• Nie Znaleziono Wyników

Interaction of afterbody boundary layer and propeller

N/A
N/A
Protected

Academic year: 2021

Share "Interaction of afterbody boundary layer and propeller"

Copied!
53
0
0

Pełen tekst

(1)

k

cI

3 SE?.

iai

ARCHIEF

Ref. PAPER 4/2 - SESSION i

lz)

Lab. v ScheepsbouwkunJe

Tech&sche Hogeschoo(

Dell L

SYMPOSIUM ON

"HYDRODYNAMICS OF SHIP AND OFFSHORE PROPULSION SYSTEMS"

HØVIK OUTSIDE OSLO, MARCH 20. 25,1977

"INTERACTION OF AFTERBODY BOUNDARY LAYER AND PROPELLER"

By

Thomas T. Huang and Bruce D. Cox

David W. Taylor Naval Ship Research and Development Center Bethesda, Maryland

(2)

ABSTRACT

Measurements of the boundary layer characteristics and pressure distribution on axisymmetric bodies with and without a stern propeller in operation are used to verify new analytical techniques for computing the

hydrodynamic interaction. It is shown that both the in-fluence of the propeller on the body pressure distribution, which causes an increase in resistance (thrust deduction),

and the interaction of the propeller with the stern boundary layer velocity profile can be predicted accurately by

in-viscid flow analyses. A new technique is proposed for cal-culating the effective wake which is an important input in

propeller design. Finally, it is concluded that the primary effect of scale is a change in the nominal wake profile which can be accurately computed using boundary-layer theory

for axlsymmetric bodies.

(3)

INTRODUCTION

Many single-screw ship propellers operate inside of thick and.possibiy separated stern boundary layers. The propeller normally actsto accelerate

the flow in the thick boundary layer which results in a decrease of

boundary-layer thickness. The effective inflow velocity distribution experienced by

the propeller depends on the interaction of the propeller and the thick stern boundary layer, and jg different from the nominal velocity

distribu-tion measured in the absence of the propeller., In addition, the upstream

suction of an operating propeller generates higher stern pressure drag and

skin-friction drag. Naval architects ref er to the drag' increase due to

pro-peller suction on the afterbody as thrust deduction. A knowledge of both the effective inflow and the added drag is essential for design of a pro-peller which meets specified propulsion requirements.

In order to focus on the physical nature of the comple,x interaction

between a propeller and a thick stern boundary layer, axisymmetric bodies were chosen for the present investigation. The geometric simplicity of the axisymmetric bodies offers considerable experimental and computational con-venience in studying the fundamental aspects of the interaction. The Laser

Doppler Velocimeter (LDV) was successfully used by Huang et al.,1 to measure

velöclty profiles very close to the propeller. The measured differences

between thé velocity profiles, stern pressure distributions, and stern shear stress distributions with nd without an operating propeller, provide the

necessary clues to the proper understanding of the af.terbody boundary layer and propeller interaction.

New and improved analytical techniques1 have been developed as' a result

of insignt gained from the experiments The first new technique computes

trie effective velocity distribution from the nominal velocity distribution

tneasuredjn the absence of a propeller. he second.improved analytical technique solves for the pressure distribution and velocity profiles over the entire body by considering the interaction between the displacement thicknéss of the boundary layer on the body and the external potential flow

(4)

in the absence óf a propeller. The third improved technique uses a lifting-surface representation of the propeller to solve för the thrust deduction añd the difference in pressure distribution with and without the propeller.

A brief derivation of the analytical techniques will be given and còm-parison of theOretical predictions and experimental datá at model Segle will

be made. The application of the present techniques to solve the problem of full-scale propeller/boundary-layer interaction on axisynetric bodies will

be discussed.

PROPELLER AND STERN BOUNDARY-LAYER INTERACTION

The bouñdary-layer velocity distribution in the absence of the

propel-ler will be called the nominal profile. The nominal profile at the

propel-ler plane of a ship model is usually measured by a standard wake rake. With a propeller operating, the flow over the afterbody is. normally accelerated.

The mutual interaction of the propeller and the thick stern boundary layer

results 'in a new resultant velocity profile. An effective velocity profile

is defined to be the resultant inflow velocity profile with propeller in

operation minus trie propeller-induced velocity profile as computed, from lifting-surface theory. The effective profile is an important input to propeller design and is essential for the correct prediction of powering,

cavitation performance, and unsteady forces. It is. therefore essential to

develop reliable and sound theoretical procedures to calculate the effective

profile.

The influence of a stern-mounted propeller on the flow field past

different bodies of revolution and, a flat plate was measured. by Hucho,2'3'4 although no attempt was, made to actually calculate the effective wake

dis-tribution. Investigations related to this subject were made by Wertbrecht,5

6 '.7

'8

flickling, Tsakonas and Jacobs, and Wald. Estimation of effective wake

9 . . lÓ . . '' 11

was made byRaestad, Nagamatsu and Sasajima, and Titof f and Qtlesnov.

The on'lyknown previous effort to theoretically addtess this problem is due to D.M. Nelson* who developed an unpublished computer program fôrcalc4ating

the effective' wake 'from the measured nominal wake and static pressure dis-tribution across the boundary layer. The method to. be presented 'below.

*.

Priv ate counïcation.

(5)

requires only the measured nominal wake distribution. Seriouè effort has been made in this work to compare the theoretical wake prèdictions withwake distributions measured by an LDV in the presence of a propeller.

The experimental evidence given in Reference 1. allows one to conclude

that the influence of propellers n stern boundary layers is significant in a region extending two propeller diameters upstream of the propeller. Up-streamof the propeller, the mean circumferential velocity, Ve, Is idén-tically equal to zero on an axlsyetric body both with and without the. propeller in operation. The following assumptions are made to derive a

theoretical approximation of the hydrodynamic interaction between a

propel-ler and a. thick stern boundary layer upstream of the propelpropel-ler: (a) the flow is axisynunetric and the fluid is incompressible; (b) the interaction

of propeller and nominal velocity profile is considered to be inviscid in nature; thus, propeller-induced viscous losses and turbulent Reynolds stresses are neglected; (c) the conventional boundary-layer assumption,

v/ax<3u/ar, is assumed to be valid för the nominal boundary layer .n

the absence of a propeller; and Cd) upstream of the propeller, no energy is added to the fluid by the propeller, and the propeller-induced velocity

field upstream of the propeller Is irrotational.

The vector equat-ion Of steady motion for an inviscid fluid is given by (see, for éxample, Ref erence.12)

-V X (A) = grad H, (1)

where is the fluid velocity, = V X is the vorticity vector, and

H.= p+-. p(.) is the total heàd. For cylindrical polar coordthätes (r,

O, x) w1th= (v,v = O, u), the radial

cOmponent Of Equation (1) may

be writteñ1 as

lau aii\ ru

i x

ri

laHd'Y

xaH

X\_

¡) WT'

where 'Y is the stream function, for an incompressible axisyetric flow

definedby .

(6)

i. a'v

u

--x r

r r

Since the flow velocities are increased due to the action of the pro-peller, stream surfaces are shifted closer to the body surface. As shown

in Figure 1, a typical stream surface moves inward from r to r while the velocity is increased from the nominal velocity u to u due to the

inter-X p

action of the propeller with the nominal velocity profile. The resultant

velocity u1 as would be measured in front of the propeller will be called

the apparent total velocity profile. The effective velocity is obtained by substracting the computed propeller-induced velocity from the apparent

total velocity.

Since no energy is added to the upstream fluid by the propeller and since we assume no change of viscous losses due to propeller-induction effects, the total pressure head within the same stream annulus remains constant with (denoted by a subscript p) and without the propeller in operation; thus, upstream of the propeller

1

3Hl(3H\

a

pk5i7)p

or from Equation (2)

The radial velocity with the propeller in operation is assumed equal to

y + y

, where y is the radial velocity without propeller and y is the

r pr r pr

circumferential average propeller-induced radial velocity. Furthermore,

5 /9u x

3v\

r 1 r p 3u p 3(vr

+v

pr) (3)

rI3r

3x1

f 3r p

(7)

the normal boundary-layer açprox1.mation avr/ « au/ar is assumed to be

valid for the nominal profile and the propeller-induced velocity field is assumed to be irrotatioñal, which requires

av 3u

Vx

. pr

pl ax ar

p

where1 is the resultant propeller-induced velocity, and ua is the circumferential-average propeller-induced axial velocity. With these two

approximations, Equation (3) yields a simple relationship for the location of the new stream surface r

p

3u

¡au

3u

1 x i i p a

Tta.r

:

\.

-

p

Within a. given Stream annulus the. massf low is constant, which at a given

axial location Of x may be written as

d'1'.á ru dr r u dr

:P p:p

Inserting this expresion Into Equation. (4) yelds

u d

xx

n u d(u - u ) (6)

p p a

Equations (5) and (6) are theTgoverning eqùatons for the propeller.

and stern bOundary-layer interaction. The nomiñal velocity profile, ui(r).

and the circumferential-average propeller-induced axial velocity profile, u (r ), can be used to obtain the new location of, the stream

ap

urfäce r

- p

(8)

are equal on the body surface. Far outside of the boundary layer the flow

is uniform without any vorticity. Then we have du O as r - in Equation (6), which implies d(u - u. ) O and u - u constant V , since u is

p a p a .s a

zero and u has to be equal to V as r - Thus, for z - , the effective

p s p

velocity is identically equal to ship speed V. The present results are,

of course, applicable to a propeller operating in open watér where du is

idéntically, equal to zero. Thus, the effective velocity for a propeller operating in open water is identically equal to V; the apparent, total velocity with the propeller in operation is equal to V5 + ua(rp). As can be Seen from EquatIon (6), the necessary condition for the effective pro-file to be different from the nominal propro-file Is'that there must be a ve-locity gradient in the nominal profile (du # O). In other wordS, the pro-. peller can only interact with a nominal profile which. has. non-zero. vorticity. The numerical solutions of the simultaneous Equations (5) and (6). are

straightforward and are given iñ Reference 1.

In current propeller design and performance prediction practice, Only

the measured model nominal velocity profile is available. The full-scale

nominal velocity profile is then estimated from the measured model nominal

profile. In order to ¿ompute the effective velocity profile either for the

model propeller or for the full-scale propeller,, the present theory can be applied in an iterative procedure:.

First, guess an effective velocity' prof:ile from the nominal ve-locity profile. One may start with the guess u(r) =.c, .u(r) where the'

constant c is assumed to be the ratio of the Taylor thrust-identity wake fraction to the .volumetric-mean wáke. fraction. .

The conventional propeller design computer programs13. can.be used

to design a propeller to operate In the estimated effective velocity

pro-file

Lep) wIth a prescribed nondimensional circulation distribution

G(r)

(9)

of. a propeller of given geometry and rpm, computer programs14'15 are

avail-able to predict G(r) for the estimated values of Ue(tp)

(c.) The circumferential-average própeller-induced áxial velocity,

Ua

at the propeller disk is then calculated by using a field-point velocity program13 which Includes lifting surface corrections. A simpler result,

which may be used as a first. approximation, canbederived from moderately

loaded lifting line theory:

ir'

r'

dG()

___._2.._ p

ir'

ir'

Ir'

d (&_.) tan

p p p

(7)

8

where G r/27rRv5. tan

= (Ue + ua)/(fZrp - u), K is the number of blades,

is the propeller angular velocity, and u is the tangential

in-duced velocity. . .

With u(r) and ua(rp) now estimated, Euations (5) and (6) are

applied to determine u(r). . .

The new effective profile becomes u (r ) u (r) - u (r.).

e p p p

ap

Steps (b) through (e) are repeated until the values of ue(rp) and G(r) converge. .

The final computed nondimensional circulation distribution G(r) for a measured nominal axial velocity profile is shown in Figure2.. It has been found that the values of the effective, axial velocity distribution

Ue(p)/'Vs and G(r) essentially converge o their final values after three

iterations. The results of computations shownin Figure 2 indicate that the ratio of the effective to nominal velocity varies with. radius, having a larger value near the hub of the propeller than near the tip.

PROPELLER-HULL INTERACTION: FRICTIONAL AND PRESSURE-DEFECT COONENTS OF THRUST DEDUCTION

It is well known in naval architecture practice that the delivéréd

(10)

of the propeller. Traditionally, the increase in resistance due to

propeller-hull interaction has been. defined in terms of the thrust deduc-tion fracdeduc-tion t,

T=R

T

where T is the delivered propeller thrust and R is the hull resistance.

The theoretical and experimental literature on thrust deduction is very

2,7,16,17,18,19,20 17

extensive. . Weinbium summarized the earlier contribu-tions of hydrodynamicists to the understanding and formulation of the

phySics. of thrust'deduction. In addition, an extensive list of references

20 on the subject can be found in the paper by Nowacki and Sharma. A

potential-flow computational scheme for calculating thrust deduction by Lagally's theorem was presented by Beveridge.16

An operatitg propeller produces an upstream suction which caúse 'a redúction in pressure on the afterbody thereby increasing the hull pressure

drag.. Simultaneously, this suction also causes an increaseof flow

veloc-ities on the 'afterbo4y and hence, additional frictional drag. Thus, the thrust deduction for a fully-submerged, axisymmetrical, self-propelled bod.y

may be divided into two cmponents,

t = tF + t,

where the pressure-defect compònent, t, is given, by

L/R: t. p 2 C TS r

JR

max p

j

(

C)

rh/RP 9 (8) 2 CTS

(11)

and the frictioúa]. componént, tF Is given by wh e r è CTS = p V5TrR

(C), - (C)

2 CTS 2 =

[(ci

(CT)b] (

£

L/R p Xh/R

i

max p e s lo r

T ,IR

o p

d(f)

p

(thrust loading coefficient based

on ship, speed)

(decrease of pressure coefficient due to propeller)

(increase of local skin friction coefficient due to propeller)

.(9)

tan dr/dx. (slope of body profile)

The pressure cOntribution to rope1lér-hull Intetaction is approxi-mated by the potential-flow approximation. of .C. The propeller-inducéd.

velocity on the hull can be computed from the propeller field-point pro-grarn)3 To determine the change in pressure due to the presènce.of the

21

propeller by the Doùglas-Neumann computer program, hull Sourçe strengths must be adjusted to cancel the normal. velocit.y induced by the propeller,

=

(V) =v

cosc-u sina

(12)

where n is theunit outward normal to the surface, v is,. the propeller-induced radialvelocity, 'andua is thé ptoe1ler-propeller-induced axial velocity.

Thus, when the velocity fields of the hull and the propeller are,

superim-posed, the normal velocity must be zeroalong the body contour. The total

pressure coefficient in the presence of the propeller is

(C) = i

-t) =

- (

+++.$)

/

u u u.

(10)

where U is the total surface potential velocity in the presence of the

propeller, which consists of u5, t.he perturbation velocity due to the bare

hull, Usn the tangential perturbation velocity which results from cancelling

the normal propetler-induced velocity, and u ' u 'cos a .+

y

sin ct, the

sp a

pr''

propeller-thduced tangential velocity o. the hull. 'The pressure coefficient' of the 5bare

'hull

is given by (Cp)b = i '- (i + u5/V5)2 Since the 'pressure-defect component of thrust deduction is of a potential-flow natue, the

value of t computed by Equation (8) will be the sameas that' compute4 from

p 16

Lagally's 'theorem in the Beveridge' formuation of interaction of hull-induced velocity on the propeller. By application of Lágaily's theory22 the interaction force arising from a lifting-line (sink disk)

represen-tation of the propeller may be written as

=

f

o u.n, dS '4K

=çS

t. ' u

'n

Jv

rh/R

fr'

'p

d fr'

'p

li

f

Equation (il) invOlves integration of the.product of hull-induced axial velocity, uh(rP) and propeller sink strength a(r) over the propeller disk.

(13)

Here we have, used Weinblum's17 finding that the propeller sink strength a is equal to -2u at the propeller disk; the lifting-linê computation of u, in Equation (7), has al8o been used The value of t can also be

obtained by integration of the product of propèller-induced axial velocity

U and hull source strength ah over the hull surface dSh,

-$

°h U a- u dS 2 i

ha

h = -C

f

VV

2 TS j s sTrR Sh p

In Equation (12) the values of U can be computed from thé propeller 13

f ield-ppint computer program using either a line or lifting-surface propeller representation, and the hull source strength ah can be computed from the Douglas potential-flow computer program.21 In Equation

(il), the hull-induced velocity at the propeller can be obtained by the method of Reference 21 and the values of G and tan 81can be obtained from the propeller lifting-line design13 r a performance prediction computer

14,15

program. It is important to note, that lifting-surface theory can be

used to compute Ua 'for Equation (12). It s therefore, possible to examine

the separate contributions of additional propeller characteristics,

includ-ing blade thickness, blade location (skew and rake)' and chordwièe varia-tion in loa4ing,, to the thrust, deducvaria-tion t. ''

The value of tF is computed hereby Equation (9) frOm values of shear

stress coefficients with and without the propeller -in operation, as derived from a boundary-layer computation.

(14)

POTENTIAL FLOW/BÖUNDARY-LAYER INTERACTION

Satisfactory predictions of turbulént boundary-layer characteristics can be made for the forward portions of a body by solving the boundary-layer equations, in either integral or differential forms. However, at the ship

stern, thethickness of the boundary layer increases rapidly, mainly due

to the diminishing cross sectional area. The thickness of the stern bound-ary layer usually .exceeds the thickness of the body. Therefore, the

trans-verse curvature of the flow in the boundary layer cannot be approximated by

the transverse curvature of the body and the pressure distribution on the body cannot be solved accurately without considering the displacement

effect of the thick boundary layer. The general approach used here is to perform the initial boundary-layer computation using the potential-flow

21

pressure distribution on the body. The initial flow calculations are then used to modify the geometry of the body and wake, by adding the local displacement thickness as suggested by Lighthlll.24 Potential-flow methods

are then used to compute the pressure distribution around the modified bod and the boundary-layer calculations are repeated using the new pressure

distribution. The basic scheme is continued until the pressure distribu-tions on the body from two successive approximadistribu-tions agree t within a

given error criterion.

25 25 27. 28

The work of Beatty, Cebeci et al., Beveridge, and Myring follow this general approach. Nakayama et al.,29 considered an even more complex model where, allowance is made foi transverse variation of static pressure across the boundary layer and a simple linear profile is assumed for the nominal velocity. A signific.nt difference df the approach of Nakayama et al.29 from the present approach is that the displacement thickness,

which is an integrated effect of the boundary layer, is not uSed to modify

the body. Instead, the potential solution is matched to the boundary-layer

calculation at the edge of the boundary layer and wake.

23

In the present work, the Douglas CS differential boundary-layer method, modified to properly accounted for the effects of transverse curvature, is used to calculate the boundary layer over the axisymmetric

13

(15)

body. The integral wake relations given by Granville3°

are used to calcu-late the flow in the wake. Since neither of the two methods properly model

the thick boundary layer in the stern/near-wake region, the calculated dis-placement body is not assumed to be valid in the region 0.95 <X/L<l.05. In this region, a fifth-degree polynòminal is used with the constants de termined by requiring that the thickness, slope and curvature be equal to

those calculated at X/L = 0.95 and 1.05. Should separation occur before

X/L = 0.95, the upstream matching is moved to the separation point. A detailed description of the potential flow/boundary-layer computation

scheme is given in Reference 1.

MODELS AND EXPERIMENTAL METHODS

Three. axisyinmetric afterbodies having afterbody length-diameter ratios

(LAID) of 4.308, 2.247, and l.484were selected for the present

experi-mental investigation (see Figure 3). Afterbody prismatic ratios

were 0.606, 0.526, and 0.416, respectively. As shown in Tablet, each afterbody was coñnected to a parallel middle L and a strèamlined forebody with bow-entrance length/diameter ratio (LEID) of 1.82. The total length of each model was fixed at a constant value of 3.066 m. Other hull par

ticulars of the three models are listed in Table 1.

TABLE t

HULL PARTICULARS FOR AXISY!»NETRIC AFTERBODIES

(L = .3.066 M,L/D l0.97)

14

Model . Afterbody 1 Afterbody? Afterbody3

LM/D 4.85 6.91 7.67

LAID . 4o31 2.25 1.48

pA 0.606 0.526 0.416

p 0.787 0.844 0.862

(16)

An existing model propeller whose diameter is 54.5 percent of the maxi-mum body diameter,was located at X/L 0.983 for the experimental investi-gation of propeller and stern boundary-layer interaction. This propeller

was designed for a wake distribution which was different from the wakes of

the three afterbodies. Thus, the iterative procedure described earlier usiñg the propeller/boundary-layer-interaction program and a propeller per-formance computer program was required to calculate the hydrodynamic char-acteristics of the propeller for the given propeller geometry, rpm, and the

specified nominal axial velocity distribution. The measured nominal axial

velocity distributions of the thteè afterbodies are shown in Figure 4. The computed nonditnensional circulations for the propeller operating in the

wakes of the three afterbodies are shown in Figure 5.

The experimental investigation was conducted in the wind tunnel of

the Anechoic Flow Facility of DTNSRDC. The wind tunnel has a 2.438 by 2.438 meters closed-jet test section with a maximum air speed of 61. metérs per second. The model was supported from below by two streamlined Struts located roughly one-third of the model length apart., The disturbances

generated by the supporting strUts were fourìd to 11e within th region below the horizontal center plane. Therefore, all the measurements were made In the vertical center plane along the upper meridian. Each stern

protruded from the wind tunnel closed-jet working section. into the anechoic

chamber (6.4 x 6.4.x 6.4 meters) located upstream o the diffuser. The

propeller was driven by a 9-kIlowatt high-speed motor mounted inside the stern of' the model. Propeller rpm was measured- by a magnetic pickup.

Mea-surëment of the following quantities was made-on each stern without and

then with a propeller operating at one or two advance coefficients.: surface shear stress measured by Preston, tubes, surface pressure distribution

sured by pressure taps, and boundary-layer axial velocity distribution

mea-sured remotely by a Laser Doppler. Velocimeter. The LDV'was located on an

optical bench in the quiescent region of' the anechòic chamber and vas

operated in a dual-beam off-axis backscatter mode Photographs of

the model and optical arrangements are shown in Figuré 6and 7. . The.

(17)

focal length of the LDV optics-was 1.5 meters and the total beam angle was 3.72 degrees. The effective probe volume for these optics was roughly eJ-lipsoidal having dimensions of about 0. 5 x 5 , and the probe volumi could

be focused in the stern boundary. layer at distances down to 2 from the full surface.

Detailed analysis of the measurement accuracies has not been made. Hwever, the standard deviations of the measured data were estimated from

repeat runs. The standard deviations of the measured static wall pressure

and shear Stress were less than 5 percent of their mean values and the Stan-dard deviation of the measured velocities was less than 2 percent of the

free-stream velocity.

COMPARISON 0F EXPERIMENTAL AND THEORETICAL RESULTS

Figure 8 shows the measured and computed valúes of pressure coefficient for After-bodies 1, ., and 3 at a length Reynolds number R = 5..9 x i,o6.

The transition is fixed at X/L = 0.015 which corresponds to the virtual. origin of turbulence of a 0.61-nun diameter trip wire located at X/L = 0.05. The results show that agreement is very good for Afterbody 1, fairly good. for Afterbody 2, and relatively poor for Afterbody 3. The measured and

computed local shear stress distributions CT are shown for After-body 1 in

Figure 9, for Afterbody 2 in Figure 10, and for Afterbody 3 in Figure 11. The agreement between measured and computed shear stresses IS good for

Afterbodies 1 and 2. For Afterbody 3, good agreement exists up to the

point of separation (X/L 0.92). After separation, the present boundary-layer computation method. breaksdown and no calculated results are

pre-sented in Figure 11 after separation. Qualitatively, the measured shear Stress takes small negative values in the separation bubble and becomes

-positive again after flow reattachment at a value f X/L àf about Ó.97.

The disagreement between measured.and computed values of,C for After-body 3 in Figure 8 was caused by the presence of shoulder separation at X/L 0,92.

The effect of the propeller on the stern pressure distribution was

com-puted by the potential-f làw approximation. As can be séen from Equation (10),

(18)

the total body surface velocity In the presence of the propeller is the sum of the free-stream velocity, the bare-hull perturbation velocity

(in-cluding the displacement effect) , the perturbation velocity as the result

of cancelliñg the propeller-induced velocity normal to the hull, and the propeller-induced velocity tàngentia1 to the hull. The computed

boundary-layer characteristics in the presence of the propeller aré based on this

surface velocity.

The measured and computed values of

CT with the propeller in operation

are shown in Figures 9 through 11. As expected, the propeller accelerates the flow at the stern resulting in an increase of shear stress. However, this increase of C is limited to the region X/L >0.90. Nô effect Is noted

at distañces larger than two propeller diameters (2%) upstream of the

pro-peller. As shown in Figure 11, the suction of the propeller does not change

the point of boundary-layer separation on Afterbody 3. The distance between the propeller plane and the. point of separation is 1.3 propeller diameters where the influence of the propeller is not strong. enough to alter the sep-aration location.

Figures 12 through 14 show the measured and computed pressure ,distri-butions with and without the propeller in operation. Again, the effect of

the propeller òn pressüre distribution is felt up to a distance of 2D

up-stream of the propeller. As can be seen from Figures 12 and 13, the mea-sured values of (C) (Cp)b and Ê = (Cp)b - (C) are in good agreement

with the computed values for Afterbodies i and 2. Figure 14 for Afterbody 3 shows that the measured values of i C are smaller than the computed values

of ¿ C aft of the separation point X/L = 0.92. These smaller measured

valuesof may be caused by the cushion effect o the separationbubble.

However, the measured values of C are larger than the computed values of

C upstream of the separation point, reflecting the pòssibie contraction

of the separation streamline due to the propeller. ' The values of t, obtaiqed by integration of the measured values of

C according to Equation (8) are given in Figures 12 through 13. The

good agreement between the measured and computed values of t shown in Table 2 for Afterbodies 1, 2, and 3 shows that the pressure. component of

(19)

TABLE 2

COMPARISON OF COMPUTED AND MEASURED THRUST DEDUCTION FRACTÏONS

thrust deductiOn can be well predicted by the potential-flow analysis propeller/hull interaction approximation. However, if there exists flow

separation at the stern, the potential-flow interaction approximátion for

t should be used with caution. Although the computed values of t, for

Afterbody 3 agrees well with the value' ôf t obtained by the integration of the measured values of C , the measured and computed values of C do

p p

not agree very well in the vicinity Of the separation point (Figure. 14).

23

The CS boundary-layer computation can be performed with the pres-sure distribution (C) as modified by the propeller. I The measured and

computed. local shear strEss distributions and the values of tF computed by integration of measured values of (using Equation (9)) over the three

afterbodies are shown in Figures 9 through 11. The measured values of t and tF for the three afterbodies at R = 5.9 x 10 aresuarized in.

- n

Table 3. . .

The agreement between the computed and measured pressure components of thrust duduction. t is satisfactory. The measured ratio t

It

is found to

Fp

be less than 5 percent for the three afterbodies at J = 1.25 and .J 1.07. Therefore, it is sufficient to estimate the value of thrust deduction fraction, t, from its preSsuré component t.

18 Computed t P Measured t p Measured .t F Measured

t.t+t

pF

Measured

t/t

Fp

Afterbody l,J1.25 0.068 0.07 0.0024 0.0724 3.4% Afterbody 2, JE1.25 0.129 0.143 0.0023 0.1453 1.6% Afterbody 2, J=l.07 0.126 0.140 0.0026 0.1426 1.9% Afterbody 3, Jul.25 0.106 0.109 0 0.109 0 Afterbody 3, J1.07 0.103 0.103 0 .0.103 0.

(20)

TABLE 3

MEASURED AND COMPUTED VALUES OF THRUST DEDUCTION FOR AN APPENDED SERIES 58 BODY (NSRDC MODEL 4620, PROPELLER 3638)

The computation procedure developed for calculating t can be applied to arbitrary bodies. As an illustrative example, the potential f low pro-peller/hull InteractIon aialyses have, been applied to a Séries 58.appended

body. The details of afterbody (NSRDC Model 4620), appendages, and propel-ler (NSRDC Model 3638) are given by Beveridge)6 The results of the

mimer-16' ical calculation are shown in Table 3. In Equation (il) Beveridge.. set

a

= -

22 where is the induced velocity at the lifting line... However, the

p a ...a'

:'

:

present metho4 properly used, ap = - 2Ua where is thé. circumférential-mean propeller indücèd axial velOcity at the sink disk. It is interesting to note that the lifting surfacé cortectións. (in. this case primarily thickness

19 o '

METHD

. THRUST DEDUCTION

Measured (Resistance and Self-Propulsion . t. + tf

Experiments)

.

0150

LiftIng Body Pressure Integration, Surface Equation (10)

t 0.135

p

Method

. /

(for Ua) Legally Theorem, Equation (12)

- .

t 0.133

Computed '

--Lifting Legally Théorem, Equations (11) t = 0.141

Line or (12) ..

Method

Beveridge,16 Equation (11) with t = 0 159

'Disk)

a

=-2tz

,

p 'a ,

(21)

since this propeller has no rake) yield a 5-percent -lower válue of t

corn-pared with the lifting-line method. The measured value of the total thrust deduction, t = t + tF obtained from towing tank resistance and

self-propulsion experiments is also shown in Table 3 for comparison. Although the value of tF was not measured or computed, one may estimate

tF 0.05 t.

It is evident that the present methods cai. be used with confidence t

pre-dict the potential flow propeller/hull interaction for arbitrary bòdiês

and propeller conf:igurations.

The mean axial velocity profiles in the boundary layers of the three

afterbodies with and without a propeller n operation vere measured by a Laser Doppler Velocinieter. The boundary-layer profiles were also computed

13

by the Douglas CS differential boundary-layer computer. program. The

pres-sure distributions calculated from the present potential flow/boundary layer iteration scheme were used to compute t-he profiles for the bare-hull.

The pressure. distributions calculated from the potentIal f làw propeller!

hull interaction approximation (Equation (10)) were used to compute. the

pro-files with the propeller in operation. Figures 15 through 17 show the

measured and computed mean axial velocity profiles immédiately upstream.of

the propeller (X/L = 0.977). The "present" approximations shown in

Fig-ures 15 through 17 for the profiles with a propeller in operation are the results computed from the inviscid propeller/stern boundary layer inter-action approximation, using the measure nominai velocity profiles as input.

As shown in Reference 1, the agreement between the measured and com-puted nominal profiles in the absence of a propeller ls.excellent for all

afterbodies in the region X'/L < 0.90. For Afterbody 1, which has a mild

adverse pressure gradient, the present potential flow/boundary-layer interaction scheme does provide a very good approximation of the stern f low (even t X/L = 0.977 as shown in 'Figure 15).

However, for Afterbody 2, which has a strong adverse pressure gradient, the measured velocities are

progressively slower than the computed velocities as the end of the stetn

is approached (Figure 16). Since the boundary-layer computation Is term-inated when flow separation is encountered, no computed results are pre-sented in Figure 17. .

(22)

The measured and computed axial apparent total velocity profiles u(r) with the propeller in operation are shown in Figures 15 through 17. Good

agreement between the measured and computed values of u is found for all

of the cases. The measured difference in axial velocities obtained by the LDV with and without the propeller in operation, u/V, and the computed propeller axial velocities, Ua/V are shown in Figure 18. The present

approximation for tu/V is in excellent agreelent with the measured data.

It is important to note that the propeller plane is et X/L 0.983 in all

cases. The measurement stations vere located at a distance of 0.12 D from

13 p

the propeller. The propeller field-point velocity program was used to calculate the circumferential-mean propeller-induced axial velocities Ua at the positions where the nominal and apparent total velocity profiles were measured by the LDV. The calculated radial distribution of circulation for

the propeller operating in the wakeS of the three different afterbodies which is used in the field-point program, has been shown in Figuré 5.

Re-call that the efective velocity prof:ile Ue is the apparent total velocity

profile minus the propeller-induced velocity profile Ua (Figure 18).

The good agreement between the measured and computed values òf shown in

Figures 15 through 18 suggests that the present inviscld approximation can be used with confidence to calculate the effective velocity profile for a propeller on an axisyetric body from the measured nominal velòcity

pro-file u and propeller-induced velocity profile ua.

..

SCALE EFFECT ON NOMINAL WAKE DISTRIBUTION

Ship models are tested at Reynolds numbers which are two orders of magnitude or more smaller than Reynolds numbers of full-scale ships. Thus,

the nominal velocities for a full-scale ship can be expected to be greater

than those measured at model scale. Most towiñg tanks have developed their own empirical apprpaches to account for the scale effect on the wake. Some

empirical methods have been presented at the ITT meetings, and a compréhen-sive review of varlous.approaches has been given by Dyne.31'32

(23)

The empirical approach proposed by Sasajima and Tanaka33 was the first attempt to extrapolate the wake distribution measured on a modèl to the

full-scale ship. The method assuméd that the magnitude of the frictional

component of. the wake varied linearly with the flat-plate frictional

coef-f icient. However, even on a fiat plate, the turbulent boundary layer does not have the necessary similarity property required for this assumption to

be valid. Furthermore, the intrinsic nature of a thick boundary layer on a ship stern cannot be properly represented by the flat-plate approximation. The thickness of the boundary layer increases rapidly at the stern as the

cross-sectional area decreases. Since the thickness of the stern boundary layer usually exceeds the radius of the body, the transverse curvature of the boundary layer flow cannot be realistically represented by the

trans-verse curvature of the body. The differences between predicted, flat-plate.

4nd axisymmetric stern boundary-layer velocity profiles on Afterbody i are

illustrated iñ Figure 19. The Douglas CS23. differential method was used to

compute the flat-plate boundary layer without considering the displacement

effect. It can be Seen from Figure 19 that the boundary-layer thickness at an axisymmétricstern. is much larger than the flat-piateboundary layer

at the same length Reynolds number.

In the following we digress briefly to .outline a derivation of the

.33

simplified wake-scaling formula used by Sasajirna and Tanaka. It is well known34 that two-dimensional laminar boundary layers possess a

similarity characteristic; for a flat plate, the longitudinal, velocities

take the following similarity form

'X'ff...Y/X\f\.ffz1\

s'.'

\/Vx/V1

\6/

Yc

F'

where x is the verse distance efficient (i.e characteristic (13)

longitudinal distance. from the leading edge, y is the trans-6 is the boundary-layer thickneàs, and is the drag co-., the 1957 TTC line). There is no corresponding similarity

for two-dimensional turbulent boundary layers. ven for

(24)

a

turbulent,boundary layer along

a

flat plate, the well-known laws Of the wall and wake cannot be brought into the form35

However, for

a

given vElue of Reynolds number, the vélocity profile of a two-dimensional flat-plate turbulent boundary layer can be crudely

approx-imated by a power law

i

u n

v

t6)

s

where n is a function of Reynolds number. In this case, the total

resistance coefficient becomes35

which implies that & is proportional to for a fixed

value

of n.

There-fore, a similarity form of the solution for the - power velocity profile

when n is fixed may be approximated by

= (ï

X

u\ u

n .6

V/ V

\ó/ (n+l)(n+2) x. (14) (15) (16)

If the velocity profilé ls.known (either from meäsurement' or computation)

at a given Reynolds number, the velocity pröfLle can be obtained for other

Reynolds numbers .f n' is assumEd to have a fixed valüe. For a given value

of u/V, Equation. (16) states that the coordinate of the velocIty ptofile

isshifted by ' . .

(25)

(.

X

s

where the subscripts m and s denote the values corresponding to the. model and full-scale ship, respectively.

Formula. (17) vas f.irst proposed by Sasajima and Tanaka.33 The value

of can be considered to include the smooth flat-plate- resistance

co-efficient and a roughness allowance

AF.

However, the. value of the power in reality changes with Reynolds number. The well-known + - power law

is only approximately valid for io6 < R < Figure 20 shóws the

theo-retically computed and empirically calcúlated scale effect on a flat plate

velocity profile. Again, the Douglas CS23 differential method was used to compute the axial velocity profiles at two length Reynolds numbers. The flat-plate émplrical method was used to calculate the profile for R =

9

- 6

-s

3 X

10: from the profile at a-lower Reynolds number R = . 6 X 10 via

Equation (17). If thé flat-plate empirical methods were adequate, theñ

Curve ii should be a' good approximation to Curve Iwhich is'not the

caé

as shown in Figure'20. The discrepancy is quité large nEar theplae.

-' As for the flat. plate,.if»the axisytnmetric nominal velocityprofile at the propeller disk is assumed to follow a'* power law, Equation (14),

can be related to' boundary-layer thickness by

2

/''u\ 'u

2.

ô I

11__-.L_

Id(L\

n ô S (x)

J

\.

VJ VS, ô

\ 6 J . (2n+1) (2n+2) S (x) o

where S(x) is the surface area. The approximated similarity form of the

i ,

-solution for the axisymmet-ric power velocity profile is. then given by

24

(17)

(18)

(26)

Here, ve have neglected the variation of. static pressure and the normal

stress across the boundary layer.

Figure 21 shows the theoretically

coin-puted and empirically calculated scale effect on the nominal axial velocity

profile at the propeller plane of Afterbody 1.

The potential f.low/

boundary-layer interaction program' was used to compute the axial velocity

profile at two length Reynolds numbers.

The flat-plate and axisytninetric

methods were used to calculate the profile for R

= 3 X

from the

profile at the lower Reynolds númber R

6 X

io6.

It is obvious from

Figure 21 that the flat-plate empirical method is not suitable for

calcu-lating the scale effect on the axisyinmetric thick-stern boundary layer.

Although the profile computed by the axisytnmetric empirical method

(Equa-tion (19)) is closer to the theoretically computed profile, it is still

not a4equate to accurately compute the scale effect on the nominal velocity

profile.

CONCLUDING REMARKS

In this paper we have described recent experimeñtal and theoretical

studies of the classical propeller-body interaction problem.

Acomprè

hensive set of boundary layer measurements is presented for thrée

axisym-metric afterbodies, obtained with and without an.operating stern propeller.

The results ate compared with 'calculations based on new theoretical analyses

and 'provide valuable insights into the manner in which the propeller mf

lu-ences the flow over the afterbody and interacts with a thick stern boundary

layer.

Comparison with experimental results shOws that the potential flow!

boundary-layer interaction program computes, accurate values of 'pressure,

shear stress, and velocity profiles f pr the fine stern.

Less satisfactory

ptedictions are obtained for the somewhat fuller stern, while only fair

predictions ate obtained 'for the fullest stern having shoulder sepáration.

In all cases the agreément is excellent

ver the forward 90-percent of the

bodies.

These results suggest that the modeling of the wake should be

modi-f ied. to, treat cases omodi-f separated modi-flow and that the present boundary-layer

equations should be further improved for the fuller stens.

(27)

It is shown that in most cases, the inviscid propeller/stern boundary-layer interaction theory predictS very well the measured total velocity

profile with the propeller in operation. The ratio of the computed

effec-tive and nominal velocities Is found to be largest at the própeller hub

and decreases toward the propeller tip. This is cOntrary to methods which

assume that. the effective uke profile is a constant multiple of the

nominal wake profile. Since the radial diStributiop of. effective wake

plays an important role in propeller propulsion and cavitation performance,

it is recommended that the present t.heory be incorporated into design

procedures for wke adapted propellers.

It is also shown that the added body resistance (thrust deduction) arises almost entirely from the pressure defect on the stern. Moreover, the distribution of this pressure defect and the thrust deduction are pre-dicted very well from potential flow coñsiderations alone. For this reason, both the thrust deduction and propeller/boundary-layer interaction, which

is essentiallyinviscid in. nature, should be independent of scale. The primary scale effect is manifested by a change in the nominal velocity profile, which in the case of axisymmetric bodies, cañ be computed. from boundary-layer theory as described in this paper.

ACKNOWLEDG1ENT

The Initial experimental work reported herein was funded uñder the David W. Taylor Naval Ship R&D Center's Independent Exploratory

Develop-ment Program, EleDevelop-ment Number 62766M. All of the subsequent investigations

were funded under the Naval Materials Command Direct Laboratory Funding,

Element Number 62543N.

The authors wish to express their gratitude to Mr. JustjnH. McCarthy

for his continuous.support and technical guidance. Dr. William.B.. Morgan is thanked for this advice in applying propeller theories. The authors would also. like to thank their co-workers, Mr. N. Santelli, Dr. H.T. Wang, Ms. N.C. Groves,. Dr. A.G. Hansen, and Mr. Thomas A. LaFone for their co-operation and contribution. The staffof the Center's Anechoic Flow

Facility are also thanked for their effective experimental support.

(28)

REFERENCES

Huang, T.T., H.T. Wang, N. Santelli, N.C. Groves," Propeller! Stern/Boundary Layer Interaction on Axisyminetric Bodies: Theory and Experiment," David W. Taylor Naval Ship Research and

Development Center Report 76-0113 (Dec 1976).

Hucho, W.-!!., "iiber den Einfluss einer Heckschraube auf die

Druckverteilung und die Grenzschicht eines Rotationskrpers

-Teil II: Untersuchungen bei hoheren Schubbelastungsgraden," Institut FUr Stromungsmechanik der Technischen Hochschule

Branunschweig, Bericht 64/45 (1965).

Hucho, W.-H., "iber den Zusammenhang zwischen Normalsog,

Rei-bungssog und dem Nachstrom bei der Strdmung um Rotation-skrper," Schiff und Hafen, Heft 10, pp. 689-693 (1968).

Hucho, W.-H., "Untersuchungen aber den Einfluss einer Heckschraube auf die Druckverteilung und die Grenzschicht Schiffsähnlicher

Krper," Ingenieur-Archiv Vol. XXXVII, pp. 288-303 (1969). Wertbrecht, H.M., "Vom Sog, ein Versuch Seiner Berechnung," Jahrbuch Schiffbautechnische Gesellschaft, Vol. 42, pp. 147-204

(1941).

Hickling, R., "Propellers in the Wake of an Axisymmetric Body," Transactions of the Royal Institute of Naval Architects, Vol. 99,

pp. 601-617 (1957).

Tsakonas, S. and Jacobs, W.R., "Potential and Viscous Parts of the Thrust Deduction and Wake Fraction for an Ellipsoid of Revolution," Journal of Ship Research, Vol. 4, No. 2, pp. 1-16

(1960).

(29)

Wald, Q., "Performance of a Propeller in a Wake and the

Inter-actionof Propeller and Hull," Journal of Ship Research,.Vol. No. 1, pp. 1-8 (1965).

Raestad, A.E.,"Estimation of Marine Propeller's Induced Effects on the Hull Wake-Scale Effect on the Hull Wake Field," Det

norske Ventas Report No. 72-3-M, Chapter 1 (1972).

Nágamatsu, T. and Sasajima, T., "Effect of Propeller Suction on Wake," Journal of the Society of Naval Architects of Japan,

Vôl. 137, pp. 58-63 (1975).

il. Titoff, LA. and Otlesnov, Yu, P., "SOme Aspects .of

Propeller-hull Interftction," Swedish-SOviet Propeller Symposium, M9scow

(1975).

Thwaites., B., Incompressible Aerodynamics, Chapter. XI, OxfOrd

University Press '(1960)..

Kerwin, J..E. and Leopold, L., "A Design Theory for Subcavitating

Theory," Transactions of the Society of Naval Architects and Marine Engineers, Vol. 72,pp. 294-335 (1964).

Cummings; D.E., "Núinerical Prediction of Propeller Charaèteristics,"

JOurnal of'Ship Research, Vol. 17, Part 3, pp. 12-18 (1973). Tsaö, S.S.-K., "Documentation of Program for the Analysis of

Per-formance and Spindle Torque f Controllable Pitch Propellers,"

MIT, Department of Ocean Engineéring Report No. 75-8 (May 1975)

Beveridge, J.L., "Analytical Prediction of Thrust Dèduction for Submersibles and Surface Ships," Journal of Ship Research, Vol. 13, No. 4, pp. 258-271 (1969).

(30)

17. Weinbium, G.P., "The Thrust Deduction," Joürnal Of American

Society of Naval Engineers, Vol. 63, pp.. 363-380 (1951).

18. Amtsberg, H., "Investigatións on the Interaction between Hull and Propeller of odies of Revolution," (in German), Jàhrbuch der Schiffbautechnischen Gesselschaft, Vol. 54, pp. ll7-52

(1960). (David Taylor Model Basin Translation 309 (1965)).

Nowacki, H., "Potential Wake and Thrust Deduction Calcultions for Ship-Like Bodies," Jahrbuch der Schiffbautechnjschen

Gesselschaft, Vol. 57, pp. 330-363 (1963)..

Nowacki, H. and Sharma, S.D., "Surface Effect in Hull Propelle.r Intéraction," The Ninth Office of Nàval Research Symposium on Naval Hydrodynamics (Aug 1972), Paris, Ftance, ACR-203, U.S.

Government Printing Office, Vol. 2, pp. 1845-1961 (1972).

Hess, J.L. and Smith, A.M.O., "Calculation of Potential Flow about

Arbitrary Bodies," Progress in Aeronautical Science.,. Vol. 8, Pérgamon Press, New York (1966).

Curnmins, W.E., "The Force, and Moment on a.. Body in a Time-Varying Potential Flow,! Journal of Ship Research, VOl. 1, pp. 7-18(1957). Cebeci, T. and Smith, A.M.O., Analysis of Turbulent Boundary

Layer, Academic Press, New York., (1974).

Lighthiil, M.J.,. "On Displacement Thickness," Journal of Fluid

Mechanics, Vol. 4, Part 4, pp. 383-392 (1958).

25 Beatty, T D , "A Theoretical Method for the Analysis and Design

of Axisyimnetric Bodies," National Aeronautics and Space Admin-istration CR-2498 (1975).

(31)

Cebecï,,T., MosinsIis,G.J., and Smith, A.M.0., "CaÎculation of Viscous Drag and Turbulent Boundary-Layer Separation on Two-Dimensional and Axisymmetric Bodies in Incompressible FlowS,"

Douglas Aircraft Report MDC J0973-.0l (1970).

Beveridge, J.L., "Pressure Distribütion on Towed and Propelled Streamlined Bodies of Revolution at Deep Submergence," David

Taylor Model Basin Report 1665 (1966).

Myring, D.F., "The Profile Drag of Bodies of Revolution, in

Sub-sonic Axisymmetric Flow," Royal Aircraft Establishment Techniçal

Report 72234 (1972).

Nakáyama, A., Patel, V.Ç., and Landweber,L., "Flow Interaction

Near the Tail of a. Bòdy of Revolution: Part 1 - Flow Exterior to Boundary Layer and Wake, Part 2 -Iterative solution for Flow Within and Exterior to Boundary Layer and Wake,' Journal of Fluids Engineering, Transactions of the American Söciety of

Mechani'al Engineers, Vol. 98, Series 1, No. 3, pp. S31-49'(1976). Granville, P.S., "The Calculation of the Viscous Drag of Bödies of Revolution," David Taylor Model Basin Report 849 (1953).

3l. Dyne, G., "On the Scale Effect on Wake and Thrust Dedùction,"

Proceedings, Thirteenth International Towing Tank Conference., Berlin, Hamburg, Wést Germany, Appendix, 6, Report of Perfòrmance Committee, Vpl. 1, 1972.

32. Dyne,, G., "A Study of the Scale Effect on Wake, PröpeUer

Cavita-tion and Vibratory Pressure at Hull of Two Tanker Mo4els," Transactions, The Society of Naval Architects and Marine

Engineers, Vol. 82, pp. l62.l85 (1974)'.

(32)

Sasajima,H. and I. Tanaka,, "On the Estimation of Wake of Ships,' Proceedings, Eleventh International Towing Tank Conference,

Tokyo, Japan, Appendix X, Performance Session, pp. 1,40-143 (1966).

Landau, L.D. and E..M. Lifshitz, Fluid Méçhanics, First Edition,' Pergamon Press, London, pp. 145-151 (1959).

Schlichting, H., Boundary-Layer Theory, Sixth Edition,

Chapter XIX, McGraw-Hill Book Co., New York (1968).

(33)

Region of Propeller lñfluence Extends to About

2 Dp Upstream

of

the Propeller

Stream Surface without

the Presence of Propeller

Stream Surface with

Propeller in Operation

Du-.-_-...

-.

_o.-9

.

-i

t

r

F-Ag.

I

Definition Sketch for Propeller

Stern Boundary Layer

Interaction

(34)

04

0.004 0.008 0.01 2 0.016 0.020

oil

,/

\\

Computed Measured Nominal

/

Circulation G by

Velocity, .A

Propeller Inverse Ua

I

Program .

1.

F

Computed by Prpeiler

/

FieidPoint,Program

Propeller:

/

/

CTS T =0.371 1

I

/

/

/

/

'N.

4 1.25 .

f

/

ue

r(r/Rp).

2lrRp V5

/

I,

.,,,'

/

N%Up

,

V-

s

/

,

Computed by the Computer Program.

/

,,

0.2

Ò4

O'5

I I.

L

Ux/Vs,LetVs,Ot Up/Vs

Figure 2Computed Nondimensional Circulation and Effective and Apparent

Axial Velocity Profiles after FoUr Iterations for a Typical Propeller

Operated ¡n.a Measured Nominal: Axidl Velocity Profile of a Typical

Axisymmetricul Body

0.6

0.7

0.8

0.9

1.0

0.9

08

0.7

0.6

r

Rp

0.5

0.3

.2

(35)

0.05

0.04

0.03

0.02

0.0I.

0.7

Afterbody

I

-Afterbody 2

Afterbody 3

.0.8

X/L

Agure 3 The Three Axisymmetric Afterbodies

0.9

(36)

1.0

0.9

.0.8

0.7

//

/

MEASURED NOMINAL AXIAL

/

/

VELOCITY PROFILES AT THE'

0.6

PLANE OF PROPELLER:

/ /

AFTERBO'Y

I .

/

/

/

05

AFTERBODY 2

AFTER800Y 3

0.2

0.1

I

'h

I J I I

I

I 0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

ux

.vs

Figure 4Measured Nominal Axial Velocity Profiles for the

(37)

0.I

- .-.

%% Afterbody I

CTS:O.371, JI.25

.

CTS:O.420,

J:I.25T'

CTS : 0.637, J: 107

R s»

i,'

.1'

'

\/

//

Afterbody 3

///

i)SN,.

.CT$:O.428,J:I.25

__

J..

0.647, J=I.07

i

o

»0

0.004: 0.008

0.012

0.016

0.020

0.024

0.028

0.032

r(r/p)

.

ir

Fig. 5 Nondimeñsional Circulation for the Pröpefler Operated in

the Wakes of Three Aftérbodies

1.0

0.9

0.8

0.7

0.6

r

0.5

0.4

0.3

0.2

(38)

(I)

ARGON

LASER

BAC KSCAT TER

MODULE

IL'

PHOTO

DETECTORS

(3)

'(4)

FOCUSING

LENSE

SPLITTER

Fig. 6 Off -Axis Dual

-

Beam Backscatter Optics

(39)

Fig. 7 - Photographic View of Afterbody

2 and Optical ßrrangement of Laser

(40)

0.3

0.2

0.I

_0i

-_0.2

_0.3

_O.4

_0.5

0.75

Afterbody

Afterbody

Afterbody

0---.

-s.---

0

/1

o

/

\

il

N

Measured Computed I

o

2

3D

Afterbody 3

Location of Flow A

Separation on

R

= 5.9x10

At terbody 3

UI

s Afterbody 3

o

/ I

I

I

/

/

/

D

/

I

-,

"h

/

'h

/

,Afterbody 2w"

o

0.80

0.85

0.90

0.95

loo

X

L

Fig. 8 - Comparison of Measured and

Computed Stern

Pressure

Distributions for the Three Afterbodies

o

cP

(41)

00025

xl, 2

frCr

)dx

?:

R2 TS P-'X

00035

0.0002

- 0.0001

0.0030 -.

L

0.0

Q Measurements

0'

06

0.7

0.8

0.9

1.0

X/L

Fig. 9 Measured and Corriputed Skin - Friction Distribution

on Afterbody I With and Without Propeller in

Operation (R

5.9 x

Q6)

max

o

o

- 0.0004

0.0003

i.9

o

I-L)

O

No Propeller

0.0020 -

Propeller J = 1.25

0.0015

Boundary-Layer Computations No Propeller

0.0010 -

- - - - Propeller J:l.25

0.0005

0.0045

Propeller

J:125

Measured Derived tF

0.0024

0 0040

Fropeller CTS 0.371

(42)

o

I I i

06

0.7

0.8

0.9

1.0

X/L

Fig. IO Measured and Computed Skin -

Friction Distribution

on Afterbody 2 With and Without

Propeller in

Operation (R=5.9 x

¡Q8)

I-.

L)

O.00O8

0.0007

0.0006

0.0003

0.0001

0.0

-0.0004

0.0002

i.?

0

0.0045

0.0040

0.0035

0.0030

0.0025

0.0020

0.0015

0.0010

0.0005

Propeller

J:I 25

J 1.07 Measured CT0 A Derived 'F

0.0023

0.0026

Propeller CTS

0.42

0.637

2

jxh

I

Q

I

I

I

,

I

I

£

-2

r(C)dx

C R Xmax

-Measurements

o

No Propeller

-

Propeller J 1.25 Propeller J 1.07 Boundary-Layer Computations

No Propiier

o

F

o

o

- - - - Propeller J:125

Propeller J:l.07

(43)

0.0045'

0.0040

0.0035

00030

0 002.5

0.0020

0.0015

00010

0.0005

o

Measured

Computed

o

Propel 1er Has No Measurable

Ef fèct on C1.

Flow Separation

Predicted

Measured

o9i8

L "'

.

tiL

I'

I

0.6

0.7

0.8

.0.9

1.0

X/L

.

Fig.:II_ Measured and Çomputed Skin-Friction

DistributIons

(44)

a

o

o

5.9.

8.8

Potèntiol Flow

0.5

. Exper'ments

No Propellei With Propeller R x 10_6

J:125

Q4

o

e

5.9

£

8.8

Computed

0.2

0-0.6

0.7

0.8

0.9

I.0

I. I

X/L

Fig. I2

Measured and ÇómputedPressure Distribution

on

fterbody

I

With and Without Propeller

in.

(45)

Q

o

_04

0.15

L 0.10

Q-Q

o

»

4 4

I__Computationst

/

0.6

0.7

0.8

0.9

X/L

1.0

Fig. 13 Measured and Computed Pressure Distribution on

Afterbody 2

With and Without Propeller in

Operation

0.4

0.2

Ex per iments

J:107

R x

5.9

8.8

No Propeller With Propeller

J:125

o

(46)

0.15

0.10

Q.

o

.0 Q.

o

%.;

0.05

a. C.,

il

J=l.25 J:l.07

s

s

L

No Dota. Computation

zio-E ,iperiments

J:125

J:i.07

O

5

5.9

--

8.8

Potential Flow

0.428 0.646

CTS

0.109 0.103

tp

rrh

= 2

j r(áC)d

P

C.RJ

maz

I

I

I

I

I

I

I

I

I

!

,1.

0,8

I -I

j

0.6

0.7

0.8

0.9

1.0

X/L

Fig. 14 Measured and Computed Pressure Distribution on

Afterbody 3

With and Without Propeller in

(47)

1.0

0.9

0.8

0.7

2

0.6

0.5

0.4

0.3

,

-

Propeller J'l.25

r,

'0.136

rmax

Measurement

B-L Calculation

No Propeller

O

o

A

ux

vs

Present Approximation

Propeller J

1.25

4/

/

I

I

'I

up

vs

0.2

0.1

o

0

0.1

02 0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0 1.1

UI

Up

Fig. 15 Measured and Computed Axial Velocity Profiles at

0.977 of Afterbody I

With and Without Propeller ¡n

Operation (R

5.9 x 10)

(48)

1.0

0.9

0.8

0.7

X

D.

E

0.4

0.3

0.2

0.I

o

0

0.1

0.2

0.3

0.4

0.5

0.6 07

0.8

0.9

1.0

-.vs.vs..

Fig. 16 - Measured and Computed Axial Velocity Profiles at

= 0.977 of At terbody 2 With and Without Propeller in

Operation (Rn.

5.9 X 106),

.

Measurement 8-L Calculation

No Propeller

Q

Propeller J'l.25

A

Propeller Jl.O7

O

Present Approximation

(ix

vs

J

1.25

IO?

o0 J 1.25 1.2 /

i0

(49)

io

0i

/

1.25

J- 1.07

o

vs o O

fiji

o

n

o

A

I I

o

0

o ,;

od

o

j

Up vs

o

0

0.1

0.2

0.3

0.4

0.5

O.

0.7

0.8

0.9

1.0

Vs

Vs

1.2

Fig. 17 Measured Axial Velocity Profile at

'0.977 of

4f terbody 3 With and Without Propeller

in

Operation (R

= 5.9 x

106)

Mesurernen?

0.9

N

ProeIier

O

Propeller

'l.2

A

Propeler J-07

o

0.7

No B-L Calculation After Separation

*06

o.

Present Approximation

0.5

0.4

03

0.2

r

;-.

-0.136

'maz o

(50)

.1.2

1.0

0.8

0.6

0.4

0.2

nax

0.04

Propeller

J'

l.25

R = 5.9 x

t

t.

't

'

u

u-u

\

?

/!..Measurement

-,

X

p-Present Approxlmatl9n,

Ue Up U.a

v""v)

.Ux

Ax

o'...

.&?

R

o

rmax

/

.

loO

-- ,..

. Computed Ua

Propeller Induced

-vs i -

:.

.

L

.Ö8

0:12

0.16

ö0

'

0.24

U0 -Vs

or-ç,--

s

Fig 18 Measured and Computed Axial Velocity Increase at

X/L = 0.977 on Afterbody 2 Dué to' the Suction of

Propeller'

' . ..

(51)

0.050

0.045

0.Ò40

0.035

0.2

0.3

0.4

0.5

0.6

O7

r

0.8

0.9

1.0

ux

vs . . i

Figure 19Fundamental Difference Between the Flat-elate and

(52)

0.020

0.018

0.016

0.014

0.012

-J

)

0.010

0.008

0.006

0.004

0.002

-TRANSITION AT X/L 0.010

FLAT-PLATE EMPIRICAL METHOD

(II)

FOR R3xI09

FROM THE PROFILE OF

R

- 6x106

m

BOUNDARYLAYER COMPUTATION

R

_6xIO6

(I)

R'3xIO

0.2

0.3

0.4

0.5

ux

vs

Figure 20Theoretically Computed and Empirically Calculated

Scale Effect on Axial Velocity Profile on Flat-Plate.

(53)

0.7

0.6

0.5

0.3

0.2

0.I

o

BOUNDARY-LAYER COMPUTATION

-

6x106

R

-

3x109

ns-.----.----EMPIRICAL METHODS FOR

R

-

3x109 FROM THE

/

s

PROFILE OF R,1

"m

6x106

/

FLAT-PLATE

,'

AXISYMMETRlC.

77v,

os

0.9

1.0

0.2

0.3

0.4

0.5

0.6

0.7

ux

vs

Figure 21 Theoretically Computed and Empirically

Calculated Scale Effect on Nominal Axial

Cytaty

Powiązane dokumenty

In smart buildings, the models used for Heating, Ventilation, and Air Conditioning energy management and for maintenance scheduling differ in scope and structure: while the models

Rys. Rozmieszczenie magazynów gazu w Niemczech, stan na 2010 rok Źródło: www.udo-leuschner.de/energie-chronik/090806.htm Magazyny w kawernach solnych Magazyny w pozostałych

Obliczone powyżej granice kontrolne mogą być wykorzystane w odniesieniu do przyszłych podzbiorów dopóty, dopóki proces się nie zmieni lub przestanie być statystycznie

w sprawie RAD 42/66 — ma zastosowanie w postępowaniu dyscyplinarnym prze­ ciwko adwokatom i aplikantom adwo­ kackim, wynika, że decydujące zna­ czenie przy

tytu ł pierwszej publikacji

Jeśli bowiem ustawodawca wyraźnie stanowi, że postępowanie dyscyplinarne toczy się nieza­ leżnie od postępowania karnego, a nadto że komisje dyscyplinarne

Rów­ nież w dniu dzisiejszym członkowie Prezydium NRA zostali przyjęci przez członka Rady Państwa prof.. dra