbulence intensity

118

2) If it is assumed that the actual distribution of k lies close to the results obtained using

the RNG k-e turbulence model (for example), then as can be seen from Fig. 5.8b,

there appears to be considerable anisotropy in the turbulence close to the inside of the

bend y/0<0.3 and in the core flow for the range 0.62<y/D<0.9. From Fig. 5.8a, there

appears to be considerable anisotropy across the duct centreplane.

Fig. 5.9 provides a comparison between the measured streamwise turbulent velocity

fluctuations and the rms turbulence intensity calculated using both the RNG k-e

turbulence model and the Standard k-e turbulence model. The two turbulence models

have been used with the Standard wall function. It can be seen from both Fig. 5.9a and

Fig. 5.9b that the measured distributions of streamwise turbulence intensity show the

effect of secondary flow-induced velocity gradients on the turbulence development, with

high levels of velocity fluctuations being associated with steep mean velocity gradients.

The contour patterns obtained using the RNG k-e model are qualitatively in better

agreement with the experimental data than those obtained using the Standard k-e model.

This may be attributed to the greater convection of turbulence by stronger secondary

flow and steeper secondary flow-induced velocity gradients over the cross-section

(which act to increase the production of turbulence).

The plotting of velocity vectors on the duct centreplane provides an understanding of

not only the magnitude and direction of the local flow direction, but also the developing

streamwise velocity profile across the duct. The results obtained using the RNG k-e and

the Standard k-e turbulence models, both using the Standard wall function, are shown in

Fig. 5.10. Examination of the figure reveals the dramatic increase in boundary layer

thickness at the inside of the bend driven by the secondary flow. It is interesting to note

that the growth of the boundary layer at the inside of the bend continues downstream of

the bend as a direct consequence of the secondary flow, but one that is decaying in its

strength as it is no longer driven by a transverse pressure gradient associated with the

bend.

Comparisons between the computed non-dimensional turbulent velocity-scale ('..fklUrer)

and the computed turbulent length-scale ( R E /D) on the duct centreplane, obtained using

119

the two turbulence models with the Standard wall function are shown in Fig. 5.11 and

Fig. 5.12 respectively. The turbulent length-scale, used throughout this chapter was

calculated as

.f = C k 312/E E )! (5.5)

Experiment RNGk-E Standard k-E

a) 8=75° plane

Experiment RNGk-E Standard k-E

b) One diameter downstream of bend

Fig. 5.9 Comparison between measured and computed turbulence intensity

120

- =0.92m/s

a)RNGk-£

- =0.92m/s

b) Standard k-£

Fig. 5.10 Comparison of velocity vectors on bend centreplane

121

---2:----

-../k/U,e! D 0.120 c 0.110 8 0.100 A 0.090 9 0.080 8 0.070 7 0,060 6 0.050 5 0.040 4 0.030 3 0.020 2 0.010

0.000

a) RNGk-E

-../k/Ure! D 0.120 c 0.110 8 0.100 A 0.090 9 0.080 8 0.070 7 0.060 6 0.050 5 0.040 4 0.030 3 0.020 2 0.010

0.000

b) Standard k-E

Fig. 5.11 Distribution of computed turbulent velocity-scale on bend centreplane

122

ljD D 0.0497 c 0.0458 B 0.0420 A 0.0382 9 0.0344 8 0.0306 7 0.0267 6 0.0229 5 0.0191 4 0.0153 3 0.0115 2 0.0076

0.0038

a) RNGk-E

IJD D 0.0488 c 0.0451 B 0.0413

A 0.0375 9 0.0338 8 0.0300 7 0.0263 6 0.0225 5 0.0188 4 0.0150 3 0.0113 2 0.0075

0.0038

b) Standard k-E

Fig. 5.12 Distribution of computed turbulent length-scale on bend centreplane

123

It can be seen from both figures that the core flow contains eddies of larger length-scale,

but smaller velocity-scale than those present in the boundary layers. These are the large

energy-containing eddies that are convected with the core flow. The greater turbulence

intensity and smaller length-scales of the eddies in the developing boundary layer at the

inside of the bend suggest smaller more dissipative eddies. Hence, there can be seen to

be a cascade of energy from the large energy-containing eddies in the core flow to the

smaller highly-turbulent dissipative eddies in the boundary layer.

It is evident from Fig. 5.12 that the relatively steep velocity gradients between the

developing boundary layer near the inside of the bend and the core flow lead to

significant turbulence production resulting in increased turbulence levels. Quantitative

differences in rms turbulence levels across the duct can be seen by examining Fig. 5.8. It

is interesting to note the large differences in computed length-scale in the core flow

between the results obtained from the two turbulence models. While the length-scale

predicted by the RNG k-e turbulence model decreases downstream over the bend, the

length-scale predicted by the Standard k-£ model remains high. This suggests that the

length-scale determining procedure (via solution of the £ equation) appears to lack

sensitivity to the effects of the streamline curvature present in the flow (this issue will

be discussed in greater length in Section 5.5). As a consequence, predicted eddy­

viscosity, the Reynolds stresses and hence total pressure loss will be higher in the core

flow when the Standard k-e model is used to model the flow.

5.3 Flow in an S-Duct

The accuracy of the CFD techniques described in Chapter 3 are assessed and

benchmarked against experimental data for flow in an S-Duct of Bansod and Bradshaw

(1972) who investigated the flow in three S-shaped ducts. Their primary objective was

to investigate and explain the formation of a region of low-velocity fluid at the duct exit.

Numerous workers have used CFD to investigate the flow in S-Ducts, but they have

adopted different approaches to the governing equations solved and the turbulence

modelling used. Among these are papers originating from the NASA Lewis group and

include the work of Towne (1984), Vakili et al (1984) and Anderson et al (1994). Other

124

works include papers by Jenkins and Loeffler (1991) and Zhao (1997). Jenkins and

Loeffler (1991) used the algebraic turbulence model of Baldwin and Lomax (1978) to

examine the flow in a compact diffusing S-Duct. Although qualitative features of the

flow were reproduced, overall agreement with experimental data was relatively poor,

especially with regard to wall static pressure distributions. The primary cause of the

poor agreement may be attributed to the limitations of the turbulence model in

modelling flow separation behaviour and other flow effects within the duct.

Towne (1984) used a parabolised form of the Navier-Stokes equations (based on the

assumption that streamwise viscous diffusion is negligible) with algebraic turbulence

modelling. Results in good quantitative agreement with experimental data for flow in

the 22.5°-22.5° S-Duct of circular cross-section of Taylor et al (1984) and the 22.5°-

22.50 S-Duct of square cross-section of Taylor et al (1982b) were obtained. The

corresponding Reynolds numbers of the flows, based on duct diameter, were 48000 and

40000 respectively. Since an algebraic mixing-length turbulence model was used, the

good agreement between the computation and experiment may be attributed to two main

factors. The first being the absence of flow separation in the inlet, due to the mild

curvature of the duct (RID=7) and hence, relatively mild streamwise pressure gradients.

The second reason is the fact that since secondary flow development is to a large extent

an inviscid-flow phenomena, therefore the relatively large boundary layer thicknesses at

the entrance to the S-Duct (ranging from 0.10 to 0.20 of the duct diameter) will have a

predominant influence on the corresponding secondary flow development. Thus the

good agreement with the experiment is not surprising.

Vakili et al ( 1984) compared computational and experimental results for flow in a 30°-

30° non-diffusing S-Duct of circular cross-section using the same parabolised Navier­

Stokes code as that of Towne ( 1984 ). The entrance Mach number of flow entering the S­

Duct was 0.6 and algebraic turbulence modelling was used. Agreement between theory

and experiment was generally good. No separation was present in the duct. Boundary

layers were thin (relative to the duct diameter) in the straight duct section upstream of

the entrance to the S-Duct and the radius of curvature of the bend (RID=5) was smaller

than that of Towne (1984). As a consequence, the accuracy of prediction is more reliant

125

on the accuracy of prediction of the boundary layer development and the effect of

streamwise pressure gradients on the development of turbulence and mean flow

structure. Therefore the limitations of the turbulence modelling used are apparent.

Zhao (1997) investigated the flow in the same non-diffusing S-Duct of circular cross­

section examined by Vakili et al (1984), but using a low-Reynolds number k-e

turbulence model. Overall agreement was reasonable, but wall static pressure

predictions deteriorated over the second bend. This was attributed to the limitations of

the k-e model in handling streamwise curvature and adverse pressure gradient. Zhao

(1997) also investigated the flow in a diffusing S-Duct using the experimental data from

the AGARD Fluid Dynamics Panel Working Group 13 numerical subgroup test case 3.

In this case, the wall static pressure was well predicted despite the failure of the

turbulence model to predict separation in the second bend of the duct at the outer wall.

This is most likely due to the diffusing nature of the flow and rising duct static

pressures, which tends to minimise the effects of flow separation on the static pressure

distribution on the duct surface. The calculated total pressure distribution at the duct exit

was in poor agreement with the experimental data. This was attributed to the poor

performance of the k-E turbulence modelling under the effects of streamline curvature

and large adverse pressure gradient (see Section 3.2.3).

It is thus evident, from the brief literature review presented above, that the application of

CFD to the prediction of the flow in S-Ducts has met with various degrees of success,

depending on the actual flow under consideration and the computational modelling used

to predict that flow.

5.3.1 Experimental Configuration

The geometry of S-Duct configuration three of Bansod and Bradshaw (1972) is shown

in Fig. 5.13b. The duct has a diameter of 150 mm and consists of two 45° bends, with

the upstream bend having RID=2.25 and the downstream bend having RID=3.5, where

R is the radius of curvature of the duct centreline and D the duct diameter. A straight

duct section of 0.5 D length separates the two bends and the downstream bend is

followed by a straight duct section of 0.5 D.

126

Flow conditions at the entrance to the S-Duct are provided by the blow-down rig shown

in Fig. 5.13a. Air from the blower is diffused into the settling chamber (where the

honeycombs are located) before blowing through the contraction unit. Inside the 150

mm straight duct section upstream of the S-Duct inlet, the air is tripped to induce

turbulence using a circumferential trip wire of 0.5 mm thickness. The Reynolds number

of the flow based on the diameter of the duct is approximately 5x 105•

duct e"ltry plane

stat1c hole

trtp w.re O·Srnrn.d 1a

a) Wind tunnel configuration

C3 R/0:2·25 followwd by R/0 =3·5 \Optinun intat.•)

b) S-Duct geometry showing measurement stations

Fig. 5.13 Experimental configuration ofBansod and Bradshaw (1972)

5.3.2 Computational Modelling of Experimental Configuration

The flow domain is modelled from the start of the 150 mm straight duct section

upstream of the S-Duct, through to the exit of the S-Duct. The meshed flow domain is

shown in Fig. 5.14. The mesh topology and the boundary condition types used in the

modelling of the flow domain are identical to those of the 90° bend described in Section

5.2 and therefore the reader is referred to Section 5.2.2 for a more detailed discussion.

127

y

~z

Fig. 5.14 31x31x71 surface grid bounding the modelled flow domain

5.3.3 Computational Simulation

A plug-flow velocity profile of 45 m/s was specified on Boundary 1 of the flow domain.

An air density of 1.293 kg/m3 and a molecular viscosity of 1.75x10-5 Ns/m2 was

specified in order to achieve the experimental flow Reynolds number.

A number of simulations were executed using the RNG k-£ turbulence model with the

Standard wall function on successively finer grids in order to determine the minimum

size of grid necessary to reduce discretisation error to acceptable levels and so ensure a

reasonably "grid-independent" solution. For all intents and purposes, it is found that this

could be achieved with a 31x31x71 grid. Calculations were judged to have converged

when the sum of the normalised residual of the pressure correction equation and the

other transport equations is reduced to below 1xl0-3•

With discretisation errors reduced to acceptable levels, simulations were subsequently

run in order to assess the relative accuracy of different turbulence model and wall

function combinations. Both the Standard k-£ and the RNG k-£ turbulence models are

examined with the Standard wall function and the TLBN wall function. The results of

128

the validations study will be presented in the next section.

5.3.4 Experimental and Theoretical Comparisons

In order to determine a suitable size of grid necessary to ensure that discretisation error

is sufficiently reduced to acceptable levels, simulations were run on a number of

successively finer grids using the RNG k-E turbulence model with the Standard wall

function. The results of the grid study are shown in Fig. 5.15, where the computed total

pressure loss coefficient (Cp) along the centreplane of the duct at Station 4 and Station 5

are compared with the corresponding experimental data in Fig. 5.15a and Fig. 5.15b,

respectively. In Fig. 5.15, distance from the lower wall (y) has been non­

dimensionalised by the duct diameter (D). The plots shown are consistent with an

upwelling of low-momentum fluid close to the wall resulting from the development of

secondary flow in the form of streamwise vortices. These vortices are responsible for the

convection of low-momentum fluid from the boundary layer into the core flow, thus

causing a loss in the average total pressure of the fluid over the duct cross-section.

1.0

0.9

0.8

0.7

0.6 ... 0.5 u

0.4

0.3

0.2

0.1

0.0

0.0

• Experiment -21 x21 x71 --31 x31 x71 ······41x41x71 --31 x31 x 106

41 X 41 X 141

0.1 0.2 0.3 0.4 0.5 y/D

a) Station 4

1.0

0.9

0.8

0.7

0.6

u 0.5

0.4

0.3

0.2

0.1

• Experiment 21 X 21 X 71

--31 x31 x71 ••• · ··41 x41 x 71 --31 x31 x 106

- 41 X 41 X 141

0.0 +--<--L..L.J...+-'-'-.1-l--f--'--L-L-L-J.--L-'-_,_y~~

0.0 0.1 0.2 0.3 0.4 0.5 y/D

b) Station 5

Fig. 5.15 Effect of grid size on computed total pressure loss coefficient

The results of Fig. 5.15a reveal that the 21x2lx71 grid is too coarse, as the change to a

3lx31x71 grid yields results in better agreement with the experimental data for

y/D<0.03 and y/D>0.12. There is some variation in the solution with increasing grid

refinement in the range of 0.03<y/D<0.12, but the solutions are essentially identical for

129

y/D>0.15. Examination of Fig. 5.15b again shows the 21x21x71 grid to be too coarse,

with the solutions on the finer grid being in better accord with the experimental data.

The solutions on the finer grids are essentially identical over the complete range of y/D.

Therefore for all intents and purposes, the 31x31x71 grid is chosen for further

computational investigation, as it yields a suitably "grid-independent" solution.

The effect of turbulence and near-wall modelling on the predicted Cp at Station 4 and

Station 5 can be seen from Fig. 5.16. In Fig. 5.16a it can be seen that both turbulence

models under-predict the peak total pressure loss at y/D=0.12. The Standard k-£

turbulence model with both wall functions and the RNG k-£ turbulence model with the

TLBN wall function over-predict the total pressure loss for y/D<0.06 and under-predict

the boundary layer thickness. The RNG k-£ model using the Standard wall function

provides the best qualitative agreement with the experimental data. At Station 5 (Fig

5.16b) the results for the Standard k-£ are relatively poor showing significant differences

between the computed and measured values of total pressure loss. The total pressure

loss is greatly over-predicted in the range 0$;y/D$;0.18-0.20 and significantly under­

predicted in the range 0.20~y/D~0.42. The results for the RNG k-£ are in much better

agreement with the experimental data.

1.0

0.9

0.8

0.7

0.6

u 0.5

0.4

0.3

0.2

0.1

• Experiment -- RNG k-e, TLBN Wall Fn. • • • • • · RNG k-e, Std. Wall Fn. -- Std. k-e, TLBN Wall Fn.

Std. k-e, Std. Wall Fn.

0.0 +-c-.L.L-'-i---'-'-'-'--+~~!1-J.i ....... '*"'::W<~

0.0 0.1 0.2 0.3 0.4 0.5 y/D

a) Station 4

1.0

0.9

0.8

0.7

0.6

c)0.5

0.4

0.3

0.2

0.1

• Experiment

-- RNG k-e, TLBN Wall Fn. • • - • • · RNG k-e, Eq. Wall Fn. -- Std. k-t, TLBN Wall Fn.

Std. k-t, Eq. Wall Fn.

0.0 -f--.-.--'---'--1f--L--'-~..L-.1-14~~Pi!:r;~ 0.0 0.1 0.2 0.3 0.4 0.5

y/D

b) Station 5

Fig. 5.16 Effect of turbulence and near-wall modelling on computed static pressure loss coefficient

Fig. 5.17 compares the measured distribution of Cp over the cross-section at Station 5

(Fig. 5.17a) with the distributions obtained using the RNG k-£ model (Fig. 5.17b) and

130

the Standard k-E model (Fig. 5.17c), both using the Standard wall function. The use of

the same wall function allows a consistent comparison with the data of the performance

of the two turbulence models. Also shown are the associated secondary flow vectors

over the duct cross-section. It is evident that the Cp contours calculated using the

Standard k-E model qualitatively are in better accord with the experimentally measured

distribution than those obtained using the RNG k-E model. However, it is seen from

Fig. 5.16b that the RNG k-E model produces a superior prediction of the centreplane

distribution of Cp. This is evident also from Fig. 5.17b. The RNG k-E model does show

a greater "tucking-in" of Cp contours toward the duct centreplane. The contour

associated with Cp =0.4 is poorly represented.

The extent of the core flow is over-predicted in the lower-half of the duct by both

models, but particularly so by the RNG k-E model. The RNG k-E model yields a more

accurate prediction of the core flow in the upper-half of the duct. Both turbulence

models over-predict the Cp of the boundary layer at the top of the duct, with the RNG k­

E turbulence model providing the greatest over-prediction.

Secondary flows and the adverse pressure gradient associated with the second-half of

the second bend will assist in the thickening of the boundary layer in the upper duct. The

difference between the results obtained using the two turbulence models and the

experimental data can be explained by examining the secondary flow behaviour over the

duct cross-section. Fig. 5.17d and Fig. 5.17e show the predicted secondary flow

behaviour for both turbulence models. In both cases, it can be seen that a streamwise

vortex is predicted near the duct centreplane. This vortex is associated with the

accumulation and up-welling of low-momentum fluid in the lower-half of the duct near

the wall, while at the same time thinning the boundary layer on the side of the duct. A

second streamwise vortex system is also evident from the figure, one which extends

from the lower-half of the duct to the centreplane on the upper-duct. This vortex

convects boundary layer fluid from the side of the duct, thus thinning the boundary layer

there and accumulating boundary layer fluid at the top of the duct. The existence of two

vortices is therefore physically realistic and to be expected from the distribution of Cp

over the duct cross-section.

131

b) Cp distribution- RNG k-e

-- - =5mls ~::-:;-' I -I I I-

I \ ' ' I I 1 \ ' ' '

\ .... ' ' I I 1 1 ' , ' , 1 \'i

\ ' ' .\ l I \ I \ \ ' ' ' \ \1 I \ \ \ ' ' ... I\ ,,,, ', ' ' ' ... ~

\\ \ \ \ I

"''',, \ \ ' ' '' ' .. ,,,, ,, \ \ ' ' ' ' I I "" \ \ ' ' \ , 0 ' ' \ '' ' ' ' \ ~ ~ ' \ \ \ ' '~',

... ' ' ' l/1 ,/ .. ..:- ~ ~ \ ~ \ \ \ -,,,,~- _....... ' ' \ \ \ Jl

- ' ' \ \ f/// ,' \ \ I I .//, -- ' \ \ \ /I/.,.,- ,' \ I I,,~.

lf/-'' \\ I ,''•·t.

{

\ \ I ,,,\ 1\ If I

I ,, I,

,. \ - I I I I t,, ,, ....... ;I 1.~,: '\-.......-~~ ~ ~

,,' , I

, _...,.-­

""',"'

/ C3S I

I I

~ 6\ ...... ·01 \ ,""' '..., ,' ·1 .... ___ ..

·2 ·3 .,

9 = 18cf

a) Cp distribution - Experiment

d) Secondary flow - RNG k-E

c) Cp distribution- Standard k-E

- =5mls

. '-I '•

e) Secondary flow- Standard k-E

Fig. 5.17 Comparison of measured and predicted flow behaviour at exit of S-Duct

132

The Standard k-E model under-predicts the intensity of the streamwise vortex in the

lower-duct resulting in insufficient accumulation of low-momentum fluid there. The

RNG k-E model on the other hand, tends to over-predict the strength of the secondary

flow resulting in the "tucking-in" of the total pressure contours toward the duct

centreplane. Both turbulence models most likely over-predict the secondary flow

associated with the upper vortex. This results in excessive accumulation of low­

momentum fluid at the top of the duct. This is particularly evident for the RNG k-E

model (Fig. 5.17b and Fig. 5.17d) which shows larger secondary flow vectors close to

the wall than the Standard k-E model.

Fig. 5.18 shows the comparison between measured static pressure coefficient (Cp) and

that calculated using the RNG k-E turbulence model with the Standard wall function, on

the duct centreplane, at the top and bottom of the duct. Cp has been plotted against arc

length along the duct centreline (S) non-dimensionalised by the duct diameter (D). fu

fact, the other turbulence and near-wall modelling combinations used give essentially

identical results when plotted in the form of Fig. 5.18. It can be seen that good

agreement has been obtained between the computational results and the experimental

data over the first bend. The static pressure is also well predicted over the top of the first

half of the second bend, but deteriorates thereafter, with static pressure under-predicted

by Cp =0.08 at the duct exit. The static pressure over the bottom of the second bend is

over-predicted as a consequence of the failure to adequately predict the onset of the

region of separation/near-separation measured by Bansod and Bradshaw ( 1972).

Further insight into the inaccuracy in calculated static pressure distribution may be

obtained by considering the distribution of calculated and measured skin friction

coefficient (Cr) at the top and bottom of the duct centreplane, as we11 as the calculated

and measured mass-weighted total pressure loss coefficient ( Cp ) over the duct cross-

section. Calculated and measured CP is tabulated in Table 5.2 for the different

turbulence model and wall function combinations examined.

133

u

u

0.4

0.3

0.2

0.1

0.0

u"" -0.1

-0.2

-0.3

-0.4

-0.5

• Experiment - Top A Experiment - Bottom

--CFD-Top - CFD- Bottom

-0. 6 ..P....L...L.J...f-'-'--L..L..j-'-'-'-'-t-'-L...I...J...f-'-'--L..L..j-L..L..L.'-t-'-.L..L.J...f-'-'-.L...L.f-L.W..'-t-'-.L..L.J...f-'-'-.L...L.f-J...J...L.J...j

6.0E-03

5.0E-03

4.0E-03

3.0E-03

2.0E-03

I.OE-03

O.OE+OO

?.OE-03

6.0E-03

5.0E-03

4.0E-03

3.0E-03

2.0E-03

I.OE-03

O.OE+OO

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 SID

q 0

0 d

Fig. 5.18 Distribution of static pressure overS-Duct centreplane

Vl q "1 q Vl 0 Vl 0 Vl d N c-4 rri rri -o:i -o:i

SID

a) Top of S-Duct

Vl q "1 0 Vl q Vl 0 Vl d C'i C'i ('f'l <"i -o:i ...j

SID

b) Bottom of S-Duct

• Experiment - Top - RNG k-£, TLBN Wall Fn. • • • • • · RNG k-£, Std. Wall Fn. -- Std. k-e, TLBN Wall Fn.

- - Std. k-e, Std. Wall Fn.

• q "1 0

..0 Vl Vl

• Experiment

-- RNG k-£, TLBN Wall Fn. • • • • • · RNG k-e, Std. Wall Fn. -- Std. k-£, TLBN Wall Fn.

Std. k-£, Std. Wall Fn.

q "1 0 Vl Vl ..0

Fig. 5.19 Distribution of skin friction on S-Duct centreplane

134

Bansod and Bradshaw ( 1972) calculated mass-averaged total pressure loss coefficient as

- 4 i21tlDn pU (P - P) Cp = --2 1

ret 2

rdrd8 1tD 0 0 PrerUref 2PrerUref

(5.6)

where suffix "ref' refers to the reference conditions, taken to be PreF 1.293 kglm3,

UreF45ms-1 and PreF1309 Pa.

Turb. Model Wall function Cp Rei. Error Absolute Error

RNGk-e TLBN 0.090 47.5% 0.029

RNGk-e Standard 0.092 50.8% 0.031

Standard k -e TLBN 0.097 59.0% 0.036

Standard k -e Standard 0.099 62.3% 0.038

Measured Cp =0.061

Table 5.2 Mass-averaged total pressure loss coefficient at duct exit

For all turbulence and near-wall modelling combinations studied here, the relative error

in the over-prediction of Cp ranges from 47.5% to 62.3%, which is a significant over­

prediction of total pressure loss. The absolute error in predictions ranges from 0.029 to

0.038. The Standard k-e model predicts greater total pressure loss than the RNG k-e

model when the same wall function is used. In both cases, the use of the TLBN wall

function lead to a slightly lower loss in total pressure than use of the Standard wall

function.

Over-prediction in total pressure loss is consistent with over-predictions in the

distribution of skin friction coefficient over the duct surface, as can be seen from Fig.

5.19a and Fig. 5.19b. In the case of the distribution of Cr over the top of the duct (Fig.

5.19a), calculated Cr agrees well with measurements over most of the duct, except over

the second half of the second bend where it is over-predicted. Interestingly, this

corresponds to the location where the static pressure at the top of the duct is under­

predicted, as mentioned above. This suggests an over-prediction of the velocity in the

core flow adjacent to the boundary layer.

At the bottom of the duct, Cr is significantly over-predicted over the first bend

135

(S/0<1.97) and is poorly predicted over the remainder of the duct except near the exit.

The poor agreement over the first bend may be attributed to the effect of streamline

curvature on the boundary layer turbulence structure and hence momentum transfer.

This issue will be discussed in greater length in Section 5.5. Although the computational

models predict a reduction in Cr upon entry to the second bend, they fail to adequately

capture the region of separation/near-separation that would be evident by a reduction in

the computed Cr to zero, or alternatively a near-zero value of Cr. As a consequence of

this failure to capture the separation/near-separation behaviour and the subsequent fall

and recovery in Cr, inaccuracies in computed downstream flow behaviour are reflected

in the incorrectly calculated distribution of Cr at the bottom of the duct over the second

bend.

Further understanding of the relative performance of the two turbulence models studied

can be gained by examining the flow behaviour on the duct centreplane, corresponding

approximately to the plane of flow symmetry. Fig. 5.20 shows the distribution of total

pressure loss coefficient over the duct centreplane for both models using the Standard

wall function. It can be seen that the RNG k-E turbulence model provides a more

realistic description of the build-up of low-momentum fluid in the bottom of the duct,

beginning after the first bend. This is not surprising considering the superior agreement

with the experimental data at Station 4 and Station 5, shown in Fig. 5.16.

The velocity on the duct centreplane for both turbulence models using the Standard wall

function (shown using velocity vectors in Fig. 5.21), provides a comparison of the

developing velocity profile as the flow moves downstream. Of particular interest is the

development of the velocity profile at the bottom of the duct over the second bend. The

smaller velocities calculated adjacent to the wall (at the bottom of the duct) using the

RNG k-E model demonstrate that model's superior ability to handle streamline curvature

and adverse pressure gradient in comparison to the Standard k-E model, which shows a

fuller velocity profile.

136

CP B 1.00 A 0.90 9 0.80 8 0.70 7 0.60 6 0.50 5 040 4 0.30 3 0.20 2 0.10

0.00

a)RNGk-E

Cp B 1.00 A 0.90 9 0.80 8 0.70 7 0.60 6 0.50 5 040 4 030 3 0.20 2 0.10

0.00

b) Standard k -£

Fig. 5.20 Computed distribution of total pressure loss coefficient on duct centreplane

Comparison between the calculated turbulent velocity and length-scales for the two

turbulence models are shown in Fig. 5.22 and Fig. 5.23 respectively. As would be

expected, the core flow contains eddies of large length-scale and low turbulence

intensity. These are the large energy-containing eddies convected by the core flow. The

137

a)RNGk-E

b) Standard k-E

Fig. 5.21 Computed velocity vectors on duct centreplane

velocity and length-scales of the core flow over the second bend are greater in the case

of the results obtained using the Standard k-E model, suggesting greater eddy-viscosity.

This results in greater Reynolds stresses and hence a greater drop in total pressure in the

core flow. This is exactly what happens as a careful study of Fig. 5.20 reveals.

The turbulence in the developing boundary layer at the bottom of the duct is

characterised by larger velocity-scales and smaller length-scales than the core flow. This

138

..Jk/Uref 0 0120 c 0.110

0.100 0090 0080 0.070 0.060 0.050 0.040 0.030 0.020 0.010 0.000

a)RNGk-E

..Jk/Urer c 0110 B 0100 A 0.090 9 0.080 8 0.070 7 0060 6 0.050 5 0.040 4 0.030 3 0.020 2 0.010

0.000

b) Standard k-E

Fig. 5.22 Computed distribution of turbulent velocity-scale on duct centreplane

indicates the presence of eddies which are not only highly turbulent but also more

dissipative. The larger velocity gradients present in the outer part of the boundary layer

result in large shear and turbulent production. This in turn results in the observed

increase in velocity-scales observed there.

Of particular interest is the distribution of turbulent length-scale throughout the flow

domain. The RNG k-£ results show smaller length-scales after the first bend than the

139

corresponding length-scale distribution obtained using the Standard k-E model. The

reduction in length-scale, as can be seen from Fig. 5.23, results in lower eddy-viscosity

and lower Reynolds stresses at the start of the second bend. This results in reduced

momentum transfer through the boundary layer and hence a greater thickening of the

boundary layer at the bottom of the duct. This is in contrast to the flow behaviour

predicted by the Standard k-E model. The explanation of this behaviour will be

presented in Section 5.5.

1/D B 0.0624 A 0.0565 9 0.0505 8 0.0446 7 0.0386 6 0.0327 5 00267 4 0.0208 3 0.0149 2 0.0089

00030

a)RNGk-E

1/D B 0.0583 A 0.0528 9 0.0472 8 0.0417 7 0.0361 6 0.0305 5 0.0250 4 0.0194 3 00139 2 0.0083

0.0028

b) Standard k-E

Fig. 5.23 Computed distribution of turbulent length-scale on duct centreplane

140

5.4 Flow in a W aterjet Inlet

In this section the accuracy of the CFD techniques described in Chapter 3 are assessed

and benchmarked against experimental data for flow in a waterjet inlet model tested in a

wind-tunnel by Roberts (1998). Although the experimental validation cases presented

earlier in this chapter provide rigorous test cases for testing turbulence models and near­

wall modelling techniques, it is crucial to assess their accuracy against actual

experimental data for a flush-type waterjet inlet. This is necessary in order to understand

the limitations of these techniques and draw conclusions regarding predictive accuracy

for different waterjet inlet geometries and flow regimes.

5.4.1 Experimental Configuration

Experimental data was obtained, as part of the Australian Maritime Engineering

Cooperative Research Centre propulsion program, from Mr Jason Roberts who

investigated a 117.5 scale-model waterjet inlet using the closed-circuit wind tunnel of

the Department of Civil and Mechanical Engineering at The University of Tasmania.

The waterjet inlet model, the geometry of which is shown in Fig. 5.24 is fabricated from

perspex and is mounted on the centreline of the side of the 1.2 m long test -section of the

wind tunnel. Details of the dimensions of the waterjet inlet model can be found in

Roberts (1998). No impeller shaft or impeller shaft housing/fairing is installed in the

model. The cross-section of the wind tunnel test-section shown in Fig. 5.25, is octagonal

in shape. As can be seen from Fig. 5.24, the experimental waterjet inlet geometry is a

realistic geometry based on existing designs and so provides an excellent validation

case. The design of the model waterjet inlet has the following geometric features:

1) Rectangular/semi-elliptical profile for the inlet opening

2) Circular inlet lip profile

3) Rectangular to circular transition section on the upper inlet

4) Bend fabricated from straight-duct segments. Roberts (1998) described this as a

"lobster-back" bend.

It must be noted that many installed waterjet inlets have a slanting inlet plane due to

deadrise on the vessel hull and operate in "thick" hull boundary layers, whereas in this

validation study, the ingested boundary layers are "thin". This should not, however,

141

detract from the validity of the conclusions reached regarding the accuracy of CFD in

predicting waterjet inlet flow behaviour. Boundary layer thickness only affects the actual

flow in the waterjet inlet, rather than the accuracy of the CFD methodology presented

herein.

Fig. 5.24 Geometry of waterjet inlet model

100 400 100

100

400

100

Fig. 5.25 Geometry of wind tunnel cross-section

The mass flow-rate and hence the IVR of the flow through the waterjet inlet is

controlled by drawing air through the inlet with an auxiliary fan , with make-up air

drawn in from the ambient surroundings in the return working-section of the wind­

tunnel as described by Roberts ( 1998).

142

5.4.2 Computational Modelling of Experimental Configuration

Inlet3D was modified to mesh the experimental geometry. In particular, the main

modifications are made to incorporate the polynomial ramp profile and the specific

rectangular to circular transition geometry on the upper part of the inlet. In these two

respects, the model waterjet inlet differed from that of the generic parametric geometry

described in Chapter 4. A smooth bend is used to represent the segmented ("lobster­

back") bend geometry present in the experimental model.

An obvious limitation of the mesh topology used by Inlet3D is the difficulty in

representing the geometry of the wind tunnel test-section (domain external to the

waterjet inlet). The mesh topology used by Roberts (1998) meshes the test-section

geometry in a more natural manner, but suffers from other limitations. It was therefore

decided to use the generic external domain produced by Inlet3D, with the underlying

assumption that the flow around the inlet is affected by the geometry and flow

conditions "close" to the inlet, that is in the immediate vicinity of the waterjet inlet. The

validity of this assumption therefore, also has an effect on the choice of boundary

conditions applied to the flow domain.

Since the flow in the waterjet inlet and external domain may be considered to be

symmetrical about the vertical plane of symmetry of the waterjet inlet, only half of the

external domain is meshed. It will be shown in later sections that the flow in the inlet

exhibits a gross swirl (evident from the data supplied for IVR=0.61) at the duct exit and

the reason for this behaviour will be explained. Since the distortion of the total pressure

distribution at the duct exit is relatively small in circumferential deviation from the

vertical centreplane of the duct, meshing only one-half of the waterjet inlet and external

domain is justified. In addition, this approach requires half of the grid size that would

otherwise be required for a given spatial discretisation of the flow domain.

The topology used by Inlet3D to mesh a wateijet inlet and a simplified flow domain

external to it was discussed in Section 4.2. Since the flow under consideration here

exhibits "stationary turbulence", the RANS equations assume an elliptic form thus

requiring the specification of Dirichlet or Neumann boundary conditions on all surfaces

143

bounding the flow domain. Table 5.3 lists the grid planes bounding the flow domain and

the Fluent boundary condition cell type applied to each boundary.

The following observations govern the choice of boundary conditions chosen:

1) Sufficiently far upstream and away from the waterjet inlet, the velocity will assume

that of the wind tunnel free-stream velocity (outside of the boundary layer on the

tunnel wall). Sufficiently far upstream, the static pressure on the tunnel side-wall will

be unaffected by the presence of the inlet ramp.

Grid Plane in Computational Space Boundary Plane ~IDln ~max Tlmin Tlmax ~IDJD ~max Fluent

1 ~=1.0 0.0 0.5 0.0 1.0 1.0 1.0 Inlet 2 ~=1.0 0.5 1.0 0.0 1.0 1.0 1.0 Inlet 3 ~=0.0 0.0 1.0 0.0 1.0 0.0 0.0 Outlet 4 11=1.0 0.0 1.0 1.0 1.0 0.0 1.0 Wall 5 TJ=O.O 0.0 1.0 0.0 0.0 0.0 1.0 Axis 6 ~=1.0 1.0 1.0 0.0 1.0 0.0 1.0 Symmetry 7 ~=0.0 0.0 0.0 0.0 1.0 0.0 1.0 Symmetry

Table 5.3 Relationship between boundary conditions and mesh topology

2) The boundary layer on the flat side surface of the wind tunnel may be approximated

as a two-dimensional flat plate boundary layer. Boundary layer data collected by

Roberts ( 1998) upstream of the waterjet inlet, at several distances from the centreline

of the working section, validates this assumption.

3) Downstream of the waterjet inlet, the mass flow-rate will be reduced relative to that

upstream of the inlet, due to the flow of mass through the inlet. If m. is the mass

flow-rate through Boundary i, then the mass flow-rate through Boundary 2 is simply

(5.7)

Eqn 5.7 expresses a global mass conservation over the flow domain. A uniform

velocity is specified over Boundary 2 with the velocity vector aligned with the x

direction. Although it is recognised that this specification of velocity is physically

unrealistic, it is found to have little effect on the flow upstream, except in the

immediate vicinity of the boundary. This is of no consequence to the waterjet inlet

flow, provided that the boundary is "sufficiently" far away from the waterjet inlet.

144

This approach does have the advantage of allowing a simple specification of mass

flow-rate through the boundary and hence an indirect specification of the mass flow­

rate through the inlet via Eqn 5.7.

4) In the pipe downstream of the exit of the waterjet inlet (in the experimental

configuration) it may be assumed that the streamwise gradients of the transported

quantities <P are much smaller than in the actual waterjet inlet. Hence streamwise

gradients may be approximated by

aq,jax = 0. (5.8)

Therefore an "outlet" (Neumann) boundary condition, where the streamwise gradients

of all transported quantities are set to zero according to Eqn 5.8, is specified as the

boundary condition for Boundary 3. Boundary 3 is located one diameter downstream of

the exit of the waterjet inlet in order to increase the validity of the assumption.

The total length, half-width (since only half of the waterjet was simulated) and depth of

the external domain are specified as 1662 mm, 300 mm and 600 mm, respectively. This

size of external domain is found to give satisfactory results in line with the above

discussion. The surface grid enclosing the simulated flow domain is shown in Fig. 5.26.

Fig. 5.27 shows the grid meshing the plane of flow symmetry.

The first two observations mentioned above are used to provide upstream boundary

conditions for the flow simulation. Since the tunnel boundary layer must be represented

on Boundary 1, boundary conditions for velocity (U), turbulent kinetic energy (k) and

turbulent dissipation (E) must be provided. A simple approach of "growing" a two­

dimensional flat-plate boundary layer in the absence of externally-imposed pressure

gradients is adopted in order to provide approximate values of U, k and E for the wind

tunnel boundary layer on Boundary 1. The boundary layer thickness is approximately

10% of the duct diameter.

5.4.3 Computational Simulation

All simulations were run with a free-stream velocity in the test-section of 22.8 ms-1, an

air density of 1.29 kg/m3 and a molecular viscosity of 1.7lxl0-5 Ns/m2. The quoted

values of air density and molecular viscosity correspond to a temperature of 273 K (see

145

Gerhart and Gross (1985), Table A.3) and are adequate for the purposes of calculation.

v

~X z

Fig. 5.26 31x41xl02 surface grid bounding the modelled flow domain

Fig. 5.27 Grid on centreplane of flow domain

TheRe based on the 600 mm depth of the test section is therefore 1.032x106• At higher

temperatures Re will be less due to decreased air density and increased molecular

viscosity dropping by 25% to 0.7546x106 if the tunnel temperature is 323 K. Griffith­

Jones (1994) found his wind-tunnel results to be fairly insensitive toRe for the order of

146

magnitude of the Re (order of 1 05) used in his experiments. The tunnel free-stream

velocity is deduced from static pressure and static pressure coefficient data supplied to

the author by Mr Roberts assuming the above air density.

Computations were initially performed on a 31x41x102 grid for a nominal IVR of 0.61

using the RNG k-e turbulence model with the TLBN wall function. The grid was

subsequently separately doubled in each computational coordinate direction in order to

determine the effect of grid refinement on the solution. The solutions obtained on the

31x41xl02 grid were used as starting solutions for these cases. Due to the limitations of

the hardware it was not possible to double the grid in all computational coordinate

directions simultaneously. Therefore it was not possible to obtain a definitive conclusion

as to how close the solutions (on the 31x41x102 grid) are to grid-independency.

Calculations are deemed to have converged when the sum of the normalised residuals of

the transport equations and the pressure correction equation fall below lx10-3.

In order to compare the effect of turbulence and near-wall modelling on the solution,

calculations were run using the Standard wall function. The effect of the choice of

turbulence model (at an IVR value of 0.61) is examined by repeating the computations

using the Standard k-e model with the TLBN wall function. Computations were then

made for IVR values of 0.68, 0.80 and 0.97 using the RNG k-e turbulence model with

the TLBN wall function in order to assess the accuracy of CFD calculation at these

values of IVR.

5.4.4 Experimental and Theoretical Comparisons

Comparison between experimental data and CFD calculations are presented for a test

case having a nominal IVR of 0.61. Comparison between measured and calculated static

pressure distributions for higher IVR are then presented. Results have been presented for

the above-mentioned range of values of IVR in order to analyse the trends in predicted

flow behaviour over the range of IVR examined.

Fig. 5.28 shows a comparison between the measured boundary layer velocity and

turbulence intensity profiles 0.5D upstream of the ramp tangency point. It can be seen

147

s E '-'

from Fig. 5.28a that the computed velocity profile is within 10% of the measured

velocity profile, but the boundary layer thickness is over-predicted by the computations.

This indicates that the boundary layer profile specified on Boundary 1 is too thick. The

values of boundary layer displacement thickness and momentum thickness calculated by

CFD were similarly larger than the experimentally-determined values. The comparison

between the measured and calculated rms turbulence intensity is poor, most likely as a

result of inaccuracies in the calculation of boundary layer velocity gradients caused by

the inaccuracy in the boundary layer velocity profile. This affects the turbulent

production and hence the turbulence intensity. It will be shown later in this section that

since the ingested boundary layer is "thin" and due to the limitations of the turbulence

models used, the effect on the solution appears to be minor.

II) 00 0 0

ot¥~~~~~~~~~

-5 -+-CFD

-10 • Experiment 10% Error Bars Shown

-15 xperiment Boundary Layer Thickness= 15 mm

0

-5

-10

-15 ,...... E

0 0 0 0 0

s 0

8 ~ ~ 8 8 ~ 8 s 0 0 0 0 0 0 0 0

1splacement Thickness =19mm >. -20 E -20 '-'

Momentum Thickness =13mm >. -25

-25 CFD

-30 Boundary Layer Thlckness=20 mm Displacement Thickness =26mm

-+-CFD

• Expenment -30

-35 -35 Momentum Thickness =19mm

-40

a) Boundary layer velocity distribution b) Boundary layer turbulence distribution

Fig. 5.28 Boundary layer velocity and turbulence distribution one-half duct diameters upstream of the waterjet inlet

Comparison between the measured and calculated distribution of Cp on the upper and

lower surface of the waterjet inlet surface, on its vertical plane of symmetry, is shown in

Fig 5.29 for the base case corresponding to IVR=0.61. On the upper inlet, the non­

dimensional arc length (SID) is measured from the ramp tangency point. On the lower

surface SID is measured from the lip trailing edge, with SID increasing as arc length is

measured around the lip and into the inlet.

The relationship between SID on the upper and lower centreplane of the waterjet inlet

and the relevant geometric features of the waterjet inlet are shown in Table 5.4. It can be

148

seen from Fig. 5.29a that very good agreement with the measured Cp data has been

achieved over the inlet ramp and the horizontal straight duct section after the bend. The

sharp "kink" in the data at SID=3.9 indicates the onset of flow separation on the upper

part of the ramp, upstream of the transition section. Cp is over-predicted in the range

3.9<SID<6.8 owing to inaccuracies in predicting the correct onset of flow separation

behaviour. Over this range the maximum deviation from the measured results is

approximately 0.1. From Fig. 5.29b, good agreement between measured and calculated

Cp is obtained over the lower inlet centreplane for SID> 1.25. Cp in the vicinity of

SID=0.8 is over-predicted, most likely as a consequence of the incorrect calculation of

flow separation on the upper inlet which leads to larger values of calculated Cp than

those measured.

Upper Centreplane SID Ramp tangency point 0.00 End of ramp I Start of upper duct transition section 4.44 End of transition I Start of bend 5.78 End of bend I Start of straight duct section 6.66 End of straight duct section 7.35

Lower Centreplane SID Lip tailing edge 0.00 End of lip 0.12 Start of bend 1.08 End of bend I Start of straight duct section 1.56 End of straight duct section 2.26

Table 5.4 Relationship between SID and the waterjet inlet geometry

The calculated static pressure distribution over the inlet lip is shown in Fig. 5.29c. Since

no static pressure tappings are placed on the actual lip profile, it is not possible to

compare the calculated distribution of Cp with experimental data. In spite of this, it can

be seen that large variations in Cp occur over the inlet lip, ranging from a maximum of

Cp= 1.0, corresponding to the stagnation point of the dividing streamline (on the

centreplane) on the upper lip to a minimum of Cp=-2.2, indicating strong suction on the

underside of the lip. Maximisation of this minimum static pressure is essential to avoid

the inception of cavitation at high vessel speed.

149

In order to test the sensitivity of the solution to changes in grid size, the grid is doubled

in each computational coordinated direction. No notable change in the solution occurred

as can be seen from Fig. 5.29c and Fig. 5.32a. A complete doubling of the gird in two or

all coordinate directions simultaneously would have been beyond the capability of the

hardware and was therefore not undertaken. While it cannot be claimed that the

computational solution is by any means "grid-independent", it appears from the above­

mentioned grid investigation and the results presented in this section, that a reasonable

level of discretisation error has been achieved.

Fig. 5.30 shows the effect of turbulence and near-wall modelling on the calculated

distribution of Cp on the inlet centreplane for an IVR of 0.61. It can be seen that the

Standard k-E turbulence model yields results that are inferior to those obtained using the

RNG k-E turbulence model, although the differences between the two results are not

large. The static pressure distributions appear to be insensitive to the wall function used.

The cross-sectional distribution of total pressure and secondary flow at the duct exit of

the waterjet inlet is of particular relevance to pump operation as it defines the operating

environment of the pump and hence the effect on pump performance and operational

efficiency of the waterjet unit. The measured distribution of total pressure coefficient

(Cp) at the duct exit is shown in Fig. 5.31a, together with the computed distribution of

Cp in Fig. 5.31c, both at an IVR of 0.61. It is interesting to note that the measured Cp

distribution is asymmetrical about the vertical plane of symmetry of the waterjet inlet,

thus indicating a slight gross swirling motion of the flow. The rotation of the total

pressure distribution about the centreplane is approximately 15° and is therefore small in

circumferential extent. This gross swirl occurs as a result of a combination of the

interaction between flow asymmetries in the inflow to the side-mounted waterjet inlet

(Roberts ( 1998) attributed this to thermal convection in the vertical direction within the

wind tunnel, caused by temperature differentials in the tunnel air), bend pressure

gradients and flow separation. Guo and Seddon (1983a) found that an S-shaped inlet

duct (centreplane lying in the horizontal plane) operating with flow separation at the

inlet lip due to an angle of incidence in the vertical plane, would develop a gross­

swirling pattern.

150

0.4

0.3

0.2

uc. 0.1

0.0

-0.1

-0.2

0.6

0.5

0.4

uc. 0.3

0.2

0.1

1.0

0.6

0.2

-0.2

u -0.6

-1.0

-1.4

-1.8

-2.2

0

0 0 0 0

• Experiment --31 X 41 X 102 • • • • • · 61 X 41 X 102 --31 X 81 X 102

31 X 41 X 203

2 4

SID

a) Upper inlet surface

• Experiment --31 X 41 X 102 • • • • • • 61 X 41 X 102 --31 X 81 X 102

- - 31 X 41 X 203

lf'l 0 C'! 111 - -SID

6

b) Lower inlet surface

lf'l 0 C'l lf'l 0 q 0 0

lf'l s 0

--31 x41 x 102 • • • • • · 61 X 41 X 102 --31 X 81 X 102

- - - 31 X 41 X 203

0 lf'l 0 lf'l 0 lf'l 0 C'l lf'l r- 0 C'l - -- - C'l C'l 0 0 0 0 0 0

SID

c) Inlet lip

0 lf'l C'l 0

Fig. 5.29 Effect of grid refinement on centre­plane static pressure coefficient

8

151

0.4

0.3

0.2

d'"0.1

0.0

-0.1

-0.2

0

0.6

0.5

0.4

uc. 0.3

0.2

0.1

1.0

0.6

0.2

-0.2

uc.-0.6

-1.0

-1.4

-1.8

-2.2 0

8 0

• Experiment -- RNGk-E, TLBNWallFn. • • • • • · RNG k-E, Std. Wall Fn. -- Std. k-E, TLBN Wall. Fn.

2 4 SID

6

a) Upper centreline

• Experiment -- RNG k-E, TLBN Wall Fn. • • • - • · RNG k-E, Std. Wall Fn. -- Std. k-E, TLBN Wall Fn.

SID

8

b) Lower inlet surface

lf'l 0 C'l lf'l 0 q 0 0

-- RNG k-E, TLBN Wall Fn. • • • • • · RNG k-E, Std. Wall Fn. -- Std. k-E, TLBN Wall Fn.

lf'l 0 lf'l 0 lf'l 0 lf'l 0 r- 0 C'l lf'l r- 0 C'l lf'l 0 - - - - C'l C'l C'l 0 0 0 0 0 0 0 0

SID

c) Inlet lip

Fig. 5.30 Effect of turbulence modelling on static pressure coefficient

a) Total pressure contours - Experiment

c) Total pressure contours - CFD

Sm!s -~

"/ /- ..... / / ..... ,,, // "' . . . '

• I I I ; "'. . . . . '' \ \ \ ,,,_ .. _, \ 1/--,•,\

, , I I I I I { (o ~ ', , t 1 I 1

\ \ \ \ ' --·· ... \\ ,'' ....... •', •I

' .. . .

' '

' .. -' ... -... ...

b) Secondary flow vectors - Experiment

d) Secondary flow vectors - CFD

Fig. 5.31 Comparison of predicted and measured flow at duct exit plane

Neglecting the effect of gross-swirl on the computed total pressure distribution in the

waterjet inlet, it can be seen that the computed distribution of Cp is in reasonably good

agreement with the measured flow behaviour, especially on the plane of flow symmetry.

Fig. 5.32 provides a comparison between the measured and computed Cp on the plane of

flow symmetry. In Fig. 5.32, Cp is plotted against the distance along the plane of flow

symmetry (y) non-dimensionalised by the duct diameter, with y increasing from the

lower duct surface to the upper duct surface. It can be seen that the computed Cp is in

excellent agreement for flow in the core (y/Ds;0.44) and the boundary layer near the

152

lower duct surface. Good agreement is also evident for the inner part of the boundary

layer near the upper duct surface (y/D;:::0.88). Over the outer part of the upper boundary

layer (0.44:s;y/D::;0.88) the computed Cp is over-predicted by as much as 11% (relative

error) of the corresponding measurements. This indicates insufficient loss of total

pressure in the upper duct.

The computed results appear to be insensitive to changes in the grid size in each

computational coordinate direction, as can be seen from Fig. 5.32a. This is in line with

the discussion presented above regarding the effect of grid size on the solution

behaviour for the centreplane Cp. Computed Cp appears to be insensitive to the wall

function used, but inferior results are obtained when the Standard k-E turbulence model

is used, as can be seen from Fig. 5.32b.

1.2

1.0

0.8

u 0.6 • Experiment ~---31 X 41 X 102

0.4 ·····61x41x102 ~--- 31 X 81 X 102

0.2 - - - - - 31 X 41 X 203 12% Error Bars Shown

0 N ~ ~ ~ ~ ~ 00 ~ ~ 0 0 0 0 0 0 0 0 0 0

y/D

a) Effect of grid size

1.2

1.0

0.8

u 0.6 • Experiment

0.4 -+--- RNG k-E, TLBN Wall Fn.

0 - N ~ ~ ~ ~ ~ 00 ~ 0 0 0 0 0 0 0 0 0 0 0 ~

y/D

b) Effect of turbulence and near-wall modelling

Fig. 5.32 Measured and computed total pressure on line of flow symmetry at duct exit

An interesting feature of the measured total pressure distribution in Fig. 5.31 is the

extent of the distortion of the contours of Cp toward the plane of flow symmetry in the

upper duct. The distribution of measured total pressure is clearly qualitatively similar to

that in an S-Duct. The "tucking-in" of Cp contours toward the plane of flow symmetry is

less pronounced for the computations. This discrepancy can be explained by considering

the secondary flow behaviour present in the duct. Fig. 5.31b and Fig. 5.31d show the

measured and computed secondary flow vectors, respectively. In the computational

results two distinct streamwise vortices can be seen, one in the upper part of the duct

close to the vertical centreplane and the other close to the wall just below the line of

153

horizontal symmetry of the duct cross-section. The upper vortex convects fluid near the

wall toward the plane of flow symmetry and is responsible for the distortion of the

contours of Cp toward this plane. The lower vortex convects fluid near the wall toward

the bottom of the duct, thus thickening the boundary layer there and displacing and

distorting the core flow. This double vortex system is evident in the experimental data

but is not as clear due to the gross swirl inherent in the flow.

It is evident from the experimental data, that the discrepancy between the measured and

computed distribution of Cp over the cross-section of the duct is caused by an under­

prediction of the strength of both vortices, but in particular, the upper vortex. This may

be attributed to the under-prediction of upstream boundary layer thickness and vorticity

due to the failure of the computation to capture the full effects associated with flow

separation within the duct.

The accurate calculation of the dimensions and shape of the inlet streamtube is of

particular importance to the accurate determination of ingested momentum and energy

fluxes for use in parametric performance models such as the one outlined in Chapter 2.

Fig. 5.33 shows a comparison between the computed and measured dimensions of the

cross-section of the inlet streamtube upstream of the waterjet inlet corresponding to an

IVR of 0.71. It can be seen from the figure that the shape of the inlet streamtube cross­

section is in good qualitative agreement with the measurement. Quantitative results are

generally within 10% relative error for streamtube width for a given vertical position.

The streamtube depth is within 5% of the measured value on the plane of flow

symmetry. The dimensions of the cross-section of the inlet streamtube are primarily

determined by the width of the inlet opening, the IVR and the upstream boundary layer

velocity profile. The discrepancies between the computations and the measurements

may be attributed to inaccuracies in the modelling of the boundary layer velocity profile

(see Fig. 5.28). Errors of the order of 10% in the boundary layer velocity profile

translate to errors of the order of 10% for the width of the streamtube. This can be seen

from a comparison of Fig. 5.33 and Fig. 5.28.

Having examined the results for a single IVR, it is interesting to examine the flow

154

behaviour over a range of IVR values for two main reasons. The first is to assess the

accuracy of CFD calculation at other IVR values, especially when little or no flow

separation is present. The second is to gain an understanding of the change in flow

behaviour with IVR. Therefore simulations were carried out for IVR values of 0.68,

0.80 and 0.97, since experimental data was available for these values. For the range of

IVR values considered, the RNG k-e turbulence model is used with the TLBN wall

function, as it was shown previously that the RNG k-e turbulence model produces

results superior to those of the Standard k-e turbulence model.

0 20 40 60 80 100 120 140 160

0

-10 • Experiment --CFD

,-., -20 I 0 % Horizontal Error Bars e e -30 5 % Vertical Error Bars Shown '-' ;:>,

-40

-50

-60 z(mm)

Fig. 5.33 Computed and measured inlet streamtube cross-section at IVR=0.71

Fig. 5.34 shows comparisons between the measured and calculated static pressure

distribution over the centreplane of the waterjet inlet with variation of IVR. It is

apparent from Fig. 5.34a and Fig. 5.34b that as IVR is increased, there is a general

decrease in static pressure within the inlet. This is of no real surprise, as higher IVR

results in higher velocities within the duct and hence decreased flow diffusion. This

results in lower average static pressures within the waterjet inlet. Agreement between

the computational results and the measurements improves with increasing IVR. This is

primarily a result of the diminishing extent of flow separation within the duct as a

consequence of the reduced diffusion of the flow. Therefore, with increasing IVR, the

limitations of the turbulence model in handling flow separation become less apparent,

hence predictive accuracy "improves".

Inaccuracies in the static pressure distribution on the upper duct centreplane at

IVR=0.80 and IVR=0.97 are the likely result of a difference in bend geometry between

155

c. u

0.4

0.3

0.2

0.1

0.0

-0.1

-0.2

0

U,.

1 2 3

0.6

0.4

0.2

0.0

-0.2

-0.4 \

-0.6

,, .............. , ,·........ _/ ..•. ,'. ·---- . -· .

4

SID

5

a) Upper inlet surface

.. _-'

6

-0.8 +'--'--'---"'-+-'--'-'--4-''-L-l.-'+-'-'-'---'---+-'-'--'-'-i-'-'--'-'---i-'--J-"-'-+'-J.....LJ.....j

•• •• •

7 8

IVR

• Expt. - 0.61 • Expt. - 0.68 .. Expt. - 0.80 • Expt. - 0.97

--CFD-0.61 - - - - - · CFD - 0.68 --CFD-0.80

CFD-0.97

IVR

• Expt. - 0.61 • Expt. - 0.68 .. Expt. - 0.80 • Expt. - 0.97

--CFD-0.61 - - - • • · CFD - 0.68 --CFD-0.80

- - - CFD - 0.97

0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00 2.25

SID

1.0

0.6

0.2

-0.2

b) Lower inlet surface

. -... --- .... -.. - .. IVR

--CFD-0.61 - - - - - · CFD - 0.68 --CFD-0.80

CFD-0.97

U,. -0.6

-1.0

-1.4

-1.8

-2.2 +'-l::LJ...j-LW-'-j-1-L..I....L+'-L..L..L..j-__LLl_-J...I-l...L.W.+'-L..L..L..j-.J....L..L..Lj-L.J....L..L+'-L..LL.l

§ 0 :g ~ 8

0 -

lf"l 0 lf"l 0 lf"l 0 N lf"l 1:'- 0 N lf"l

N N N 0 0 0 0 0 0 0 0 0 0

SID c) Inlet lip

Fig. 5.34 Comparison of measured and computed static pressure coefficient on waterjet centreplane with IVR

156

the computational model and the corresponding experimental model. The computational

model assumes a smooth continuous bend, whereas the experimental geometry has a

"lobster-back" bend. Roberts (1998) notes that the effect on the wall static pressure

distribution of the sharp changes in curvature associated with the lobster-back bend

becomes significant as IVR is increased and boundary layers in the duct become thinner.

For the range of IVR considered, the static pressure distribution over the initial part of

the inlet ramp (S/DS1.5) is essentially independent of the IVR. From Fig. 5.34 it can be

seen that the minimum static pressure on the ramp is greater than the minimum static

pressure on the lip for the geometry studied here. It may therefore be expected that

ramps having smaller radii of curvature will result in lower minimum static pressures

and a greater likelihood of cavitation inception. Thus the benefits of having ramps with

large radii of curvature are obvious.

The calculated and measured distribution of static pressure coefficient over the lower

duct surface for the range of IVR considered are shown in Fig. 5.34b. It can be seen that

the computed results are in close agreement with the experimental data. Fig. 5.34c

shows the variation of computed Cp with IVR in the vicinity of the inlet lip and reveals

several interesting trends with increasing IVR:

1) The magnitude of SID for the position of the stagnation point of the dividing

streamline on the plane of flow symmetry decreases. In other words, the stagnation

point moves from the upperside of the lip down to the underside of the lip.

2) The minimum static pressure on the underside of the inlet lip increases.

3) The minimum static pressure on the upperside of the lip decreases.

As discussed previously, the minimum static pressure in the vicinity of the inlet lip has

profound implications for cavitation inception at high-speed when dynamic pressures

are large.

The flow through the waterjet inlet can be examined by plotting streaklines on the plane

of flow symmetry, of the simulated flow domain as shown in Fig. 5.35. The divergence

or convergence of the streaklines represents flow diffusion or acceleration, respectively.

The shape of the dividing streamline and the stagnation point on the inlet lip can be

157

a) IVR=0.61

b) IVR=0.68

c)IVR=0.80

d)IVR=0.97

Fig. 5.35 Computed streaklines on centreplane

158

inferred from the relevant streaklines. While diffusing flow on the plane of flow

symmetry is evident in all cases, the extent of flow diffusion increases with decreasing

IVR. In addition, the upstream depth of the inlet streamtube decreases. From the

computed results, it is the author's observation that as IVR decreases a more rapid

diffusion of flow occurs further aft toward the inlet lip.

It must be noted that the converging streaklines near the duct centreline in the horizontal

duct section occur due to spurious numerical behaviour associated with the axis

boundary and do not reflect physical reality. This spurious behaviour will become

evident with the presentation of further results.

Velocity vectors showing the magnitude and direction of the local velocity on the plane

of flow symmetry are shown in Fig. 5.36 for the range of IVR values examined. The

region of low velocity associated with separation/near-separation can be clearly seen in

the upper duct for IVR=0.61 and to a lesser extent for IVR=0.68. As IVR increases, the

flow becomes more uniform. The spurious numerical behaviour associated with the axis

boundary becomes more evident with increasing IVR.

Fig. 5.37 and Fig. 5.38 show the computed distribution of Cp and the secondary flow

behaviour over the duct cross-section, respectively, for the range of IVR values

considered. It must be noted again, that the spurious numerical behaviour associated

with the axis boundary at higher IVR is evident. This does not affect the results to any

real extent, since the effects are confined to an area of only 1-1.5% of the duct cross­

sectional area. Several interesting flow features are apparent from an examination of

both figures. As IVR increases, the size of the core flow (taken here to be C~0.95)

increases. The boundary layer on the top and side of the duct becomes thinner, whereas

the boundary layer on the bottom of the duct thickens. As IVR increases, the streamwise

vortex in the upper inlet close to the plane of flow symmetry decreases in intensity and

disappears. The lower vortex situated on the side of the duct near the horizontal line of

symmetry of the cross-section, intensifies and increases in extent as can be seen from

the enlargement of the secondary flow vectors in Fig. 5.38. This intensified vortex helps

convect low-momentum fluid away from the side of the duct and so thins the boundary

layer there.

159

- =228m/s

------a)NR=0.61

- =228m/s

--b) IVR=0.68

- =22.8m/s

--

c)NR=0.80

- =22.8m/s

---

d)NR=0.97

Fig. 5.36 Computed velocity vectors on centreplane

160

a) IVR=0.61 b) IVR=0.68

c) IVR=0.80 d)IVR=0.97

Fig. 5.37 Variation of total pressure at duct exit with IVR

From the discussion of the results presented in Fig. 5.32, it is likely that although the

centreplane distribution of Cp may be reasonably well predicted over the range of IVR

examined, the strength of the two streamwise vortices may be under-predicted,

especially the upper vortex. If this is the case, this will result in an insufficient "tucking­

in" of the contours toward the plane of flow symmetry and insufficient distortion of the

contours on the side of the duct.

161

- -- -'-.::- ....... " - ..... ' I .; -- ......

' ' : - ::-::::-'

-=1mls

' - ' ' ' '' I\ \\

I 1 ,, '' I I I ,,

' ' ' ' ' I I // ' ' ' I I I ,,,,

,, ' I I I I 111,,',,", /I ;I II''

,, I I I I I ',,,',,,'~~ ~~ '; J I I/

, I I I /_ I I I/~-,

/ ll!'/f { { :-:, 11 11 1 I 1111, I

IIIII II ,, , ,, I I I \ - I I I 11 I I I 1 I \ '!

1111 r' Ill I\\'·-1 I I I I I I I I \ \' I I I I II I II I I I I I\ -, I t I t I ' \ \1 ',.... I I

11

I 1

1 ' 1 1 t

1

1.,.

1 I I 1 I I I 1 1 I -1 I I I "' ' 0 I . ' .

a) NR=0.61

c)NR=0.80

b)NR=0.68

d)NR=0.97

Fig. 5.38 Variation of secondary flow at duct exit with IVR

The variation of "k!Urer and Rr./D on the plane of flow symmetry is shown in Fig. 5.39

for the range of IVR considered. In the absence of detailed experimental measurement it

is not possible to assess the accuracy of these results. In view of the complexity of the

flow and the limitations of the turbulence modelling used, questions are raised as to the

accuracy of the results obtained for these quantities. Griffith-Jones (1994) in his

experimental investigation of flow in a flush-type waterjet inlet, indicated turbulence

intensities of up to 24% in the upper duct at the duct exit. This value is greatly in excess

of turbulence levels presented herein for the equivalent position. Despite the likely

162

inaccuracies inherent in Fig. 5.39, much can still be learned about the generic turbulence

behaviour and the trends in turbulence behaviour with IVR. Unfortunately, spurious

numerical behaviour in the vicinity of the axis boundary and near the outlet boundary

act to distort the results in the vicinity of these locations and therefore care must be

taken in interpreting these results. The following conclusions can be reached regarding

the turbulence behaviour in the waterjet inlet:

1) Large turbulence levels occur in the vicinity of the inlet lip due to the large viscous

shear forces that exist there, resulting in large production of k and hence large values

of turbulence intensity. The peak levels of turbulence occur on the underside of the

lip for IVR values of 0.61, 0.68 and 0.80. For IVR=0.97, peak turbulence levels occur

on both the underside and upperside of the inlet lip. Large velocity gradients exist as

a result of the acceleration of the flow away from the stagnation point, resulting in a

highly sheared boundary layer whose turbulence intensity is increased significantly

on encountering an adverse pressure gradient after the static pressure minima on the

lip surface (Fig. 5.34).

2) Downstream of the inlet lip in the external flow, the boundary layer thickens more

rapidly with decreasing IVR due to greater adverse pressure gradient. This is evident

from the increasing area covered by the contours of turbulent velocity-scales.

Turbulent length-scales remain small as the eddies in the boundary layer are small

and dissipative.

3) As flow diffuses into the inlet from the upstream boundary layer, the turbulence

spreads as a result of flow diffusion and turbulent diffusion. Turbulence is intensified

as a result of the adverse pressure gradients encountered in the upper part of the

waterjet inlet resulting from flow diffusion and the bend. Overall turbulence levels

inside the waterjet inlet decrease with increasing IVR. This is due to the presence of

more favourable pressure gradients resulting in a reduction of length-scales and

turbulence intensity. The large velocity and length-scales present in the upper part of

the inlet imply large highly turbulent energy-containing eddies creating large eddy­

viscosity and therefore large losses in total pressure.

4) The flow in the lower-half of the duct above the wall boundary layer features low

levels of turbulence and small turbulent length-scales, this region may be considered

to be the "potential core".

163

.Jklu,., B 0100 A 0090 9 0080 8 0070 7 0080 6 0050 5 0040 4 0030 3 0020 2 0010

0000

.Jk!U,.,

I! 0240

ljD 0220

c:::: A 00328 0200

9 00295 0180

8 00262 0160

7 00229 0140

6 00197 0120

5 00164 0100

4 00131 (Expanded V1ew of L1p) 3 00098 2 00066

00033

a) NR=0.61- Standard k-E

.JkiU .. , B 0100 A 0090 9 0080 8 0070 7 0.080 6 0050 5 0040 4 0030 3 0020 2 0010

0000

.Jk!U,.,

I,!D

li 0220 c 0200

A 00318 0180

9 00286 0160 8 00254

0140 7 00222

0120 6 00191

0100 5 00159 4 00127 (Expanded V1ew of L1p)

00095 2 00064

00032

b) IVR=0.61 - RNG k-E

Fig. 5.39 Computed turbulent velocity and length-scale on centreplane

164

..Jk!U,., B 0100 A 0090 9 0080 8 0070 7 0080 6 0050 5 0040 4 0030 3 0020 2 0010

0000

1/1)

~~ 0.220 c: 0.200

A 00322 0180 9 00290 0180 8 00257

0140 7 0.0225 0120 6 00193

0100 5 00161 4 00129 (Expanded V1ew of L1p) 3 00097 2 00064

00032

c) IVR=0.68

..Jk!U,., B 0100 A 0090 9 0080 8 0070 7 0060 6 0050 5 0040 4 0030 3 0020 2 0010

0000

..Jk!U,.,

1/D [ 0.220 c: 0.200

A 00305 0180 9 00275 0160 8 00244

0140 7 00214

0120 6 00183

0100 00153

4 00122 (Expanded View of Up) 3 00092 2 00061

00031

=-=

d) IVR=0.80

Fig. 5.39 (cont.)

165

vklu,., B 0100 A 0090 9 0090 8 0070 7 0080 6 0050 5 0040 4 0.030 3 0020 2 0010

0000

1/D [ 0.220 c 0.200

A 00226 0180 9 00203

0160 8 00181

0140 7 00158

0120 6 00135

0100 5 00113 4 00090 (Expanded Vrew of Lrp) 3 00068 2 00045

00023

e) IVR=0.97

Fig. 5.39 (cont.)

It is interesting to examine the differences in turbulent velocity and length-scales

predicted by the Standard k-E turbulence model and the RNG k-E turbulence model as

can be seen from a comparison of Fig. 5.39a with Fig. 5.39b. The Standard k-E predicts

larger turbulent velocity and length-scales throughout the flow domain, thus predicting

larger eddies of greater turbulence intensity and hence greater eddy-viscosity within the

flow domain. These observations are consistent with the results obtained for the

previous two validation cases. Of particular interest, is the difference in the magnitude

of the turbulent velocity-scales predicted by both models in the vicinity of the inlet lip.

The Standard k-E model predicts a larger region of high-intensity turbulence, thus

indicating less sensitivity to the effects of streamline curvature on the turbulence.

The calculated mass-averaged and area-averaged total pressure coefficients over the

cross-section of the duct exit, are shown in Fig. 5.40, for the four IVR values

considered. Mass-averaged total pressure coefficient is defined as

166

_ 2 11t1D/2 U p Cp = --

2 -- 1 2

rdrd8 1tD o o Urer 2PUrer

The area-averaged total pressure coefficient is defined as

Cp = ~ rlt rD/2 I p 2 rdrde 1tD Jo Jo 2 pU ref

(5.9)

(5.10)

It can be seen that as IVR increases, the average total pressure over the cross-section

increases, thus indicating a reduction in the total pressure lost in the duct. This is

expected, since the losses associated with flow separation disappear at higher IVR

values. The values of mass-averaged Cp are higher than the corresponding area­

averaged value because the regions of higher Cp (and higher streamwise velocity) carry

more of the mass flow and hence a more energetic flow. This of course results in a

higher Cp over the cross-section.

Mass-averaged Cp is of particular interest for use in waterjet parametric performance

models (see Chapter 2) as it in essence describes an energy recovery efficiency, whereas

area-averaged Cp may be preferred for design purposes as it implicitly contains a

measure of the non-uniformity of the total pressure distribution at the duct exit. It is

interesting to note that there is relatively little variation in mass-averaged CP , while the

change in area-averaged Cp is more significant.

0.86

0.84

u 0.82

~ 0.80

-+---Mass -Averaged 0.78 -A- Area -Averaged

0.76 0 10 0 10 0 10 0 10 0

"' "' 1'- 1'- 00 00 0\ 0\ C! 0 0 0 0 0 0 0 0

IVR

Fig. 5.40 Change in average total pressure coefficient over duct exit with IVR

167

5.5 Discussion of Results

In this section the accuracy of the computational results presented in the preceding

sections for the three validation studies undertaken, will be discussed in relation to the

capabilities and limitations of the computational modelling used.

Since the same turbulence and near-waH modelling are used in all three validation cases

and the same general trends in results observed, it was decided that it is more

appropriate to discuss the accuracy of the computational results in a unified manner and

refer to each validation case as appropriate, for reasons of compactness and to avoid

repetition of the discussion.

Discussion shaH be focused on three main issues associated with computational

simulation of the experimental configurations examined. These are the boundary

conditions used, the size and quality of the numerical grid and the turbulence and near­

wall modelling adopted. It is principally this latter issue of turbulence and near-wall

modelling which will be shown to have the greatest impact on the accuracy of the

solutions obtained, as a consequence of the limitations of the models used.

5.5.1 Boundary Conditions

Inaccuracies in representing the velocity upstream of the duct in the case of the 90° bend

of Enayet et al (1982) and the waterjet inlet of Roberts (1998) will affect the

development of streamwise vorticity in the duct and hence the downstream flow

behaviour. This can be qualitatively deduced from the Squire-Winter formula (Eqn 5.1).

Therefore in these cases, inaccurate boundary condition representation is clearly a

source of error which will be reflected in the results obtained, although the magnitude of

these errors is difficult to quantify in view of the other sources of error present in the

computational modelling procedure.

In the modelling of the 90° bend, it is likely that the under-prediction of the core

velocity by all turbulence and near-wall modelling options examined (as can be seen

from Fig. 5.5, Fig. 5.6 and Fig. 5.7) may be directly attributed to inaccuracies in the

velocity profile upstream of the bend. It is difficult to quantify the error associated with

168

incorrectly modelled boundary conditions in the case of flow in the waterjet inlet, but

inaccuracies in the dimensions of the cross-section of the inlet streamtube are reflective

of this. There was insufficient data available in Bansod and Bradshaw (1972) to allow

the accuracy of the boundary conditions adopted in the study presented in Section 5.3 to

be assessed.

In the case of the waterjet inlet of Roberts (1998), inaccurate representation of the

geometry is another source of error. An obvious example of this is the representation of

the "lobster-back" bend with a smooth bend. This is found to lead to inaccuracies in the

static pressure distribution over the duct bend at larger IVR, when flow separation is

absent.

5.5.2 Grid Size and Quality

The solutions obtained on the 4lx4lx81 grid for the 9~0 bend and the 31x3lx71 grid

for the S-Duct are deemed to be "grid-independent" and so errors associated with the

discretisation of the governing equations should have a negligible effect on the solution.

Both grids possess excellent orthogonality and the gridlines in the streamwise (K)

coordinate direction are aligned with the primary flow, therefore minimising cross­

stream diffusion. In the case of the 90° bend, gridlines are spaced evenly in the radial (J)

direction. This was done in order to place the cell-centre of the wall-adjacent cells in the

range of y + =30 at entry to the duct and so allow optimum use of wall functions to be

made.

Larger variations in the spacing of radial gridlines are necessary for the S-Duct in order

to resolve the greater range of turbulent length-scales present in the flow. Since the

solution obtained for the S-Duct may be considered to be "grid-independent", errors

associated with the variation in radial gridline spacing and hence variation in cell size,

may be considered to be negligible. Care was also taken to ensure that the cell-centres of

the wall-adjacent cells were placed in the range of y+=30-40 so that appropriate use of

wall functions could be made. It may thus be concluded that the meshes for the 90° bend

and the S-Duct are of a high quality and are in essence a negligible source of solution

error.

169

The situation is however different for the 31 x41 x 102 grid used for simulation of the

flow domain associated with the waterjet inlet. Section 5.4.4 shows that although the

refinement of the grid in each computation coordinate direction does not result in a

change in solution, it cannot be claimed that the solutions obtained on this grid are

"grid-independent". Therefore errors arising from discretisation and associated with

growth in cell size may be present in the solution. Other sources of errors arising from

cross-stream diffusion and cell non-orthogonality will be present, particularly in areas of

extreme cell skewness. The magnitude of these errors is difficult to quantify but must be

considered when examining the solution. The quality of the grid produced by Inlet3D

was discussed in greater detail in Section 4.5.

5.5.3 Turbulence Modelling

In all of the three validation cases considered, the use of the RNG k-e turbulence model

produced results in better accord with the experimental data than those obtained using

the Standard k-e turbulence model, although the limitations of the RNG k-e model are

also apparent. The flows considered in this chapter contain the following features:

1) Adverse pressure gradient in the stream wise direction

2) Streamline curvature

3) Flow separation/near-separation.

It is well known from the literature that the Standard k-e model performs poorly under

adverse pressure gradient. Wilcox ( 1993) used perturbation analysis to show that the

length-scale predicted by the Standard k-e model under adverse pressure gradient tends

to be too large in the near-wall region. Rodi and Scheuerer (1986) showed that under

adverse pressure gradient, the coefficient of the rate of production of e, Ce1 in the e

equation is under-predicted and must therefore be increased in value if the predicted

length-scale is to be made independent of pressure gradient. As a consequence, the

production of e is under-predicted and the length-scale determined by the e equation

rises more steeply near the wall than in the case of zero pressure gradient. Eddy­

viscosity is therefore over-predicted and the tendency of the flow to separate suppressed.

Flows with streamline curvature contain what Bradshaw (1973) termed "extra rates of

170

strain" which modify the structure of the turbulence to a much greater extent than

suggested by the extra production terms that appear in the Reynold stress equations.

Bradshaw (1973) noted that "extra rates of strain", unrelated to mean strain rates, cause

changes in the higher-order structural parameters of the turbulence, which subsequently

cause changes in the Reynolds stresses and so modify the structure of the turbulence.

Therefore in "complex" turbulent flows, such as those studied here which contain "extra

rates of strain" due to curvature and flow separation, the Boussinesq hypothesis must

necessarily fail, as Wilcox (1993) noted.

Interesting phenomena associated with streamline curvature include the "collapse of

turbulence" for flow over a convex surface and the augmentation of turbulence for flow

over a concave surface, accompanied by the formation of streamwise longitudinal

vortices in the boundary layer (see Bradshaw (1973)). Gillis and Johnston (1983)

building on previous efforts, investigated two-dimensional flow affected by strong

convex curvature and demonstrated the "collapse of turbulence" phenomenon as well as

a substantial reduction in skin friction coefficient (relative to the corresponding flat plate

boundary layer value) on the convex surface.

The over-prediction of the skin friction coefficient on the first bend, at the bottom of the

S-Duct, by all turbulence and near-wall modelling combinations (Fig. 5.19b) may be

directly attributed to a failure of the turbulence modelling to properly account for the

effects of streamline curvature on the development of the turbulent boundary layer over

this convex surface. This, therefore, results in a failure to predict the general decrease in

Cr relative to the corresponding flat plate value.

The flows considered in the validation study presented here contain a combination of

streamline curvature, adverse pressure gradient in the streamwise direction and/or flow

separation. Therefore the k-E turbulence modelling has to account for all of these effects

simultaneously, rather than individually. Comparison between the centreplane

distributions of turbulent velocity and length-scales obtained using the Standard k-E and

RNG k-E turbulence models for the three validation cases, clearly show that the

distributions obtained using the RNG k-E turbulence model exhibit a greater sensitivity

171

to the effect of streamline curvature. This results in reduced turbulent velocity and

length-scales as would be expected, considering the effects of "extra rates of strain" on

turbulent flow behaviour.

The superior performance of the RNG k-E turbulence model can be attributed primarily

to the rate of strain term R (Eqn 3.32), in theE equation which effectively modifies the

constant C2£ in that equation such that

C* = C +ell 113(1- 11 I 'Tio) £2 £2 1+~113

(5.11)

where 11 =SkI E and S = ~s.Js•J . Thus c:2 becomes a function of strain-rate, thus

making the destruction of E term, c:2 £2 /k, a function of strain rate. As a result, strain

rate acts to augment the destruction of E, thus augmenting E and reducing k. The effect is

therefore a reduction in both the turbulent velocity and length-scale. This explains the

apparent sensitivity of the RNG k-E model to streamline curvature, with a general

reduction in turbulent velocity and length-scales relative to those calculated using the

Standard k-E turbulence model.

5.5.4 Near-Wall Modelling

It can be seen from the figures presented in previous sections (which compare the two

wall functions for a given turbulence model with the experimental data) that the use of

either wall function leads to results that are either identical or equivalent for the flows

considered. For example, in the case of the 90° bend, both wall functions give

essentially identical results for the streamwise velocity profile on the duct centreplane

one diameter downstream of the bend, as shown in Fig. 5.5b. For the streamwise

velocity distribution at the 8=75° bend in the plane (Fig. 5.5a) the use of the TLBN wall

function yields more accurate results for y/D<0.25, but is less accurate for

0.25<y/D<0.4. Therefore the results may be considered as being equivalent.

For the distribution of total pressure coefficient (Cp) over the plane of flow symmetry at

the duct exit of the waterjet inlet (Fig. 5.32b) the results obtained using the two wall

functions are identical. The difference between the results obtained using the two wall

172

functions is more marked in the case of the S-Duct. As can be seen from Fig. 5.16, both

models give equivalent results in relation to the experimental data.

So what can be concluded regarding the different wall functions? Clearly the use of the

TLBN wall function appears to offer no advantage over the use of the Standard wall

function as far as calculated mean velocity and pressure profiles (for the validation cases

considered) are concerned. The concept of turbulent equilibrium (Gk=E) and the law of

the wall in the inner region of the boundary layer (outside of the buffer layer) is a

reasonable approximation for most of the developing boundary layer in the validation

cases considered, but this assumption is known to break down in regions of flow

separation and strong adverse pressure gradient.

The results of Driver and Seegmiller (1985) for the separated flow behind a backward­

facing step, show that the concept of turbulent equilibrium is inaccurate for separated

flow. Driver (1991) used laser-Doppler velocimetry to examine two flows experiencing

adverse pressure gradient, with one flow remaining attached and the other separated. He

found that the law of the wall breaks down for p+>0.05 (where p+=(Jl/u~) dp/ds) and

that the assumption of turbulent equilibrium in the near-wall region is inaccurate under

strong adverse pressure gradient. It is therefore clear that the assumptions inherent in the

near-wall modelling are a further source of error in the case of strong adverse pressure

gradient and flow separation.

One would expect that under conditions of zero streamwise pressure gradient (dp/ds=O),

the two wall functions would yield identical results. Interestingly this is not the case.

Under zero pressure gradient, the expressions for the law of the wall obtained from the

Standard wall function (Eqn 3.50) and the TLBN wall function (Eqn 3.53) are the same,

but the expressions obtained for Gk and E are different.

Considering Gk , subtraction of Eqn 3.57 (TLBN function) from Eqn 3.51 (Standard

wall function) gives

(5.12)

173

where y: =11.225. If y+=41.5 then both wall functions calculate the same Gk. If y+<41.5

the TLBN function under-predicts Gk relative to the Standard wall function. For

y+>41.5, the opposite is true.

Considering E, subtraction of Eqn 3.58 (TLBN wall function) from Eqn 3.52 (Standard

wall function) gives

(5.13)

Eqn 5.13 is a function of both kp and y; and the behaviour of each wall function with

respect to the other is therefore dependent upon y+ for a given value of kp.

It is clear from Eqn 5.12 and Eqn 5.13 that both wall functions do not give the same

turbulent production and dissipation under zero pressure gradient. It can therefore be

concluded that the use of the different wall functions will yield different results for the

development of turbulence in the flow, even under zero pressure gradient! Furthermore,

the difference in results obtained from the validation cases examined, when using the

two wall functions with the same turbulence model, may be attributed to differences in

turbulence prediction in the wall-adjacent cells.

5.6 Closure

Three experimental validation studies were conducted in order to examine the accuracy

of the Standard k-E and the RNG k-E turbulence models implemented in Fluent. The

Standard wall function and a Two-Layer-Based Nonequilibrium (TLBN) wall function

were used with both models for the purposes of near-wall flow modelling. Experimental

data was obtained for the flow in a 90° bend of Enayet et al (1982), S-Duct

configuration 3 of Bansod and Bradshaw ( 1972) and the model flush-type waterjet inlet

of Roberts et al (1998). The last validation study is of particular relevance in assessing

the likely accuracy of solutions presented in the remainder of this thesis (for flows in

waterjet inlets). The first two validation cases help provide a rigorous analysis of the

capabilities of CFD in modelling the flow physics associated with duct flow.

174

The following conclusions may be drawn as to the modelling approaches examined:

1) In all of the cases examined, the RNG k-E turbulence model provides solutions of

superior accuracy to the Standard k-E turbulence model. The deficiencies of the

Standard k-E turbulence model are well known from the literature. The rate of strain

term (R) in the E equation of the RNG k-E turbulence model provides a better

modelling of flows containing "extra rates of strain" such as streamline curvature,

thus leading to more accurate results.

2) For the validation cases considered, both the Standard wall function and the TLBN

wall function lead to either identical results or results of comparable accuracy for

mean flow quantities such as velocity and pressure.

3) The failure of both models to adequately handle flow separation is due to an

inaccurate prediction of the onset of flow separation and the "extra rates of strain"

phenomenon, as well as a break-down of the assumptions underlying the use of wall

functions in separated flow. This leads to inaccuracies in the computed flow and is

particularly evident in the computational results obtained for the S-Duct and waterjet

inlet.

4) Additional sources of error include the modelling of the boundary conditions for the

goo bend and the waterjet inlet. An additional source of error is the quality of the grid

used for simulation of the waterjet inlet. It is also possible that discretisation error

may be another source of error in the results obtained for the waterjet inlet on the

31x4lxl02 grid. The 41x41x81 grid used for simulation of the experimental

geometry associated with the goo bend and the 31x31x71 grid used for simulation of

the S-Duct provide reasonably "grid-independent" solutions.

5) The computed turbulence behaviour was discussed for the purposes of comparing the

two turbulence models examined. While it is thought that the results reflect the

qualitative flow behaviour, the accuracy of the computed velocity and length-scale

profiles obtained could not be validated due to a lack of experimental data.

Overall, reasonably good agreement between computed velocity and pressure

distributions and the experimental data was obtained for the three cases considered

when the RNG k-E model is used. The agreement between computed results and the

experimental data improves when flow separation is absent from the flow.

175

It can thus be concluded that CFD can be used as an effective analysis tool for

simulating the flow in flush-type waterjet inlets, producing credible results of

reasonably good accuracy for pressure and velocity distributions within the waterjet

inlet. Care must be exercised in the interpretation of results when it appears that flow

separation may be present.

Finally, due to its superior accuracy to the Standard k-E model, the RNG k-E turbulence

model is chosen for use in the simulation of waterjet inlet flow on a 31 x41 x 102 grid in

combination with the TLBN wall function, for the work presented in the remainder of

this thesis.

176

Chapter 6 Boundary Layer Investigations

Two main issues will be addressed in this chapter. The first is the issue of selecting

suitable parameters necessary for quantifying various aspects of the hydrodynamic

performance of the waterjet inlet. This provides the basis for quantifying the effect of

geometric design variations on the hydrodynamic performance of the waterjet inlet and

so provides a necessary foundation for the work presented in this chapter, the parametric

design space investigations in Chapter 7 and the waterjet inlet optimisation example in

Chapter 8.

The second issue involves the effect of the thickness of the upstream boundary layer on

the flow within the waterjet inlet. It will be shown in this chapter, that the thickness of

an upstream boundary layer has a significant influence on the flow inside the waterjet

inlet and hence on its hydrodynamic performance. It may thus be expected that the same

waterjet inlet installed in different vessels will exhibit differing hydrodynamic

performance. This is due to the differences in ingested momentum and energy fluxes

resulting from the upstream hull boundary layer, amongst other factors.

In Chapter 2, a parametric model of waterjet thrust and efficiency was developed.

Subsequently, a parametric analysis was presented which examined the effect of

ingested hull boundary layer fluid on waterjet performance. It was concluded that

ingestion of fluid from the hull boundary layer has a beneficial effect on overall waterjet

efficiency. In this chapter, the emphasis is on a detailed analysis of how ingestion of

varying amounts of boundary layer fluid can affect the internal hydrodynamics of the

waterjet inlet and the quality of flow delivered to the pump.

Furthermore, the issue of upstream boundary layer thickness is also of relevance to

waterjet inlet optimisation, as an inlet optimised for a particular thickness of upstream

boundary layer may give inferior performance when the flow conditions upstream of the

177

inlet change. Therefore, the hydrodynamic performance of a waterjet inlet and hence an

optimised waterjet inlet design is a function of the boundary conditions of the flow

domain of interest. In other words, the optimum waterjet inlet shape may change

depending on the flow conditions in the vicinity of the inlet and the mass flow-rate

through the waterjet inlet.

In Section 6.1 parameters used to quantify the various aspects of the hydrodynamic

performance of the waterjet inlet (which shall be termed the "hydrodynamic

performance parameters" in the remainder of this thesis) are presented and the rationale

behind their selection discussed. The waterjet inlet geometry, CFD modelling and CFD

simulation used for the investigations presented in this chapter are outlined in Section

6.2. The CFD modelling methodology outlined in Section 6.2 also fonp.s the basis of the

generic flow modelling used in subsequent chapters. In Section 6.3, results are presented

of an investigation of the effect of the thickness of the upstream boundary layer on the

hydrodynamic performance of the waterjet inlet. The results of this investigation are

discussed in detail in Section 6.4. A summary of the main points of this chapter and the

conclusions of the boundary layer investigation are presented in Section 6.5, together

with their implications for propulsion system design.

6.1 Assessment of Hydrodynamic Performance

In order to determine the hydrodynamic performance of the waterjet inlet, it is necessary

to identify the most important parameters affecting hydrodynamic performance and

quantify them. This systematic approach is a fundamental requirement for optimisation

of the waterjet inlet, as it allows the objective function of the optimisation problem to be

appropriately related to the hydrodynamic performance of thewaterjet inlet design.

Of hydrodynamic importance are:

1) The presence of flow separation in the inlet

2) The presence of cavitation in the waterjet inlet

3) The recovery of total pressure at the duct exit

4) The distortion of the flow field at the duct exit

5) The turbulent velocity and length-scales at the duct exit

178

6) Drag forces acting on the inlet (Section 2.1.2)

7) The volume of water entrained in the waterjet inlet

8) The vertical forces acting on the waterjet inlet

9) The dimensions of the inlet streamtube

These aspects of hydrodynamic performance are described below.

6.1.1 Cavitation

When the static pressure in the waterjet inlet falls below the vapour pressure of water,

water will cavitate (boil) forming a two-phase liquid/gas flow of water and water vapour

(steam) bubbles. These bubbles of water vapour are convected downstream with the

mean flow and implode as the static pressure increases above the vapour pressure of the

water. The implosion of these water vapour bubbles cause localised pressure waves

which can pit and erode metal surfaces. Cavitation is therefore an undesirable

phenomenon that must be avoided, or at least minimised, through careful geometric

design of the waterjet inlet.

Other undesirable effects of cavitation will manifest themselves through reduced

waterjet efficiency via the terms of Eqn 2.38. Cavitation at the inlet lip below the

stagnation line, will cause an increased lip loss thrust fraction ( t L) as a consequence of

the altered pressure field. Cavitation inside the waterjet inlet will reduce the inlet

efficiency (Tlin) and may lead to a reduction in the pump's rotative efficiency (TlRot).

Further complications arise due to the unsteady time-dependent nature of the cavitating

flow.

Areas of inlet cavitation can be identified primarily by an examination of the cavitation

number distribution over the surface of the waterjet inlet. The inception of cavitation

could also occur in the interior of the flow domain, such as in the cores of strong

vortices, located for example in the comers of rectangular duct sections. However, for

all intents and purposes, the primary sources of cavitation are generally likely to be low

pressure regions in the vicinity of the duct surface resulting from high velocity flow over

these surfaces. Hence examination of the pressure distribution over the duct surface

alone, may be deemed to be an acceptable approach to identifying areas of cavitation

179

inception. Cavitation number is defined here as

cr= Patm -pgH+p-pv 1 uz 2P ref

(6.1)

where cr is the cavitation number, Patm the atmospheric pressure, Uref the reference (free­

stream) velocity, p the static pressure calculated by Fluent (which neglects the

contribution to the pressure field arising from hydrostatic pressure), H the height above

the water-level for the point under consideration, g the gravitational constant and Pv the

vapour pressure of water for a given ambient temperature.

An alternative approach for examining the likelihood of waterjet inlet cavitation, is to

use the distribution of the static pressure coefficient over the surface of the waterjet inlet

in order to determine the corresponding pressure distribution. Eqn 6.1 can then be

applied to calculate the cavitation number distribution. The static pressure coefficient is

defined in the usual manner as

C = p-Pref p 1 u2

2P ref (6.2)

where Cp is the static pressure coefficient and Prer the reference static pressure. This

approach allows the results to be extended more easily to other vessel speeds. For the

purpose of inlet optimisation for a given vessel speed, it is only necessary to consider

the cavitation number distribution, rather than the static pressure coefficient distribution.

In this chapter, however, attention will be focused mainly on the use of Cp in order to

generalise the results obtained to other vessel speeds.

Since the governing equations solved by Fluent do not account for the cavitation

phenomenon, the flow behaviour predicted by the governing equations will not be in

accord with reality when cr becomes negative. This may be viewed as either

advantageous or disadvantageous, depending on the perspective adopted. The clear

disadvantage of the flow modelling used, is that the actual extent of cavitation in a flow

cannot be determined. Therefore, the severity and extent of flow disruption caused by

cavitation cannot be quantified. On the other hand, what is actually sought is not the

extent or severity of waterjet inlet cavitation, but rather whether cavitation will occur in

the first place, since it is desirable that the waterjet inlet be free from cavitation. In

180

addition, since the governing equations are solved for a continuous single-phase flow,

the static pressure distribution (unmodified by cavitation) over the waterjet inlet allows

areas of peak negative pressures to be identified and correlated with the underlying

geometry. Indeed, one of the advantages of computational simulation is that physically

impossible situations can be examined.

6.1.2 Inlet Total Pressure Losses

The loss of total pressure in the waterjet inlet provides a measure of its total pressure

recovery efficiency. A non-dimensional total pressure recovery efficiency (11) may be

defined as

(6.3)

where Pref is the average reference total pressure, P0

the average total pressure at the

duct exit and the overbar denotes an average quantity. The average total pressure at the

duct exit may be determined either on a mass-averaged basis as

1 i21t rDo/2 Po = Jc PU rdrdO

AOUO 0 0 (6.4)

or on an area-averaged basis as

(6.5)

where Uo is the volumetrically-averaged velocity flowing through the duct exit, P the

total-pressure at a point on the cross-section and the subscript o denotes the duct exit

location. The integrations of Eqn 6.4 and Eqn 6.5 are formulated in polar coordinates as

this provides for the most convenient evaluation of the double integrals over the cross­

section of the duct exit. Similarly, a mass-averaged or area-averaged total pressure loss

coefficient (~) may be defined as

~=1-11 (6.6)

The reference total pressure is taken to be the total pressure of the free-stream condition.

With a reference static pressure of 0 Pa the reference total pressure is then given by

(6.7)

It must be noted that as a consequence of using this definition of reference total

pressure, the value of ~ will not only be affected by the actual viscous losses within the

181

waterjet inlet itself, but also by the ingestion of boundary layer fluid. The greater the

quantity of boundary layer fluid ingested into the inlet, the greater ~will appear to be.

The alternative approach would be to determine a mass-averaged or area-averaged

reference total pressure by performing an integration of the total pressure over the cross­

section of the inlet streamtube at the ramp tangency point. This reference total pressure

may be evaluated on a mass-averaged basis as

2 JhJW/2 Pref = -- PU dzdy A1U1 o o

(6.8)

or alternatively on an area-averaged basis as

_ 2 JhJW/2 pref = - p dzdy

A1 o o (6.9)

where U 1 is the volumetrically-averaged velocity of the mass flux through the cross­

section of the inlet streamtube and A1 is the total cross-sectional area of the inlet

streamtube (see Fig. 2.2) of maximum depth hand total width W at the ramp tangency

point.

For a given thickness of upstream boundary layer and vessel speed, the total pressure at

the duct exit is of interest since this is the total pressure which affects the net positive

suction head (NPSH) of the flow available to the pump. It is therefore simpler and more

logical to base the reference total pressure on the free-stream condition rather than to

evaluate the reference total pressure by performing an integration of the total pressure

over the cross-section of the inlet streamtube. This latter approach would require a

detailed investigation of the dimensions and shape of the cross-section of the inlet

streamtube, which is clearly impractical for design and optimisation studies, where a

large number of simulations are required.

6.1.3 Flow Distortion at the Duct Exit

Distortion of the velocity and static pressure field upstream of the impeller can lead to

impeller vibration, cavitation (see Aartojarvi (1995)) and possibly fatigue failure of the

impeller shaft due to the cyclic nature of the non-uniform blade loading as the shaft

rotates. All of these effects can be directly attributed to the flow behaviour in the

immediate vicinity of individual blades, as the blades pass through a non-uniform flow

182

field. There are essentially three main causes for a distorted flow field at the pump inlet

in a turbulent flow field. These are:

1) Time-averaged (steady-state) distortiOn of the pressure field and the axial velocity

field resulting directly from the geometry of the waterjet inlet and upstream boundary

conditions.

2) Time-averaged secondary flow velocities (swirl), resulting from the geometry of the

waterjet inlet acting upon the upstream vorticity caused by boundary layer flow (see

Chapter 5).

3) Distortion of the flow field about its time-average due to the level of turbulent

fluctuations present in the flow.

The interaction of the distorted turbulent flow field with the pump will cause a reduction

in the rotative efficiency of the pump (llRot) and hence a reduction in overall waterjet

efficiency (Chapter 2).

Analogues can be made between the above-mentioned problem and:

1) The performance of pumps in a non-uniform flow-field

2) The interaction of ship propellers in a hull wake

3) Aircraft inlet flow and gas turbine compressor interaction

On the subject of the interaction of ship propellers in a hull wake, Tsakonas et al (1967)

and Breslin (1970) expressed the wake upstream of a ship's propeller as a Fourier Series

of axial and tangential velocities in their study of propeller-induced vibratory forces in

ships. Lewis (1963) also examined the subject of propeller-induced vibratory forces in

the hull structure resulting from the operation of a propeller in the non-uniform wake aft

of the hull. It is clear from the literature, that a rotating propeller or impeller operating in

a non-uniform flow field will excite vibratory harmonics, not only in itself, but also in

the surrounding structure. In the case of a waterjet, it may thus be expected that

vibratory harmonics will be induced in the impeller and the waterjet inlet duct. The

amplitude and frequency of these harmonics will depend on many factors and requires a

detailed study, which is beyond the scope of the work presented in this thesis.

There are many references in the aeronautical literature dealing with the subject of non-

183

uniform inlet duct flow upstream of a gas turbine compressor and its relationship to

compressor stall and engine surge. Momentary total pressure distortion levels caused by

the turbulent nature of the inlet flow can be of a magnitude and duration sufficient to

cause compressor stall. Seddon & Goldsmith ( 1985) provided a brief review of the

subject and cited relevant references. Melick et al (1975) also discussed this issue.

Melick and his co-workers modelled the inlet turbulence as a random distribution of

discrete vortices being convected downstream by the mean flow. Tangential velocity

and local pressure gradients associated with each vortex, when superimposed on the

mean flow, produce the measured pressure fluctuations. In order to use the theory of

Melick et al (1975), it is necessary to obtain the root mean square (rms) level and power

spectral density (PSD) function of the inlet total pressure fluctuations from experimental

data. The most probable maximum distortion level under particular operating conditions

is stated as being dependent on:

1) Rms level of total pressure fluctuations

2) PSD function of the total pressure fluctuations

3) Steady-state distortion level

4) Length of time at the operating condition

5) Frequency response characteristics of the engine.

The work of Melick et al (1975) was further extended by Brilliant & Bauer (1977) and

Melick et al (1978). Brilliant & Bauer found the work of Melick et al (1975) to be in

good agreement with experimental results.

Angular swirl in S-shaped inlet ducts is another flow problem that can arise. Again there

has been a significant amount of work on this subject in the aeronautical literature. The

angle of swirl (in the absence of compressor inlet guide vanes) can become sufficiently

large to stall the aircraft gas turbine compressor, resulting in engine surge (see Seddon

and Goldsmith (1985)). Seddon described how swirl has been shown (by wind tunnel

testing) to be a secondary effect of flow separation. Swirl behaviour in S-shaped ducts is

caused by a complex interaction of pressure gradients associated with bends, pressure

gradients associated with incidence, flow separation and the duct geometry, as Guo and

Seddon (1983a) and Guo and Seddon (1983b) demonstrated experimentally.

184

It must be noted that high inlet total pressure recovery is not a guarantee of low flow

distortion at the duct exit and low inlet total pressure recovery does not necessarily

entail a high distortion level. Distortion, therefore, is a feature of inlet flow that must be

evaluated parallel to the mean pressure recovery.

The uniformity of the pressure and velocity fields at the duct exit can be quantified via

the use of distortion coefficients, many definitions of which are presented in the

aeronautical literature. According to Seddon and Goldsmith ( 1985), the usual definition

of distortion coefficient in the United Kingdom is

(6.10)

where P is the mean total pressure at the engine face and P9 the mean total pressure in

the 'worst' sector of angular width e of the engine face. Dc(S) does allow some degree

of circumferential distortion to be assessed, but does not allow assessment of the degree

of radial distortion. An alternate form of distortion coefficient is

(6.11)

where P max , P mm and Po are the maximum, minimum and mean total pressures at the

cross-section under consideration, respectively.

Rather than using the definitions for inlet distortion present in the aeronautical literature,

it is decided to use a statistical approach and define inlet distortion in terms of the

standard deviation of the quantity under consideration, whether this be static pressure,

total pressure, or velocity. This approach suffers from the disadvantage that there is no

distinction made between radial and circumferential distortion. Ideally, however, a

uniform distribution of total pressure is sought at the duct exit, so it can be argued that

both radial and circumferential distortion are equally unacceptable. A clear advantage of

this approach is that it provides (in a single number) a concise statistical representation

of the level of distortion about the mean flow at the duct exit. The standard deviation of

total pressure over the cross-section of the duct exit of the waterjet inlet is therefore

defined as

185

(6.12)

In Eqn 6.12, the square of the deviation between the total pressure measured at point i

on the cross-section (P1) and the average total pressure acting over the cross-section (P0 )

is weighted by the incremental area (A1) surrounding point i. The square of the total

pressure deviations are summed over the N measuring points on the cross-section. For

the purposes of practical CFD evaluation of De, the N measuring points correspond to

grid cell-centre locations, with the total pressure being evaluated from the value of total

pressure stored at the corresponding cell-centre location.

Distortion coefficient may therefore be defined in terms of total pressure coefficient as

according to Seil et al ( 1997), where the total pressure coefficient Cp is defined as

C = P-Pref p 1 u2

2P ref

(6.13)

(6.14)

From the results and discussion presented in Chapter 5, it is clear that the pump will

ingest a distorted flow field. Therefore, it is likely that the distortion of total pressure at

the pump inlet will have the greatest impact on pump performance. As a consequence of

this, no measure is made of the amount of swirl in the inlet. The standard deviation of

total pressure distortion can be related to the distortion coefficient by

1 2 O"p =Dc(-pUref)

2 (6.15)

It should be noted that the distortion coefficient defined by Eqn 6.13 is based on the

time-averaged flow behaviour assuming "stationary turbulence". An understanding of

the magnitude of the rms turbulent velocity fluctuations present in the flow can be

determined from the CFD solution by examining the distribution of turbulent kinetic

energy over the duct exit. The rms turbulent velocity fluctuations can be determined as

u' =Jfk (6.16)

It must be noted that the rms value of the turbulent velocity fluctuation is only an rms

value and does not reflect a maximum or minimum value, which relies on the

"statistics" of the turbulence, as discussed previously.

186

6.1.4 Internal Volume ofWaterjet Inlet

The weight of water entrained in the waterjet inlet may be considered as representing

either a loss of vessel buoyancy or an additional mass that must be carried by the vessel.

Apart from the consideration of the weight of water entrained in the waterjet inlet, the

size of the waterjet inlet also affects the weight of the inlet ducting and the weight of the

structure surrounding the waterjet inlet. Larger waterjet inlets take up more space in the

vessel, space that could be utilised for other purposes. The internal volume of the

waterjet inlet thus primarily affects the hydrodynamic performance of the vessel. Since

there exists a distinct link between the volume of the waterjet inlet and the vessel

hydrodynamics, it is necessary to keep waterjet inlet volume to a minimum. Hence it

may be argued that large waterjet inlet volumes should be considered as representing

poor hydrodynamic performance, as the hydrodynamic performance of the vessel is

ultimately affected. The volume of the waterjet inlet may be non-dimensionalised by the

inlet throat diameter as

V* = V /D 3 (6.17)

thus allowing the volume of the waterjet inlet to be scaled with throat diameter.

6.1.5 Vertical Forces acting on the Waterjet Inlet

The net vertical force (Fv) acting on the waterjet inlet duct is composed of forces due to

static pressure and shear stress and may be evaluated as

(6.18)

where p is the static pressure, dA the differential surface area vector, ::C the shear stress

vector and ] the unit vector in the vertical (y) direction of a Cartesian coordinate

system. The vertical force may be expressed in non-dimensional form as a lift

coefficient

(6.19)

where Cpy is the lift coefficient, AT the area of the inlet throat and Q the volumetric

flow-rate through the inlet. A positive lift coefficient implies a net lifting force on the

waterjet inlet, whereas a negative lift coefficient implies a "suck-down" effect. The

vertical force acting on the waterjet inlet will affect the trim of the vessel and thus

187

constitutes a waterjet-hull interaction effect, if insufficient bottom plating is present on

the vessel hull to counter the lifting force. Van Terwisga ( 1996) discussed the issue of

so-called "lifting forces" caused by the waterjet inlet and showed that for a waterjet

mounted on an infinite flat plate in a potential flow, there is no net lifting force on the

combined waterjet inlet and plate. Van Terwisga ( 1996) noted that lifting forces could

arise from a reduction in bottom plating on the vessel hull, but showed this effect to be

small (under free-stream conditions) for the geometry he investigated. Despite this, a

study of the vertical forces acting on the waterjet inlet duct is still worthwhile, in order

to examine changes in the vertical momentum of the fluid ingested by the waterjet inlet.

6.1.6 Dimensions of the Inlet Streamtube

Knowledge of the dimensions and cross-sectional shape of the inlet streamtube allows

the flux of mass, momentum and energy ingested by the inlet to be determined and

hence provides for accurate calculation of the waterjet thrust and efficiency. A

knowledge of the dimensions of the inlet streamtube is also essential, in order to avoid

the placement of vessel appendages within the volume of the inlet streamtube upstream

of the inlet. This is due to the fact that appendages will generate a turbulent wake and so

affect the quality of the flow entering the inlet.

6.2 Computational Modelling and Simulation

In this section the CFD modelling which forms the basis of the flow simulations used to

obtain the results presented in Section 6.3 is discussed, together with the geometry of

the waterjet inlet investigated. It must be noted that Section 6.2.2 and Section 6.2.3

describe the generic flow simulation methodology used in subsequent chapters, although

it must be emphasised that the CFD simulations used to obtain the results presented in

Chapter 7 and Chapter 8 are for the same thickness of upstream boundary layer.

At the cruise condition, where the vessel is expected to spend most of its time during

transit, a flush-type waterjet inlet will ingest fluid from the boundary layer of the vessel

hull. As discussed in Chapter 2, the ingestion of boundary layer fluid reduces the

momentum drag associated with the inlet. This leads to an increase in vessel thrust and

waterjet efficiency. Due to the length of modem large high-speed vessels (see Trillo

188

(1994)), it can be expected that the boundary layer thickness upstream of the waterjet

inlet will be of the order of the throat diameter (D) of the inlet. This contrasts with

subsonic aircraft inlet flows, where the upstream boundary layer is thin in comparison

with the dimensions of the inlet opening. The amount of boundary layer fluid ingested

will depend on the width of the inlet and the volumetric flow-rate through the waterjet.

At the cruise condition, which is currently likely to be in the range of 35-55 knots,

(Trillo (1994)), there is a significant net positive suction head available at the pump

inlet. This is caused by the large dynamic pressure resulting from the motion of the

vessel. Waterjet inlets operate at IVR values less than unity in order to allow significant

static pressure recovery. Since many large waterjet installations are of approximately

constant cross-sectional area from the inlet throat to the pump inlet, significant diffusion

of the flow will occur upstream of the inlet throat. The static pressure at the pump inlet

and therefore the amount of flow diffusion from the flow conditions upstream of the

waterjet inlet to the pump inlet will depend, of course, on IVR. The determination of a

suitable IVR is therefore of importance to the design of the waterjet propulsion system.

Van Terwisga (1996) stated a cruise IVR of 0.62, while Griffith-Jones (1994) suggested

a design IVR of 0.7 for the Hamilton 211 jet unit. Kashiwadani (1997) optimised his

waterjet inlet for an IVR of 0.75. The ultimate choice of cruise IVR depends on a design

compromise based on the requirements of the pump, the vessel speed, constraints on the

inlet geometry and the hydrodynamic performance of the waterjet inlet. The choice is

therefore somewhat arbitrary, usually over a range between 0.6 and 0.8. It was therefore

decided to focus on the lower end of this range and so an IVR of 0. 6 was taken as being

representative of the cruise condition.

6.2.1 Waterjet Inlet Geometry

The geometry of the waterjet inlet used for the simulations presented herein is based on

the author's generic geometry described in Chapter 4 and is shown in Fig. 6.1. The

values of the geometric parameters describing this geometry, as well as attributes of the

geometry, are listed in Table 6.1. For a description of the generic parametric geometry,

the reader is referred to Fig. 4.3. The geometry of the waterjet inlet design described

here, is based on a duct diameter of 600 mrn, which is a realistic size used in modem

189

waterjet propulsion units such as the KaMeWa 60S (see Trillo (1994)).

Geometric Parameters Symbol Value Diameter of inlet throat D 600mm Angle of inclination of inlet to horizontal plane a 25° Radius of inlet lip RL 20mm Height of pump centreline above inlet opening H 600mm Radius of curvature of centreline of duct bend Ro 1600mm Length of horizontal duct section downstream of bend LH 400mm Ratio of duct exit area to throat area NAT 1.00 Height of inlet lip centreline above inlet opening plane HL 40mm Angle of inclination of raised lip profile 'Y 7.5°

Geometric Attributes Duct volume - 0.719 m3

Surface area - 5.358 m2

Radius of curvature of inlet ramp - 6424mm Length of inlet opening - 2977mm Overall length of waterjet inlet - 4176mm

Table 6.1 Description of waterjet inlet

Fig. 6.1 Waterjet inlet geometry

6.2.2 Computational Modelling of Flow Domain

Inlet3D is used to mesh the waterjet geometry described in Section 6.2.1 and the default

generic semi-eiJipsoidal shape of the external flow domain. Since the flow in the

waterjet inlet and external domain may be considered to be symmetrical about the

vertical plane of symmetry of the waterjet inlet, only half of the flow domain is meshed.

190

This follows the same approach used in Chapter 5 and neglects the possibility of there

being any gross swirl in the flow as is the case for the waterjet inlet model of Roberts

( 1998) at an NR=0.61. This is a reasonable assumption, as the gross swirl measured by

Roberts is coincidental and a function of the construction of the model. The gross swirl

present, therefore, cannot be generalised to arbitrary waterjet inlet geometries.

Unlike the validation study of the model waterjet inlet presented in Chapter 5, where

there is a need to adjust the boundary conditions to account for the wind tunnel cross­

section, the computational simulations presented in this and subsequent chapters are

based upon a full-scale waterjet operating on a flat plate in a large expanse of water.

Hence, the boundary conditions of the flow domain are thus driven by slightly different

considerations.

The topology used by Inlet3D to mesh a waterjet inlet and a simplified flow domain

external to it, has already been discussed in Section 4.2 and is the same as that used for

meshing the waterjet inlet and external flow domain in Chapter 5. Since the flow under

consideration here exhibits "stationary turbulence", the RANS equations assume an

elliptic form, thus requiring the specification of Dirichlet or Neumann boundary

conditions on all surfaces bounding the flow domain. Table 6.2 lists the grid planes

bounding the flow domain and the Fluent boundary condition cell type applied to each

boundary.

Grid Plane in Computational Space Boundary Plane ~mm ~max 11mm 11max Smm l:max Fluent

1 l;=l.O 0.0 0.5 0.0 1.0 1.0 1.0 Inlet 2 l;=l.O 0.5 1.0 0.0 1.0 1.0 1.0 Inlet

3 l;=O.O 0.0 1.0 0.0 1.0 0.0 0.0 Inlet

4 11=1.0 0.0 1.0 1.0 1.0 0.0 1.0 Wall

5 11=0.0 0.0 1.0 0.0 0.0 0.0 1.0 Axis

6 ~=1.0 1.0 1.0 0.0 1.0 0.0 1.0 Symmetry

7 ~=0.0 0.0 0.0 0.0 1.0 0.0 1.0 Symmetry

Table 6.2 Relationship between boundary conditions and flow domain mesh for generic waterjet inlet simulation

The following assumptions and considerations govern the choice of boundary conditions

191

chosen for the flow domain:

1) Sufficiently far upstream of the waterjet inlet, the flow will be at the free-stream

velocity (vessel speed) relative to the waterjet inlet. This provides the basis for the

specification of Boundary 1 as a Dirichlet boundary condition where the free-stream

velocity, turbulent kinetic energy and turbulent dissipation are specified over this

boundary (outside the boundary layer).

2) Sufficiently far downstream of the waterjet inlet, the flow will assume the ambient

pressure of the water. This assumption provides the necessary basis for the selection

of boundary conditions for the specification of Boundary 2. Boundary 2 is therefore

set as a Dirichlet boundary condition with a static pressure of 0 Pa (neglecting

hydrostatic pressure).

3) Downstream of the waterjet inlet, the mass flow-rate will be reduced relative to that

upstream of the inlet. This is due to the flow of mass out of the external flow domain

and into the inlet. If Ih, is the mass flow-rate through boundary i, then the mass flow­

rate through Boundary 2 is simply

(6.20)

Eqn 6.20 expresses a global mass conservation over the flow domain and also shows

a net diffusion of flow from Boundary 1 to Boundary 2. It may thus be expected that

the average static pressure (neglecting hydrostatic effects) over Boundary 1 will be

less than Boundary 2.

4) Boundary 3 is located one duct diameter downstream of the exit of the waterjet inlet

as in Chapter 5. Rather than specifying an "outlet" (Neumann) boundary condition as

before, Boundary 3 is now specified as a Dirichlet boundary condition having a

uniform velocity distribution. With Boundary 2 specified as a Dirichlet boundary

condition of constant static pressure (neglecting hydrostatic pressure), it is necessary

to specify Boundary 3 as a velocity Dirichlet boundary condition in order to satisfy

global conservation of mass according to Eqn 6.20.

Furthermore, it was found that the location of Boundary 3 one diameter downstream

of the duct exit is sufficient to minimise the influence of the constant velocity

distribution over this boundary, on the upstream flow. Using this type of boundary

condition for Boundary 3, the NR of the flow through the waterjet inlet can be

192

accurately set by simply specifying the velocity over Boundary 3.

The total length, half-width (since only half of the waterjet is simulated) and depth of

the external domain are specified as 8000 mm, 2000 mm and 3000 mm, respectively.

The dimensions of the external domain are arbitrarily chosen. The ratios of these

dimensions relative to the duct diameter, are greater than the corresponding ratios for

the external domain of the waterjet simulations presented in Chapter 5. The external

domain chosen for the simulation of the waterjet inlet in Chapter 5 was found to be large

enough such that the specification of the boundary conditions over Boundary 1 and

Boundary 2 has no adverse affect on the solution. In fact, it was found that the flow into

and inside of the waterjet inlet is governed more by the flow conditions closer to the

inlet opening. It may be concluded that the size of the external domain chosen here is

adequate for accurate simulation. The surface grid enclosing the simulated flow domain

is shown in Fig. 6.2. The grid meshing the plane of flow symmetry of the modelled flow

domain is shown in Fig. 6.3.

y

~z

Fig. 6.2 31x41x102 surface grid bounding the simulated flow domain

193

Fig. 6.3 Grid on centreplane of flow domain

waterjet inlet, the simple approach of "growing" a two-dimensional flat plate boundary

layer profile (in the absence of externally imposed pressure gradients) is adopted in

order to provide approximate values of U, k and E on Boundary 1. It must be noted that

this approach to boundary layer representation is a gross simplification of an actual hull

boundary layer, which is subject to pressure gradients and streamwise curvature. For all

intents and purposes, this is a reasonable approximation considering the geometric

simplicity of the external flow domain modelled.

6.2.3 Computational Simulation

In seeking to investigate the effect of upstream hull boundary layer thickness on the

hydrodynamics of waterjet inlet flow, the question as to what upstream boundary layer

thicknesses are representative of actual vessel hull boundary layer development arises.

The formula of Wieghardt ( 1972)

8 = 0.27L(ReL)-116 (6.21)

can be used to approximate the growth of a turbulent boundary layer on a vessel hull

form for the high Reynolds numbers typically encountered at full scale (Steen and

Minsaas (1995)). In Eqn 6.21, 8 is the boundary layer thickness, L the wetted-length

over which the boundary layer grows and ReL the Reynolds number based on L. Fig. 6.4

shows the growth in thickness of the hull boundary layer (for a kinematic viscosity of

194

9.0x10-7 m2s-1), with wetted-length upstream of a waterjet inlet, for a vessel cruise speed

of 40 knots.

0.8

0.7

0.6

0.5 ,..-... s '-' 0.4 t-0

0.3

0.2

0.1

0.0

0 10 20 30 40 50 60 70 80 90 100

L(m)

Fig. 6.4 Growth of hull boundary layer with wetted-length

It can be seen from Fig. 6.4 that for a waterjet inlet having a throat diameter of 600 mm,

the boundary layer thicknesses are of the same order as the throat diameter. Hence the

waterjet inlet will operate in the presence of a "thick" hull boundary layer. Furthermore,

for a wetted-length of 77.3 m upstream of the inlet, the boundary layer has a thickness

equal to the throat diameter of the inlet in the present case! It is therefore necessary to

consider a range of non-dimensional boundary layer thicknesses (OlD) corresponding to

typical wetted-lengths upstream of the waterjet inlet on modern high-speed vessels. An

examination of the vessel length data contained in Trillo (1994 ), implies a range of o/D

up to and beyond unity (for D=600 mm).

In view of the likely range of boundary layers encountered for a waterjet inlet installed

in different vessels, eleven CFD simulations were run spanning a range of non­

dimensional boundary layer thickness from 0/D=O.O to 0/D=l.O, with equal increments

of boundary layer thickness (.!\0/D=O.l) between each case. The free-stream velocity is

specified as 20.58 ms-1, corresponding to an equivalent vessel speed of 40 knots.

Upstream boundary layer velocity profiles are plotted in Fig. 6.5.

195

I 0

09

0.8

"!! ~ 07 -.. ~

06

05

04 0 0

• • !II! I X lK 0

+

Ai I ~ ~

-0

t • li!i i5

l>!D +00.01A02

X 0 3 X 0.4 0 0 5

+06 -07-08

0 0 9 [] 1.0

\1)

0

y/D

"' 0 r-0

00 0

Fig. 6.5 Upstream boundary layer velocity profiles

0

Since water is simulated, the fluid density is specified as 1000 kg/m3 and the molecular

viscosity as 9.0xl0·4 Ns/m2• As a cruise IVR of 0.6 is simulated, a velocity of 12.35 ms·1

is specified over Boundary 3. The Reynolds number in the 600 mm diameter throat of

the inlet duct, based on this outlet velocity, is 8.23x106• This is of the order of 107 and is

therefore a very high Reynolds number. This value is also large in comparison to the

Reynolds number of the flow in the waterjet inlet model studied in Chapter 5, which is

more of the order of 105-106.

All computations are performed using the 31x41xl02 grid size which was found to give

adequate results from the validation study of Chapter 5. The two-equation RNG k-E

turbulence model is used to provide general turbulence closure in conjunction with the

Two-Layer-Based Nonequilibrium (TLBN) wall function for near-wall turbulence

closure. Calculations were deemed to have converged when the sum of the normalised

residuals of the transport equations and the pressure correction equation (Section 3.6.2)

fell below 1x10·3 (Fluent's default convergence criteria).

6.3 Results

In this section the computational results relating to the CFD simulations discussed in

Section 6.2 are presented and the details of the computed flow behaviour are discussed.

A more detailed discussion of the results, relating the observed flow behaviour to the

upstream momentum and energy fluxes ingested by the waterjet inlet is reserved until

196

Section 6.4.

Fig. 6.6 and Fig. 6.7 show the variation, with boundary layer thickness, of the static

pressure coefficient (Cp) on the upper and lower surface of the waterjet inlet (on its

vertical plane of flow symmetry), respectively. On the upper inlet, the non-dimensional

arc length (SID) is measured from the ramp tangency point. On the lower inlet surface,

SID is measured from the lip trailing edge with SID increasing as arc length is measured

into the inlet. This follows the convention introduced in Chapter 5. The relationship

between SID and the geometric features of the waterjet inlet is shown in Table 6.3. An

examination of both figures reveals a trend of decreasing static pressure within the inlet

as the thickness of the upstream boundary layer increases. This trend is to be expected,

as reduced dynamic pressures (corresponding to the increased fluid retardation as the

boundary layer thickens) lead to a reduction of possible static pressure recovery within

the waterjet inlet.

"-u

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0.0

-0.1

-0.2

0

BID --o.o --0.2 - - - - - - 0.4 --0.6 --0.8 ----- 1.0

2 3 4 5 6 7 SID

Fig. 6.6 DistributiOn of static pressure coefficient on upper waterjet surface at centreplane

0.8

0.7

0.6

0.5

u"" 0.4

0.3

0.2

0.1

--o.o --0.2 - - - - - • 0.4 --0.6 --0.8 ----- 1.0

0. 0 +'-'--'---f-'--'-'+.LL.Lf-W-I..j-J-J~-'-'-f..L..J-L+-'-L-4-'-'-'--f-J'-'-'--l

~ ~ 00 ~ ~ ~ ~ 00 0 N ~ 0 0 0 ~ ~ - - ~ N N N

SID Ftg. 6.7 Distribution of static pressure coefficient on

lower waterjet mlet surface at centreplane

On the inlet ramp for SID~3, there is little variation in the static pressure coefficient

with increasing BID, although it is interesting to note that with increasing BID, static

pressure coefficient increases. This latter point may be attributed to the decreased

velocity in the immediate vicinity of the ramp surface as BID increases. It is also

interesting to note that the distribution of Cp over the inlet ramp for SID~3, for the range

of BID examined, is similar to that of the geometry of Roberts (1998) and may be

attributed to the relatively large radii of curvature of both geometries.

197

Upper Centreplane SID

Ramp tangency point 0.00 End of ramp I Start of inclined duct section 4.67 End of inclined duct section I Start of bend 5.14 End of bend I Start of straight duct section 6.53 End of straight duct section 7.19

Lower Centreplane SID

Lip tailing edge 0.00 Start of lip 0.26 End of lip I Start of inclined duct section 0.34 End of inclined duct section I Start of bend 0.82 End of bend I Start of straight duct section 1.76 End of straight duct section 2.43

Table 6.3 Relationship between SID and the waterjet inlet geometry

Of particular relevance is the relationship between the static pressure distribution over

the inlet lip and the likelihood of the onset of lip cavitation. Fig. 6.8 reveals interesting

trends regarding the variation of Cp in the vicinity of the inlet lip.

The following trends are apparent from this figure:

1) The location of the stagnation point (indicated by the local maximum in Cp) of the

dividing streamline on the plane of flow symmetry moves closer to the centre of the

lip (S/0=3.0) from the upper-lip.

2) The magnitude of the stagnation pressure decreases with increasing 8/D. This is

primarily a result of the reduced velocity in the vicinity of the inlet lip, resulting from

1.0

0.5

c. -0.5 u

-1.0

-1.5

-2.0

0.0

BID --o.o --0.2 •••••• 0.4 --0.6 --0.8 ----- 1.0

0.1 0.2

SID

0.3 0.4

Fig. 6.8 Distribution of static pressure coefficient in the vicinity of the inlet lip

198

the reduced dynamic pressure of the upstream flow.

3) The minimum static pressure on the underside of the inlet lip (located in the range

0.26<SID<0.28) increases with increasing 8/D. The rate of reduction in the

magnitude of static pressure is largest at small 8/D as can be seen from Fig. 6.8 and

Fig. 6.9. The increase in minimum static pressure is substantial over the range of 8/D

investigated and can be seen, from Fig. 6.9, to range from Cp=-2.48 to Cp=-0.94.

4) The minimum static pressure on the uppers ide of the inlet lip decreases with

increasing 8/D. This is a direct consequence of the location of the stagnation point on

the inlet lip. As the stagnation point moves toward the centre of the lip (with

increasing 8/D) the flow accelerating from the stagnation point into the inlet

negotiates a larger turning angle which results in greater velocity and hence lower

static pressure on the duct surface.

The above-mentioned trends reveal the significant impact that the thickness of the

upstream boundary layer has on the flow in the vicinity of the inlet lip and hence the

cavitation performance of the waterjet inlet. Boundary layer ingestion may thus be seen

as being beneficial in raising minimum static pressures at the lip in the cruise condition.

Operation of the inlet in the presence of an upstream boundary layer therefore improves

the margin against cavitation inception (if cavitation is absent), or alternatively

eliminates cavitation. Fig. 6.10 shows the maximum vessel speed at which cavitation

-0.7

-0.9

-1.1

Q. -1.3 u

-1.5 s:: ~ -1.7

-1.9

-2.1

-2.3

-2.5

0 d d

8/D ~ ~ v ~ ~ ~ 00 ~ q d d d d d d d d

Fig. 6.9 Variation of minimum lip static pressure coefficient with boundary layer thickness

199

29

27 ,....._ "' 25 ..... 0 s:: ~

23 '-"

~ s ~ 21

19

17 0 ~ ~ v ~ ~ ~ 00 ~ q d d d d d d d d d d

8/D

Fig. 6.10 Variation of lip cavitation inception with vessel speed and boundary layer thickness

will begin for a given upstream boundary layer thickness and reveals a greater speed

range for cavitation-free operation (of the waterjet inlet) when the upstream boundary

layer is thick. Cavitation inception was calculated from Eqn. 6.1 by setting the

cavitation number to zero and solving for the reference velocity using the static pressure

data of Fig. 6.9. It must be noted that Reynolds number effects have been neglected in

the calculation of the data plotted in Fig. 6.9.

The variation of the dimensions and shape of the cross-section of the inlet streamtube,

upstream of the waterjet inlet, is shown in Fig. 6.11 for four upstream boundary layer

thicknesses corresponding to 6/D=O, 0.3, 0.7, 1.0. This range is chosen as it spans the

range of BID investigated with approximately equal differences between the values.

Indeed, these four cases are investigated in greater detail in the remainder of this

chapter. It can be seen that as the thickness of the upstream boundary layer grows, so do

the dimensions of the inlet streamtube. The cross-sectional shape is retained, but the

depth and width of the cross-section grow with increasing 6/D. It is interesting to note

that the rate of growth of the width of the streamtube cross-section decreases with

increasing 6/D and it appears that the widths for the cases of 6/0=0.7 and 6/D=l.O are

identical. This will be discussed at greater length in the next section.

0 r- 00

d 0 d 0.00

-0.05 BID -0.10 ---0.0

--0.3 -0.15 --0.7

e -0.20 ----- 1.0 >.

-0.25

-0.30

-0.35 -- --___ _ -0.40

ziD

Fig. 6.11 Growth of the cross-section of the inlet streamtube with boundary layer thickness

The streakline plot of Fig. 6.12 shows the generic diffusing flow into the waterjet inlet

for the geometry consideration here. In this case the flow corresponding to 6/D=O has

been shown.

200

Fig. 6.12 Computed streaklines on centreplane (OID=O.O)

It can be seen from the velocity vector plots on the centreplane of the modelled flow

domain (Fig. 6.13), over the range of OlD investigated, that no flow separation is evident

in the inlet. In Chapter 5, the limitations of the two-equation RNG k-E turbulence

modelling in capturing the full effects of flow separation are discussed and related to the

assumptions underlying the construction of this model. Never the less, the model is able

to predict near-separation behaviour at a nominal IVR of 0.61. In the cases examined

here, it is clear that there is no separation/near-separation behaviour as an examination

of the velocity vectors plots reveal.

It is thus evident that the waterjet inlet examined here offers an improvement in design

over the waterjet inlet of Roberts (1998), which is based on a generic industrial design.

This improved hydrodynamic performance may be attributed to the circular inlet throat

used in the author's parametric geometry. This acts to converge the diffusing flow on

the upper part of the ramp toward the centreplane, thus resulting in larger local

velocities near the upper inlet surface and hence a fuller velocity profile. This in tum

results in a greater resistance to flow separation. The ramp of the waterjet inlet of

Roberts ( 1998) is flat in the transverse direction. The flow in the vicinity of the ramp is

therefore more two-dimensional in nature and develops in an analogous manner to two­

dimensional flow over a convex surface in the presence of an adverse pressure gradient.

The flow in the geometry considered here is clearly three-dimensional on the ramp

surface. Fig. 6.14 shows the distribution of total pressure coefficient over the cross­

section of the duct exit with increasing OlD. It can be seen that the average total pressure

over the cross-section decreases with increasing OlD.

201

- =206m/s

---

- =206m/s

---

- =206m/s

---

- =206m/s

---

------------- --- - ------~-----~-- ------------------------

a) o/D=O.O

--------~-

b) o/D=0.3

-~~----~----------

c) o/D=0.7

-~-

d) o/D=l.O

Fig. 6.13 Computed velocity vectors on centreplane

202

a) OID=O.O b) o/D=0.3

c) o/D=0.7 d) o/D=l.O

Fig. 6.14 Distribution of total pressure over the cross-section of the duct exit

This is due to an increasing ingestion of low-momentum boundary layer fluid. The

distribution of total pressure over most of the cross-section of the duct exit is fairly

uniform at 0/D=O.O, except in the wall boundary layer. Hence, there is a low distortion

of total pressure over the duct cross-section. The distribution of total pressure over the

203

duct cross-section can be seen to be very non-uniform for 8/D=0.3, but becomes more

uniform with increasing 8/D. This trend may be quantitatively expressed through the

total pressure distortion coefficient (De) defined by Eqn 6.13. It can be seen from Fig.

6.15 that as 8/D increases from zero, De rises steeply to reach a maximum around

8/D=0.2, before decreasing with increasing 8/D. The rate of decrease of De, dDcfd(8/D),

decreases with 8/D and no further reduction of De is evident for 8/D>0.9 over the range

of 8/D examined.

The distribution of total pressure coefficient on the centreplane of the duct exit is shown

in Fig. 6.16. In Fig. 6.16, y/D is the non-dimensional distance from the surface of the

lower duct The above-mentioned trends in the magnitude and uniformity of the total

pressure distribution can be seen. As expected, the distributions of total pressure over

the cross-section of the duct exit (shown in Fig. 6.14 and Fig. 6.16) show larger total

pressure toward the lower duct. This is due to the higher velocities in this region

resulting from the bend and the thinner upstream boundary layer on the wall of the

lower duct.

0.12

0.11

0.10

0

0 0.09

0.08

0.07

0 N M V ~ ~ ~ 00 ~ ~ 0 0 0 0 0 0 0 0 0 0 -

BID

Fig. 6.15 Variation of distortion coefficient with boundary layer thickness

1.0

0.9

08

0.7

0.6

u 0.5

0.4

0.3

0.2

0.1

0.0 0 0

··. .. .. .. ~ .............. ..

BID --0.0 - ••. - • 0.3 --0.7 ----- 1.0

N M V ~ ~ ~ 00 ~ ~ 0 0 0 0 0 0 0 0 0

y/D

Fig. 6.16 Distribution of total pressure over the duct exit on centreplane

It is interesting to examine the variation of the secondary flow behaviour at the duct

exit, shown in Fig. 6.17. It can be seen from Fig. 6.17 that all cases shown exhibit one

primary streamwise vortex (for each half of the duct) extending from the upper inlet

204

near the vertical plane of symmetry, down to the plane of flow-symmetry in the lower

inlet. A gross secondary flow directed from the lower duct to the upper duct can also be

seen. As the thickness of the upstream boundary layer increases, the magnitude of the

secondary flow appears to decrease.

a) BID=O.O b) BID=0.3

-=2mls

' '

'' '' I I

11

'''I I Ill I

"'•• ., ', ""•·)''')' \, '': 1111 11 I ) / I

Jlllllll Ill ,' I /1 I I

If(// II / I

II I I

1111 II ,'

1 1 I 11

I I I I

c) BID=0.7 d) BID=l.O

Fig. 6.17 Secondary flow at duct exit

205

The computed secondary flow behaviour shown here differs from the computed

secondary flow for the waterjet inlet of Roberts (1998), shown in Chapter 5. The

difference is due to the absence of a second streamwise vortex located near the plane of

flow symmetry in the upper duct (c.f. Fig. 5.31). In fact, the secondary flow behaviour

resembles that in a bend of small turning angle and circular cross-section. The

difference between the secondary flow here and that reported in Chapter 5 may be

attributed to the influence of the waterjet inlet geometry, in particular the difference in

the development of upstream secondary flow in the inlet region. The development of

secondary flow is more pronounced in the waterjet inlet of Roberts (1998) due to the

vertically aligned sidewalls of the inlet. The curved sides of the parametric inlet

considered here make it more difficult for strong cross-stream vorticity to develop. As a

consequence of this, the secondary flow upstream of the bend is weaker and hence its

development is dominated by bend pressure gradients and so it resembles the secondary

flow in a bend of small bend angle.

The distribution of non-dimensional turbulent velocity-scale (Vk!Uref) and turbulent

length-scale (ReiD) on the centreplane of the flow domain is shown in Fig. 6.18 for the

cases corresponding to ()/D=O and <>ID=l.O. Unfortunately, in the figure there appears to

be some spurious numerical behaviour in the vicinity of the duct centreline for <>ID= 1.0.

This is caused by the axis boundary condition type. Despite this, the trends in predicted

flow behaviour are clearly evident.

At a nominal <>ID=O.O, it is interesting to note how the turbulence spreads within the

upper part of the inlet from an upstream boundary layer of almost negligible thickness.

While the convex curvature of the inlet ramp should act to "collapse" the turbulence,

diffusion of the flow and the adverse pressure gradient encountered by the flow on the

upper ramp surface and first half of the bend (the outer bend being a concave surface)

act to spread and intensify the turbulence in the upper inlet. This results in energetic

eddies of relatively large velocity and length-scales. Hence it may be expected that

relatively large total pressure losses will occur in the upper part of the inlet. This is

exactly the case as can be seen from an examination of Fig. 6.14 and Fig. 6.16. The

balance of the flow in the waterjet inlet forms an essentially "in viscid core".

206

"k/U"' A 0100

9 0090

8 0080 7 0070

6 0060

5 0050 4 0040

3 0030

2 0020

0010

"k!U,.,

~~ 0060 0050 0040 0030 0020 0010

ljD (Expanded V1ew of L1p) A 00129

9 00116 8 00103 7 00090

6 00077

5 00064 4 00051

3 00039

2 00026

00013

a) 3/D=O.O

,Jw, .. A 0100

9 0090

8 0080 7 0070 6 0060 5 0050 4 0040 3 0030

2 0020 2 2

0010

"k!U .. ,

[ 0200 0183 0167 0150 0133 0117 0100

1/D (Expanded V1ew of L1p) A 00475

9 00427 8 00380

7 00332

6 00285

5 00237 :A;..: 4 00190 - • 3 00142 9

2 00095

00047 ---b) 3/D=l.O

Fig. 6.18 Computed turbulent velocity and length-scales on waterjet centreplane

207

What is particularly interesting is the behaviour of the turbulence in the vicinity of the

inlet lip, where new boundary layers will grow on the surface of the inlet lip

downstream from the stagnation point. A question of interest is the ability of the

turbulence model to accurately predict the correct laminar-to-turbulent transition

behaviour of the developing boundary layers on the lip. It is not possible to quantify the

accuracy of the turbulence model in predicting boundary layer transition without a

detailed and comprehensive experimental validation study. It is however likely that the

turbulence model will predict transition behaviour at Reynolds numbers at least an order

of magnitude too low (see Wilcox (1993)). Never-the-less, examination of the predicted

behaviour of the turbulence in the vicinity of the inlet lip provides valuable qualitative

insight.

The extremely small length-scales in the vicinity of the inlet lip indicate highly

dissipative eddies. An initial examination of the turbulent velocity-scales in the vicinity

of the inlet lip in Fig. 6.18a appears to show lower levels of turbulence than is the case

in Fig. 5.39b. This is, however, deceptive. The Reynolds number of the flow in the

present case is of the order of 107 whereas for the simulations of Chapter 5, the

Reynolds number is more of the order of 105-106• Hence there is a greater separation of

length-scales between the integral scale of the turbulence and the Kolmogorov

microscale of the dissipative eddies. Thus the region of greatest turbulent production

and hence velocity-scale, is closer to the wall and is less easily seen. It therefore appears

as though the turbulent velocity-scale in the vicinity of the inlet lip is smaller than for

the waterjet inlet flow of Chapter 5. In addition, at higher Reynolds number, developing

boundary layers are generally thinner as can be seen from a comparison of flow in the

lip region for Fig. 6.18 and Fig. 5.39b. It is also evident that the adverse pressure

gradient on the underside of the inlet lip acts to increase the turbulence intensity.

The following trends in the turbulence behaviour may be noted as the thickness of the

upstream boundary layer increases:

1) The flow throughout the waterjet inlet becomes turbulent and the overall magnitude

of the turbulent velocity-scale increases as a result of the diffusion of the upstream

turbulence into the inlet.

208

2) Overall turbulent length-scales throughout the waterjet inlet increase in magnitude

owing to the ingestion of larger, more energetic eddies present in the upstream

boundary layer.

3) The turbulence levels in the vicinity of the inlet lip increase on both the underside and

upperside of the lip. Velocity and pressure gradients in the vicinity of the lip act to

augment the existing turbulence levels present in the upstream boundary layer.

Fig. 6.19 shows the variation of total pressure recovery efficiency (Eqn 6.3) with the

upstream boundary layer thickness. It is clear that as o/D increases, there is a reduction

in 11· The rate of reduction of 11, d11/d(o/D) decreases with increasing o/D. The primary

reason for this reduction in 11 is due to the ingestion of low momentum boundary layer

fluid which reduces the energy of the ingested fluid. This will be discussed at greater

length in the next section.

The reduced momentum flux upstream of the waterjet inlet (with increasing o/D) results

in a reduction in non-dimensional lift coefficient on the inlet as shown in Fig. 6.20. This

will also be discussed at greater length in Section 6.4.

0.95

0.90

0.85

0.80

0.75

0.70

0.65

0. 60 -j-LW..Lj-l-1-'-'-j-U..L.L.j-L.l..U.f'U-U..f-'-'-'-'-J-U-U+'-'-'-'-f-'--W"+'-'-'-'-1

0 ~ ~ ~ ~ ~ ~ 00 ~ q d d d d d d d d d d

Fig. 6.19 Reduction of the total pressure recovery efficiency with boundary layer thickness

209

0.35

0.30

0.25

~ 0.20

u 0.15

0.10

0.05

0. 00 -j-LWU-j-I-L.J..J..f-L.LL.Lf-l..LL.Lj-'-'--'-'-f-U-'-'-J-'-U"+'-'-'-'-1-'-l.W.fJ-'-'-'-l

0 ~ ~ ~ ~ ~ ~ 00 ~ q d d d d d d d d d d

BID F1g. 6.20 ReductiOn in lift coefficient With

boundary layer thickness

6.4 Discussion of Results

The change in flow behaviour in the waterjet inlet with ingestion of fluid from boundary

layers of increasing thickness may be primarily attributed to the changes in the upstream

momentum and energy fluxes ingested by the inlet. In Chapter 2 the notion of a

momentum flux coefficient (Cm) and an energy flux coefficient (Ce) were introduced in

order to account for the effect of upstream boundary layer velocity and pressure fields

on the ingested flow of momentum and energy, when compared with a corresponding

"free-stream" momentum and energy flux over the cross-section of the inlet streamtube.

Neglecting the effect of hydrostatic pressure, for the flow domain considered here, Cm

may be evaluated as

1 JW/2Jh Cm = . (pU 2) dydz

mUrer o o (6.22)

Similarly, Ce may be evaluated as

1 JW/2Jh ce = . I 2 (p+fpU2 )Udydz (m-zpUrer) o o

(6.23)

where W is the total width of the cross-section of the inlet streamtube and h is its depth.

Roberts (1998) and Griffith-Jones (1994) clearly showed that the dimensions and shape

of the cross-section of the inlet streamtube change with IVR. They also showed that the

shape of the cross-section is neither rectangular, as suggested by the ITTC (see

ITTC( 1987)) or elliptical, although there is a general resemblance to this latter shape. In

order to obtain a suitable analytic approximation to the shape of the cross-section of the

inlet streamtube, regression analysis is used to fit an nth-order polynomial of the form

(6.24)

to the data. In Eqn 6.24, a1 are the coefficients and b1 the exponents of the ith term and z

as defined in Fig. 6.11. It was found that polynomials of sixth order gave excellent

correlation with the streamtube cross-sectional data shown in Fig. 6.11.

Using a power-law velocity profile of the form

~=(r)l'n uref 0

(6.25)

to represent the velocity distribution over the cross-section of the inlet streamtube, the

210

following expressions for Cm result

U JW/2 em=~ (y-28/(n+2))dz Q 0

U ( )2/n( ) W/2 = ~ _! _n_ f y<n+2)/ndz Q 8 n+2 Jo

(6.26)

Similarly Ce becomes

U JW/2 ce =~ (y-38/(n+3))dz Q 0

U ( 1 )3/n ( ) W/2 = ~ - _n_ f y<n+3)/ndz

Q 8 n+3 Jo

8:s;h

(6.27)

8>h

The variation of Cm with 8/D is shown in Fig. 6.21, calculated for the actual streamtube

cross-sectional shape and the corresponding Cm distribution if an equivalent rectangular

streamtube cross-section is calculated for:

1) The same volumetric flow-rate through the waterjet inlet.

2) The same depth of the inlet streamtube as the corresponding CFD result.

A similar plot for Ce is presented in Fig. 6.22. It is interesting to note that the

depth/width ratio of the equivalent rectangular streamtube will be approximately 0.22

for a 8/D of zero, which represents the minimum depth/width ratio over the range of 8/D

examined.

The variation of Cpy, the minimum static pressure on the inlet lip and the general static

pressure distribution on the duct surface, may be correlated with the Cm, as these

quantities ultimately depend on the ingested momentum flux. The variation of the 11.

may be correlated with Ce. as 11 is primarily dependent on the ingested energy flux. The

correlation between Cm with Cpy and Cp is presented together with the correlation

between Ce and 11 in Table 6.4. The correlation coefficient is calculated as

N L (X, - Jlx )(Y, - Jlv) j=l Pxv =....:...._ _______ _ (6.28)

where PxY is the correlation between the two data sets X and Y, Jl the mean of the data

set given by

211

(6.29)

(for set X), N the number of elements of the data set and the standard deviation crx, is

defined by

(6.30)

It can be seen that the correlation between the respective data sets, shown in Table 6.4,

is good. This therefore indicates a clear link between the effect of ingested momentum

and energy flux on the hydrodynamic performance of the waterjet inlet.

Data Set 1 Data Set 2 PxY Cp Cm 0.997

Cpy Cm -0.974

11 Ce 0.999

Table 6.4 Data set correlation

1.00 1.00 ~Actual

~Actual -Rectangle

0.95 -Rectangle 0.90

0.90 0.80

0.85 " e u u

0.80

0.75

0.70

0 N ~ ~ ~ ~ ~ ~ ~ 0. d d d d d d d d d d

BID Fig. 6.21 Reduction in momentum flux coefficient

with boundary layer thickness

0.70

0.60

0 N ~ v ~ ~ ~ ~ ~ ~ d d d d d d d d d d

Fig. 6.22 Reduction in energy flux coefficient with boundary layer thickness

The results of Fig. 6.21 and Fig. 6.22 reveal two results of primary importance. The first

is that the assumption of a rectangular cross-section of the inlet streamtube will lead to

significant errors in the calculation of ingested momentum and energy flux. This is

particularly detrimental to the calculation of waterjet thrust or efficiency when using a

212

parametric model, such as the one presented in Chapter 2. These results therefore

highlight the necessity of an accurate determination of the shape of the cross-section of

the inlet streamtube so that an accurate integration of the velocity profile can be made

and hence accurate calculation of ingested momentum and energy flux.

The second trend that is apparent is the decreasing rate of change of Cm and Ce with olD.

fu other words, both dC.Jd(O/D) and dCJd(o/D) decrease with increasing BID. This is

evident from Fig. 6.21 for Cm and Fig. 6.22 for Ce , where these quantities decrease in an

almost linear manner for O<h, but then follow a much reduced rate of decrease driven by

a power-law. The reader is also referred to Eqn 2.44 to Eqn 2.47, which further

illustrates this point for an inlet streamtube of rectangular cross-section. Since there

exists a good correlation between Cm , Ce and the data, it is therefore of no surprise that

the greatest change in the results occurs for small o/D (eg. O/D<0.3), with a decreasing

rate of change in the results as olD increases. fu a similar manner, the rate of change of

the dimensions of the inlet streamtube, in particular its width, decrease as olD increases.

This is evident from Fig. 6.11.

So what causes this decreasing variation of Cm and Ce with BID? The primary reason

may be attributed to the boundary layer velocity profile becoming "fuller" with

increasing boundary layer thickness. Since the cross-section of the inlet streamtube will

be completely immersed in the boundary layer fluid for o/h~l, the change in velocity

across the cross-section decreases with increasing boundary layer thickness. Assuming

o/h~1 and a power-law velocity profile (Eqn 6.25), the rate of change of the streamwise

velocity with boundary layer thickness may be given by

acu;u ref) = _..:!._1_(~)1

--vn ()() n ()2 y

Alternatively Eqn 6.31 may be written as

()(U/U ref) = -Cjo~+lfn ()()

(6.31)

(6.32)

where C is a constant dependent on n and the selected y position. Therefore as the

boundary layer thickens, the rate of decrease of the streamwise velocity is reduced and

hence the rate of reduction of Cm and Ce decreases.

213

This also helps to explain the variation of distortion coefficient with 8/D, as shown in

Fig. 6.15. The distortion coefficient grows rapidly to reach a maximum at 8/D:::::0.2,

before falling at a reducing rate. At low 8/D (Oih<0.6) the waterjet inlet ingests a

combination of free-stream fluid with a growing proportion of boundary layer fluid,

hence Qb1/Q increases. If one considers the velocity profile across the cross-section of

the inlet streamtube, there is a large variation in the y direction. It may be considered

that the waterjet inlet ingests two "layers" of fluid. Although there is some mixing of

these two "layers" by the secondary flow present in the duct, a large distortion of the

total pressure at the duct exit is a direct result of the large variation in streamwise

velocity profile over the cross-section of the inlet streamtube. For 8/h~0.6, Qb1/Q tends

to unity and for Qbi/Q=1, d(U/Urer )/dO decreases over the whole cross-section of the

inlet streamtube as 8/D increases. Hence the waterjet inlet will ingest a more uniform

distribution of fluid, resulting in a more uniform distribution of total pressure at the duct

exit, albeit with a lower average value.

The results for 11, shown in Fig. 6.19, show how crucial it is that the thickness of the

boundary layer be accounted for when quoting this efficiency, otherwise quoted results

will be misleading. The discrepancy between 11 and Ce may be accounted for as viscous

losses within the waterjet inlet. This may be denoted as 111n according to the definition of

Eqn 2.32 for the parametric model of waterjet efficiency presented in Chapter 2. The

variation of 111n with 8/D is shown in Fig. 6.23. The figure reveals that there is little

variation over the range of 8/D examined. This is possibly due to the lack of flow

separation within the waterjet inlet, with 111n varying between 0.912 and 0.965. There is,

however, some uncertainty in these values, due to an uncertainty in the calculation of Ce.

The power-law velocity profile fit to the actual velocity profile over the cross-section of

the inlet streamtube was within 2% relative error. Inaccuracies are therefore translated to

a maximum possible relative error of 6% in the calculation of Cm, thus leading to an

uncertainty in the results presented in Fig. 6.23.

If it is assumed that these values are reasonably accurate, then the values of 11 quoted in

the literature, typically around 0.8 (Verbeek (1992) quotes 0.83), either account for very

214

large viscous losses (eg. flow separation in the waterjet inlet) or the ingestion of

boundary layer fluid causing a reduction in Ce and hence 11·

0.98

0.96

0.94

~ 092

0.90

0.88 6% Error bars shown

0 N M v ~ ~ ~ 00 ~ 0 0 0 0 0 0 0 0 0 0 0

0/D

Fig. 6.23 Variation of inlet efficiency with boundary layer thickness

6.5 Closure

Parameters that can be used to assess the hydrodynamic performance of the waterjet

inlet were presented in Section 6.2. These provide the basis, not only for interpreting the

results of the boundary layer investigations presented in this chapter, but also for the

design and optimisation-related investigations presented in Chapter 7 and Chapter 8,

respectively. Of particular interest are:

1) Minimum static pressure, expressed either as a cavitation number by Eqn 6.1, or as a

static pressure coefficient by Eqn 6.2

2) Total pressure recovery efficiency (Eqn 6.3) which can be expressed either on a mass­

averaged or an area-averaged basis

3) Distortion coefficient, quantifying the non-uniformity of the total pressure at the duct

exit, can be calculated using Eqn 6.13

4) Turbulent velocity and length-scales at the duct exit.

5) Drag forces acting on the inlet.

6) Internal volume of the waterjet inlet, Eqn 6.17.

7) Vertical forces acting on the waterjet inlet, Eqn 6.19.

8) Dimensions and shape of the cross-section of the inlet streamtube.

215

The generic CFD modelling methodology outlined in Section 6.2, forms the basis of the

computational simulation of this and subsequent chapters. The methodology is based on

the solution of the Reynolds-averaged Navier Stokes equations with two-equation RNG

k-E turbulence modelling and the use of a Two-Layer-Based Nonequilibrium wall

function for near-wall turbulence closure.

The effect of the thickness of the upstream boundary layer on the hydrodynamic

performance of the waterjet inlet was investigated for 11 boundary layer thicknesses

ranging from OID=O to O/D=1, where 0 is the boundary layer thickness and D the

diameter of the throat of the waterjet inlet. The geometry of the simulated waterjet inlet

is described in Section 6.2.1 and has a throat diameter of 600 mm.

The following conclusions may be drawn regarding the effect of increasing boundary

layer thickness on the hydrodynamic performance of the waterjet inlet, for the specific

waterjet inlet geometry and IVR investigated:

1) There is an overall reduction in static pressure within the duct

2) The minimum static pressure on the surface of the waterjet inlet occurs on the

underside of the inlet lip and increases significantly with o, thus reducing the

likelihood of cavitation inception

3) The stagnation point moves closer to the centre of the lip

4) Vertical forces acting on the waterjet inlet decrease

5) The width and depth of the inlet streamtube grow

6) Total pressure recovery decreases

7) The distortion coefficient (Eqn 6.13) increases rapidly from 0/D=O, reaches a

maximum around O/D=0.2 and then decreases, indicating an increasingly uniform

distribution of total pressure at the duct exit.

8) There is an overall reduction in the magnitude of the secondary flow at the duct exit

due to a reduction in upstream secondary flow caused by a decrease in flow diffusion

upstream of the inlet throat.

9) Overall turbulent velocity and length-scales increase in the duct

1 0) A lack of flow separation within the waterjet inlet for the range of boundary layer

thicknesses examined, may be attributed to the use of a throat of circular cross-

216

section and is thus an improvement over conventional waterjet inlet geometries, such

as the waterjet inlet geometry examined in Chapter 5.

The first four points may be correlated with the reduction in upstream momentum flux

flowing across the cross-section of the inlet streamtube (with increasing boundary layer

thickness), while the sixth point may be correlated with the reduction in energy flux

across this boundary. The behaviour of the distortion coefficient is linked to the rate of

change of the boundary layer velocity distribution over the cross-section of the inlet

streamtube. The greatest rate of change of the results presented in this chapter occurs

when upstream boundary layer thicknesses are small. This is because the rate of change

of ingested momentum and energy fluxes are greatest for these values.

Therefore in conclusion, it is clear that the upstream boundary layer thickness can have a

marked influence on the hydrodynamics of a waterjet inlet, affecting such aspects as

cavitation inception and the quality of the flow (total pressure distribution and swirl

characteristics) delivered to the waterjet pump and hence the pump performance. It is

therefore necessary to design the waterjet propulsion unit to give adequate

hydrodynamic performance, not only over a range of operating IVR values, but also over

a range of possible upstream boundary layer thicknesses. This issue is particularly

relevant when installing a given waterjet propulsion unit design in different hull forms.

It also highlights the fact that the waterjetlhull system should really be designed as an

integral unit rather than separately.

217

Chapter 7 Design Subspace Investigations

The parametrically-defined generic flush-type waterjet inlet presented in Chapter 4

represents an infinite number of possible waterjet inlet geometries and, hence, flow

solutions. It would be ideal if the hydrodynamic performance parameters (\j11) presented

in Chapter 6, are known analytic functions of the waterjet inlet geometry such that

\j11=f{ X )=f{XJ,X2, ... ,X0 ) {7.1)

where X1 are the components of X , the vector of design variables and f is an explicit

function linking \j11 with X . If this were the case, investigation of the hydrodynamic

performance of the waterjet inlet would be a simple process allowing easy optimisation

of the waterjet inlet geometry. This unfortunately is not the case, as the \j11 are unknown

functions of X and hence the hydrodynamic performance of the waterjet inlet over its

design hyperspace is unknown.

Since \j11 are unknown functions of X , one approach may be to attempt to find an

analytic relationship between \j11 and X . If flow computation were fast and inexpensive,

it could be possible to compute the \j11 at several points in each direction of the design

hyperspace, attempt to fit an analytic approximation to the data (using regression

analysis) and so establish the link between \j11 and X . With m design points in each

dimension, the total number of computations required would be

(7.2)

where N is the total number of computations and n is the number of components of X . For the generic waterjet inlet geometry presented in Chapter 4, n is 8. If 5 points are

investigated in each dimension of the design hyperspace, 390625 computations would

be required. Clearly the total number of computations required quickly becomes

prohibitive for anything but the smallest values of m and n. This approach is therefore

infeasible.

218

On the other hand, investigation of a two-dimensional subspace of the design

hyperspace is a manageable problem. Since such an investigation is limited to the

variation of only two components of X , with the other components held at constant

values, the total number of computations is thus reduced to

(7.3)

Therefore, two-dimensional subspaces of the design hyperspace can be investigated and

an explicit analytic relationship between 'tf1 and X established. For a generic geometry

described by n parameters, there are

N - n! 2-

(n-2)!2! (7.4)

possible two-dimensional subspace combinations that can be investigated. For the

generic geometry presented in Chapter 4 (n=8), there are 28 possible two parameter

subspace combinations. Of particular relevance are those subspaces containing

geometric components that have the greatest effect on flow performance. This issue will

be discussed at greater length in Section 7 .2.

The focus of this chapter is dedicated to the investigation of three two-dimensional

subspaces of the design hyperspace at the vessel cruise condition (IVR=0.60) with a

boundary layer upstream of the waterjet inlet. The motivations behind this study are:

1) The large amount of data collected from such investigations allows the flow

behaviour in the waterjet inlet to be correlated with the underlying geometry via an

understanding of the physical mechanisms governing the flow behaviour.

2) The examination of the sensitivity of the flow in the waterjet inlet to variations in

parameter values.

3) To illuminate the most important hydrodynamic issues associated with inlet design.

4) The design data collected for a two-dimensional design subspace allow an

optimisation to be made in that subspace. In other words, the waterjet inlet could be

designed from a "two-parameter series".

It must be noted from the outset, that the data presented in this chapter is dependent

upon IVR and the upstream boundary layer thickness (Chapter 6) and is therefore not

219

applicable to other NR values or different upstream boundary layer profiles. Griffith­

Jones (1994) noted that the Hamilton Jet model 211 operates at a cruise NR of 0.70.

Therefore for the cruise NR of 0.60 considered here, it may be expected that suction on

the underside of the inlet lip is greater (especially in the case of no internal flow

diffusion) and hence the elimination of cavitation at this location at high vessel speeds

poses a more challenging problem.

An overview of this chapter is presented below. In Section 7.1 the methodology used to

investigate two-dimensional subspaces of the design hyperspace is discussed. The

computational modelling and simulation procedure is outlined in Section 7 .2.

Computational results are presented for the three two-parameter design subspaces

investigated in Section 7 .3. The physical mechanisms underlying the trends in flow

behaviour are investigated in detail in Section 7.4. Finally, the conclusions of the design

subspace investigations are summarised in Section 7.5.

7.1 Investigation Methodology

The basic aim of the investigation methodology presented herein is to find an explicit

relationship between \jl1 and X over an examined two-dimensional subspace of the

design hyperspace. The methodology may be summarised as:

1) CFD flow solutions are obtained for a "grid" of design points. In other words, two­

components of the vector of design variables are varied with the other components

held constant. The grid is composed of equi-spaced points in each component

direction.

2) The \jl1 are determined from the flow solutions corresponding to the respective design

points.

3) Non-linear least-squares regression is used to fit a bicubic patch surface to the data

obtained for \jl1 in order to construct an explicit relationship between \jl1 and X .

4) The correlation coefficient is calculated in order to examine whether a good

correlation has been obtained between the surface fit to the data and the data itself.

The approach to non-linear least-squares regression is described in greater detail below.

It was originally thought that a bicubic patch of the form

220

(7.5)

would provide a suitable analytic approximation to the data over the range of parameter

values considered. As will be shown later, this was indeed the case. In Eqn 7.5, u and w

represent coordinates on a unit square such that

(7.6)

Furthermore, if x1 and x2 represent any two components of X comprising the sub-space

under investigation, then u and w may be defined as

x1- X nun u = _.:______::::::...._ xmax- X nun

x2 -X nun w = ----:=--__....;;;;=--

X max -X nun

(7.7)

where subscripts min and max denote the lower and upper values of x1 and x2 on the

grid of design points investigated in that subspace.

It must be noted that the bicubic surface patch also accounts for surface fits of order less

than three and so can be used to represent biquadratic and bilinear surfaces. Non-linear

least -squares regression is used to determine the numerical coefficients of Eqn 7.5. Non­

linear least-squares regression involves the minimisation of the least square error

between '1'1 and '¥1 over the range of data examined. The least-square error between '1'1

and 'Pi may be written as

1 N 2

E = 2 L ('I'.,J- 'P.) j=l

(7.8)

where E is the least-square error and N the number of data points (N can be calculated

using Eqn 7.3).

In order to minimise the least-square error represented by Eqn 7.8, any optimisation

technique can be used. Minimisation of E is essentially an unconstrained optimisation

for the vector ofbicubic patch coefficients (ak,I) which may be written as

(7.9)

The unconstrained quasi-Newton algorithm of Liu and Nocedal (1989) is used to effect

the minimisation of E. Information relating to quasi-Newton methods can be found in

Gillet al (1981).

221

Once a is calculated, the correlation between the bicubic patch surface fit to the data

and the actual data can be calculated using the relationship for correlation coefficient

given by Eqn 6.28. A correlation coefficient of unity implies an exact match between the

surface fit to the data and the data itself, whereas correlation coefficients less than unity

imply an inexact surface fit to the data, with increasing disparity evident as the

correlation coefficient decreases in magnitude.

7.2 Computational Simulation

In this section, an overview of the CFD simulation methodology used to generate the

results of Section 7.3 is presented. The same boundary conditions, turbulence

modelling, near-wall modelling and size of external domain used for the simulations of

Chapter 6, are also used here for reasons of consistency. The boundary conditions for

velocity, turbulent kinetic energy and turbulent dissipation which are applied to

Boundary 1 are computed here using the same methodology as described in Chapter 5

and Chapter 6. The boundary layer velocity profile used in the computational

simulations to generate the results presented here is shown in Fig. 7 .1. This boundary

layer is based on a non-dimensional thickness of 0/0=0.8.

1.0

0.9

0.8

'i! ::J 0.7 ..... ::J

0.6

0.5

0.4 0 C'l (<) v lrl \0 r-- 00 0\ ~ d d d d d d d d d d

y/D

Fig. 7.1 Upstream boundary layer velocity profile

Three two-dimensional (two-parameter) design subspaces of the design hyperspace

spanned by the vector of design variables ( X ) are examined. Several key design

questions provide the motivation behind the choice of the selection of the subspaces

examined. These are:

222

1) The possibility of using a steeper inlet (larger a) than used by many conventional

designs and the effect of inlet steepness on 'lf1•

2) The effect of the radius of the inlet lip (RL) on the minimum static pressure on the lip

and other aspects of hydrodynamic performance.

3) The effect of using a raised-lip profile on the flow in the vicinity of the inlet lip and

the effect on the resulting minimum static pressure on the lip, as well as other aspects

of hydrodynamic performance.

4) The possibility of any hydrodynamic benefit resulting from an internal diffusion of

the waterjet inlet flow. In other words, achieving part of the flow diffusion between

the inlet throat and the duct exit, rather than relying on external flow diffusion

upstream of the inlet throat.

As a consequence of the above considerations, the following two-parameter design

subspaces and geometric constraints or selected:

1) Design Subspace 1 - RUD and a

2) Design Subspace 2 - y and HUD

3) Design Subspace 3 - AJ AT and RUD

0.025~UD:::;0.1 , 20o:::;a:::;35°

7.5o:::;y:::;15°, o.o5:::;Huo:::;o.1

1.0:::;AJAT:::;1.6, 0.025:::;RUD:::;0.1

The design parameter values of the generic waterjet inlet geometries used for the

purposes of two-parameter design subspace investigations are listed in Table 7.1

Geometric Parameter Symbol 1 2 3 Throat Diameter D 600mm 600mm 600mm Angle of inclination of inlet to a Varied 25° 25° Radius of inlet lip RL Varied 15mm Varied Height of pump centreline above base H 600mm 600mm 600mm Radius of curvature of centreline of duct Ro 1200mm 1200mm 1200mm Length of horizontal duct section LH 400mm 400mm 400mm Ratio of duct exit area to throat area AJAT 1.0 1.0 Varied Height of inlet lip centreline above base HL Rt.!2 Varied Rt.!2 Angle of inclination of raised lip profile 'Y oo Varied oo

Width of inlet opening - 700mm 700mm 700mm

Table 7.1 Waterjet inlet geometric data for design subspace investigations

All calculations were initiated using an initial "guess" velocity of 20.58 ms-1 in the x

223

direction throughout the flow domain, thus allowing the computations to be run as batch

processes after initial problem set-up. Once all computations have converged, the 'lf1 are

calculated and the methodology outlined in Section 7.1 is used to determine a suitable

bicubic patch fit to the data. In all cases, a correlation coefficient of unity was obtained

thus indicating a perfect match between the bicubic-patch approximation to the data and

the data itself.

In order to investigate the first design subspace (Section 7.3.1), a total of sixteen CFD

computations was executed, with four computations run in each parameter direction. For

the second design subspace (Section 7.3.2) a total of twelve computations was executed,

with four computations in the y parameter direction and three in the HL parameter

direction. A total of sixteen computations was executed for the investigation of the third

design subspace (Section 7.3.3), with four computations run in each parameter direction.

The total number of computations executed for each design subspace is deemed to be

adequate provided that the variation of the '1'1 over the design subspaces investigated is

smooth. From the results obtained in Section 7.3, this does appear to be the case.

Furthermore, the expectation will be for a smooth variation of the 'lf1 over the design

subspaces in the absence of any major flow separation effects.

If the 'lf1 are not smoothly varying over the range of parameter values comprising the

subspace under investigation, a significantly larger number of computations will be

required to resolve any potential discontinuities, as a result of the larger number of

simulations required in each parameter direction. Due to the quadratic nature ofEqn 7.3,

it is obvious that a doubling of the number of computations in each parameter direction

leads to a four-fold increase in the total number of computations required. The presence

of discontinuities also complicates the form of the analytic surface required to provide a

good correlation with the data. If discontinuitiues are suspected in the 'lf17 an alternative

approach may be to select points in the vicinity of the suspected discontinuity, determine

the values of the '1'1 and then progressively resolve the discontinuity by further addition

of points in the appropriate locations.

224

7.3 Results

The results of the three two-parameter design subspace investigations are presented in

this section, with the results of each subspace investigation presented separately as an

individual subsection. It must be stressed that the results presented in this section reflect

the boundary conditions and IVR used. Therefore arbitrary generalisation of these

results to other flow conditions associated with different upstream boundary layers

and/or IVR values cannot be undertaken. The results primarily represent a means by

which an understanding can be developed as to the physical mechanisms underlying the

observed results. This is ultimately of the greatest value, as an understanding of the

principles governing the hydrodynamic performance of a waterjet inlet allows the

evolution of better hydrodynamic design.

7.3.1 Variation of Lip Radius and Inlet Inclination

The variation of O'mm and minimum Cp over the two parameter design subspace is shown

in Fig. 7 .2a and Fig. 7 .2b respectively. In all cases examined, the minimum static

pressure is found to occur on the underside of the inlet lip. It is clear from the contour

plots that there is a trend of increasing static pressure with increasing RL and a., with

both parameters having a significant influence on the result. Using the results presented

in Appendix A.l, there is an increase in minimum Cp over the range investigated of 44%

of the minimum value, which represents a marked increase. The negative values of O'mm

over the design subspace indicate that cavitation will be present on the underside of the

inlet lip at a vessel speed of 40 knots.

In order to examine why increasing RL and a. affect the minimum static pressure on the

underside of the inlet lip, the static pressure distribution in the vicinity of the inlet is

shown in Fig. 7.3 for five cases corresponding to perturbations in RL and a. about the

point Rr/1)=0.05, a.=25°. The convention of previous chapters of measuring the non­

dimensional arc length (SID) from the lip trailing edge into the inlet is adopted here. The

relationship between SID and the underlying waterjet inlet geometries on the

centreplane is presented in Table 7 .2.

225

35 0

325

30 0

~

'<.-- 27.5 i:$

25 0

22 5

20.0 0 LO 0 LO 0 LO LO ,.._ 0 C\1 LO ,.._ C\1 C') LO <0 ,.._ CX) 0 0 0 0 0 0 ci ci ci ci ci ci

RJD

a) Minimum cavitation number

35.0

32.5

30.0

c 27.5 i:$

25 0

22.5

20.0 0 LO 0 LO 0 LO LO ,.._ 0 C\1 LO !Xi C\1 C') LO <0 ,.._ 0 0 0 0 0 0 ci ci ci 0 ci ci

RJD

b) Minimum static pressure coefficient

0 0 0 .... 0

0 g .... ci

F

E D

c B

A

9

8

7

6

5

4

3

2

F

E D c B

A 9

8

7

6

5

4

3

2

crmln

-0660

-o 714

-0769

-0824

-0878

-0933

-0.987

-1.042

-1 096

-1.151

-1 205

-1.260

-1 315

-1.369

-1.424

Min.CP -1.129

-1.183

-1.238

-1.293

-1.347

-1.402

-1 456

-1.511

-1.565

-1 620

-1.675

-1.729

-1.784

-1.838

-1.893

Fig. 7.2 Design subspace 1 - Variation of hydrodynamic performance over subspace

226

DC F 7 835E-2

E 7.769E-2

D 7 702E-2

c 7 636E-2

B 7.570E-2

A 7 504E-2

9 7 437E-2

8 7 371E-2

7 7 305E-2 c 6 7 239E-2 ts

5 7.172E-2

4 7106E-2

3 7.040E-2

2 6.974E-2

6.907E-2

225

RJD

c) Distortion coefficient

CFY F 0.2553

E 0.2383

D 0.2214

c 0.2045

B 01875

A 0.1706

9 0.1537

8 0.1368

7 0.1198

6 0.1029

5 0.0860

4 00690

3 0.0521

2 0.0352

0.0182

RJD

d) Non-dimensional lift coefficient

Fig. 7.2 (cont.)

227

e) Area-averaged total pressure recovery effi . ICiency

350-1-~---------..-11

I l

~..---~ - ........... 0 LO 0

:e ..... CXI

0 0

0 0 ci ci ci

RJD

f) Non-dimensional . waterJet inlet volume

F I g. 7.2 (cont.)

228

ll F 06723

E 06721

D 0.6718

c 06716

8 06713

A 06711

9 0.6708

8 06706

7 0.6703

6 0.6700

5 0.6698

4 0.6695

3 0.6693

2 06690

0.6688

v· F 3.774

E 3.697

D 3.619

c 3.541

B 3.464

A 3.386

9 3308

8 3.231

7 3.153

6 3075

5 2.998

4 2.920

3 2.842

2 2.765

2.687

All of the cases shown in Fig. 7.3 display the usual variation of static pressure over the

lip surface for an NR=0.60, with the stagnation point located on the upperside of the lip

highlight and the minimum static pressure located on the underside of the lip. The

magnitude and angular locations (9) of the stagnation points and points of minimum

static pressure for the five cases, are listed in Table 7.3. Angular location is measured

from the lip trailing edge.

1.00

0.50

0.00

u"'" -0.50

-1.00

-1.50

. . .

I . . . . . . . . . --- RdD=0.05 , a=25°

RdD=0.025 , a=25°

-- RdD=0.075 , a=25°

RdD=0.05 , a=20°

RdD=0.05 , a=30°

-2.00 +--'---'-------'---'--t----'----'------'----'--t-----'---'-L--....___-t---'---'---'-------'---1

0.00 0.05 0.10

SID

0.15 0.20

Fig. 7.3 Distribution of static pressure coefficient in the vicinity of the inlet lip

Rr/D 0.05 0.025 0.075 0.05 0.05

a 25° 25° 25° 20° 30° Upper Centreplane SID SID SID SID SID

Ramp tangency point 0.000 0.000 0.000 0.000 0.000 Start of inclined duct section 4.665 4.443 4.887 6.000 3.749 Start of bend 5.289 5.179 5.398 6.913 4.160 Start of straight duct section 6.380 6.270 6.489 7.785 5.469 End of straight duct section 7.047 6.937 7.156 8.452 6.136

Lower Centreplane Lip trailing edge 0.000 0.000 0.000 0.000 0.000 Start of inclined duct section 0.135 0.068 0.203 0.140 0.131 Start of bend 0.760 0.804 0.714 1.052 0.542 Start of straight duct section 1.415 1.459 1.369 1.576 1.327 End of straight duct section 2.081 2.126 2.036 2.242 1.994

Table 7.2 Relationship between SID and the underlying waterjet inlet geometry

229

A study of Fig. 7.3 and Table 7.3 reveals an increasing minimum static pressure on the

underside of the inlet lip for the cases corresponding to increasing lip radii at a=25°.

This may be partially attributed to the differences in stagnation point location, since the

peak minimum static pressures occur at approximately the same angular location on the

lip. With increasing lip radius, the stagnation point moves closer to the centre of the lip

and the flow in the vicinity of the inlet lip is turned through smaller angles, hence peak

negative pressures are reduced in magnitude. The stagnation points corresponding to the

cases of a constant lip radius of Rr./D=0.05 (RL=30 mm), are located at approximately

the same angular location. Similarly, the angular location of the static pressure minima

for these cases is also located at approximately the same angular location, yet the

magnitude of the peak negative pressure on the underside of the lip decreases with

increasing a ! This therefore suggests that the angle of incidence of the flow (relative to

the lip profile) in the immediate vicinity of the inlet lip is the factor affecting the

magnitude of the peak minimum static pressure on the lip.

Maximum Minimum

Rr./D a Cp 9 Cp 9 89 0.025 25° 0.874 123.8° -1.748 21.0° 102.8° 0.050 25° 0.860 108.0° -1.480 19.5° 88.5° 0.075 25° 0.855 103.0° -1.369 20.6° 82.4° 0.050 20° 0.861 107.6° -1.662 19.4° 88.2° 0.050 30° 0.856 108.2° -1.309 20.8° 87.4°

Table 7.3 Magnitude and location of maximum and minimum static pressures on lip

The streakline plots of Fig. 7.4 show the location of the stagnation points on the lip and

the incidence of the approaching flow. The movement of the stagnation point toward the

centre of the lip with increasing lip radius can be clearly seen from Fig 7 .4. It must be

noted that the intersection of the streakline with the lip geometry, such as in Fig. 7 .4a,

may be attributed to inaccuracies in the calculation of the streakline trajectory by the

visualisation software used. It is also evident that as a is increased the incidence of the

incoming flow is decreased. It may thus be concluded that the angular location of the

stagnation point on the inlet lip and the angle of incidence of the flow in the vicinity of

the dividing streamline have a significant effect on the magnitude of the minimum static

pressure on the inlet lip. These issues will be discussed in greater depth in Section 7 .4.

230

b) RJD=0.025 , a=25° c) RJD=0.075, a=25°

a) RJD=0.050 , a=25°

d) RJD=0.050 , a=20° e) RJD=0.050, a=30°

Fig. 7 4 Streaklines in the vicinity of the inlet lip

231

The distribution of static pressure coefficient over the surface of the upper inlet on the

plane of flow (and geometric symmetry) is shown in Fig. 7.5 for the same five cases.

The static pressure distributions are similar despite the variation in lip radius for a=25°.

This indicates that the flow in the vicinity of the inlet lip has negligible effect on the

flow in the vicinity of the upper inlet for the cases examined. Increasing a for

RJD=0.05 (RL=30 rnrn) leads to a more compact inlet and lower minimum static

pressures on the ramp as the radius of curvature of the ramp is decreased. The minimum

static pressure on the inside of the duct bend also decreases with increasing a (and

hence increased bend angle), as shown in Fig. 7 .6.

0.40

0.30

0.20 /

/ /

I

'a=25° / RdD=0.05 "" I

u 0.10 ------ RdD=0.025 , a=25° I /

/ RdD=0.075 , a=25° 0.00 /

/ /

RdD=0.05 'a=20° /

-0.10 -- RdD=0.05 'a=30°

-0.20

0 2 4 6 8 10

SID

Fig. 7.5 Distribution of static pressure coefficient on upper inlet at centreplane

0.50

0.45

0.40

0.35

0.30

u"" 0.25

0.20

0.15

0.10

0.05

0.00

0.20 0.45

' ' ' ' \ \

RL/D=0.05 ' a=25°

RdD=0.025 , a=25°

RdD=0.075 , a=25°

RdD=0.05 , a=20°

RdD=0.05 , a=~3~0°_..~~-:...--

0.70 0.95 1.20 1.45 1.70 1.95 2.20 2.45

SID

Fig. 7.6 Distribution of static pressure coefficient on lower inlet surface at centreplane

232

The behaviour of the distortion coefficient reveals a general trend toward a more

uniform distribution of total pressure at the duct exit with increasing RL and a., as can be

seen from Fig. 7 .2c. The distribution of total pressure over the cross-section of the duct

exit is shown in Fig. 7.7 for five cases. Noting the results of the boundary layer

investigations of Chapter 6, in particular the total pressure distributions of Fig. 6.14, it is

apparent that the total pressure distributions here are primarily influenced by the

ingested energy flux from the boundary layer.

The distortion coefficient varies by 11.8% of the minimum value (see Appendix A.1)

over the range of computed values. This is not a large variation and may be attributed to

the boundary layer and secondary flow development within the waterjet inlet. Fig. 7.8

shows that increasing RL does not appear to have much effect on the magnitude of the

secondary flow at the duct exit. Therefore, the differences in uniformity of flow at the

duct exit are caused by differences in upstream streamwise boundary layer development.

Increasing a. results in an increase in the magnitude of the secondary flow, this being

evident from the flow secondary close to the duct wall as can be seen from Fig. 7 .8. The

increased secondary flow results from greater cross-stream pressure gradients caused by

larger bend angles. The secondary flow thus acts to convect low-momentum fluid at the

top and sides of the duct toward the bottom of the duct and so leads to a slight increase

in the uniformity of the total pressure distribution across the cross-section of the duct

exit. This secondary-flow-driven convection is evident from Fig. 7.7e which shows a

greater build up of low momentum fluid at the bottom of the duct and higher total

pressures at the top of the duct when compared to Fig. 7.7d.

It is interesting to examine how the vertical forces on the waterjet inlet change over the

design subspace. As can be seen from Fig. 7 .2d, there is a clear trend of increasing CFY

with increasing RL and a.. The observed trend may be partially attributed to the change

in vertical momentum flux through the waterjet inlet. Discussion of this issue is

reserved until Section 7 .4.

The plots of the results for area-averaged 11 shown in Fig. 7 .2e, reveal some interesting

points. An average magnitude of 11 of 0.67 over the two-parameter design subspace

233

b) RJD=0.025 , a=25° c) RJD=0.075, a=25°

a) RJD=0.05 , a=25°

d) RJD=0.05 , a=20° e) RJD=0.05, a=30°

Fig. 7. 7 Distribution of total pressure coefficient over the cross-section of the duct exit

234

I I I I

I I I I I I

111 I \I I II II

1111 II I I 11 I I

""1 \ II I I Ill I I I I IIIII') Ill Ill II I~ ~

uul11 ))} I/ lfllfll , , / I

11 1 I I I 1111 11 I I

I I I I I I I II I I I

I

I

- =1 m/s

I '

I '

I I I I I I I I I I //

I I I / //, I I , ;,:::;

I I II ~- ~ I I ' /. I I 1 -,_:::........:::;

.:::-«

b) Rr/1)=0.025 , a=25°

I I I

I 1 I I\ I

II II \I \I \

11 II \I I 1\\ I II

I I

I

11111 II II I 111 II I I

) ) ) ) ! ' II Ifill , , I

II I I I 1111, ,I II /

I till II I

I I I II II

I I I I I I

I I I

- =1 m/s

d) Rr/1)=0.05 , a=20°

I I I \

I 1 I\

I 1 11 \1

I\ II II II I II II I I

II II I 1111 1 11 I I 1

111 11 I I ""Ill I) ) I I~ :

,1) ~~ ) I : 1111111 / , / I

-=1m/s

I '

I '

I '

I '

1l 1 I I I 11 11 11 I I 1

II I II II I I II I

I I I I I I I I I I /

1 I I I 1 _, ~ I I I I I , ;..:::;

I 1 - h I I ' /. I

a) Rr/1)=0.05 , a=25°

I I

I I I 1

It I l

I\ I I I I I

'• II II 111

11 I I

II I I

I

'"1' II II I 1111 11 I II I

)) ) ) ; ! lflllll / , / I

1l 1 I I I 11 11 11 I I 1

111 I 1 1 II II I

I I I I I I I

I I I " I I I I' ~ h I I I - h

I ' /.

', ~ .&ii --""

-=1m/s

c) Rr/1)=0.075, a=25°

I I I I

I I 1 I I

It I I I 11 I

It I II I II II II II I

1111 1 11 I I ,,1 II I

""'))Ill II ,, II

111111 I ) ) j Ill I I

lfl 1 I I I 1111 II I

Ill I II

1111 II I I I I

I I I I I I I I I

I I I I I I I I I 1 I I I ~ ~-.

- =1 m/s

e) Rr/1)=0.05, a=30°

Fig. 7.8 Secondary flow at the duct exit

235

indicates that the effect of boundary layer ingestion is the primary source of total

pressure loss (see Chapter 6). The dimensions and shape of the cross-section of the inlet

streamtubes, shown in Fig. 7.9 for the five cases examined in detail, are essentially

identical. This suggests that the dimensions of the cross-section of the inlet streamtube

are governed by the width of the waterjet inlet, the mass flow-rate through the inlet and

the upstream boundary layer. Since the widths of the waterjet inlet geometries examined

here are identical and since the same upstream boundary layer is used as a boundary

condition for all simulations, it may be concluded that the energy flux coefficient (Ce) is

the same for all cases. Ce is calculated as 0.708 using Eqn 6.27. Furthermore, the lack of

significant variation in 11 over the range of RL and a suggests that viscous losses within

the waterjet inlet are essentially identical for the different geometries examined.

It can be seen from Fig. 7.2f that increasing a and decreasing RL results in a smaller,

more compact waterjet inlet, as would be expected. A decrease in volume of 27% is

achievable by moving from Rr./1)=0.1, a=20° to Rr./1)=0.05, a=35°. This represents a

dramatic decrease in waterjet inlet volume and therefore highlights the necessity of

ensuring the absence of flow separation on the inlet ramp, in order to achieve a more

compact waterjet inlet design.

0 0

0.00

-0.05

-0.10

-0.15 e -0.20 >.

-0.25

-0.30

-0.35

-0.40

RdD=0.05 , a=25°

- - - - - - RdD=0.025 , a.=25°

RdD=0.075 , a.=25° ------

RdD=0.05 , a.=20° ---

R,/0=0.05 , a=30°

00 0

ziD

Fig. 7.9 Cross-section of the inlet streamtube

236

7.3.2 Variation of Lip Profile

The variation of O'mm and minimum Cp over the two-parameter design subspace is

shown in Fig. 7.10a and Fig. 7.10b respectively. As was noted in Section 7.3.1 for all

cases examined, the minimum static pressure is found to occur on the underside of the

inlet lip. It is clear from the contour plots that there is a trend of increasing static

pressure with increasing y and HL, with both parameters having a significant influence

on the results obtained. Using the results presented in Appendix A.2, there is an increase

in the minimum Cp over the range investigated of 57.6% of the minimum value. This

represents a large increase. In fact, a minimum Cp of -1.33 represents a marked increase

over a minimum Cp of -1.75 for a lip radius of 15 mm but without the raised-lip profile.

The benefits of using a raised lip profile, are therefore clearly evident. This further

highlights the importance that must be given to the design of the inlet lip. The negative

values of O'mm over the two-parameter design space indicate that cavitation will be

present on the underside of the inlet lip at a vessel speed of 40 knots, although the

maximum value of O"mm of -0.096 is close to a cavitation-free operation.

In order to examine how changing the shape of the lip profile (by increasing y and HL)

affects the minimum static pressure on the underside of the inlet lip, the static pressure

distribution in the vicinity of the inlet region is shown in Fig. 7.11 for five cases

corresponding to perturbations in y and HL about the point y=10°, Hr./D=0.075. The

relationship between SID and the underlying waterjet inlet geometry on the upper and

lower centreplane locations is presented in Table 7 .4. The trend of increasing minimum

static pressure on the underside of the inlet lip, with increasing y and HL can clearly be

seen from the pressure distributions of Fig. 7 .11. It is interesting to note that there is

relatively little variation in Cp over most of the inclined surface downstream of the lip,

except in the immediate vicinity of the lip itself, or at the "inlet/hull" interface. The

increasingly sharp decrease in Cp at the inlet/hull interface with increasing y, may be

attributed to the sudden change in geometry at this location.

237

crmln 0100 F -o 146

E -0197 0095

D -0247

0090 c -0297

B -0348

0085 A -Q398

9 -0.448

0080 8 -0499

7 -0549 0

0.075 ~ 6 -Q599 :c 5 -o.650

0070 4 -0.700

3 -0750 0.065 2 -0 801

-0.851 0.060

0.055

0050 0 LO ,...:

a) Minimum cavitation number

Min. CP 0100 F -0 616

E -Q.666 0.095

D -0.716

0090 c -0.766 B -o 817

0.085 A -0.867

9 -0.917

0.080 8 -0.968

7 -1.018 0

0.075 ~ 6 -1.068 :c 5 -1.119

0.070 4 -1 169

3 -1.219 0065 2 -1.270

1 -1.320 0.060

0.055

0.050 0 LO ,...:

b) Minimum static pressure coefficient

Fig. 7.10 Design subspace 2 - Variation of hydrodynamic performance over subspace

238

0 ~ J:

0 ~ J:

c) Distortion coefficient

d) Non-dimensional lift coeff" . tctent

Fig. 7.10 (cont.)

239

F

E

D

c B

A

9

8

7

6

5

4

3

2

F

E

D

c B A

9

8

7

6

5

4

3

2

DC 7 418E-2

7 376E-2

7.334E-2

7.292E-2

7 250E-2

7 208E-2

7166E-2

7124E-2

7 082E-2

7.040E-2

6 998E-2

6.956E-2

6 913E-2

6 871E-2

6 829E-2

CFY 01710

0.1629

01549

01468

0.1388

01308

0.1227

0.1147

0.1066

0.0986

0.0905

0.0825

00744

0.0664

0.0584

11 0100 F 06718

E 06715 0095

D 06713

0090 c 06710

B 06708

0085 A 06705

9 06702

0080 8 06700

7 06697 Cl

6 0.6694 -:.. 0075 :c 5 0.6692

4 06689

3 0.6686

2 06684

1 0.6681

y (0)

e) Area-averaged total pressure recovery efficiency

v . 0.100 F 3.262

E 3.256 0.095

D 3.250

0090 c 3 244

8 3.238

0.085 A 3.232

9 3.226

0.080 8 3.220

7 3.214 Cl

6 3208 -:.. 0.075 :c 5 3.202

4 3.195

3 3189

2 3.183

1 3177

y (0)

f) Non-dimensional waterjet inlet volume

Fig. 7.10 (cont.)

240

1.00

0.50

0.00 c. u

-0.50

-1.00

-1.50

0.00

'

0.10

' I \ I

I

I I I I I I

I I I I

1/

/\ I \ I ,_

I I I I I I I

0.20 0.30

SID

.. .. ..

0.40

-- y=l0° 'HdD=0.075

· · • ·- • y=7.5° , HdD=0.075

-- y=12.5°, HdD=0.075

- - - - - - y= 10° , HdD=0.025

-- y=10° , HdD=O.lO

0.50 0.60

Fig. 7.11 Distribution of static pressure coefficient in the vicinity of the inlet lip

'Y 100 7.50 12.5° 100 100

HJD 0.075 0.075 0.075 0.05 0.1 Upper Centreplane SID SID SID SID SID

Ramp tangency point 0.000 0.000 0.000 0.000 0.000 Start of inclined duct section 4.676 4.676 4.676 4.559 4.792 Start of bend 5.294 5.294 5.294 5.237 5.352 Start of straight duct section 6.385 6.385 6.385 6.328 6.443 End of straight duct section 7.051 7.051 7.051 6.995 7.110

Lower Centreplane Trailing edge of lip profile 0.000 0.000 0.000 0.000 0.000 Start of lip 0.290 0.385 0.234 0.146 0.434 Start of inclined duct section 0.353 0.449 0.296 0.209 0.497 Start of bend 0.972 1.067 0.914 0.888 1.057 Start of straight duct section 1.626 1.722 1.569 1.542 1.712 End of straight duct section 2.293 2.389 2.236 2.209 2.379

Table 7.4 Relationship between SID and the underlying waterjet inlet geometry

If y is sufficiently large, the minimum static pressure on the surface of the inlet would

occur at the inlet/hull interface, with this location becoming a potential source of

cavitation. Therefore, a limitation of the author's generic geometry has been exposed

and a suitable change in profile is required in order to ensure a smoother transition from

the inlet to the hull. For all intents and purposes, the existing parametric geometry is

241

adequate, provided that the static pressure at the inlet/hull interface is high enough to

avoid the inception of cavitation. It must be noted that the results for CJmm and minimum

Cp presented in Fig. 7.10 are based on the minimum static pressure on the surface of the

waterjet excluding the inlet/hull interface. This is because the primary focus of the

design subspace investigations presented here are based on identifying and

understanding the effect of the geometry on the flow in the region of the inlet lip and

how this flow can be improved.

The magnitude and angular locations (8) of the stagnation points and points of minimum

static pressure on the inlet lip, for the five cases examined in Fig. 7.11, are listed in

Table 7.5. Angular location is measured from the vertical, with 8 increasing in the

clockwise direction. It must be noted in the interpretation of the results of Table 7.5, that

the values for 8 quoted are calculated using nodal values of Cp and so reflect the

discretisation of the geometry of the waterjet inlet. Hence, the listed values of Cp and

their angular locations may not exactly correspond to the actual maximum or minimum

static pressures. There is thus a small amount of error introduced.

Maximum Minimum

'Y Hr/D Cp 8 Cp 8 ~8

7.5° 0.075 0.852 111.9° -1.010 23.1° 88.8° 10.0° 0.075 0.860 111.0° -0.939 18.8° 92.2° 12.5° 0.075 0.835 113.1° -0.784 17.9° 95.2° 10.0° 0.05 0.857 112.6° -1.216 20.6° 92.0° 10.0° 0.10 0.836 108.5° -0.728 17.5° 91.0°

Table 7.5 Magnitude and location of maximum and minimum static pressures on lip

From the results of Table 7.5, it can be seen that there are only minor changes in the

angular location of the stagnation points with increased y for Hr/D=0.075. The angular

locations of the static pressure minima are within 1 ° of each other for y= 1 oo and

y=-12.5°, but the result for y=7.5° appear to be somewhat anomalous. There also appears

to be little variation in the locations of the stagnation point and static pressure minima

with increasing HL for y= 10°. These results suggest that the primary mechanism for the

changes in minimum static pressure with changes in both parameters are due to the way

in which the raised lip profile affects the incidence of the upstream flow in the vicinity

242

of the inlet lip. The effect of the profile of the lip on the flow in the vicinity of the inlet

lip (with changes in y and HL) can be seen from the streakline plots of Fig. 7.12. Fig.

7.12 shows how a lip profile with increasing values of y and HL appear to direct the flow

upstream of the lip in such a way as to minimise the perturbation of the lip on the

surrounding flow. This whole issue will be discussed at greater length in Section 7 .4.

The distribution of static pressure coefficient over the surface of the upper inlet on the

plane of flow symmetry is shown in Fig. 7.13 for the above five cases. The static

pressure distributions are essentially identical, with slight variations being due to the

small differences in geometry. It is therefore evident that the profile of the inlet lip has a

negligible influence on the pressure distribution in the upper part of the waterjet inlet,

for the cases examined here. This indicates that the profile of the inlet lip region affects

only the flow in its vicinity. The profile of the inlet lip does not appear to affect the

minimum static pressure on the inside of the duct bend either, as can be seen from Fig.

7.14.

It is difficult to discern any clear trends in the behaviour of the distortion coefficient

from the contour plot of Fig. 7.10c. There does appear, however, to be a region of

minimum distortion centred around "(=12.75°, HJD=0.90. The distortion coefficient

varies by 8.1% over the two parameter design space, but this is not a large variation. It is

therefore thought that the distortion coefficient is influenced by the streamwise

boundary layer development within the waterjet inlet, resulting from the geometry of the

inlet region.

An examination of Fig.7.10d reveals a clear trend in the non-dimensional lift

coefficient. Increasing either y or HL results in a greater vertical force on the waterjet

inlet, although the trend tends to be more pronounced in the HL direction. As was found

to be the case for the first design subspace examined in Section 7.3.1, the trend in Cpy

may be attributed to the change in the vertical momentum of flow through the waterjet

propulsion unit. Again, a detailed discussion of this issue is reserved until Section 7 .4.

The plot of area-averaged 11 is shown in Fig. 7 .lOe. There is a variation of less than

0.3% about an average-value of 11 of0.67. In Section 7.3.1 it is shown that the ingestion

243

1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1

of boundary layer fluid is the primary source of total pressure loss, with Ce being the

same for all cases. These results apply here and hence it may be concluded that the

viscous losses within the different waterjet inlets are essentially identical, for all intents

and purposes.

c. u

0.40

0.30

0.20

y=100 ,HdD=0.075 0.10

· · · • · • y=7 .5° , HdD=0.075

0.00 y=12.S0, HdD=0.075 ------ y=10° ,HdD=0.025

-0.10 y=10° ,HdD=0.10

-0.20

0 2 3 4 5 6 7

SID

Fig. 7.13 Distribution of static pressure coefficient on upper waterjet inlet at centreplane

0.50

0.45

0.40

0.35

0.30

uc. 0.25

0.20

0.15

0.10

0.05

--y=10° ,HdD=0.075

------ y=7S 'HLID=0.075

--- y=12.5°' HLID=0.075

------ y=10° , HdD=0.025

--y=10°

0.00 -j--'----L.--'--'------t-----'---'-----'---'--+---'----'---'-~--t~-'--'---'---1

0.60 1.10 1.60

SID

2.10 2.60

Fig. 7.14 Static pressure coefficient on lower waterjet inlet at centreplane

245

8

It can be seen from Fig. 7.1 Of that increasing HL increases the volume of the waterjet

inlet, as would be expected. Similarly, one would expect that increasing y would result

in a decrease in duct volume. This, however, is not the case, as Fig. 7.1 Of reveals an

oscillatory behaviour for waterjet inlet volume as y is varied at constant HU:O. The

variation of waterjet inlet volume with y is caused by the transition of the geometry (in

the vicinity of the inlet lip) from the centreplane to the side of the inlet. This is therefore

a direct consequence of the transfinite interpolation surface mesh generation process

outlined in Chapter 4.

7.3.3 Variation of Lip Radius with Internal Diffusion

The variation of O"mm and minimum Cp over the two-parameter design subspace is

shown in Fig. 7.15a and Fig. 7.15b respectively. For all cases examined, the minimum

static pressure occurs on the inlet lip. It is evident from both plots that AJ AT has the

greatest effect on the minimum static pressure with RL having only a minor influence.

There is a trend of increasing minimum static pressure with AJAT up to AJAT=l.38,

when the trend reverses. Using the results presented in Appendix A.3, there is an

increase in minimum Cp of 72% of the minimum value, over the range of values

investigated. This therefore represents a large increase.

The negative values of O"mm obtained for the design subspace indicate the presence of

cavitation at a vessel speed of 40 knots, although a maximum tabulated value of O"mm of

-0.022 (see Appendix A.3) is close to a cavitation-free condition.

In order to explain the behaviour of the minimum static pressure on the inlet lip, it is

necessary to examine the flow in the vicinity of the inlet lip. As noted above, since

AJ AT has the greatest effect on the minimum Cp, attention will therefore be focused on

establishing a link between AJ AT and the minimum Cp on the lip. It will be shown

below that this primarily results from a movement of the location of the stagnation point

and the change in the incidence of the flow in the lip region. For the design space

investigations presented in this chapter, an inlet throat diameter of 600 mm is used for

all waterjet inlet geometries and the IVR is held constant at 0.6. In Chapter 1, IVR was

defined as the ratio of the velocity at the duct exit to the free-stream velocity. Therefore

246

crmln 0.100 F -0102 0095 E -0184

0090 D -0.265

0.085 c -0347

B -0.428 0.080 A -0 510 0.075 9 -0.591

0.070 8 -Q672

0 0.065 "":...

7 -0754

6 -0.835 a: 0.060 5 -0.917

0055 4 -0.998

0050 3 -1 080

0.045 2 -1.161

-1 242 0.040

0.035

0.030

0.025 0

a) Minimum cavitation number

Min.CP 0.100 F -0.571 0095 E -o.653

0.090 D -0.734

0.085 c -0.816

B -Q897 0.080 A -0.979 0.075 9 -1.060

0070 8 -1 141

0 0.065 "":...

7 -1.223

6 -1.304 a: 0.060 5 -1 386

0.055 4 -1 467

0.050 3 -1.549

0.045 2 -1 630 -1.712

0.040

0.035

0030

0.025 q

AJAr

b) Minimum static pressure coefficient

Fig. 7.15 Design subspace 3 - Variation of hydrodynamic performance over subspace

247

0 ~ a:

0100

0095

0090

0085

0 080

0075

0.070

0065

0.060

0.055

0050

0045

0040

0035

0.030

0025 ~ .....

AJAr

c) Distortion coefficient

0.025 I:.I....L.J.J..L...I-LU...LL...L..I.....IL....I.....L....I.L....L.J..I..-11.....1.-A......L-'--L-L...<I'--I-I......I...o.::t........I_..L.-1-..L....:I

~ ..... ..... .....

d) Non-dimensional lift coefficient

Fig. 7.15 (cont.)

248

DC F 1 021E-1

E 1 OOOE-1

D 9.800E-2

C 9 597E-2

B 9.394E-2

A 9191E-2

9 8.988E-2

8 8 785E-2

7 8.582E-2

6 8.379E-2

5 8.176E-2

4 7 973E-2

3 7.770E-2

2 7 567E-2

7.364E-2

CFY F 0 1538

E 0 1264

D 00990

c 0.0715

B 0.0441

A 00167

9 -O.D108

8 -0.0382

7 -0.0656

6 -0.0931

5 -0.1205

4 -0.1479

3 -0 1754

2 -0.2028

1 -0.2302

0025~~~ 0 ~ ~ M V ~ ~ ~ ~ ~ ~ ~ ~ ~

AjAr

e) Area-averaged total pressure recovery efficiency

0.025~~~

~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~

AjAr

f) Non-dimensional waterjet inlet volume

Fig. 7.15 (cont.)

249

increasing AJ AT results in a greater mass flow-rate through the waterjet inlet and larger

velocities throughout the waterjet inlet. Hence, it is more appropriate to base NR on the

volumetrically-averaged velocity through the throat of the inlet (rather than the duct

exit), when internal diffusion is present within the duct. This new definition of NR may

be denoted as NRT. Increasing AJAT must necessarily increase NRT as shown in Fig.

7.16. IVRT thus varies from 0.6 at an AofAT of unity to 0.96 at an AofAT of 1.6.

1.00

0.95

0.90

0.85

!! 0.80 .......

0.75

0.70

0.65

0.60

1.0 1.1 1.2 1.4 1.5 1.6

Fig. 7.16 Variation of throat inlet velocity ratio with internal diffusion

The magnitude and angular locations of the stagnation points and points of minimum

static pressure on the inlet lip are listed in Table 7.6 for five cases corresponding to the

point AofAT=l.2, Rr/D=30 mm and perturbations about this point. It is clear that the

stagnation point moves from the upperside of the lip, to the centre of the lip and then to

the underside of the lip as NRT is increased. This follows the general trend of Chapter

5, for increasing NR. The streakline plots of Fig. 7.17 show that as AJAT (and NRT) is

increased up to a value of 1.4, the flow becomes more symmetric about the centre of the

lip. This is also evident from Fig. 7 .18, which shows the distributions of Cp in the

vicinity of the inlet lip for the five cases examined. The relationship between SID and

the underlying geometry of the waterjet inlet is tabulated in Table 7.7. It can be seen

from Fig. 7.18 that as NRT increases, the minimum static pressure on the underside of

the inlet lip increases, whereas the minimum static pressure on the upperside of the inlet

lip decreases. The profiles appear to become more symmetric about the arc length

corresponding to 8=(1t-a)/2. The profile ofCp for AJAT=l.4 is almost symmetric about

250

c) AJAT=l.4, Rt/D=0.050

a) AJAT=l.2, Rt/D=0.050

e) AJAr-1.2, Rt/D=0.075

Fig. 7.17 Streaklines in the vicinity of the inlet lip

251

the arc length corresponding to (1t-a)/2, (S/0=0.068). The minimum static pressure in

this case occurs on the upperside of the lip. Furthermore, this result is close to the

maximum, minimum Cp, of the design subspace. This suggests a clear link between a

symmetric distribution of pressure about the half-angle of the lip profile and the

minimum static pressure. Further discussion of this issue is reserved until Section 7 .4.

Maximum Minimum

AJAr Rr}D Cp e Cp e L\8 1.0 0.050 0.860 108.0° -1.480 19.5° 88.5° 1.2 0.050 0.888 92.5° -1.011 14.4° 78.1° 1.4 0.050 0.895 78.2° -0.622 150.0° -71.8° 1.2 0.025 0.905 101.7° -0.982 14.9° 86.8° 1.2 0.075 0.876 92.7° -1.011 14.5° 78.2°

Table 7.6 Magnitude and location of maximum and minimum static pressures on lip

1.00

0.50 I

I

0.00 c..

u --- AofAT=l.2 'RLID=0.05 -0.50 - - - - - - AJAT=1.0 ' RLID=0.05

--- AofArl.4 , RLID=0.05 -1.00 ------ AJAT=l.2, RJD=0.025 . , , . , ·. .. -1.50

-- AJArl.2, RLID=0.075

0.00 0.05 0.10 0.15 0.20 0.25

SID Fig

. 7.18 Distribution of static pressure coefficient in the vicinity of the inlet lip

It is noted above that for AJAr> 1.38, the minimum Cp at the lip begins to decrease,

reversing the trend of increasing minimum Cp with AJAr. This trend reversal is caused

by a continued downward movement of the stagnation point, with a growing asymmetry

in the pressure distribution about the angular bisector of the lip profile. The minimum

static pressure will thus occur on the upperside of the inlet lip, with an increasing

minimum static pressure on the underside of the inlet lip.

252

A/AT 1.2 1.0 1.4 1.2 1.2 RUD 0.050 0.050 0.050 0.025 0.075

Upper Centreplane SID SID SID SID SID Ramp tangency point 0.000 0.000 0.000 0.000 0.000 Start of inclined duct I ramp 4.665 4.665 4.665 4.443 4.887 Start of straight duct section 6.380 6.391 6.402 6.287 6.500 End of straight duct section 7.047 7.057 7.069 6.953 7.167

Lower Centreplane Lip trailing edge 0.000 0.000 0.000 0.000 0.000 Start of inclined duct I ramp 0.135 0.135 0.135 0.068 0.203 Start of straight duct section 1.415 1.412 1.404 1.457 1.368 End of straight duct section 2.081 2.079 2.071 2.124 2.035

Table 7.7 Relationship between SID and the underlying waterjet inlet geometry

Fig. 7.19 shows the distribution of static pressure coefficient over the surface of the

upper waterjet inlet on the plane of flow symmetry for the above five cases. Although

the static pressure at the duct exit is similar for all cases, the maximum static pressure at

the outside of the duct bend decreases with increasing internal flow diffusion. Indeed, as

AJ AT is increased, static pressures in the upper waterjet inlet fall as a result of higher

velocities in the waterjet inlet. The same trend is also noticeable in Fig. 7.20 from the

static pressure distributions over the lower surface of the waterjet inlet at the plane of

flow symmetry.

0.40

0.30

0.20

c. 0.10 u

0.00

-0.10 -----------0.20

0 2 3

I

4

SID

5

-- AJAT=l.2 , RJD=0.05

- - - - - - AJAT=l.O , RJD=0.05 -- AJArl.4, RJD=0.05 ______ AJAT=l.2, RJD=0.025

__ AJAT=l.2 , RdD=0.075

6 7 8

Fig. 7.19 Distribution of static pressure coefficient on upper waterjet inlet at centreplane

253

0.50

0.40 . . . . . 0.30

0.20 c. u

0.10

0.00

-0.10

-0.20

0.25

· .. . ' "•, .

', ' ' . '\ '·.. .. \ .. ·- -· -·.

\

0.75 1.25

SID

--- AJAT=l.2 , RdD=0.05 --- ·-- AJArl.O, RdD=0.05 --- AJAT=l.4, RdD=0.05 ------ A0 /AT=l.2, RdD=0.025 __ AJAT=l.2, RdD=0.075

1.75 2.25

Fig. 7.20 Distribution of static pressure coefficient on lower waterjet inlet at centreplane

As can be seen from Fig. 7.19, the distinct kinks in the curves, located at the inlet throat,

for AJAT values of 1.2 and 1.4 are caused by a lack of derivative continuity between the

ramp and the diffusing duct section of the waterjet inlet. This is a direct consequence of

the linear variation of duct area with centreline arc length. It is thus necessary to select a

more suitable polynomial profile for the cross-sectional area variation of the diffusing

duct. In other words, one that allows a matching of derivatives between the inlet section

and the diffusing duct section.

Fig. 7.15c shows a general increase in distortion coefficient with increasing AJAT and

only a very minor variation of De with RL. Hence, De is essentially a function of AJAT.

The increase in total pressure distortion at the duct exit is large, with De increasing by

37.3% of its minimum value over the design subspace. The increased distortion of total

pressure at the duct exit is evident from an examination of the total pressure distribution

at the duct exit shown in Fig. 7 .21. It can be seen from Fig. 7.21 that the increasing

internal diffusion within the duct leads to a dramatic thickening of the boundary layer at

the bottom of the duct, due to an increasingly adverse pressure gradient and the

accumulation of boundary layer fluid at the bottom of the duct as a result of increased

secondary flow. At AJAT=l.6 it was found that the flow in the bottom of the duct had

separated.

254

c) AJAT=1.4, RrJD=0.05

Fig. 7.21 Distribution of total pressure coefficient over the cross-section of the duct exit

255

Fig. 7.22 shows the secondary flow at the duct exit for the five cases examined. The

effect of flow diffusion is evident from the increased sideward component of the

secondary flow vectors. With increasing AJ AT the secondary flow is significantly

intensified as a consequence of greater cross-stream pressure gradients associated with

the duct bend. For AJAT values greater than unity, the outer radius of the bend is

increased relative to the outer radius of the bend at an AJAT of unity, whereas the inner

radius is decreased relative to its value at an AJ AT of unity. These changes in bend radii

are the fundamental cause for the increase in cross-stream pressure gradients associated

with the duct bend and hence the stronger secondary flow at the duct exit.

The thickening of the boundary layer at the bottom of the duct partially accounts for the

increase in total pressure distortion. The other factor accounting for the increased

distortion is related to the quality of the flow ingested by the inlet. An increase in AJAT,

results in an increased mass flux across the cross-section of the inlet streamtube and

hence the depth of the cross-section grows while there is only a slight increase in its

width, as can be seen for the three cases examined in Fig. 7.23. This results in a greater

variation of velocity across the depth of the cross-section of the inlet streamtube and

hence the ingestion of a less uniform flow. This causes a greater distortion of total

pressure at the duct exit. This is a corollary of the argument put forward in Chapter 6

(c.f. Section 6.4), explaining the reduction of De with the ingestion of fluid from

upstream boundary layers of increasing thickness.

There is a clear trend in CFY with RL or AJ AT. As RL is increases CFY is also increased

over the design subspace. This follows the same trend as was noted for the design

subspace investigated in Section 7.3.1. Increasing AJAT leads to a "suck-down" effect

where CFY becomes negative and then increasingly negative, although the rate of change

of CFY, d(CFY)Id(AJAT), can be seen to decrease with increasing AJAT. Increasing

AJ AT results in a greater mass flow-rate through the waterjet inlet. This in effect

increases the velocity throughout the waterjet inlet, thus leading to lower static pressures

on the duct surface, as is evident from an examination of Fig. 7.19 and Fig. 7 .20. Thus a

suck-down effect results for sufficiently largeAJAT.

256

- =1 m/s

I I

I I I I

I I I I I I I

I I I I I I I I I I

111 II I I 11 1 1 I I

IIIII Ill II II I I ' 11 1 11 I I 1

)));'') i: 1111

111 11 I / I

1111

11 I I 111 II 1 1 1

I I 1 1 1111 I

1 1 I I

I I I I I I I I I I I I I I

I I

I I

I I I 1 I

I I I I I I

I I 1 I 1 I

- =1 m/s

I I I I II I II I I -'~

II 1 I 11 I I 1 ' I 11 1 I I , - '\l

lilrrr Ill II II I I II I I , _,~

IIIII) 1)111 Ill ~ II I~

1111111 II ) )l ; : '\\ ""II 1/ I / : : _- ''I 1111 II" / I '- 'll I I I I II / /- 'It

1 111 II/-.,, I I t- 1 /A Ill 1,.1

I I I I I I I I I I /,

I I I I ' /

I I " I ' ' _;;.dii l~-

I I

I I I I

I 1 I I I I I

I I II I II I I 1 I

111 11

I I ''I I I

.... ,, \ ,, II

- =1 m/S

)\, II I

·;;,))}), ,,

1111

" I I I 111

11 I I 1/ I I I

I 1 1 1 1 I 1

I I I I I II I

I I I I II ~ 1 I I I 1 , ~

I 1 1

1 I \ :_ ~

I I \ -.....::;:

~~~ ---a) AJAr=l.2, Rr/D=0.05

Fig. 7.22 Secondary flow at duct exit

257

-=1m/s

I I I I ,' I I I I I I I

I I I I I I I I I I 11 11 1 1 I I 1 .--

1 1 I I I --IIIII II 11 11 1

1

II I I _, -~ 111111 II II I I /

IIIIIIIII;IIIJIJ~ )\ /~ ~ j / / --<

""'''l /l ; : / IIIII ,, I / -' J1 till ;I I / _, - ' \

I I / _, \ ,,,,, / / / -- '1

;I / _, ' l tll /---1

I / - 'L I I I I /, - 'I

I/ - ' I•

~j/"_--:_tt I I ' _, .I­I ' -~-_.,.;;

c) AJAr= 1.4 , Rr}D=0.05

I

I I I

, I

I

I I I I II II I

II I I II II 'rl I

I Iii If II II II

- =1 m/s

""'')'lilt Ill II I

)

~ I I

!IIIII Iff/ ) jl j I , ''I 1111 II I I -,\1

11/ / I I ' -I I I II I I , 'lJ IIIII / I '--'11

II I ' 1 II I I I I I - IJ

I ,, / -I'~ I JII 1,1~

I I , IJ. I I I I ~;Zoi

I I • I £IIi I I I I ... ~ I ( , _;;~ l --

0 c:i

00 c:i

0. 00 +--r-..-r-.+-r-r-r-r--t--T---r-o---.--t-.---r-.---.-h-,--,....--1-r-..-r+-r-r-r-r-+-r--,-,-..-+.,..,.-.,..,.-+-r..-r~

-0.05

-0.10

-0.15

-0.20

~ -0.25

-0.30

-0.35

-0.40

-0.45

-0.50

--- A0/AT=l.0, RdD=0.05

AJAT=l.2 , RdD=0.05 ___ AJAT=l.4 , RdD=0.05

ziD

Fig. 7.23 Variation of the cross-section of the inlet streamtube with internal diffusion

The variation of 11 over the two-parameter design space is shown in Fig. 7.15e. Unlike

the other two design subspaces investigated, there is a clear trend of increasing 11 with

increasing Aol AT and RL, up to AJ AT= 1.48 when the trend begins to reverse. The

greatest influence on 11 is clearly due to AoiAT. The variation of 11 over the design

subspace is approximately 3.5% of the minimum value. This variation is relatively

small, although significantly larger than the variations over the other two design

subspaces. There are essentially two conflicting factors which result in the above­

mentioned trend. The first is due to the nature of the ingested flow and its effect on Ce.

For example, Ce increases from 0.708 to 0.744 for the cases examined in Fig. 7.23. An

increase in AoiAT requires an increased mass flux across the cross-section of the inlet

streamtube and hence the depth of the cross-section grows with a small change in its

width. The greater depth of the inlet streamtube results in the ingestion of boundary

layer fluid of higher velocity and hence greater energy, thus increasing Ce. This will

beneficially lead to a larger 11 at the duct exit. The second factor, counteracting the

beneficial effects of the greater Ce of the ingested flow, is the increased viscous loss

within the waterjet inlet duct. There are two main causes for these viscous losses. The

first is due to the larger overall velocities within the waterjet inlet leading to increased

pressure drop. The second results from the loss in total pressure associated with the

258

thickening and eventual separation of the boundary layer in the lower duct. This latter

point is the likely reason for the trend of decreasing 11 for A/ AT> 1.48.

It is evident from Fig. 7 .15f that larger values of either RL or A/ AT result in a larger

waterjet inlet, with AJ AT having the greatest effect on duct volume. The total variation

in volume over the two-parameter design space is 20.4% of the minimum volume.

7.4 Discussion of Results

In this section the physical mechanisms underlying the observed trend in the results for

the minimum static pressure on the inlet lip and the vertical forces acting on the waterjet

inlet are discussed at greater length. The former issue is of particular relevance to the

design of the inlet lip for the avoidance of cavitation inception at high vessel speed.

7.4.1 Hydrodynamics of Inlet Lip

The results of Section 7.3 show that the minimum static pressure is affected by the

following aspects of the waterjet inlet geometry:

1) The radius of the inlet lip (RL)

2) The shape of the lip profile (y, HL)

3) The internal diffusion within the waterjet inlet (AJAT).

In order to maximise the minimum static pressure on the inlet lip, it is necessary that the

flow in the vicinity of the inlet lip be symmetrical about a line through the centreline of

the lip and bisecting the angle of the lip profile, as shown in Fig. 7 .24. In other words,

the dividing streamline must be aligned at an angle A (to the horizontal), where A is

given by

A=( a-y)/2 (7.10)

This is almost achieved for the case shown in the streakline plot of Fig. 17 .c. When the

angle (~) of the dividing streamline in the immediate vicinity of the inlet lip deviates

from A, the minimum static pressure decreases. Therefore, the greater the deviation of ~

from A, the lower the minimum static pressure on the inlet lip will be.

259

+ve A-13 Half-Angle ofL1p Profile

Lip Profile

Fig. 7.24 Definition of angle of lip profile and dividing streamline

In order to examine whether a correlation exists between the minimum static pressure

on the inlet lip and the deviation between A and ~. the results of the fifteen cases

examined in detail in Section 7.3 are plotted in Fig. 7.25. Relevant data used as the basis

of Fig. 7.25 are tabulated in Appendix A.4. It can be seen that a clear correlation exists

between the minimum static pressure on the inlet lip and the deviation of the incident

flow from A.

-0.6 • -0.8

-1.0

c. -1.2 u

-1.4 • Circular lip

-1.6 A Raised Lip Profile

0 5 10 15 35 40 45 50

Fig. 7.25 Relationship between incident flow and minimum static pressure on the lip

From Fig. 7.25 it is apparent that the minimum Cp on the inlet lip is dependent upon A-~

and the shape of the lip profile. For both lip profiles a linear trend is evident, thus

indicating a simple linear relationship between the minimum Cp on the lip and A-~. The

steeper curve of the raised lip profile may be attributed to the generally higher static

260

pressures on the underside of the inlet lip arising from the inclined surface between the

lip and the inlet/hull interface. It is thus evident that in the design of the lip profile of a

waterjet inlet, the flow in the vicinity of the inlet lip must be directed in such a way as to

minimise its deviation from the bisector of the lip profile angle (A.) of the lip profile if

the minimum static pressure on the inlet lip is to be maximised.

It is found from the two-parameter design space investigations of Section 7.3.2 that a

raised lip profile proves to be beneficial in increasing the minimum static pressure on

the inlet lip, primarily as a result of its ability to minimise A.-~. It is also found that there

is a trend toward lower peak negative pressures if either y or HL is increased. Using the

results of Appendix A.2, the minimum static pressure on the inlet lip is plotted against a

lip profile parameter (yHr/D) in Fig. 7 .26. Here, y is measured in degrees for

convenience. It is evident from the figure that a distinct relationship exists between the

minimum Cp and yHr/D. This suggests that yHr/D governs the hydrodynamic

performance of the inlet lip over the design subspace investigated in Section 7.3.2. A

fourth-order polynomial of the form

Cp=5.02(yHr/D)4 -19.03(yHr}D)3 +25.02(yHr}D)2 -12.94(yHr/D)+0.94 (7.11)

-0.5

-0.6

-0.7

-0.8

-0.9 c.

u -1.0

-1.1

-1.2

-1.3 • -1.4

0.3 0.5 0.6 0.8 0.9 1.1 1.2 1.4 1.5 yl-ldD

Fig. 7.26 Effect of lip profile parameter on minimum static pressure coefficient

applicable to the range of data shown in Fig. 7 .26, is found to provide a good least­

squares regression fit to the data. It may be concluded that an inlet lip profile with a

larger yHr/D exerts a greater influence on the flow in the vicinity of the inlet lip and acts

261

to direct the flow over the lip in such a way as to minimise A.-~ and thereby increase the

minimum static pressure on the lip.

It is found from the two-parameter design space investigations of Section 7 .3.3, that

increasing the internal diffusion within the waterjet inlet (via AJAT) leads to increased

minimum Cp on the inlet lip for AJAr<l.38, after which the trend is reversed.

Increasing AJAT and hence IVRT, changes the shape of the inlet streamtube, due to the

greater mass flux ingested and so acts to direct the flow in such a way as to minimise A.-

~- At an AJ AT of 1.38, where the minimum Cp on the lip is maximised over the design

subspace, IVRT is 0.828. For a waterjet having an AofAT of unity, IVR and IVRT are

identical. It was found from the results presented herein, that increasing IVRT can have a

beneficial effect on the cavitation performance of the inlet lip. This therefore suggests

that operating a waterjet inlet having an Aof AT of unity at a higher IVR can have a

beneficial effect on the cavitation performance of the inlet lip.

7.4.2 Vertical Forces acting on the Waterjet Inlet

The vertical forces acting on the waterjet inlet may be partially attributed to the net

changes in vertical momentum of the flow through it. In the following discussion,

reference is made to the control volume of Fig. 7.27, which is similar to the one used for

F

Fig. 7.27 Control volume for vertical force analysis

262

the derivation of waterjet thrust in Chapter 2. fu this case A1 has been moved to

correspond with the ramp tangency point (hence A3=0) of the inlet and A6 is positioned

at the duct exit.

If the waterjet inlet is not inclined to the horizontal, then the net vertical force caused by

momentum changes acting on the control volume must necessarily balance the weight of

the entrained water. Mathematically this may be expressed as

(7.12)

where } is the unit vector in the vertical direction, V the volume of the control volume,

dA the differential surface area and subscript CV denotes the control volume.

Alternatively Eqn 7.12 may be written as

(7.13)

where the force vectors, F. , represent the integral contributions of static pressure and

shear stress over the control volume area and may be evaluated as

F. =-JpdA-f ~dA A, A,

(7.14)

fu Eqn 7.14 dA represents the differential surface area vector pointing out of the control

volume and ~ is the shear stress vector. Thus, the hydrodynamic forces acting on the

solid surface of the control volume in the vertical direction may be written as

(7.15)

The total vertical force acting on the waterjet inlet, however, includes the contribution

of the area associated with the inlet lip and is given by

(7.16)

where Fy is the total vertical force acting on the waterjet inlet. Using Eqn 7.15, Eqn 7.16

may be also be written as

(7.17)

Thus the vertical force acting on the waterjet inlet is a function of the vertical forces

acting on the inlet streamtube and the inlet lip.

The vertical forces acting on the inlet streamtube and inlet lip are however implicit

functions of the geometry of the waterjet inlet. The work presented in this thesis has

263

shown how changes in the geometry of the waterjet inlet and the IVR (or IVRT) affect

the flow into the inlet and hence the shape of the inlet streamtube. Similarly, these

changes also alter the flow in the vicinity of the inlet lip. It is therefore apparent that the

vertical forces acting on the waterjet inlet will change with geometry or IVR.

As a result of the design subspace investigations of Section 7.3, Cpy is found to

generally increase with increases in RL, a, y and HL. On the other hand, the trend is for

Cpy to decrease with increasing AofAT. Except for this latter trend of decreasing Cpy

with increasing AofAT, the trend in the behaviour of Cpy is similar to the trend in

minimum Cp with changes in the design parameters for the first two design subspaces

investigated.

Since the minimum Cp on the inlet lip can be correlated with A.-~, Cpy must similarly

correlate with A-~, thus suggesting a clear link between A-~ and the change in vertical

forces acting on the waterjet inlet. This relationship is shown in Fig. 7.28 for the ten

cases examined in detail in Section 7.3.1 and Section 7.3.2. It must be noted that the

relationships between Cpy and A.-~ for the two design subspaces are distinct and

different. This difference arises from two causes. The first being the generic geometry of

0.16

• •• 0.14

0.12 • 0.10 •

i::: 0.08 u 0.06

• 0.04 • Design Subspace 1 • 0.02 • Design Subspace 2

0.00

15 20 25 30 35 40 45 50

A.-~ (o)

Fig. 7.28 Relationship between incident flow and vertical forces on waterjet inlet

the inlet which affects the shape of the inlet streamtube and therefore the change in the

vertical momentum of the flow through the waterjet inlet. The second is the lip profile

264

used, as the geometry of the inlet lip profile will inevitably affect F7 in Eqn 7.17. The

relevant data used as the basis of the plot of Fig. 7.28 is tabulated in Appendix A.S.

A qualitative investigation of the vertical forces acting on the waterjet inlet, the results

of which are presented in Appendix A.6, was undertaken in order to determine the

relative contributions of the components of the waterjet inlet to the overall force

balance. The waterjet inlet is treated as being two-dimensional and the pressure forces

acting on the upper and lower waterjet inlet are calculated according to

(7.18)

where (Fy )1 is the net vertical force (per unit width) acting on the ith section of the

waterjet inlet geometry, Sx the component of arc length in the X direction and s2 and sl

are the upper and lower limits of the range of arc lengths over which the integration is

carried, respectively. Note that the contribution to the vertical force arising from skin

friction has been omitted, as the primary interest here is on the changes in the pressure

forces.

The following conclusions can be drawn from the study:

1) For all cases examined, the net vertical force on the inlet ramp is negative

2) The net vertical force on the inlet lip is negative and represents between 10%-20% of

the net vertical force acting on the lower waterjet inlet for the design subspaces

investigated in Section 7.3.1 and Section 7.3.2.

3) The outside of the duct bend represents the largest contribution to the vertical force

on the upper waterjet inlet.

4) For the design subspaces investigated in Section 7.3.1 and Section 7.3.2, the relative

change in vertical force is greatest in the lower waterjet inlet, thus adding weight to

the argument that it is the angle of the flow into the waterjet inlet that affects the

vertical force via changes in the vertical momentum of the flow and lip suction.

265

7.5 Closure

Three two-parameter subspaces of the design hyperspace spanned by the vector of

design variables of the author's parametric waterjet inlet geometry are investigated in

order to examine how changes in the geometry of the waterjet inlet affect its

hydrodynamic performance.

The design subspaces investigated are:

1) RJD and a - Radius of the inlet lip with inlet steepness

2) y and Hr/D - Profile of the inlet lip

3) A/AT and Rr/D -Internal diffusion with radius of the inlet lip

All CFD simulations were run for a waterjet inlet operating at an IVR of 0.6 with a free­

stream velocity of 20.58ms-1 (40 knots vessel speed) corresponding to a vessel cruise

condition. A boundary layer of non-dimensional thickness of BID=O.S was used to

simulate the ingestion of fluid from a thick upstream hull boundary layer.

The following conclusions may be drawn from the investigations presented herein:

1) The minimum static pressures on the surface of the waterjet inlet occurred in the

vicinity of the inlet lip and are sufficiently low to cause cavitation in all cases

examined.

2) The minimum static pressure on the inlet lip is largely dependent on the angle

between the angle of the bisector of the profile of the inlet lip (A) and the angle of the

dividing streamline in the vicinity of the inlet lip (~). Large A.-~ results in low

minimum static pressure on the inlet lip. When A.=~, the flow is symmetrical about

the inlet lip profile and the minimum Cp on the lip is maximised.

3) A raised lip profile acts to direct the flow in the vicinity of the inlet lip more

symmetrically over the lip and so tends to minimise A.-~. Larger values of the

parameter yHr/D are beneficial in increasing the minimum Cp on the inlet lip when a

raised lip is used.

4) The net vertical force acting on the waterjet inlet results from a net change in the

vertical momentum of the flow through the waterjet inlet and vertical forces acting

on the inlet lip below the stagnation line. This was shown to be a function of the

266

geometry of the inlet streamtube, the inlet lip and is found to correlate with A.-~.

Generally, larger RL, HL and a resulted in greater positive vertical forces on the

waterjet inlet. The lower overall static pressures in the waterjet inlet, due to the

increased mass flow-rate as a result of from internal diffusion (AoiAT), lead to a

"suck-down" effect.

5) There are no distinct trends and little overall variation in the area-averaged 11 for the

first two design subspaces investigated, thus suggesting that viscous losses within the

waterjet inlet are approximately equal. The overall level of 11 is determined primary

by the energy flux coefficient (Ce) of the ingested flow. For the third design subspace,

there is a noticeable trend of increasing 11 with AoiAT up to Ao1AT=l.48 when the

trend reverses. This trend reflects a conflict between the increasing Ce of the ingested

flow (primarily by virtue of increased streamtube depth) and the increasing viscous

losses within the waterjet inlet.

6) Variation of RL, a, HL or y leads only to relatively small variations in the distortion

coefficient (De) at the duct exit. Increasing the internal diffusion (AJ AT) lead to

significant changes in De. This is due to a combination of the quality of the ingested

flow and the thickening of the boundary layer in the bottom of the duct by virtue of

adverse pressure gradients and secondary flow.

7) Smaller, more compact waterjet inlets are favoured by large a, small RL and small

HL. Increasing the internal diffusion (AoiAT) within the waterjet inlet leads to an

increase in duct volume at constant throat diameter.

267

Chapter 8 Optimisation of Waterjet Inlet Design

In this chapter a methodology for the optimisation of a parametrically-defined generic

flush-type waterjet inlet (in the absence of hull form) is presented. The methodology is

applied to the optimisation of an inlet for the maximisation of cavitation number on the

surface of the inlet at the cruise condition (selected IVR of 0.60). The main aim of this

chapter is to outline the optimisation methodology and to demonstrate its effectiveness

in improving the hydrodynamic performance of a genericwaterjet inlet geometry.

Seil et al ( 1997) noted the following qualities of an ideal waterjet inlet:

"An ideal waterjet inlet would be free of cavitation and flow separation in the inlet,

deliver a spatially uniform distribution of velocity and pressure to the pump with the

minimum of total pressure losses and contain the minimum weight of entrained water.

An ideal inlet should also affect the flow around the vessel hull in such a way as to

minimise the thrust deduction factor".

In addition, turbulence present in the flow should have no adverse effect on pump

performance. Implicit in the above statement by Seil et al ( 1997) is the notion that the

"idealised" waterjet inlet must perform well over the complete range of IVR

encountered in its operation.

The performance of the ideal inlet can be expressed mathematically usmg the

hydrodynamic performance parameters introduced in previous chapters as

O"mm>>O, tJ=O, t$;0, (~, Dc,V*, ...JkiUref, lJD)-70 (8.1)

where O"mm is the minimum cavitation number on the surface of the waterjet inlet, tj is

the lip loss thrust deduction fraction (see Chapter 2), t the thrust deduction fraction of

the waterjet-hull system, ~ the total pressure loss factor of the waterjet inlet, De the

distortion of total pressure at the duct exit, v* the non-dimensional duct volume of the

waterjet inlet, ...JkiUrer the velocity-scale of turbulence in the inlet and fJD the

268

corresponding turbulent length-scale.

In reality of course, flush-type waterjet inlets do not possess the characteristics of the

ideal waterjet inlet and therefore there must be a compromise between the various

aspects of the hydrodynamic performance of the waterjet inlet, in order to achieve a

practical waterjet inlet design. Since a parametric description of a flush-type waterjet

inlet involves of the order of ten parameters (8 for the author's geometry), one cannot

possibly achieve an "optimum" (the term being used loosely in this case) waterjet inlet

design based purely on experience. Even if one is to achieve a practical waterjet inlet

design of superior hydrodynamic performance, it cannot be claimed that the best

possible design has been achieved. The question is thus posed as to what the optimum

shape of the waterjet inlet is, for a given set of criteria and what the means are to

achieve this shape. Herein lie the issues at the heart of waterjet inlet optimisation.

As noted in Chapter 1, the optimisation of the hydrodynamic performance of waterjet­

propelled vessels requires an optimisation of the complete waterjet inlet/hull geometry

over a range of NR values. This poses a very complex hydrodynamic problem.

Therefore, the simplified problem of waterjet inlet optimisation in the absence of hull

form, but with the effect of hull form being represented by a boundary layer upstream of

the waterjet inlet, is examined in this chapter.

The optimisation methodology presented here is based on the approach of treating the

waterjet inlet optimisation as a formal mathematical optimisation problem where a

mathematical objective function, expressing favourable design criteria, is to be

minimised (or maximised) within a set of constraints placed upon the geometry of the

waterjet inlet. The actual approach to optimisation used here is what Frank and Shubin

(1992) call a "Black-Box Method", in which an optimisation routine is coupled with an

existing CFD code (Fluent). The CFD code provides the flow solution data necessary

for calculation of the objective function. One advantage of this approach is that the CFD

analysis code can be used without modification and hence there is no need to modify

complicated discretisation schemes within the code. The disadvantage of such an

approach is, however, that there is a high associated computational cost. This cost grows

linearly as the number of design variables increases. Frank and Shubin ( 1992) further

269

noted that one mitigating factor of this limitation is that the solution of perturbed

analysis problems is considerably less computationally expensive than solving arbitrary

problems. It will be shown in Section 8.2 that this approach does result in significant

computational savings.

As discussed in Chapter 1, the use of CFD provides a necessary foundation for the cost­

effective optimisation of waterjet inlet design. Furthermore, the flow in the waterjet inlet

can only be simulated realistically using computational techniques that account for

viscosity and turbulence. Thus it is necessary to solve the Reynolds-averaged Navier

Stokes (RANS) equations with a suitable turbulence closure. From the experimental

validation studies presented in Chapter 5, it is clear that the RNG k-£ turbulence model

can be used to provide suitable turbulence closure of the RANS equations. Thus, the

optimisation methodology presented in this chapter is under-pinned by flow simulations

in which the RANS equations are solved using the RNG k-£ turbulence model.

In Section 8.1 the optimisation methodology used for the optimisation of flush-type

waterjet inlets is presented. The computational modelling of the flow domain and a

discussion of the simulations undertaken to optimise the generic flush-type waterjet inlet

geometry are presented in Section 8.2. It must be noted that the optimisation undertaken

is for constant NR, but not constant flow-rate. In Section 8.3, the hydrodynamics of the

optimised waterjet inlet geometry (with respect to cavitation performance at the cruise

condition) are compared with the hydrodynamics of the initial geometry. The results of

the waterjet inlet optimisation are discussed in Section 8.4. The conclusions of the work

presented in this chapter are summarised in Section 8.5 and the hydrodynamic design of

the optimised waterjet inlet is compared with current industrial designs.

8.1 Optimisation Methodology

The optimisation methodology presented in this section is based on a formal

mathematical approach to optimisation where an objective function describing the

hydrodynamic performance of the waterjet inlet is to be minimised (or maximised, since

this is the equivalent problem), subject to a set of linear and non-linear constraints on

the geometry of the waterjet inlet. This may be mathematically expressed as

270

minimise f(X), X e 9\" subject to the constraints

c.(X)=O, i=1,2, ... ,m'

c.(X)~O, i=m'+1, ... ,m

(8.2)

using the generic form outlined in Gillet al (1981). Eqn 8.2 states that the mathematical

objective function (f) is to be minimised by finding a suitable set of variables (X ) of

which f is either an explicit or implicit function, subject to a set of constraints (c1) on the

components of X .

In the work presented herein, X represents a vector of design variables

(8.3)

whose components are the geometric parameters of the generic flush-type waterjet inlet

geometry described in Chapter 4. The description of this generic geometry X may thus

be written as

(8.4)

where:

1) a. - The angle of inclination of the inlet to the horizontal plane

2) RL- Radius of the inlet lip

3) H - Height of the pump centreline above the base of the inlet

4) R0 - Radius of curvature of the centreline of the duct bend

5) LH - Length of the horizontal duct section downstream of the bend

6) AJ AT - Ratio of duct exit area to throat area

7) HL- Height of the centreline of the inlet lip above the inlet opening plane

8) y - Angle of inclination of the raised-lip profile

In order to allow an optimal shape of waterjet inlet to be geometrically scaled arbitrarily

(provided that Reynolds number effects on the inlet flow are small) so as to increase the

applicability of an optimum waterjet inlet shape to other sizes of waterjet inlet, the

components of X may be non-dimensionalised to give

271

X= [a, RL/D, H/D, R 0 /D, LH/D, A 0 /AT, HL/D, yf (8.5)

where Dis the diameter of the circular throat section of the generic inlet and a and"{ are

measured in radians. This approach does not however, guarantee scaling for 0", as the

static pressure varies linearly with elevation due to gravitational effects. Concerns

about Froude number scaling are unnecessary since attention is focused purely on an

isolated waterjet inlet flow here.

The geometric constraints on the components of X are

a:

RUD:

HID:

RoiD:

LHID:

AofAT:

HLID:

y:

(a)I:s;a:s;(a)z

(RiiD)I:s;RUD:s;(RUD)z

HID=(HID)I

(Ro/D)I:s; RofD:s;(Ro/D)z

(LH/D)I:s; LHID:s;( LH/D)z (8.6)

(AJAT)I:s; AJAT:s;(AJAT)z

(HiiD)I:s; HUD:s;(H-Ro(l-cosa)-(RL +D/2)sin(7t/2-a))/D

("{)J:s;y:s;(y)z

where subscripts 1 and 2 refer to the lower and upper constraints on the particular

geometric parameter.

The objective function (f) describing the hydrodynamic performance of the waterjet inlet

may be written as

(8.7)

where f is an explicit function of the hydrodynamic performance parameters ('1ft) which

provide a measure of the different aspects of the hydrodynamic performance of the

waterjet inlet. The 'lf1 are, however, unknown functions of the inlet geometry, hence

(8.8)

Therefore, f is an implicit function of X1 and so f is ultimately a function of the waterjet

inlet geometry,

f = f('lf 1(XpX 2 , ••• ,xn),'lf 1(XpX2 , ••• ,xn), ... ,'lfN(xpX2 , ••• ,xn)) (8.9)

The choice of explicit relationship between f and 'lf1 is arbitrary, but a suitable choice is

272

N

f = :Laz,-J(az,'I'Jb, (8.10) 1=1

where a2,_ 1, a2, and b, are constants. Clearly f is a non-linear function of 'I'• which

becomes linear if b,=l. Eqn 8.10 is the generic form adopted in Seil et al (1997). Other

functional relationships between f and 'I'• are of course possible, but it is the author's

view that the generic form used in Eqn 8.10 offers the greatest generic usefulness.

Since the focus of this chapter is on the optimisation of the waterjet inlet geometry with

the aim of eliminating cavitation on the underside of the inlet lip and maximising

overall static pressure on the duct surface for the cruise condition, Eqn 8.10 takes the

following form

f = -crmm. (8.11)

Minimisation of Eqn 8.11 actually corresponds to a maximisation of O"mm· Optimisation

based on Eqn 8.11 clearly ignores other factors of hydrodynamic performance, but

focuses attention on the issue of inlet cavitation. This allows a greater understanding of

how a synergistic combination of design parameters can improve cavitation performance

and facilitates a greater understanding of how the waterjet inlet geometry affects the

flow within it.

An alternative approach to the optimisation of the waterjet inlet, with regard to its

cavitation performance, is to solve the inverse design problem. Given the desired static

pressure distribution on the duct surface (for example, the pressure distribution on the

duct centreplane), a suitable geometry is found that minimises an objective function of

the form

(8.12)

is found. In Eqn 8.12, c represents the least-square error between the computed static

pressure distribution on the duct surface (p) and the desired distribution (Pspec). Jameson

(1995) suggested this approach for the optimisation of airfoil shapes.

8.1.1 Overview of Optimisation Procedure

In this section the sequence of steps involved in waterjet inlet optimisation are

273

discussed. The overall optimisation cycle is shown in Fig. 8.1 and is described below. In

order to begin the optimisation, an initial design that satisfies the geometric constraints

on its geometry is selected. The mesh bounding the modelled flow domain containing

the waterjet inlet geometry is then generated.

The mesh generation procedure follows that outlined in Chapter 4 and is performed

manually, as the process of obtaining a suitable grid is necessarily iterative, guided by

the experience of the user. Since waterjet inlet optimisation requires the examination of

a large number of different geometries, the mesh must be suitably modified for each

new geometry in order to ensure that sufficient cells are placed in regions of large flow

gradients and features of the waterjet inlet geometry (in particular the inlet lip) are

adequately represented.

Note that Fig. 8.1 does not imply that only one mesh is generated during the grid

generation process. The number of meshes generated depends on the number of

evaluations of the objective function for each iteration of the optimisation routine. If an

optimisation algorithm based on the evaluation of gradients of the objective function is

used (eg. quasi-Newton methods, see Gillet al (1981)), evaluation of the gradient vector

would require n+ 1 or 2n flow simulations depending on whether the gradients are

calculated using forward differences or central differences, respectively. Calculation of

second derivatives using central differences for evaluation of the Hessian matrix in

Newton methods requires 2n+ 1 function evaluations. If direct search methods are used

(see Gottfried and Weisman (1973), Section 3.3), such as in this thesis, the results of

only one flow simulation are required during each iteration of the optimisation cycle

shown in Fig. 8.1.

After the mesh generation cycle, a CFD analysis using Fluent is undertaken in order to

obtain a converged solution for the flow in the modelled domain. The results analysis

program (Analysis) reads flow data extracted from Fluent and data relating to the cell­

face area vector components of the surface of the waterjet inlet mesh produced by

Inlet3D. Output from Analysis contains such 'lf1 as crmm. De. 11 and Cpy. Using the values

of '1'1 , the optimisation routine calculates the value of the objective function (f) decides

whether the criteria for convergence of the optimisation algorithm has been met and

274

determines the next design point if the optimisation algorithm has not converged. If the

optimisation algorithm has converged, or the number of iterations of the optimisation

cycles exceeds a certain maximum limit, the result of the optimisation is returned.

• •

Volume Surface Area

Surface grid cell-face area

vectors

l X,

Gnd GeneratiOn (lnlet3D)

l (x,y,z)

CFD Analysis (Fluent)

l (p,P0 ,U,V,W,tw)

Results Analysis (Analysis)

Optimisation Routine

(Analysis)

l Converged?

No

Fig. 8.1 Flow chart of optimisation methodology

275

x.

8.1.2 Optimisation Algorithm

The optimisation algorithm chosen for use in the optimisation of the waterjet inlet is the

algorithm of Sherif and Boice (1993), which is essentially a development of the "Pattern

Search" algorithm of Hooke and Jeeves (1961). The Sherif-Boice Algorithm (SBA) is a

direct-search method which does not rely on the calculation of either the gradient vector

or the Hessian in determining the next suitable design point. It rather relies purely on the

calculation and comparison of values of the objective function. The SBA is simple,

robust, computationally efficient and allows easy implementation of constraints.

Furthermore, since it does not involve the calculation of partial derivatives of the

objective function, it can be used effectively when the objective function contains

discontinuities, thus furthering its appeal. There is, however, no guarantee that the

solution obtained using the SBA does in fact correspond to the global minimum of the

objective function. This is the case for most optimisation algorithms.

The SBA is presented in the form of a flow-chart in Fig. 8.2. In the SBA there are

essentially two types of movements made in the design hyperspace. The first is termed

an "exploratory move", where a perturbation in the components of X is made about a

base point and the objective function is evaluated. This is done in order to determine

whether a reduction in f has occurred and so determine a direction of decreasing f.

Should a reduction in f occur, then the location where this reduction occurs becomes the

new base point. The second type of move is termed a "pattern" move and is essentially

an extrapolation from the current and previous base points. The rationale behind this

being that the vector describing the difference between the base points describes a

direction of decreasing f and so moving in this direction should result in a reduction of f.

The SBA shall be explained in greater detail below and proceeds as follows:

Initialisation

An initial solution vector X0 , satisfying the constraints of the problem, is first chosen

and the value of the objective function f(X 0 ) is then determined at this point. A

perturbation vector

(8.13)

276

No

No

Input Pattern Search

Calculate Base Point

Exploratory Moves

Set New Base Point

Pattern Move

Exploratory Moves

No

Reduce Step Size

Yes

Return Optimised Geometry

No

Fig. 8.2 Sherif-Boice Algorithm (Sherif and Boice (1993))

277

is defined where the components of this vector correspond to the size of the

perturbations in the corresponding components of X , such that

(8.14)

Exploratory Move

A perturbation is made in X I of the ith base point, such that X = [X I + f. I ' X 2 ' ••• 'X n r and the objective function is evaluated at this point. If f(X) < f(X.), then this X

becomes the new base point. If f(X) ~ f(X.), then another perturbation is made in x1,

but in the opposite direction, such that X = [X I - tp Xz ' ••• 'X n r. If f (X) < f (X.) then

this new value of X becomes the base point, otherwise the above procedure is repeated

for each component of x. until f is minimised. If a new base point cannot be found

after perturbation of all components of x. , then the step sizes of the components of the

perturbation vector are halved, i.e. e -7 E"/2 and the above process is repeated.

Pattern Move

If a base point is found after an exploratory move, then a pattern move is made in which

a new solution vector is determined from an extrapolation of the current and previous

base points. The new solution is calculated as

x = x.-1 +2(X. -x.-1) (8.15)

where X is the new point resulting from the pattern move, x. the current base point

determined from the exploratory move and X1

_ 1 the previous base point. The value of f

is determined at X. Should f(X) < f(X.) then X is taken as the new base point and

exploratory moves are made around the new base point. If on the other hand,

f(X) ~ f(X.), then the pattern move is judged to have failed and x. is retained as the

current base point. An exploratory move is then made around X. in search of a new

base point.

Termination

When exploratory moves around a base point continually fail and every component of

e is reduced below the convergence criterion (limiting step size), the algorithm

278

terminates. The objective function has found a local minima and X is at its local

optimum value.

Depending on the characteristics of the solution space, one may expect that the initial

rate of convergence of the SBA will be relatively large when the starting solution is far

from the solution vector corresponding to a minimum of f and the step size ( £1) of each

component of the perturbation vector is large. It is, however, clear that as the local

minimum value of f is approached, the rate of convergence will become slower and will

require a large number of exploratory moves and hence evaluations of f to obtain a

converged solution ( £, = ec)2, where £c,1 is the limiting step size). The stricter the

convergence criterion (minimum step size) is, the larger will be the number of

exploratory moves that may be expected before convergence is achieved. While this

presents no problem when finding the location of minima for analytic functions, this

issue assumes great importance here due to the computational expense involved in

obtaining RANS solutions (in order to evaluate the objective function). Therefore some

liberty must be taken and the criteria associated with convergence relaxed. In other

words, the tolerance or minimum step size associated with convergence must

necessarily be increased in order to avoid an excessive number of exploratory moves in

the vicinity of the optimum solution. For the purposes of practical design, this may be

deemed to be an acceptable approach.

8.2 Computational Simulation and Optimisation

In this section an overview of the actual optimisation, carried out in order to maximise

the minimum static pressure on the surface of the waterjet inlet, is presented. The same

boundary conditions, external domain size and turbulence modelling as used for the

CFD flow simulations presented in Chapter 6, are also used for the flow simulations

here. The same upstream boundary layer velocity and turbulence data applied to

Boundary 1 of the simulations of Chapter 7, is also applied on Boundary 1 for the

optimisation-related CFD simulations presented herein.

The flow solution from the previous exploratory move or base point is used as an initial

flow solution for subsequent simulation, as each new flow computation generally

279

corresponds to a minor perturbation of the previous geometry. This leads to a dramatic

decrease in computational time and it was found that converged flow solutions could

generally be obtained within 50-100 iterations, which was significantly less than the

600-700 iteration necessary to obtain a converged solution ab initio (with an initial

solution for the velocity field specified as the free-stream velocity). When it is found

that an exploratory move corresponds to a point that has already been evaluated, the

objective function was evaluated using the previously calculated objective function

value for that point. This therefore saves computational effort and hence time. Since the

SBA is essentially a sequential algorithm moving from point to point, it is therefore

impossible to run several computations in parallel. This contrasts with gradient-based

methods where flow simulations used for the evaluation of objective function gradients

can be run in parallel. Thus while total computational effort may be comparable for both

methods, the sequential nature of the SBA will result in a longer overall time taken for

the optimisation process. This is a clear disadvantage of the SBA.

All computations were performed using the dimensional form of X given by Eqn 8.4. A

five-parameter optimisation is carried out using the SBA. In other words, five geometric

parameters are given the freedom to vary while the others are held constant. The

parameters held constant are H, LH and R0 . In an actual waterjet installation, the height

of the pump centreline above the inlet opening (H) is likely to be determined by the

draught of the vessel and is therefore set equal to one duct diameter. This appears to be a

typical dimension for most installed waterjet (see Trillo (1994)). For the purposes of

maximising the minimum static pressure on the surface of the waterjet inlet, the length

of the horizontal duct section downstream of the duct bend (LH) is unlikely to have any

impact on the flow in the inlet region (where the minimum static pressures occur).

Furthermore, although an increase in LH does reduce the distortion of total pressure at

the pump inlet, it may be argued that the extra duct volume incurred (with attendant

structural weight, lost internal space within the vessel and reduced vessel buoyancy)

does not justify increasing LH beyond about one duct diameter downstream of the bend.

The radius of the duct bend (Ro) is fixed in order to focus the attention of the analysis

(presented in Section 8.4) on the geometric parameters associated specifically with the

inlet region, as well as to reduce the computational effort required for the optimisation.

The influence of R0 on the optimum solution is therefore unknown.

280

The constraints on the geometric parameters, together with the initial step size and

termination criteria for the SBA are tabulated in Table 8.1. Some liberty is taken in

selecting the convergence criteria in order to avoid an excessive number of iterations

near the optimum.

Parameter Geometric Constraint £1 CJ,C

a 2oo::;;a::;;35° 20 0.25°

RL 10 mm ::;;R~ 60 mm 3mm 0.375 mm

H H=600mm - -

Ro Ro=1600mm - -LH LH=400mm - -

AofAT l.O::;;AJAT::;;l.4 0.05 0.00625

HL (Rv2)::;;H~H-Ro(l-cosa)-(RL +D/2)sin(1t/2-a) 3mm 0.375 mm

'Y oo::;;-y::;; 11 ° 10 0.125°

Table 8.1 Geometric constraints, step size and convergence criteria

8.3 Results

In this section the results of the waterjet inlet optimisation are presented and details of

the geometry and hydrodynamics of the optimised waterjet inlet are discussed in relation

to the initial geometry. Furthermore, aspects of the optimisation are also examined. A

discussion of the correlation between the hydrodynamic behaviour of the waterjet inlet

and the underlying geometry is reserved until Section 8.4. A discussion of the

convergence behaviour of the optimisation algorithm in the current application is also

reserved until Section 8.4

Table 8.2 lists the geometric parameters describing the initial waterjet inlet geometry

and the optimised geometry. The geometry of both waterjets is shown in Fig. 8.3. The

following differences are evident between the two geometries:

1) The optimum inlet is "steeper" than the initial geometry.

2) The radius of the inlet lip is smaller for the optimised geometry.

3) The optimised inlet has an increasing cross-sectional area between the inlet throat and

duct exit. The diffusion is small, with the exit area being 5% greater than the throat

281

area.

4) The height of the inlet lip centreline has increased slightly.

5) The optimised geometry has an internal volume 8.78% less than the initial geometry.

Geometric Parameter/ Attribute Symbol Initial Optimum

Throat diameter D 600mm 600mm

Angle of inclination of inlet to horizontal plane a. 25° 30.5°

Radius of inlet lip RL 20mm lOmm

Height of pump centreline above base of inlet H 600mm 600mm

Radius of curvature of centreline of duct bend Ro 1600mm 1600mm

Length of horizontal duct section LH 400mm 400mm

Ratio of duct exit area to throat area AJAT 1.00 1.05

Height of inlet lip centreline above base of inlet HL 40mm 43mm

Angle of inclination of raised-lip profile y 110 110

Width of waterjet inlet - 700mm 700mm

Total length of waterjet inlet - 4175.5 3566.2

Length of inlet opening - 3078mm 2564mm

Internal volume v 0.7233 m3 0.6598 m3

Table 8.2 Comparison of initial and optimised waterjet inlet geometries

a) Initial Geometry

b) Optimised Geometry

Fig. 8.3 Geometry of initial and optimum waterjet inlets

282

Starting from the initial geometry, convergence (see Section 8.2) was achieved in 79

iterations of the SBA as shown in Fig. 8.4. Details of the optimisation can be found in

Appendix B.l. Here, one iteration is considered to be one evaluation of the objective

function. The minimum cavitation number on the surface of the waterjet inlet is

increased from -0.684 for the initial geometry to 0.284 for the optimised geometry.

Thus, not only is cavitation inception in the waterjet inlet eliminated, but also the

minimum static pressure on the surface of the waterjet inlet is maximised. It is

interesting to note that while the minimum static pressure occurs on the underside of the

inlet lip for the initial geometry, the minimum static pressure occurs on the side of the

inlet for the optimised geometry (at 45% of the length of the inlet opening from the

ramp tangency point).

The difference between the distribution of cavitation number on the surface of the

waterjet inlet for the two cases can be seen from Fig. 8.5. The low-pressure region of

peak negative cavitation number on the underside of the inlet can be clearly seen for the

initial geometry. For the optimised geometry, minimum static pressures have been

raised and so this region has disappeared. The distributions of cr also reflect the

diffusion of flow into the inlet and the effect of bend pressure gradients.

0.3

0.2

0.1

0.0

-0.1

~ -0.2 ~

-0.3

-0.4

-0.5

-0.6

-0.7

0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80

Iterations

Fig. 8.4 Variation of minimum cavitation number with iteration of the algorithm of Sherif and Boice ( 1993)

A comparison of the computed distribution of static pressure coefficient over the upper

surfaces of the waterjet inlets at their vertical plane of symmetry is shown in Fig.8.6.

283

The convention of previous chapters of measuring the non-dimensional arc length (SID)

from the ramp tangency point is adopted here again. The relationship between SID and

Initial Geometry

crmln

=jB 140

A 119

9 098

8 077

7 056

6 035

5 014

4 -007

3 -Q28

2 -049 Optimised Geometry -Q70

Fig. 8.5 Distribution of cavitation number on the surface of the waterjet inlet

c.. u

0.40 --Optimum

0.30 --Initial

0.20

0.10

0.00

-0.10

-0.20 +-'-~'-'--/-'--'--'----'-1-'--'-'--'-lf--'-..L..L--'-----t---'-"-'-~--'--'--'--'-+-'-'----'-'-l--'-'--'--'-l

0 2 3 4

SID 5 6 7 8

Fig. 8.6 Distribution of static pressure coefficient on upper inlet surface at centreplane

284

the underlying geometry for the initial and the optimised geometries are shown in Table

8.3. Over the initial part of the inlet ramp (S/D:s;2.2), static pressures are lower for the

optimised geometry, by virtue of the smaller radius of curvature of the ramp. The

maximum difference between ramp Cp for the two cases is -0.041.

Initial Optimum

Upper Centreplane SID SID

Ramp tangency point 0.00 0.00

End of ramp I Start of inclined duct section 4.67 3.65

End of inclined duct section I Start of bend 5.15 3.87

End of bend I Start of straight duct section 6.53 5.56

End of straight duct section 7.19 6.22

Lower Centreplane

Lip profile trailing edge 0.00 0.00

Start of lip 0.18 0.29

End of lip I Start of inclined duct section 0.26 0.33

End of inclined duct section I Start of bend 0.73 0.56

End of bend I Start of straight duct section 1.68 1.71

End of straight duct section 2.35 2.38

Table 8.3 Relationship between SID and the underlying waterjet inlet geometry

There is a difference in Cp at the duct exit of 0.03 between the two cases and the

optimised geometry exhibits a flatter pressure profile for S/D;:::4. The primary difference

between the two cases may be attributed to the presence of a region of near-separation

on the outside of the duct bend. This can also be seen by an examination of the plot of

skin friction coefficient (Cr) for the optimised geometry, along the upper duct on the

vertical plane of symmetry as shown in Fig. 8.7. A minimum value of Cr of 1.05x104

occurs at S/D=3.98, which is just downstream of the start of the bend. From the

conclusions drawn in Chapter 5 regarding the ability of the RNG k-E turbulence model

to predict the correct onset of flow separation, it is likely that the flow would separate in

a real waterjet inlet of the same geometry. Thus the combination of adverse pressure

gradients, associated with the diffusion of flow into the inlet and the outside of the duct

bend, acts to thicken the boundary layer and force flow separation.

285

I 6E-03

I 4E-03

I 2E-03

I OE-03

u 8 OE-04

6.0E-04

40E-04

2.0E-04

OOE+OO

0 2 3 4 5 6 7 SID

Fig. 8.7 Distribution of skin friction coefficient on upper inlet surface at centreplane of optimised waterjet inlet geometry

- =206m/s

--~-----~-- -- --::::...

------a) Initial Geometry

- =206m/s

----b) Optimised Geometry

Fig. 8.8 Computed velocity vectors on centreplane

Although there is a region of flow separation/near-separation in the upper duct for the

case of the optimised geometry, it can be seen from the vector plots of Fig. 8.8 that this

region is relatively small in cross-stream extent and is mainly confined to the outside

length of the bend.

286

A comparison of the computed distributions of Cp over the lower surface of the waterjet

inlets at their vertical plane of symmetry is shown in Fig. 8.9.

c. u

I 0 --Optnrused 08 -lmttal 06

0.4

02

0.0

-0.2

-04

-0.6

-0 8

-1.0

-1.2

00 0.1 0.2 SID

a) Inlet lip

050

040

0.30 c.

u 020

010

0.3 04

--Optmused -lmttal

~ ~ 00 0 N ~ ~ 00 0 N ~

0 0 0 - - - - - ~ ~ ~ SID

b) Lower inlet surface

Fig. 8.9 Distribution of static pressure on lower waterjet inlet at centreplane

The non-dimensional arc length SID follows the convention of previous chapters and is

therefore measured from the trailing edge of the lip profile into the inlet. It can be seen

from the pressure profiles of Fig. 8.9a that the overall static pressure is higher over the

sharper lip of the optimised geometry. It can also be seen that the location of the

stagnation point has moved from the upperside of the inlet lip for the initial geometry, to

around the centre of the inlet lip for the optimised geometry (refer to Table 8.3). It is

evident from the streakline plots of Fig. 8.10, that the lip profile of the optimised

a) Initial Geometry b) Optimised Geometry

Fig. 8.10 Flow in the vicinity of the inlet lip

geometry improves the flow in the vicinity of the lip by reducing the angle (A.-B)

287

between the angle of the bisector of the angle lip profile (A) and the angle of the

dividing streamline (p).

It must be noted that the pressure profiles presented in Fig. 8.6 and Fig. 8.9 are

calculated from grid node values, whereas the calculation of the objective function in

the optimisation uses cell-centre values. Thus the small region of low static pressure

associated with the inlet/hull interface at the trailing edge of the lip profile is not

captured. This can be seen from Fig. 8.9a where a minimum value of Cp of -0.269

occurs at the lip trailing edge of the optimised inlet, whereas a minimum cell-centred

value of Cp of -0.185 is calculated for the inlet surface. This fact does not in any way

detract from the validity of the optimisation presented herein (provided that no

cavitation will occur at the trailing edge of the inlet). Rather, it highlights the need to

modify the parametric geometry in order to achieve a smooth transition between the

inclined surface of the lip profile and the hull, so that static pressures at the inlet/hull

interface may be maximised. This is analogous to the way in which a larger radius of

curvature of the inlet ramp results in higher static pressure over the ramp surface.

Furthermore, the focus of the optimisation presented here is primarily directed at

maximising the static pressure in the vicinity of the inlet lip to avoid cavitation and

minimise inlet drag, erosion and other adverse downstream effects.

In Fig. 8.9b the static pressure distribution over the inside of the duct bend and

horizontal duct section is shown for the initial and optimised geometries. It can be seen

that the distribution of static pressure coefficient for the optimised geometry is lower

than that for the initial geometry (a maximum difference of Cp=O.lO), due to the higher

velocities in the lower duct region. Higher velocities in the lower duct region, in the case

of the optimised geometry, may be primarily attributed to a combination of increased

mass flow-rate, smaller inner bend radius and a thicker boundary layer in the upper duct,

relative to the initial geometry. It is also evident from Fig. 8.9b that the reduced length

of inclined straight duct between the throat of the inlet and the duct bend (for the

optimised geometry), results in a larger pressure gradient in the range 0.4~S/D~0.6. In

other words, the pressure gradient associated with the inside of the bend acts over a

shorter distance, (by virtue of the bend being located closer to the lip region) hence a

288

larger gradient results.

The dimensions and shape of the cross-section of the inlet streamtube for the initial

geometry and the optimised geometry are shown in Fig. 8.11. The width of the cross­

section of the inlet stream tube is identical for both cases, but the depth of the streamtube

is greater for the optimised geometry as a result of the greater mass flux through the

cross-section. The equal widths of the inlet streamtube for the two geometries, again

suggest that the width of the inlet streamtube is a function of the width of the inlet

opening and the boundary layer thickness (see Chapter 6). This is supported by the

results of Chapter 6 and Chapter 7.

0 0

0.00

-0.05

-0.10

-0.15

e -0.20 >.

-0.25

-0.30

-0.35

-0.40

Fig. 8.11

Geometry

--Imtla1 --Optimum

00 0

z/D

Dimensions and shape of cross-section of inlet streamtube

The improvement in static pressure distribution over the surface of the optimised

geometry does however, come at the expense of a decrease in total pressure recovery at

the duct exit and an increase in the distortion coefficient. The mass-averaged total

pressure recovery efficiency for the initial geometry is 0.676 whereas that for the

optimised geometry is 0.670. This represents a relative decrease of 0.89% and may be

attributed primarily to the effects of flow separation/near-separation within the

optimised geometry. For all intents and purposes, such a difference may be considered

to be negligible.

The distribution of total pressure coefficient over the cross-section of the duct exit is

shown in Fig. 8.12 for the initial and optimised geometries. It can be seen from the

figure that the optimised inlet exhibits a greater distortion of total pressure over the duct

289

exit, with higher total pressures in the lower duct and lower total pressures in the upper

duct when compared to the result for the initial geometry. The greater distortion of total

pressure in the case of the optimised geometry may be attributed primarily to the effect

of the upstream region of near-separation/separation on the outside of the bend. This

results in a region of low velocity and hence low total pressure at the top of the duct, as

can be seen from the vector plots of Fig. 8.8. Quantitatively, De is increased from 0.073

for the initial geometry to 0.119 for the optimised geometry. The effect of this increase

on the rotational efficiency of a pump is therefore likely to be adverse, but the

quantification of this effect is pump dependent and beyond the scope of the work

presented in this thesis.

a) Initial Geometry b) Optimised Geometry

Fig. 8.12 Distribution of total pressure coefficient over the duct exit

The secondary flow at the duct exit is shown in Fig. 8.13. The optimised waterjet inlet

shows a significantly greater secondary flow which may be attributed to the increased

angle of the duct bend and internal flow diffusion. This is in accord with the general

results of Chapter 7 which show that an increase in a. and AJAT, results in greater

secondary flow at the duct exit.

290

-=1rnls

' '

' '' I o

I I I

I IJI

II

'''•,, '•, ', ', IIJ,,, '•, ', ', I I

"'Dill II I I I '

)

11 1/ 1 I I

} I I I '

nut tl"' ,1 / : ; • 1/illll 11 I I '

1111 II I I ' 1111 Ill II II ' I

I I I I ' I Ill II I I /~ I I I I I I I I Jtfi

I ' ~~

/.;

I ' I o

a) Initial Geometry

-=1rnls

' :: ' : .:

I I , ,:-I I I I ,~ _.:::< I I I I I I _.:::::

I I I ' ~ I I I I -I 1 I I I / ,,'\l 1111 II II I' ... ,~

II I I -IIIII II II II I '\~

111111 1111 Ill II I I ,,~

;Y) ll': ': w ""''llrrl II II 11 I I I I "" 11 r 11 1 1 1 11 I I I I I ,II 11 II I ' /I

11 1 1 1 ' , I II I I I 1 ' I I I I I I I I I '- -/

I I I /

I 1 I 1

1 I

11 \ ':...

I I I I\~~ \ \ ' -_..,.;; ,~.::_--' - ~ ;:::;;:

b) Optimised Geometry

Fig. 8.13 Secondary flow vectors at the duct exit

Although the vertical force acting on the total waterjet inlet is larger for the optimised

geometry (2451 N) than for the initial geometry (2350 N), the non-dimensional lift

coefficient (CFY) is actually larger for the initial geometry (0.1097) than for the

optimised geometry (0.1039). This represents a relative decrease in CFY of 5.3%. It may

therefore be concluded that the increased mass flux through the waterjet inlet is

responsible for the increased vertical forces on the optimised geometry. The

hydrodynamic results presented above have been tabulated in Table 8.4 for the benefit

of the reader.

Hydrodynamic Performance Symbol Initial Optimum Minimum static pressure Min. Cp -1.153 -0.185 Minimum cavitation number O'mm -0.684 0.284 Distortion coefficient De 0.0729 0.119 Vertical force acting on waterjet inlet Fv 2350N 2451 N Non-dimensional lift coefficient CFY 0.1097 0.1039 Mass-averaged total pressure recovery efficiency T) 0.676 0.670 Area-averaged total pressure recovery efficiency T) 0.668 0.648 Volumetric flow-rate Q 3.480 m3/s 3.652 m3/s Internal Volume v 0.723 m3 0.660m3

Non-dimensional volume v* 3.349 3.054

Table 8.4 Table of hydrodynamic results

291

8.4 Discussion of Results

In this section two main issues associated with the waterjet inlet optimisation will be

discussed. The first issue is the relationship between the geometry of the optimised

waterjet inlet and the corresponding flow behaviour. The reasons why the underlying

geometry improves the hydrodynamic performance of the optimised waterjet inlet are

explored. The second issue is the convergence behaviour and the rate of convergence of

the Sherif-Boice Algorithm (SBA) when used for the current application.

8.4.1 Correlation Between Flow and Geometry

The optimised waterjet inlet geometry represents a synergistic combination of geometric

parameters. It is thus difficult to isolate specific aspects of the hydrodynamic

performance and correlate these with the underlying geometry. Never-the-less, such

correlations can still be made with a careful examination of the results. The correlations

made between the flow behaviour and the underlying geometry, as a result of the design

space investigations of Chapter 7, provide a valuable knowledge base from which to

correlate the hydrodynamic performance of the optimised waterjet inlet with the

geometric parameters describing it.

In order to investigate the relationship between the geometry of the waterjet inlet and the

corresponding flow, a correlation is sought between the change in minimum cavitation

number and the geometric parameters (both as functions of the iteration number of the

SBA). In other words, the correlation of the individual data sets of the plots presented in

Fig. 8.14 (excluding Fig. 8.14a) with Fig. 8.14a. The correlation coefficients calculated

using Eqn 6.28 are presented in Table 8.5 in the order of decreasing magnitude of the

correlation coefficient.

Parameter PxY RL -0.9221

a. 0.7303

AJAT 0.4292

HL 0.3747

y -0.0882

Table 8.5 Correlation between geometric parameters and minimum cavitation number

292

It must be noted that negative values of the correlation coefficient reflect an inverse

trend. For example, the negative values of the correlation obtained for RL indicate that

as RL is increased, a decrease in O"mm occurs. The almost total lack of correlation

between y and the trend in O"rmn may be primarily attributed to the initial value of y being

set at its optimum value. Since y was set at its upper limit at the start of the optimisation,

it was also at its optimum value and hence the small changes made to y during the

exploratory and pattern moves had little effect on O"mm as the SBA progressed. It may

thus be concluded that in the design of the inlet it is beneficial to specify y at as large a

value as practicable so as to increase the lip profile parameter (yHJD), provided that

there are no adverse flow effects (such as cavitation) on the waterjetlhull interface. The

reader is referred to Section 7.4.1 for a discussion of the lip profile parameter. The

results of Fig. 8.14e suggest that increased HL and hence (yHJD) is beneficial, but the

optimum value of HL is dependent upon the values of the other parameters as can be

seen from the correlations of Table 8.6.

As shown in Chapter 7, the internal diffusion within the waterjet inlet duct (between the

throat and the duct exit) has the effect of changing NRT (the NR based on the

volumetrically-averaged velocity at the inlet throat) and therefore reducing A.-~. This in

turn causes an increase in the minimum static pressure on the underside of the inlet lip.

Hence it is not surprising that the optimum value of A/AT should be greater than unity.

X y Pxv a RL -0.6072

a 'Y -0.4759

a A/AT 0.5562

a HL 0.6416

RL 'Y 0.0357

RL AJAT -0.2870

RL HL -0.2412

'Y AJAT -0.4506

'Y HL -0.6440

AJAT HL 0.5909

Table 8.6 Correlation between geometric parameters

293

03

02

0 I

0.0

-0 I c e -0 2 t)

-0.3

-04

-0 5

-0 6

-0 7

0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80

Iterations

a) Minimum cavitation number

23 22 21 20 19

8 18 g 17

..J 16 ~

15 14 13 12 11 10

0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80

lteratmns

b) Radius of inlet lip

35

34

33

32

c 31

c::s 30

29

28

27

26

25

0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80

Iterations

c) Angle of inlet inclination

Fig. 8.14 Convergence of optimisation algorithm

294

1 15

1 10

1 05

100~~~~~~~~~~~~~~~~~~~~~~~~~

0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80

lterattons

d) Area ratio

0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80

lterattons

d) Height of centreline of inlet lip

11.0

10.5

6' -;:::: 10.0

9.5

0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80

lterattons

e) Angle of inclination of raised-lip profile

Fig. 8.14 (cont.)

295

The value of a. clearly has a relatively strong influence on O'mm • It is also not surprising

that the optimised geometry has a larger value of a. than the initial geometry, given the

results of Chapter 7 (for the first design subspace investigated) which showed that

increasing a. results in a reduction of A.-13 and hence an increase in the minimum static

pressure. Despite this trend, the optimum values of AJ AT and a. are dependent, not only

on each other, but also on the other parameters.

The strong correlation between O'rrun and the radius of the inlet lip and the weak

correlation shown in Table 8.6 between RL and the other geometric parameters, (a. being

the exception) suggests that RL is an influential parameter affecting the pressure

distribution on the surface of the inlet lip and the optimum value of RL is largely

independent of the other parameters. The optimum value of RL corresponds to the lower

geometric constraint on this parameter. The trend in the behaviour of RL is therefore

toward a pointed lip. This is in direct contrast to the results of Chapter 7 which show a

trend of increasing RL to be beneficial in increasing the minimum static pressure on the

lip surface by maximising A.-j3. In this case, a different physical mechanism causes the

observed increase in O'mm with decreasing RL (at small RL). As the radius of the inlet lip

is decreased, the region of stagnated and low-velocity fluid surrounding the stagnation

point of the lip tends to "envelop" the lip and so leads to higher average static pressures

on the lip.

Since RL and yare found to converge to their optimum values after 30 and 31 iterations

respectively, it was decided to investigate the range of values over which the remaining

parameters (a., HL, A/AT) would give acceptable O'mm , when RL andy are set at their

optimum values. The results for the iterations of the SBA at which RL and y are at their

optimum values, are presented in Appendix B.2. It can be seen from Appendix B.2 that

the maximum deviation in O'mm over the range of values tabulated is only 0.063,

representing a relative deviation of 22% of the optimum value. Thus all values listed in

Appendix B.2 have a O'rrun >0.2 and so represent cavitation-free geometries under the

specified flow conditions. The results of Appendix B.2 are plotted as two-parameter

design subspaces in Fig. 8.15.

296

52

50

• 48

,....._ e § 46 • ...l ::r: • 44 • • • • • •

42 • • 40

29 30 31 32 33 an

1.10 ..... T

1.08 • 1.06 •

E-

~ • • • • • < 1.04 • • 1.02

1.00 ....

29 30 31 32 3: a (o)

52

50

• 48

,....._ e § 46 • =f • 44 • • • • • •

42 • • 40

1.00 1.02 1.04 1.06 1.08 1.1(

A./AT

Fig. 8.15 Two parameter subspaces

297

It can be seen from the relative location of the points in Fig. 8.15 that the following

ranges of values should give good cavitation-free performance:

1) a : 29°~a~33°

2) HL : 40mm~HL~52 mm

3) AJAT : l.OO~AJAT~l.lO

It is therefore evident that there is a certain design tolerance and flexibility that can be

made in selecting the dimensions of the waterjet inlet geometry, resulting from the use

of a lip of small RL.

8.4.2 Convergence Behaviour of the Optimisation Algorithm

The convergence behaviour of the SBA can be seen from Fig. 8.14a. The convergence

of the geometric parameters toward their optimum values is also evident from the other

plots presented in Fig. 8.14. The initial rate of convergence of the SBA is large during

the first 12 iterations, but then decreases dramatically as the geometry moves from a

region where O"mm changes rapidly, to a region where there is a much smaller variation in

O"mm with further changes in the waterjet inlet geometry. Since a and RL are well

correlated with O"mm , these two geometric parameters essentially "drive" the rapid

change of O"mm over the first 12 iterations, with the other parameters having a lesser

effect.

It is interesting to note (from Fig. 8.14) when the geometric parameters reached their

final values (excluding unsuccessful exploratory and pattern moves). The number of

iterations before the geometric parameters reached their optimum values is:

1) AJ AT : 25 iterations

2)RL : 30 iterations

3) 'Y : 31 iterations

4) HL : 42 iterations

5) a : 63 iterations

Hence the last 38 iterations are taken up primarily by exploratory moves resulting from

two reductions in step size and a final adjustment in a. Had the convergence criteria

been stricter, the number of iterations to convergence would have been increased. It is

thus evident that the selection of the minimum step-size can have a marked effect on the

298

number of iterations required before convergence is achieved. This issue has already

been briefly discussed in Section 8.1.2 for the general application of the SBA to CFD­

related optimisation.

8.5 Closure

The optimisation algorithm of Sherif and Boice ( 1993) was used to maxnmse the

minimum cavitation number (Eqn 6.1) on the surface of the author's parametric waterjet

inlet geometry (Chapter 4). A waterjet inlet of 600 mm was optimised at a free-stream

velocity of 20.58 ms-1 (40 knots vessel speed) with a boundary layer of non-dimensional

thickness OID=0.8 upstream of the inlet. From the optimisation study presented in this

chapter, conclusions may be drawn regarding the design of waterjet inlets for the

avoidance of inlet cavitation and the suitability of the Sherif-Boice algorithm (SBA) as

an optimisation tool.

An analysis of the results of the optimisation study presented in this chapter provides

valuable information regarding the influence of the different aspects of the waterjet inlet

geometry on the pressure distribution on the surface of the inlet, in particular in the

vicinity of the inlet lip and ramp. Thus building on the foundational work presented in

Chapter 7, useful conclusions can be made to guide the design of waterjet inlets at the

cruise condition. The following conclusions are thus drawn from the optimisation study

presented herein:

1) The optimised waterjet inlet design represents a synergistic combination of geometric

design parameters.

2) The increase in minimum cavitation number on the surface of the inlet is driven

largely by RL, a and the lip profile and there exists a range of parameter values at

which acceptable performance can be obtained.

3) Increasing the lip profile parameter (yHr./D) favours large static pressures on the inlet

lip by virtue of a reduction in the angle (A.-P) between the angle of the bisector of the

angle of the lip profile (A.) and the angle of the dividing streamline <P) relative to A..

Decreasing RL at smaller values of RL can increase the static pressure on the inlet lip.

This is due to the fact that there is a tendency for the region of stagnated or low­

velocity flow to "envelop" the lip, thus increasing overall static pressures on the lip.

299

4) A small amount of effective internal flow diffusion from the inlet throat to the duct

exit will increase the NR at the throat and hence reduce A.-~. This can be beneficial.

Alternatively an inlet with no internal flow diffusion can be run at higher NR to

achieve the same effect.

5) Although previous chapters have shown that the author's generic design offers an

improvement over conventional industrial designs, by virtue of the employment of a

circular inlet throat, if a becomes too large the flow in the inlet will begin to

separate. This will degrade the total pressure recovery efficiency and the uniformity

of flow at the duct exit. Therefore there is an effective hydrodynamic limit on a.

It is interesting to note that the conclusions drawn above relate well with the better

designs of waterjet inlet used in industry, in particular those conclusions relating to the

design of the inlet lip (RL,HL,y). Perhaps the reluctance to move to steeper inlets from

the range of a (20°-25°) typically used by industry, is due to the greater likelihood of

ramp flow separation on conventional designs. Although conventional inlet designs are

easy to fabricate, they do not offer the best resistance to ramp flow separation.

The hydrodynamic performance of the waterjet inlet is also a function of NR. Since

NR affects A.-f3 and hence the pressure distribution over the inlet lip, simply increasing

the operational NR of an existing inlet may eliminate cavitation (if it exists), improve

the margin against lip cavitation (if absent), or minimise the extent of cavitation on the

inlet lip. For conventional flush-type waterjet inlet designs, the relatively large radius of

curvature of the inlet ramp makes ramp cavitation unlikely.

It must be noted that the conclusions presented above are only applicable to the cruise

condition (NR=0.60). The manoeuvring condition (NR=oo) may give a totally different

shape of waterjet inlet, most likely having a larger radius of inlet lip. Roberts (1998)

used CFD to show that the use of an inlet lip of larger radius increased the total pressure

at the duct exit for the manoeuvring condition. The optimised waterjet inlet design with

its relatively sharp lip is likely to result in spillage at lower NR or significant internal

flow separation in the manoeuvring condition. Therefore an actual waterjet inlet must

reflect a compromise between the cruise and manoeuvring conditions, or at least give

300

adequate performance for manoeuvring. What is really required is a lip profile optimised

for both the cruise and manoeuvring conditions.

The following conclusions are made regarding the suitability of the SBA for waterjet

inlet optimisation:

1) The algorithm allows easy implementation of geometric constraints and is robust.

2) A marked improvement in minimum cavitation number is evident with only a few

iterations of the algorithm.

3) The use of the flow solution from a previous exploratory point or base point, as an

initial solution for a subsequent CFD computation, can dramatically decrease the

time required to obtain a flow solution and hence the computational effort of the

optimisation process.

4) The employment of strict convergence criteria will result in an excessive number of

iterations (and CFD computations) before convergence is finally achieved.

5) The sequential nature of the algorithm extends the time taken for the overall

optimisation process (on multi-processor machines) as CFD computations cannot be

run in parallel, as is the case for gradient-based optimisation methods. On single

processor machines this point is irrelevant.

It is therefore concluded that the optimisation algorithm of Sherif and Boice ( 1993) is an

effective tool for optimising the hydrodynamic design of waterjet inlets on a single­

processor computer, provided that the number of design variables to be varied during

the optimisation process is small (eg. 5-10).

301

Chapter 9 Conclusions and Recommendations

The objective of this research was to use computational fluid dynamic (CFD) as a tool

for analysing and optimising the design of marine waterjet propulsion unit inlets, in

order to gain a greater understanding of the hydrodynamics of waterjet inlet design. CFD

has been used effectively as a tool to develop a greater understanding of how the

geometric design of a flush-type waterjet inlet affects the resultant flow. This therefore

allows possible improvements in design to be identified. The design of the inlet lip was

found to be of particular importance to the elimination of cavitation at the cruise

condition (low inlet velocity ratio).

The conclusions of this research are therefore summarised in Section 9 .1. Section 9.2

contains a list of recommendations relating to the improvement of waterjet inlet design,

as well as recommendations for future CFD investigation of waterjet inlet flows.

9.1 Conclusions

The conclusions, of the research presented in this thesis, are summarised below.

9.1.1 Parametric Model

1) A lip loss thrust deduction fraction (tL) is introduced into the equation for waterjet

thrust to account for drag on the inlet lip resulting from flow separation and

cavitation, both of which act to decrease lip suction and hence thrust.

2) Theoretically, the ingestion of fluid from an upstream hull boundary layer acts to

increase the efficiency of the waterjet. This is primarily a result of increased

propulsor thrust resulting from a lower momentum drag associated with flow into the

inlet.

302

3) The optimum jet velocity ratio (JVR) for maximum efficiency decreases with the

ingestion of fluid from a boundary layer of greater thickness.

4) Pump rotative efficiency appears to have a significant effect on overall waterjet

efficiency for a given JVR. This indicates the importance of ensuring that the quality

of the flow delivered to the waterjet pump does not have an adverse effect on pump

operation and so decrease propulsive efficiency.

9.1.2 CFD Modelling

1) Good agreement has been obtained between CFD prediction and experimental data

for the three validation cases considered, particularly when flow separation is absent

from the flow. When flow separation is present, the predictive accuracy of the flow

computations was decreased.

2) CFD computations based on the solution of the Reynolds-averaged Navier Stokes

equations, with two-equation k-E turbulence modelling, can therefore be used as an

effective analysis tool for use in the design of waterjet inlets. The limitations of the

CFD modelling, in particular the turbulence and near-wall modelling, must be

recognised and considered in the interpretation of computational results.

3) Inaccuracies in the computational results obtained for the three validation cases are

influenced by a number of factors. These include inaccuracies in the modelling of the

boundary conditions, possible discretisation error (for the case of the waterjet inlet)

and the turbulence and near-wall modelling used.

4) The use of the RNG k-E turbulence model consistently gave more accurate results

than the Standard k-E model for the three validation cases considered. This

improvement in flow prediction, obtained via use of the RNG k-E model, is primarily

attributed to the rate of strain term in the equation for E. When the flow is subject to

rapid shear and streamline curvature, the inclusion of the rate of strain term results in

smaller predicted eddy-viscosity than when using the Standard k-E model. Therefore,

the effects of streamline curvature are more accurately modelled. This results in an

303

improvement in the prediction of mean flow behaviour.

5) The limitations of the turbulence models examined and their underlying assumption

of isotropic eddy-viscosity, arising from the hypothesis of Boussinesq, become

apparent when flow separation effects are modelled. This leads to inaccuracies in

predicted mean flow quantities. The assumptions underlying the use of conventional

wall functions breakdown under strong adverse pressure gradient and flow

separation. This clearly adds further inaccuracy to the modelling of the onset of flow

separation and separated flow behaviour.

6) The topology of the structured grid used for meshing the waterjet inlet and a simple

flow domain external to it, produces a good quality of mesh in terms of minimising

the number of skewed cells in the flow domain and ensuring orthogonality of grid

lines at the wall. The use of a single-block structured mesh does, however, impose

limitations on the detail of the geometry that can be meshed. In other words, the

geometry to be meshed must be simplified and features such as the impeller shaft

housing/fairing omitted. Furthermore, extension of the flow domain to include the

vessel hull becomes difficult.

9.1.3 Design Subspace Investigation and Optimisation Methodologies

1) CFD offers the only cost-effective means by which awaterjet inlet can be optimised.

2) Systematic investigation of the hydrodynamic design of the waterjet inlet and its

optimisation require a parametric description of a generic waterjet inlet geometry.

This allows a systematic and logical variation of the geometry in order to correlate

the hydrodynamics of the flow with the underlying geometry.

3) Grid generation typically represents the "bottle-neck" in the CFD analysis cycle.

Since design and optimisation-related work require a large number of CFD

computations, it is therefore necessary to automate the mesh generation process as

much as is possible. This is necessary in order to minimise the amount of time

devoted to mesh generation.

304

4) Investigation of two-parameter subs paces of the parametric hyperspace associated

with a generic parametrically-defined waterjet inlet geometry allows useful

hydrodynamic information (relevant to that generic design) to be gathered.

Furthermore, a direct correlation between the hydrodynamics and the underlying

geometry becomes apparent, once sufficient data has been accumulated and analysed.

5) The direct-search optimisation algorithm of Sherif and Boice (1993) can be used as a

tool for the optimisation of waterjet inlet design. This is due to the algorithm's

robustness, relative efficiency and the ease by which geometric constraints (placed on

the generic parametric geometry) can be implemented. Some liberty must be taken in

the specification of the convergence criteria of this algorithm, in order to avoid an

excessive number of iterations. This algorithm is best suited for use on single­

processor machines due to its sequential nature.

6) The computational expense of the large number of CFD computations required for

waterjet inlet optimisation, is partially mitigated by using previous solutions as initial

solutions for new calculations.

9.1.4 Effect of Upstream Boundary Layer

1) The thickness of the boundary layer upstream of the inlet was found to have a

significant effect on the flow in the waterjet inlet.

2) The minimum static pressure at the inlet lip increases in the presence of thicker

upstream boundary layers. This is primarily attributed to the decreased momentum of

the flow in the vicinity of the inlet lip.

3) For a given mass flux ingested by the inlet, both the width and the depth of the cross­

section of the inlet streamtube increase, as a result of lower average velocities over

the streamtube cross-section.

4) There is a decrease in the total pressure of the fluid at the duct exit with increasing

upstream boundary layer thickness. This is correlated with a reduction in the ingested

305

energy flux across the cross-section of the inlet streamtube.

5) The distortion of the total pressure distribution at the duct exit is strongly dependent

on the thickness of the upstream boundary layer. Except for relatively thin upstream

boundary layers, there is a general trend toward decreasing distortion with increasing

boundary layer thickness for the author's generic waterjet inlet geometry. This result

may not necessarily apply to other geometries.

6) Accurate representation of the actual shape and dimensions of the cross-section of the

inlet streamtube is essential for accurate calculation of ingested momentum and

energy fluxes. The assumption of a rectangular or semi-elliptical profile will lead to

inaccurate calculation of these quantities.

9.1.5 Waterjet Inlet Design

The conclusions presented below are obtained from an analysis of data for waterjet

inlets simulated at an inlet velocity ratio (IVR) of 0.6 in the presence of a thick upstream

boundary layer. Therefore, care must be taken when seeking to generalise these results

to other IVR values.

1) At low IVR values (eg. IVR=0.60) the minimum static pressure occurs on the

underside of the inlet lip, with a risk of flow cavitation.

2) The minimum static pressure on the inlet lip is largely dependent upon the deviation

between the angle of the bisector of the lip profile (A) and the angle of the dividing

streamline (~), as discussed in Chapter 7. Therefore, the larger the magnitude of A.-~,

the lower the minimum static pressure on the lip will be. An objective of inlet lip

design must therefore be to minimise the magnitude of the angle of attack of the local ?

flow relative to the lip (ie. minimise A.-~).

3) The use of a raised-lip profile (where the bottom of the inlet lip is raised above the

base of the inlet opening) was found to be a beneficial design feature for increasing

the minimum static pressure on the lip, by virtue of directing the flow over the lip in

306

such a way as to decrease A-~. The minimum static pressure on the lip is found to

correlate with a lip profile parameter, yHt/D.

4) Increasing the steepness of the inlet (a) was also found to be an effective means of

reducing the magnitude of A-~ and leads to a smaller, more compact inlet. The

disadvantage of this is a greater likelihood of flow separation on the upper inlet ramp

if a becomes too large.

5) When A-~ is minimised (by virtue of a raised-lip profile), inlet lips of smaller radii

appear to be beneficial in raising the minimum static pressure on the lip. This effect

arises because the region of stagnated and low velocity fluid surrounding the

stagnation point on the lip tends to envelop the inlet lip and so leads to higher static

pressures on the lip surface. At high IVR, as in a low-speed or manoeuvring

condition, a sharper lip profile is likely to be detrimental to inlet performance leading

to flow separation within the inlet. Furthermore, at very low IVR, a sharper lip profile

is likely to lead to flow separation on the outside of the inlet lip (spillage drag). These

issues have not been explored and so no further comments can be made.

6) A moderate increase in the IVR of the inlet (for example from 0.6 to 0.8) can provide

an effective means of minimising the magnitude of A-~, thus increasing the minimum

static pressure on the lip.

7) The author's generic waterjet inlet geometry requires a smoother transition between

the inlet and hull (downstream of the lip near the centreplane) when a raised lip

profile is used, in order that this location does not become a potential source of

cavitation at high vessel speed.

8) The author's generic waterjet inlet design, which employs a circular inlet throat,

offers greater resistance to flow separation in the upper inlet than conventional inlet

designs (which use a flat curved ramp surface). The author's generic geometry,

however, is likely to be more difficult to fabricate than conventional designs by

virtue of the doubly-curved nature of the inlet surface.

307

9) A larger radius of curvature of an inlet ramp results in a higher static pressure on the

ramp surface and hence a greater resistance to flow cavitation. This is beneficial at

high vessel speed, but comes at the expense of a greater volume of the waterjet inlet.

10) The distortion of total pressure at the duct exit appears to depend on the influence of

the duct bend or internal flow diffusion (if present) acting on the boundary layer

developed upstream of the bend. In all cases examined, a region of low total pressure

occurs in the upper part of the duct at the duct exit, whereas a region of higher total

pressure occurs in the lower duct. It is therefore doubtful whether a synergistic

combination of geometric parameters can be found that can significantly reduce the

distortion of total pressure at the duct exit, within acceptable geometric limits. The

effect of these distributions on pump performance is unknown and beyond the scope

of the work presented in this thesis.

11) Partial diffusion of the flow inside the waterjet inlet, between the inlet throat and

duct exit, appears to offer no hydrodynamic advantage, except as a means of

minimising A.-p. The adverse effects of internal flow diffusion include boundary

layer thickening, flow separation (with sufficiently large internal diffusion) and

increased distortion of total pressure at the duct exit. Another adverse effect is a

stronger secondary flow within the duct, by virtue of greater cross-stream pressure

gradients associated with the duct bend.

12) The dimensions of the cross-section of the inlet streamtube were shown to be

dependent upon the width of the inlet opening, the mass flux ingested by the

waterjet inlet, as well as the thickness of the upstream boundary layer.

13) The vertical forces acting on the waterjet inlet were shown to be dependent upon the

change in the vertical momentum of the flow through the waterjet inlet, vertical

forces acting on the inlet lip (below the stagnation line), and hydrostatic forces

arising from the volume of water entrained by the waterjet inlet. The change in the

vertical momentum of flow through the waterjet inlet is related to the geometry of

the inlet streamtube which is influenced by such factors as the geometry of the inlet,

308

the ingested mass flux and the upstream boundary layer profile.

9.2 Recommendations

The recommendations for future research and waterjet inlet design, arising from the

conclusions reached as a result of the work presented in this thesis, are summarised

below.

9.2.1 CFD Modelling

1) The ability to mesh the complete geometry of the waterjet inlet, without geometric

simplifications and the omission of components (such as the impeller shaft

housing/fairing), necessitates the use of CFD techniques that solve the Reynolds­

averaged Navier Stokes equations on unstructured grids. The use of unstructured grid

technology also allows the meshing of any waterjet-hull system. It is therefore

recommended that future research be directed toward the application of unstructured

CFD technology to waterjet inlet analysis, in order to include the impeller shaft

housing/fairing in the flow simulation. Furthermore, unstructured CFD technology

also offers the possibility of a complete waterjet flow simulation including the

impeller shaft housing, pump impeller, stator blades and the nozzle itself.

2) Although the RNG k-E turbulence model represents an improvement over the

Standard k-E model, there is scope for using turbulence models that offer greater

predictive accuracy for flows subject to adverse pressure gradients and turbulent flow

behaviour unrelated to mean strain rates (such as streamline curvature and separated

flow).

3) The benefits of using an automated approach to the meshing of the flow domain were

discussed as a means of reducing the CFD analysis cycle time. This is crucial when a

large number of computations are required. It is therefore recommended that future

work with unstructured grid technology be focused on automating the grid generation

process as far as is practicable, in order to obtain suitable meshes for CFD

simulation.

309

9.2.2 Waterjet Inlet Design

1) The waterjet inlet was optimised for maximum static pressure on the inlet surface at a

cruise condition corresponding to an IVR of 0.6 (Chapter 8). The performance of this

optimised waterjet inlet at the low speed/manoeuvring condition (IVR>>1) is

unknown. It is however likely that there will be significant flow separation at the inlet

lip as flow enters the inlet from underneath the lip. Satisfactory hydrodynamic

performance at this condition may give a totally different lip profile and inlet shape.

Therefore, an actual waterjet inlet must ideally reflect a compromise between the

cruise and low speed/manoeuvring condition, or at least give adequate performance

for manoeuvring. Investigation into this issue is therefore recommended.

2) A single generic lip profile was investigated. The generic profile investigated

essentially consisted of a circular lip vertically offset from the inlet opening. This

generic profile is by no means optimum and its limitations were shown. It is therefore

recommended that further research be undertaken in order to develop generic inlet lip

profiles offering improved resistance to the inception of cavitation.

3) In order to minimise the angle of attack of the local flow in the vicinity of the inlet

lip, relative to the inlet lip (ie. minimise A.-p), over a range of cruise IVR values, an

adjustable (moveable) inlet lip design could be used. An auxiliary inlet area could be

used for the low-speed manoeuvring condition (in order to minimise flow separation

at the inlet lip). Barr and Etter (1974) suggested the use of a variable-area inlet with a

movable inlet lip and ramp in order to provide satisfactory operation over a range of

IVR values. They noted that by using a moveable lip, the angle of attack of the flow

relative to the lip can be minimised thus leading to reduced lip drag and the use of

smaller lip radii.

4) In view of the conclusions reached regarding the distortion of total pressure at the

duct exit, it is the author's view that the only effective means of reducing the

distortion of the total pressure distribution at the duct exit is by mechanical means.

English (1994) suggested a freely-rotating turbine positioned on the impeller shaft

housing/fairing as a means of effecting a "homogenisation" of the velocity field

310

upstream of the pump. Seddon and Goldsmith ( 1985) discussed a similar idea for

aeronautical applications. While this idea theoretically sounds good, there are,

however, concerns about turbine cavitation. Further investigation of different

mechanical means of effecting flow homogenisation upstream of the waterjet pump

is therefore recommended.

5) The use of a circular inlet throat and a waterjet inlet shape that tends to "converge"

the flow toward the throat of the inlet should offer better resistance to flow separation

on the inlet ramp than conventional inlet designs. This was shown to be the case for

the author's generic geometry. Such an inlet design does, however, come with greater

fabrication difficulty. Since conventional inlet designs (used on large high-speed

vessels) tend to be easier to fabricate, an alternative approach for use on conventional

inlets with a curved flat ramp, would be to use a boundary layer re-energisation

technique such as tangential blowing. Burley and Hwang (1983) discussed this

approach as a means of increasing the separation-free operation of subsonic V/STOL

aircraft inlets. Griffith-Jones (1994) suggested further investigation into the potential

of this method of boundary layer control.

Therefore research into boundary layer re-energisation techniques is recommended,

as this can ultimately lead to improvements in the hydrodynamic performance of

existing designs or to steeper, more compact waterjet inlet designs.

311

References

Aartojarvi, R. (1995) "Noise from Waterjet Propulsion," Proceedings of the Third International Conference on Fast Sea Transportation (FAST'95), Ltibeck-Travemtinde, Germany, vol.2, 1157-1166.

Alexander, K. V. and van Terwisga, T. (1993) "Recent Work on Waterjet-Hull Interaction Effects," Proceedings of the 9th International High Speed Surface Craft Conforence, Singapore.

Alexander, K., Coop, H. and van Terwisga, T. (1994) "Waterjet-Hull Interaction: Recent Experimental Results," Transactions of the Society of Naval Architects and Marine Engineers (SNAME), vol. 102, 87-105.

Allison, J. L. (1993) "Marine Waterjet Propulsion," Transactions of the Society of Naval, Architects and Marine Engineers (SNAME), vol. 101,275-335.

Allison, J. L. ( 1996) "CFD Simulation Saves $100,000 in Design of New marine Landing Craft," BLA Quarterly Digest, Band, Lavis & Associates, Inc., vol. 6, Sept 1996.

Allison, J. and Dai, C. (1994) "Steering and Reversing Gear for Very Large Waterjets," Proc. RINA International Symposium on Waterjet Propulsion-Latest Developments, The Royal Institute of Naval Architects (RINA), London, 1-2 Dec, paper no. 7.

Allison, J. Land Goubault, P. (1995) "Waterjet Propulsion for Fast Craft- Optimized Integration of Hull and Propulsor," Proceedings of the Third International Conference on Fast Sea Transportation (F AST'95), Lubeck-Travemtinde, Germany, vol.2, 1117-1130.

Anderson, B. H., Reddy, D. R.and Kapoor, K. (1994) "Study on Computing Separating Flows within a Diffusing InletS-Duct," Journal of Propulsion and Power, vol. 10, no. 5, Sept-Oct 1994, 661-667.

Baldwin, B.S. and Lomax, H. (1978) 'Thin-Layer Approximation and Algebraic Model for Separated Turbulent Flows," AIAA Paper 78-257, Huntsville, AL.

Bansod, P. and Bradshaw, P. (1972) "The Flow in S-shaped Ducts," The Aeronautical Quarterly, The Royal Aeronautical Society, vol. 23, May 1972, 131-140.

Barr, R. A. and Etter, R. J. (1974) "Selection of Propulsion Systems for High Speed Advanced Marine Vehicles," AIAA, 74-334.

Boussinesq, J. (1877) "Theorie de l'Ecoulement Tourbillant," Mem. Presentes par Divers Savants A cad Sci. Inst. Fr., vol. 23, 46-50.

312

Bradshaw, P. (1967) "The Turbulence Structure of Equilibrium Boundary Layers," J Fluid Mechanics, vol. 29, part 4, 625-645.

Bradshaw, P. (1973) "Effect of Streamline Curvature on Turbulent Flow", AGARD-AG-169.

Brandau, J. H. ( 1967) "Aspects of Performance Evaluation of Waterjet Propulsion Systems and a Critical Review of the State-of-the-Art," AIAAISNAME Advance Marine Vehicle Meeting, Norfolk, Virginia, May 22-24.

Breslin, J. P. (1970) "Theoretical and Experimental Techniques for Practical Estimation of Propeller-Induced Vibratory Forces," Transactions of the Society of Naval Architects and Marine Engineers (SNAME), vol. 78,23-40.

Brilliant, H. M. and Bauer, C. A. (1977) "Comparison of Estimated with Measured Maximum Instantaneous Distortion using Flight Data from an Axisymmetric Mixed Compression Inlet," AIAA 77-876.

Burley, R. R. and Hwang, D.P. (1983) "Investigation of Tangential Blowing Applied to a Subsonic V/STOL Inlet,", J of Aircraft, vol. 20, no. 11, Nov. 1983, 926-934.

Cebeci T. and Smith, A.M. 0. (1974) "Analysis of Turbulent Boundary Layers," Ser. in Appl. Math. & Mech., vol. XV. New York: Academic Press.

Chambers, T. L. and Wilcox, D. C. (1977) "Critical Examination of Two-Equation Turbulence Closure Models for Boundary Layers," AIAA Journal, vol. 15, no. 6, 821-828.

Coles, D. E. and Hirst, E. A. (1968) Computation ofTurbulent Boundary Layers- 1968 AFOSR-IFP-Stanford Conference. Vol. 11 California: Stanford University.

Coop, H. G. and Bowen, A. J. (1993) "Hull-Waterjet Interaction Mechanisms: Theory and Validation," Proceedings of the 2nd International Conference on Fast Sea Transportation (FAST'93), Yokohama, Japan, vol.1, 855-866.

Dai, C., Kerr, C., Nguyen, P. and Wang, H. C. (1995), "A Novel Flush Inlet Design Design Methodology for Waterjet Propulsion," Proceedings of the Third International Conference on Fast Sea Transportation (FAST'95), Ltibeck-Travemtinde, Germany, vol.2, 1367-1378.

Doragh, R. A. (1963) "Magnetohydrodynamic Ship Propulsion using Superconducting Magnets," Transactions of the Society of Naval Architects and Marine Engineers (SNAME), vol. 71, 370-386.

Doss, E. D. and Geyer, H. K. (1993) "MHD Seawater Propulsion," Journal of Ship Research, vol. 37, no. 1, Mar 1993, 49-57.

Driver, D. M. (1991) "Reynolds Shear Stress Measurements in a Separated Boundary Layer," A.IAA-91-1787.

313

Driver, D. M. and Seegmiller, H. L. (1985) "Features of a Reattaching Turbulent Shear Layer in Divergent Channel Flow," AIAA Journal, vol. 23, no. 2, Feb 1985, 163-171.

Dubbel- Taschenbuch fur den Maschinenbau (1990), edited by W. Beitz and K.-H. KUttner, 17 Edition. Berlin: Springer-Verlag.

Dyne, G. and Lindell, P. (1994) "Waterjet Testing in the SSPA Towing Tank," Proc. RINA International Symposium on Waterjet Propulsion - Latest Developments, The Royal Institute of Naval Architects (RINA), London, 1-2 Dec, 1994, paper no. 2.

Enayet, M. M., Gibson, M. M., Taylor, A.M. K. P. and Yianneskis, M. (1982) "Laser­Doppler Measurements of Laminar and Turbulent Flow in a Pipe Bend," International Journal of Heat and Fluid Flow, vol. 3, no. 4, Sept. 1982, 213-219.

English, J. W. (1974) "Hydrodynamic Considerations in the Design of Condenser Cooling Water Systems for Large Ships," Trans. RINA, vol. 116, 31-46.

English, J. W. (1994) "Practical Considerations on Waterjet Inlets with Flush Intakes," Proc. RINA International Symposium on Waterjet Propulsion -Latest Developments, The Royal Institute of Naval Architects (RINA), London, 1-2 Dec, 1994, paper no. 11.

Eriksson, L. E. (1982) "Generation of Boundary-Conforming Grids around Wing-Body Configurations using Transfinite Interploation," AIAA Journal, vol. 20, no. 10, Oct 1985, 1313-1320.

Eriksson, L. E. (1985) "Practical Three-Dimensional Mesh Generation using Transfinite Interpolation," SIAM J. Sci. Stat. Comput., Society for Industrial and Applied Mathematics (SIAM), vol. 6, no. 3, July 1985,712-741.

Eriksson, L. E. and Orbekk, E. ( 1989) "Application of Mesh Generation to Complex 3-D Configurations," AGARD CP 464.

Etter, R. J., Krishnamoorthy, V. and Scherer, J. 0. (1982) "Model Testing of Waterjet Propelled Craft," Proceedings of the 19th American Towing Tank Committee (ATTC), 783-806.

Fletcher, C. A. J. (1987a) Computational Techniques for Fluid Dynamics. Vol. 1. Berlin: Springer Verlag.

Fletcher, C. A. J. (1987b) Computational Techniques for Fluid Dynamics. Vol. 2. Berlin: Springer Verlag.

Fluent Inc. (1996) Fluent V4.4 Users Guide. Lebanon, New Hampshire: Fluent Inc.

F~rde, M., 0rbekk, E. and Kubberud, N. (1991) "Computational Fluid Dynamics Applied to High Speed Craft with Special Attention to Water Intake for Water Jets," Proceedings of the 1st International Conference on Fast Sea Transportation (FAST'91), Trondheim, Norway, vol.1, 69-89.

314

Frank, P. D. and Shubin, G. R. (1992) "A Comparison of Optimization-Based Approaches for a Model Computational Aerodynamics Design Problem,", Journal of Computational Physics, vol. 98, 74-89.

Gallin, C. M., Phipps, J. H., Stevenson, J. and van Terwisga, T. (1995) "Technically and Economically Optimized Fast Ships Propulsion Systems from 18000 to 30000 kW," Proceedings of the Third International Conference on Fast Sea Transportation (FAST'95), Liibeck-Travemiinde, Germany, vol.2, 725-753.

Gany, A. (1993) "The Bubbly Waterjet - a Promising High Speed Marine Propulsion Concept," Proceedings- The Ninth High Speed Surface Craft Conference, 9-11 Mar. High Speed Surface Craft Associates.

Gerhart, P. M. and Gross, R. J. (1985) Fundamentals of Fluid Dynamics. Reading, Massachusetts: Addison-Wesley.

Gill, P. E., Murray, W. and Wright, M. H. (1981) Practical Optimization. New York: Academic Press.

Gillis, J. C. and Johnston, J. P. (1983) "Turbulent Boundary-Layer Flow and Structure on a Convex Wall and its Redevelopment on a Flat Wall," Journal of Fluid Mechanics, vol. 135, 123-153.

Gordon, W. J. and Hall, C. A. (1973) "Construction of Curvilinear Co-ordinate Systems and Applications to Mesh Generation," International Journal for Numerical Methods in Engineering, vol. 7, 461-477.

Gottfried, B. S. and Weisman, J. (1973) Introduction to Optimization Theory. Englewood Cliffs, New Jersey: Prentice-Hall.

Griffith-Jones, G. J. (1994) Investigation of Incompressible Flow through an Intake Duct with Applications to Waterjet Propulsion. Christchurch, New Zealand: University of Canterbury, Dept. of Mechanical Engineering.

Guo, R. W. and Seddon, J. (1982) "An Investigation of the Swirl in an S-Duct," Aeronautical Quarterly, Feb 1982, 25-58.

Guo, R. W. and Seddon, J. (1983a) "The Swirl in an S-duct of Typical Air Intake Proportions," Aeronautical Quaterly, Royal Aeronautical Society, vol. 34, part 2, May 1983, 99-129.

Guo, R. W. and Seddon, J. (1983b) "Swirl Characteristics of an S-shaped Air Intake with Horizontal and Vertical Offsets," Aeronautical Quaterly, Royal Aeronautical Society, vol. 34, part 2, May 1983, 130-146.

Haglund, K., Svensson, R. and Bjorheden, 0. (1982) "Design and Testing of a High­Performance Water Jet Propulsion," Proceedings of the Second Symposium on Small Fast Warships and Security Vessels, The Royal Institute of Naval Architects, paper 17, May 1982, 223-237.

315

Hanjalic, K. and Launder, B. E. (1980) "Sensitizing the Dissipation Equation to Irrotational Strains," Transactions of the ASME, Journal of Fluids Engineering, vol. 102, Mar 1980, 34-40.

Harloff, G. J., Smith, C. F., Bruns, J. E. and DeBonis, J. R. (1993) "Navier-Stokes Analysis of Three-Dimensional S-Ducts," Journal of Aircraft, vol. 30, no. 4, Jul-Aug. 1993, 526-533.

Ratte, R. and Davis, H. J. ( 1967) "Selection of Hydrofoil Waterjet Propulsion Systems," Journal ofHydronautics, vol. 1, no. 1, 12-22.

Hess, J. L. and Smith, A.M. 0. (1964) "Calculation ofNonlifting Potential Flow about Arbitrary Three-Dimensional Bodies," Journal ofShip Research., vol. 8, no. 2, 22-44.

Hewins, E. F. and Reilly, J. R. (1940) "Condensor Scoop Design," Transactions ofthe Society of Naval Architects and Marine Engineers (SNAME), vol. 48,277-304.

Hinze, J. 0. (1975) Turbulence. Second Edition. New York: McGraw-Hill.

Hooke, R. and Jeeves, T. A. (1961) "Direct Search Solution of Numerical and Statistical Problems," Journal of the Association/or Computing Machinery, vol. 8, no. 2, 212-229.

Hoshino, T. and Baba, E. (1984) "Self-propulsion Test of a Semi-Displacement Craft Model with a Waterjet Propulsor," Journal ofthe Society of Naval Architects of Japan, vol.155, 50-57.

Hutchinson, B. R. and Raithby, G. D. (1986) "A Multigrid Method Based on the Additive Correction Strategy," Numerical Heat Transfer, vol. 9, 511-537.

ITTC (1987) "Report of the High-Speed Marine Vehicle Committee - Waterjet Test Procedures," International Towing Tank Confernece 1987, 275-313.

Jameson, A. (1995) "Optimum Aerodynamic Design Using Control Theory," Department of Mechancial and Aerospace Engineering, Princeton University, Princeton, New Jersey.

Jenkins, R. C. and Loeffler, A. L. (1991) "Modelling of Subsonic Flow Through a Compact Offset Inlet Diffuser," AIAA Journal, vol. 29, no. 3, Mar 1993,401-408.

Jones, W. P. and Launder, B. E. (1972) "The Prediction of Laminarization with a Two­Equation Model of Turbulence," International Journal of Heat and Mass Transfer, vol. 15, 301-314.

Kashiwadani, T. (1985) "The Study on the Configuration ofWaterjet Inlet (1st Report)," Journal of the Society of Naval Architects of Japan, vol. 157, 121-140.

Kashiwadani, T. (1986) "The Study on the Configuration of Waterjet Inlet (2nd Report)," vol. 159, 118-128.

316

Kashiwadani, T. (1997) "On the Design Method of Waterjet Inlet for SES 'Meguro 2' ," Proceedings of the Fourth International Conference on Fast Sea Transportation (FAST'97), Sydney, Australia, vol.2, 677-684.

Kitchen, R. A. and Sedlock, D. (1983) "Subsonic Diffuser Development of Advanced Tactical Aircraft," AIAA-83-1168.

Lakshminarayana, B. (1986) "Turbulence Modeling for Complex Shear Flows," AIAA Journal, vol. 24, no. 2, 1900-1917.

Launder, B. E. and Spalding, D. B. (1974) "The Numerical Computation of Turbulent Flows," Computer Methods in Applied Mechanics and Engineering, vol. 3, 269-289.

Launder, B. E., Priddin, C. H. and Sharma, B. I. (1977) "The Calculation of Turbulent Boundary Layers on Spinning and Curved Surfaces," Transactions of ASME, Journal of Fluids Engineering, vol. 99, Mar 1977,231-239.

Leonard, B. P. (1979) "A Stable and Accurate Convective Modelling Procedure Based on Quadratic Upstream Interpolation," Computer Methods in Applied Mechanics and Engineering, vol. 19, 59-98.

Leschziner, M.A. and Rodi, W. (1981) "Calculation of Annular and Twin Parallel Jets using Various Discretisation Schemes and Turbulence-Model Variations," Tansactions of ASME, Journal of Fluids Engineering, vol. 103, Jun 1981, 352-360.

Levy, J. (1965) "The Design of Water-Jet Propulsion Systems for Hydrofoil Craft," Marine Technology, vol. 2, Jan 1965, 15-25,41.

Lewis, F. M. (1963) "Propeller Vibratory Forces," Transactions of the Society of Naval, Architects and Marine Engineers (SNAME), vol. 71,293-326.

Little, B. H. and Trimboli, W. S. (1982) "An Experimental Investigation of S-Duct Diffusers for High-Speed Prop Fans," AIAA-82-1123.

Liu, D. C. and Nocedal, J. (1989) "On the Limited Memory BFGS Method for Large Scale Optimization," Mathematical Programming, vol. 45, no. 3, 503-528.

Melick, H. C., Ybarra, A. H. and Bencze, D. P. (1975) "Estimating Maximum Instantaneous Distortion from Inlet Total Pressure RMS and PSD Measurements," AIAA 75-1213.

Melick, H. C., Ybarra, A. H. and Bencze, D. P. (1978) "Estimating Maximum Instantaneous Distortion from Inlet Total Pressure RMS Measurements," AIAA 78-970.

Mortenson, M. E. (1985) Geometric Modelling. New York: John Wiley & Sons Inc.

Motora, S., Takezawa, S. and Tamama, H. (1991) "Research and Development of Superconducting Magnetohydrodynamic Ship Propulsion," !MAS 91 High Speed Marine Transportation, 11-13 Nov 1991, Paper 6, 49-58.

317

Mount, J. S. (1965) "Effect of Inlet Additive Drag on Aircraft Performance," Journal of Aircraft, American Institute of Aeronautics and Astronautics, vol. 2, no. 5, May 1965, 374-378.

Naser, J. A. (1990) Flow Prediction in Valve/Port Assemblies, PhD Thesis, Dept. of Mechanical Engineering, Imperial College of Science, Technology and Medicine, The University of London.

Okamoto, Y., Sugioka, H. and Kitamura, Y. (1993) "On the Pressure Distribution of a Water Jet Intake Duct in Self Propulsion Conditions," Proceedings of the 2nd International Conference on Fast Sea Transportation (FAST'93), Yokohama Japan, vol.1, 843-854.

Owen, P. M. (1962) "The Prospects for Magnetohydrodynamic Ship Propulsion," Journal of Ship Research, vol. 5, no. 4, Mar 1962,43-51.

Patankar, S. V. (1980) Numerical Heat Transfer and Fluid Flow. New York: Hemisphere.

Patankar, S. V. and Spalding, D. B. (1972) "A Calculation Procedure for Heat, Mass and Momentum Transfer in Three-Dimensional Parabolic Flows," International Journal of Heat and Mass Transfer, vol. 15, 1787-1806.

Purnell J. G. (1976) "The Performance Gains of using Wide, Flush Boundary Layer Inlets on Water-Jet Propelled Craft," Research and Development Report (Report PAS-75-45), Propulsion and Auxiliary Systems Department, David W. Taylor Naval Ship Research and Development Center, Annapolis.

Roache, P. J. (1976) Computational Fluid Dynamics. Albuquerque, New Mexico: Hermosa Publishers.

Roberts, J. L. (1998), The Influence of Hull Boundary Layers on Waterjet Intake Performance, PhD Thesis to be submitted in 1998. Department of Civil and Mechanical Engineering, The University of Tasmania.

Rodi, W. and Scheuerer, G. (1986) "Scrutinizing the k-E Turbulence Model under Adverse Pressure Gradient Conditions," Transactions of ASME, Journal of Fluids Engineering, vol. 108, Jun 1986, 174-179.

Rowe, M. ( 1970) "Measurements and Computations of Flow in Pipe Bends," J Fluid Mechanics, vol. 43, part 4, 771-783.

Roy, S. M. (1994) "The Evolution of the Modem Waterjet Marine Propulsion Unit," Proc. RINA International Symposium on Waterjet Propulsion -Latest Developments, The Royal Institute of Naval Architects (RINA), London, 1-2 Dec, paper no.1.

Schlichting, H. (1979) Boundary Layer Theory. Seventh Edition. New York: McGraw­Hill.

318

Schultz, W. B. (1974a) "Boeing Jetfoil Progress Report," International Hydrofoil Society, 291-299.

Schultz, W. B. (1974b) "Key Problems Associated with Developing the Boeing Model 929-100 Commercial Passenger Jetfoil," International Hydrofoil Society, 301-306.

Seddon, J. (1984) "Understanding and Countering the Swirl in S-ducts: Tests on the Sensitivity of Swirl to Fences," Aeronautical Journal, Royal Aeronautical Society, vol. 88, no. 874, Apr 1984, 117-127.

Seddon, J. and Goldsmith E. L. (1985) Intake Aerodynamics. London: Collins Professional and Technical Books.

Seil, G. J., Fletcher, C. A. J. and Doctors, L. J. (1995) "The Application of Computational Fluid Dynamics to Practical Waterjet Propulsion System Design and Analysis," Proceedings of the Third International Conference on Fast Sea Transportation (FAST'95), Liibeck-Travemiinde, Germany, vol.2, 1379-1390.

Seil, G. J., Fletcher, C. A. J., and Doctors, L. J. (1997) "Optimisation of Waterjet Inlets using Computational Fluid Dynamics," Proceedings of the Fourth International Conference on Fast Sea Transportation (FAST'97), Sydney, Australia, vol.1, 59-64.

Sherif, Y. S. and Boice, B. A. (1993) "An Efficient Algorithm for Solving the Unconstrained Nonlinear Multivariable Minimization Problems," Advances in Engineering Software, vol. 17, 29-37.

Smith, R. E. (1982) "Algebraic Grid Generation," Numerical Grid Generation editied by J. Thompson, 137-170.

Sorenson, R. L. (1982) "Grid Generation by Elliptic Partial Differential Equations for a Tri-Element Augmentor-Wing Airfoil," Numerical Grid Generation edited by J. Thompson, 653-665.

Spalding, D. B. (1961) "A Single Formula for 'the Law of the Wall'," Journal of Applied Mechanics, vol. 28, Sept 1961,455-457.

Squire, H. B. and Winter, K. G. (1951) "The Secondary Flow in a Cascade of Airfoils in a Nonuniform Stream," Journal of the Aeronautical Sciences, Apr 1951, 271-277.

Steen, S. and Minsaas, K. J. (1995) "Experiences from Design and Testing of Waterjet Inlets for High Speed Craft," Proceedings of the Third International Conference on Fast Sea Transportation (FAST'95), Liibeck-Travemiinde, Germany, vol.2, 1255-1270.

Svensson, R. (1987) "Water Jets- Propulsion for Naval Vessels,", Navy International, Jul/Aug 1987,411-417.

Svensson, R. (1991) "A Description of the Water Jets Selected for Destriero," Proceedings ofthe 1st International Conference on Fast Sea Transportation (FAST'91), Trondheim, Norway, vol.2, 1169-1184.

319

Svensson, R. (1993) "Experience from Operation of Large Water Jet Units," Proceedings of the 2nd International Conference on Fast Sea Transportation (FAST'93), Yokohama, Japan, vol.2, 1409-1421.

Svensson, R. (1994) "Water Jet Propulsion of High-Speed Passenger Vessels," Shipbuilding Technology International '94, edited by R. Burnett, London, Sterling Publications Limited, 155-160.

Swallom, D. W., Sadovnik, I., Gibbs, J, Gurol, H, Nguyen, L. V. and van den Bergh, H. H. (1991) "Magnetohydrodynamic Submarine Propulsion Systems," Naval Engineers Journal, The Society of Naval Engineers, vol. 103, no. 3, May 1991, 141-157.

Taylor, A.M. K. P., Whitelaw, J. H. and Yianneskis, M. (1982a) "Curved Ducts with Strong Secondary Motion: Velocity Measurements of Developing Laminar and Turbulent Flow," Trans. ASME, Journal of Fluids Engineering, vol. 104, Sept 1982, 350-359.

Taylor, A.M. K. P., Whitelaw, J. H. and Yianneskis, M. (1982b) "Developing Flow in S-Shaped Ducts I- Square Cross-Section Duct," NASA CR 3550.

Taylor, A.M. K. P., Whitelaw, J. H. and Yianneskis, M. (1984) "Developing Flow inS­Shaped Ducts II- Circular Cross-Section Duct,", NASA CR 3759.

Tennekes, H. and Lumley, J. L. (1972) A First Course in Turbulence. Massachusetts: MIT Press.

Thomas, P. D. (1982) "Numerical Generation of Composite Three Dimensional Grids by Quasilinear Elliptic Systems," Numerical Grid Generation edited by J. Thompson, 667-686.

Thompson, J. F., Warsi, Z. U. A. and Mastin, W. C. (1985) Numerical Grid Generation -Foundations and Applications. New York: Elsevier.

Tomeman, G. (1994) "The Importance of High Manoeuvrability for Efficient Operation", Proc. RINA International Symposium on Waterjet Propulsion - Latest Developments, The Royal Institute of Naval Architects (RINA), London, 1-2 Dec., 1994, paper no. 13.

Towne, C. E. (1984) "Computation of Viscous Flow in Curved Ducts and Comparison with Experimental Data," AIAA-84-053I.

Trillo, R. L. (1994) Jane's High-Speed Marine Craft 1993-94. Twenty-Sixth edition. Surrey, Coulsdon: Jane's Information Group.

Tsakonas, S., Breslin, J. and Miller M. (1967) "Correlation and Application of an Unsteady Flow Theory for Propeller Forces," Transactions of the Society of Naval, Architects and Marine Engineers (SNAME), vol. 75 158-193.

320

Vakili, A., Wu, J. M., Hingst, W. R. and Towne, C. E. (1984) "Comparison of Experimental and Computational Flow in aS-Duct," AIAA-84-0033.

Van Terwisga, T. (1996) Waterjet-Hu/1 Interaction, PhD Thesis, Technical University of Delft.

Venturini, G. (1980) "Waterjet Propulsion in High Speed Surface Craft," High-Speed Surface Craft Confernece, Sussex, UK, June 24-27, 1980, 125-142.

Verbeek, R. (1992) "Application of Waterjets in High-Speed Craft," Hydrodynamics: Computations, Model Tests and Reality. Elsevier Science Publishers B.V. 1992, 133-138.

Vinokur, M. (1983) "On One-Dimensional Stretching Functions for Finite-Difference Calculations," Journal ofComputational Physics, vol. 50, 215-234.

Voulon, S. and Wesselink, A. F. (1995) "Manoeuvrability of Waterjet-Propulsed Passenger Ferries," Proceedings of the Third International Conference on Fast Sea Transportation (FAST'95), Liibeck-Travemiinde, Germany, vol.2, 1131-1156.

Wald, Q. R. (1970) "Analysis of the Integral Pumpjet," Journal of Ship Research, vol. 14, Dec 1970, 307-316.

Ward-Smith, A. J. (1963) "The Flow and Pressure Losses in Smooth Pipe Bends of Constant Cross Section," Journal ofthe Royal Aeronautical Society, vol. 67, Jul 1963, 437-450.

Wellborn, S. R., Reichert, B. A. and Okiishi, T. H. (1994) "Study of the Compressible Flow in a Diffusing S-Duct," Journal of Propulsion and Power, vol. 10, no. 5, Sept-Oct 1994, 668-675.

White, F. M. (1991) Viscous Fluid Flow. Second edition. New York: McGraw-Hill.

Whitelaw, J. H. and Yu, S. C. M. (1993) "Velocity Measurements in an S-Shaped Diffusing Duct," Experiments in Fluids, vol. 15, 364-367.

Widmark, C. and Gustafsson, L. T. (1997) "3-Dimensional CFD Calculations on a Complete Watetjet Unit," Proc. RINA International Conference Power, Performance & Operability ofSma/1 Craft, The Royal Institute of Naval Architects (RINA), 15-16 Sept 1997, Southamton, paper no. 5.

Wieghardt, K. (1972) "Viscous Resistance," Resistance Session, International Towing Tank Conference, edited by S. Schuster and M. Scmiechen, vol. 1, 83-95, Berlin.

Wilcox, D. C. (1993) Turbulence Modelling for CFD. La Caiiada, California: DCW Industries, Inc.

Wilcox, D. C. and Sharma, T. L. (1977) "Streamline Curvature Effects on Turbulent Boundary Layers," AIAA Journal, vol. 15, no. 4, Apr 1977, 574-580.

321

Y akhot, V. and Orszag, S. A. ( 1986) "Renormalization Group Analysis of Turbulence," Journal ofScientific Computing, vol. 1, no. 1, 3-51.

Yakhot, V., Orszag, S. A., Thangam, S., Gatski, T.B. and Speziale, C. G. (1992) "Development of Turbulence Models for Shear Flows by a Double Expansion Technique," Phys. Fluids A, vol. 4, no. 7, 1510-1520.

Yang, C. K., Lee, Y. B. and Choi, H. S. (1995) "A Numerical Analysis of the Flow Around the Waterjet Inlet," Proceedings ofthe Third International Conference on Fast Sea Transportation (FAST'95), Liibeck-Travemiinde, Germany, vol. 2, 1391-1401.

Youngs, J. (1994) "The Jet Age," Popular Mechanics, Feb 1994.

Zhao, Y. (1997) "Computational Modelling of 3D Compressible Flows with a LRN k-E Turbulence Model," Computational Fluid Dynamics Journal, vol. 5, no. 4, Jan 1997, 487-508.

322

Appendix A Design Subspace Data

In this appendix, the results that form the basis of the design subspace investigations and

analysis of Chapter 7 are tabulated for reasons of completeness.

A.l Tabulated Results for Design Subspace 1

(X (0) RJD O"mm Mm.Cp TJ De CFY V* 20 0.025 -1.451 -1.920 0.6712 7.6892E-02 0.0098 3.6389

20 0.050 -1.197 -1.666 0.6693 7.8946E-02 0.0484 3.7037

20 0.075 -1.089 -1.558 0.6687 7.6903E-02 0.0940 3.7685

20 0.100 -1.009 -1.478 0.6698 7.5260E-02 0.1353 3.8519

25 0.025 -1.283 -1.752 0.6726 7.6826E-02 0.0408 3.1481

25 0.050 -1.032 -1.501 0.6711 7.5794E-02 0.0940 3.1944

25 0.075 -0.920 -1.389 0.6712 7.4748E-02 0.1408 3.2500

25 0.100 -0.856 -1.326 0.6712 7.4073E-02 0.1813 3.3056

30 0.025 -1.100 -1.569 0.6718 7.3413E-02 0.0845 2.8519

30 0.050 -0.869 -1.338 0.6711 7.2343E-02 0.1465 2.8981

30 0.075 -0.796 -1.265 0.6712 7.0237E-02 0.1890 2.9444

30 0.100 -0.731 -1.200 0.6697 7.1624E-02 0.2353 3.0000

35 0.025 -0.938 -1.408 0.6716 7.3081E-02 0.1094 2.6481

35 0.050 -0.738 -1.207 0.6701 7.3802E-02 0.1762 2.6944

35 0.075 -0.655 -1.124 0.6715 7.0552E-02 0.2302 2.7315

35 0.100 -0.605 -1.074 0.6701 6.8790E-02 0.2722 2.7685

Table A. I Hydrodynamic results for Design Subspace 1

323

A.2 Tabulated Results for Design Subspace 2

HJD y(o) O"mm Mm Cp 11 De CFY V* 0.050 7.5 -0799 -1.268 0.6698 7.3544E-02 0.0608 3.1944

0.050 10 -0.865 -1.334 0.6717 7.3100E-02 0.0598 3.1759

0.050 12.5 -0.731 -1.200 0.6713 7.4503E-02 0.0785 3.1944

0.050 15 -0.621 -1.090 0.6712 7.3101E-02 0.0618 3.1759

0.075 7.5 -0.701 -1.170 0.6702 7.0880E-02 0.1021 3.2315

0.075 10 -0.512 -0.981 0.6707 7.0760E-02 0.1090 3.2130

0.075 12.5 -0.332 -0.801 0.6720 6.9280E-02 0.1178 3.2222

0.075 15 -0.322 -0.791 0.6711 7.2036E-02 0.1293 3.2037

0.100 7.5 -0.577 -1.046 0.6689 7.0776E-02 0.1330 3.2685

0.100 10 -0.336 -0.805 0.6687 7.2981E-02 0.1477 3.2407

0.100 12.5 -0.286 -0.755 0.6710 6.8903E-02 0.1639 3.2593

0.100 15 -0.096 -0.565 0.6680 7.4493E-02 0.1790 3.2407

Table A.2 Hydrodynamic results for Design Subspace 2

A.3 Tabulated Results for Design Subspace 3

RdD AJAT O"mm Min. Cp 11 De CFY V* 0.025 1.0 -1.283 -1.752 0.6726 7.6826E-02 0.0408 3.1481

0.050 1.0 -1.032 -1.501 0.6711 7.5794E-02 0.0940 3.1944

0.075 1.0 -0.920 -1.389 0.6712 7.4748E-02 0.1408 3.2500

0.100 1.0 -0.856 -1.326 0.6712 7.4073E-02 0.1813 3.3056

0.025 1.2 -0.565 -1.034 0.6864 7.7927E-02 -0.1336 3.3497

0.050 1.2 -0.543 -1.016 0.6853 7.6418E-02 -0.1107 3.3888

0.075 1.2 -0.543 -1.012 0.6843 7.6237E-02 -0.0521 3.4323

0.100 1.2 -0.545 -1.014 0.6831 7.4897E-02 -0.0163 3.4843

0.025 1.4 -0.022 -0.491 0.6946 8.4415E-02 -0.2136 3.5296

0.050 1.4 -0.153 -0.622 0.6925 8.4964E-02 -0.1769 3.5594

0.075 1.4 -0.235 -0.704 0.6908 8.6254E-02 -0.1440 3.5936

0.100 1.4 -0.294 -0.763 0.6890 8.6573E-02 -0.1140 3.6364

0.025 1.6 -1.216 -1.685 0.6936 9.9535E-02 -0.2439 3.7099

0.050 1.6 -1.133 -1.602 0.6907 1.0300E-01 -0.2130 3.7306

0.075 1.6 -1.086 -1.555 0.6890 1.0406E-01 -0.1849 3.7554

0.010 1.6 -1.082 -1.552 0.6890 1.0170E-01 -0.1577 3.7917

Table A.3 Hydrodynamic results for Design Subspace 3

324

A.4 Lip Flow and Minimum Lip Static Pressure

A (o) ~ (0) A-~ (o) Mm.CP

RV0=0.025 a=25° 12.5 -35.0 47.5 -1.752

RdD=0.050 a=25° 12.5 -23.9 36.4 -1.501

RV0=0.075 a=25° 12.5 -20.0 32.5 -1.389

RdD=0.050 a=20° 10.0 -33.0 43.0 --1.666

RdD=0.050 <X=30° 15.0 -17.0 32.0 -1.338

HV0=0.075 y=7.5° 8.75 -18.5 27.25 -1.170

HdD=0.075 y=10.0° 7.5 -13.5 21.0 -0.981

Htf0=0.075 y=12.5° 6.25 -11.0 17.25 -0.801

HdD=0.050 y=10.0° 7.5 -25.0 32.5 -1.334

Htf0=0.100 y=10.0° 7.5 -10.0 17.5 -0.805

RdD=0.050 AJAT=l.O 12.5 -24.4 36.9 -1.501

RdD=0.050 AJAT=l.2 12.5 0.0 12.5 -1.013

RdD=0.050 AJAT=l.4 12.5 10.5 2.0 -0.622

RV0=0.025 AJAT=l.2 12.5 0.0 12.5 -1.034

RV0=0.075 AJAT=l.2 12.5 0.0 12.5 -1.012

Table A.4 Variation of minimum static pressure on the inlet lip with lip flow

A.5 Lip Flow and Vertical Forces

A (o) ~ (0) A-~ (o) CFY RdD=0.025 a=25° 12.5 -35.0 47.5 0.0408

RdD=0.050 <X=25° 12.5 -23.9 36.4 0.0940

RdD=0.075 <X=25° 12.5 -20.0 32.5 0.1408

RdD=0.050 a=20° 10.0 -33.0 43.0 0.0484

RV0=0.050 <X=30° 15.0 -17.0 32.0 0.1465

HV0=0.075 y=7.5° 8.75 -18.5 27.25 0.1021

HdD=Q.075 y=10.0° 7.5 -13.5 21.0 0.1090

HdD=0.075 y=12.5° 6.25 -11.0 17.25 0.1178

HdD=0.050 y=10.0° 7.5 -25.0 32.5 0.0598

HVO=O.IOO y=l0.0° 7.5 -10.0 17.5 0.1477

Table A.S Variation of vertical force on the waterjet inlet with lip flow

325

A.6 Study of Vertical Forces on Waterjet Inlet

The vertical forces acting on the different geometric features of the waterjet inlet were

calculated, in order to allow their relative importance to be assessed. In order to study

these forces, the waterjet inlet was treated as being two-dimensional and the static

pressure coefficient integrated over the upper and lower waterjet inlet. The vertical

forces acting on different sections of the geometry of the upper and lower surfaces of the

waterjet inlet were calculated according to

(Fy ), = fc dS I D 1 u2 D p x 2P ref

(A.l)

In Eqn A.l, Fy is the vertical force acting on the section of geometry under

consideration (eg. inlet lip), Cp is the static pressure coefficient, D the throat diameter of

the inlet and Sx represents the projected arc length in the x direction. The integral of Eqn

A.l therefore represents a non-dimensional vertical force per unit width of a two­

dimensional waterjet inlet. The results of the study are tabulated below.

A.6.1 Design Subspace 1

a. (0) 25 25 25 20 30

RL(mm) 15 30 45 30 30

Horizontal Duct 0.2229 0.2220 0.2223 0.2185 0.2230

Bend 0.3944 0.3907 0.3898 0.3094 0.3849

Inclined Duct 0.1986 0.1730 0.1475 0.2679 0.1607

Ramp -0.1751 -0.1566 -0.1372 -0.0938 -0.1762

Total 0.6409 0.6291 0.6225 0.7020 0.5923

Table A.6 Vertical force contribution- Upper waterjet inlet

a. (0) 25 25 25 20 30

RL(mm) 15 30 45 30 30

Horizontal Duct -0.1754 -0.1744 -0.1748 -0.1737 -0.1757

Bend -0.0600 -0.0580 -0.0581 -0.0547 -0.0427

Inclined Duct -0.2319 -0.1850 -0.1408 -0.2665 -0.1173

Lip -0.0483 -0.0720 -0.0968 -0.0844 -0.0930

Total -0.5156 -0.4893 -0.4704 -0.5793 -0.4287

Table A.7 Vertical force contribution- Lower waterjet inlet

326

a (o) 25 25 25 20 30

RL(mm) 15 30 45 30 30

Honzonta1 Duct 34.78% 35.29% 35.72% 31.12% 37.64%

Bend 61.54% 62.10% 62.63% 44.08% 64.98%

Inclmed Duct 30.99% 27.50% 23.70% 38.16% 27.12%

Ramp -27 32% -24.89% -22.04% -13.36% -29.75%

Total 100.0% 100.0% 100.0% 100.0% 100.0%

Table A.8 Relative vertical force contribution - Upper waterjet inlet

a (o) 25 25 25 20 30

RL(mm) 15 30 45 30 30

Horizontal Duct 34.02% 35.64% 37.16% 29.99% 40.97%

Bend 11.64% 11.85% 12.35% 9.45% 9.96%

Inchned Duct 44.98% 37.81% 29.92% 45.99% 27.37%

Lip 9.37% 14.70% 20.57% 14.57% 21.70%

Total 100.0% 100.0% 100.0% 100.0% 100.0%

Table A.9 Relative vertical force contribution - Lower waterjet inlet

A.6.1 Design Subspace 2

y(o) 7.5 10 12.5 10 10

HL(mm) 45 45 45 30 60

Honzontal Duct 0.2186 0.2226 0.2200 0.2229 0.2207

Bend 0.3942 0.3928 0.3976 0.3915 0.3866

Inclined Duct 0.1711 0.1718 0.1726 0.1885 0.1561

Ramp -0.1681 -0.1717 -0.1644 -0.1734 -0.1728

Total 0.6158 0.6154 0.6258 0.6294 0.5906

Table A.IO Vertical force contribution- Upperwaterjet inlet

y(o) 7.5 10 12.5 10 10

HL(mm) 45 45 45 30 60

Horizontal Duct -0.1721 -0.1739 -0.1735 -0.1753 -0.1729

Bend -0.0605 -0.0600 -0.0627 -0.0576 -0.0571

Inclined Duct -0.1957 -0.1937 -0.1965 -0.2144 -0.1737

Lip -0.0689 -0.0632 -0.0633 -0.0681 -0.0558

Total -0.4973 -0.4907 -0.4960 -0.5154 -0.4596

Table A. II Vertical force contribution - Lower waterjet inlet

327

y(o) 75 10 12.5 10 10

HL(mm) 45 45 45 30 60

Honzontal Duct 35.49% 36.17% 3515% 35.41% 37.37%

Bend 64.02% 6382% 63.53% 62.19% 65.46%

Inclined Duct 27.79% 27.92% 27.58% 29.95% 26.44%

Ramp -27.29% -27.91% -26.26% -27.54% -29.26%

Total 100.0% 100.0% 100.0% 100.0% 100.0%

Table A.l2 Relative vertical force contribution - Upper waterjet inlet

y(o) 7.5 10 12.5 10 10

HL(mm) 45 45 45 30 60

Horizontal Duct 34.62% 35.43% 34.98% 34.01% 37.63%

Bend 12.18% 12.22% 12.64% 11.18% 12.43%

Inclined Duct 39.35% 39.47% 39.62% 41.61% 37.79%

Lip 13.86% 12.88% 12.76% 13.21% 12.15%

Total 100.0% 100.0% 100.0% 100.0% 100.0%

Table A.13 Relative vertical force contribution - Lower waterjet inlet

328

AppendixB Optimisation Data

B.l Tabulated Results for Optimisation Algorithm

Eva!. Gmm Location Case No. a (o) RL(mm) 'Y (0) HL(mm) AJAT I -0.684 lip I 25.00 20.00 11.00 40.00 1.0000

2 -0.594 hp 2 27.00 20.00 11.00 40.00 1.0000

3 -0.591 hp 4 27.00 23.00 11.00 40.00 1.0000

4 -0.517 lip 5 27.00 17.00 11.00 40.00 1.0000

5 -0.582 hp 6 27.00 17.00 10.00 40.00 1.0000

6 -0.499 lip 7 27.00 17.00 11.00 43.00 1.0000

7 -0.344 hp 9 27.00 17.00 11.00 43.00 1.0500

8 0.026 lip 10 29.00 14.00 11.00 46.00 1.1000

9 0.080 lip 11 31.00 14.00 11.00 46.00 1.1000

10 0.054 lip 13 31.00 17.00 11.00 46.00 1.1000

II 0.191 hp 14 31.00 11.00 11.00 46.00 1.1000

12 0.228 side 15 31.00 11.00 10.00 46.00 1.1000

13 0.230 side 16 31.00 11.00 10.00 49.00 1.1000

14 0.220 side 18 31.00 11.00 10.00 49.00 1.1500

15 0.216 hp 19 31.00 11.00 10.00 49.00 1.0500

16 0.214 side 20 35.00 10.00 9.00 55.00 1.1500

17 0.255 side 21 33.00 11.00 10.00 49.00 1.1000

18 0.252 side 22 33.00 14.00 10.00 49.00 1.1000

19 0.254 side 23 33.00 10.00 10.00 49.00 1.1000

20 0.253 side 24 33.00 11.00 11.00 49.00 1.1000

21 0.251 side 25 33.00 11.00 9.00 49.00 1.1000

22 0.250 side 26 33.00 11.00 10.00 52.00 1.1000

23 0.252 side 27 33.00 11.00 10.00 46.00 1.1000

24 0.239 side 28 33.00 11.00 10.00 49.00 1.1500

25 0.263 side 29 33.00 11.00 10.00 49.00 1.0500

26 0.242 side 30 35.00 11.00 10.00 49.00 1.0500

27 0.235 side 31 35.00 11.00 10.00 49.00 1.0500

28 0.277 lip 32 31.00 11.00 10.00 49.00 1.0500

29 0.182 lip 33 31.00 14.00 10.00 49.00 1.0500

30 0.280 side 34 31.00 10.00 10.00 49.00 1.0500

3I 0.280 side 35 31.00 10.00 I I.OO 49.00 1.0500

32 0.276 side 36 31.00 10.00 I 1.00 52.00 1.0500

33 0.281 side 37 31.00 10.00 11.00 46.00 1.0500

34 0.272 side 38 31.00 10.00 11.00 46.00 1.1000

35 0.251 lip 39 31.00 10.00 11.00 46.00 1.0000

36 0.240 lip 40 29.00 10.00 11.00 43.00 1.0500

329

37 0.266 side 4I 33.00 1000 II.OO 46.00 I 0500

38 0.274 hp 42 29.00 10.00 II.OO 46.00 I.0500

39 0.200 hp 43 3I.OO 13.00 II.OO 46.00 I.0500

40 0.27I side 44 31.00 10.00 10.00 46.00 1.0500

4I 0.28I side 45 31.00 10.00 11.00 49.00 1.0500

42 0.283 side 46 31.00 10.00 11.00 43.00 1.0500

43 0.272 side 47 31.00 10.00 11.00 43.00 1.1000

44 0.221 hp 48 31.00 10.00 11.00 43.00 1.0000

45 0.256 lip 49 31.00 10.00 Il.OO 40.00 1.0500

46 0.268 side 50 33.00 10.00 11.00 43.00 1.0500

47 0.240 lip 40 29.00 10.00 11.00 43.00 1.0500

48 0.186 lip 51 31.00 13.00 11.00 43.00 1.0500

49 0.247 hp 52 31.00 10.00 10.00 43.00 1.0500

50 0.281 side 37 31.00 10.00 11.00 46.00 1.0500

51 0.256 lip 49 31.00 10.00 11.00 40.00 1.0500 52 0.272 side 47 31.00 10.00 11.00 43.00 1.1000

53 0.221 lip 48 31.00 10.00 11.00 43.00 I.OOOO

54 0.273 side 53 32.00 10.00 11.00 43.00 1.0500

55 0.249 lip 54 30.00 10.00 I 1.00 43.00 1.0500

56 0.181 hp 55 31.00 11.50 I1.00 43.00 1.0500

57 0.282 side 56 31.00 10.00 10.50 43.00 1.0500

58 0.282 side 57 31.00 10.00 11.00 44.50 1.0500

59 0.254 lip 58 31.00 10.00 11.00 41.50 1.0500

60 0.277 side 59 31.00 10.00 I 1.00 43.00 1.0750

61 0.268 hp 60 31.00 10.00 11.00 43.00 1.0250 62 0.280 side 61 31.50 10.00 11.00 43.00 1.0500

63 0.284 side 62 30.50 10.00 11.00 43.00 1.0500

64 0.235 hp 63 30.50 10.75 11.00 43.00 1.0500

65 0.270 lip 64 30.50 10.00 10.75 43.00 I.0500

66 0.279 hp 65 30.50 10.00 11.00 43.75 1.0500

67 0.252 lip 66 30.50 10.00 11.00 42.25 1.0500

68 0.274 lip 67 30.50 10.00 11.00 43.00 1.0625

69 0.232 lip 68 30.50 10.00 11.00 43.00 1.0375

70 0.284 side 62 30.50 10.00 11.00 43.00 1.0500

71 0.283 side 46 31.00 IO.OO 11.00 43.00 1.0500

72 0.249 hp 54 30.00 10.00 11.00 43.00 1.0500

73 0.235 lip 63 30.50 10.75 11.00 43.00 I.0500

74 0.270 hp 64 30.50 10.00 10.75 43.00 I.0500

75 0.279 lip 65 30.50 10.00 11.00 43.75 1.0500

76 0.252 lip 66 30.50 10.00 11.00 42.25 1.0500

77 0.274 hp 67 30.50 10.00 11.00 43.00 1.0625

78 0.232 lip 68 30.50 10.00 11.00 43.00 1.0375

79 0.284 side 62 30.50 10.00 11.00 43.00 1.0500

330

B.2 Results for Optimum Lip Radius and Lip Profile Inclination

Eva!. <Jmm LocatiOn Case No. a (o) RL(mm) 'Y (0) HL(mm) AJAT 31 0.280 Side 35 31.00 10.00 11.00 49.00 1.0500 32 0.276 side 36 31.00 10.00 11.00 52.00 1.0500

33 0.281 side 37 31.00 10.00 11.00 46.00 1.0500

34 0.272 Side 38 31.00 10.00 11.00 46.00 1.1000

35 0.251 hp 39 31.00 10.00 11.00 46.00 1.0000

36 0.240 hp 40 29.00 10.00 11.00 43.00 1.0500

37 0.266 Side 41 33.00 10.00 11.00 46.00 1.0500

38 0.274 hp 42 29.00 10.00 11.00 46.00 1.0500

41 0.281 side 45 31.00 10.00 11.00 49.00 1.0500

42 0.283 side 46 31.00 10.00 11.00 43.00 1.0500 43 0.272 side 47 31.00 10.00 11.00 43.00 1.1000

44 0.221 hp 48 31.00 10.00 11.00 43.00 1.0000

45 0.256 lip 49 31.00 10.00 11.00 40.00 1.0500 46 0.268 side 50 33.00 10.00 11.00 43.00 1.0500

47 0.240 lip 40 29.00 10.00 11.00 43.00 1.0500

50 0.281 side 37 31.00 10.00 11.00 46.00 1.0500 51 0.256 lip 49 31.00 10.00 11.00 40.00 1.0500 52 0.272 side 47 31.00 10.00 11.00 43.00 1.1000

53 0.221 lip 48 31.00 10.00 11.00 43.00 1.0000

54 0.273 side 53 32.00 10.00 11.00 43.00 1.0500 55 0.249 lip 54 30.00 10.00 11.00 43.00 1.0500

58 0.282 side 57 31.00 10.00 11.00 44.50 1.0500

59 0.254 lip 58 31.00 10.00 11.00 41.50 1.0500

60 0.277 side 59 31.00 10.00 11.00 43.00 1.0750 61 0.268 hp 60 31.00 10.00 11.00 43.00 1.0250

62 0.280 side 61 31.50 10.00 11.00 43.00 1.()500

63 0.284 side 62 30.50 10.00 11.00 43.00 1.0500

66 0.279 lip 65 30.50 10.00 11.00 43.75 1.0500

67 0.252 lip 66 30.50 10.00 11.00 42.25 1.0500

68 0.274 lip 67 30.50 10.00 11.00 43.00 1.0625

69 0.232 lip 68 30.50 10.00 11.00 43.00 1.0375

70 0.284 side 62 30.50 10.00 11.00 43.00 1.0500

71 0.283 side 46 31.00 10.00 11.00 43.00 1.0500

72 0.249 lip 54 30.00 10.00 11.00 43.00 1.0500

75 0.279 lip 65 30.50 10.00 11.00 43.75 1.0500

76 0.252 hp 66 30.50 10.00 11.00 42.25 1.0500

77 0.274 hp 67 30.50 10.00 11.00 43.00 1.0625

78 0.232 hp 68 30.50 10.00 11.00 43.00 1.0375

79 0.284 side 62 30.50 10.00 11.00 43.00 1.0500

331