• Nie Znaleziono Wyników

A steady - state simulation of small amplitude wind-generated waves

N/A
N/A
Protected

Academic year: 2021

Share "A steady - state simulation of small amplitude wind-generated waves"

Copied!
100
0
0

Pełen tekst

(1)

ARCHIEF

Lab.

v. Scheepsbouwkunde

Technische Hogeschool

August 1968

Delft Technical Report No. 94

A Steady -State Simulation of Small

Amplitude Wind-Generated

Waves

by

Ronald

F. Ott

En Yun Hsu

Robert L. Street

This research was sponsored by

National

Science Foundation

Grant GK-736

and

Office of Naval Research

under

Contract Nonr 2 25( 71), Task NR 062-320

Distribution of this document is unlimited.

Department of CIVIL MINTGIINTEETtING

ST.A.NFORD T_TNIVER.SITY

(2)

Department of Civil Engineering Stanford University

Stanford, California

A STEADY-STATE SIMULATION OF SMALL AMPLITUDE WIND-GENERATED WAVES

by

Ronald F. Ott

En Yun Hsu

and

Robert L. Street

Technical Report No. 94

This research was sponsored by National Science Foundation

Grant GK-736

and

Office of Naval Research Contract Nonr-225(71), NR 062-320

Reproduction in whole or in part is permitted for any purpose of

the United States Government.

Distribution of this document is unlimited.

(3)

ABSTRACT

The primary purpose of this investigation was to extend and to

improve the results of Zagustin, et al., on steady-state, experimental

simulation of wind blowing over small amplitude waves. Experiments were

performed in a laboratory open-channel flume in which a belt driven at a constant speed and following the shape of the wave reproduced the moving boundary conditions found at an air-water interface. Water was used as

a working fluid.

Preparatory to a description of the present results, the important features of Miles' inviscid, linearized shear-flow model are reviewed.

Then, the experimental results of Zagustin, et al., are discussed and the apparatus is described.

Two sets of new experimental results are presented. The first set

was a re-run of the Zagustin experiments in which the moving belt formed one wall of the open channel. Emphasis was placed on increasing the scope

and accuracy of the previous results. In the second set of experiments,

a cover-plate was added to the open flume to minimize the free-surface disturbances. In both sets of experiments, a pitot-static probe was used

to obtain velocity and pressure data. The measured velocity profiles in

the flow above the wavy surface were used together with Miles' theory to

obtain theoretical pressure distributions for the region near the wavy

surface. These were compared to the experimental distributions.

Within the limitations of the experimental model, the data verified Miles' inviscid mechanism of energy transfer from a shear-flow

(4)

TABLE OF' CONTENTS 10 10 11, 12 13 12, 12 v 4 13 13 .3' 8

*A; 4

13

14 $. 15 1 15 : s- 4 17

3.8.1,

Least-Squareg

Fit of Velocity Profiles ; 17

3.8.2,

Calculation of Theoretical Pressure Distribution J7

. DISCUSSION OF EXPERIMENTAL RESULTS . , 6 * 6 A 1 * ;

19

4.1.

Velocity Profiles

* , ,

.

* . .

. ,

. # :

6 A 40 0 4" .. 4 19

4.2.

Pressure Profiles

AA4'

...7, ,.., 20

4.2.1.

Comparison with Miles TheoreticAl Distribution 20' i

4,2.2.

Pressure Decay . . .

* . 0

. A A ,e A W

,

22

4 3.

Streatline Patterns . .

40.44 *

w e 6

4 .0404 4

23 1 5. 'CONCLUSIONS, Page 1

1.

INTRODUCTION * o *

.

*

* 4 -1

REWIEWOF THE EXPERIMENTS OF ZAUSTIN, ET AL4

2.1.

The, Steady-State Concept., o 4

2.2,

The Shear=Flow Model .

* *

o, ;1 * *oL A to' 4

The Previous Experiments .

2.4,

,Discussion of Experiments in 13 74 4

3.

SMALL AMPLITUDE WAVE EXPERIMENTS

3.1. Introduction. .

3.2.

The Apparatus 0.. 0 0

,

11 0 11, 0 1_1 A

3.3.

Instrumentation . io7 ;i" A Al AA 111 ,*

3.4.

Special Probe Arrangement. A ,

3.5.

Experiments in an Open Flume

*

3.5.1.

Velocity Measurements

3.5.2,

Pressure Measurements.

3.6.

Experiments in a, Covered Flume . *

6 4

31 6

3.6.1.

Flume Cover-Plate 4i4

:-AAA*6

3.6.2.

Velocity Measurements .

3.6.3.

Pressure Measurements. *

,

0 4 * e e

4 * *

3.7. Experimental Errors r. .0 iv v 74 ot c It.

3.8-

Data Reduction Methods * A 76..;i1 4, 4 0

* * 25

3

3

6

(5)

TABLE OF CONTENTS (continued)

Page

REFERENCES 27

APPENDIX . Computer Program to Calculate Theoretical

(6)

TABLES

Number Page

Summary of Velocity Measurements (without cover-plate) . . 42 Summary of Pressure Measurements (without cover-plate) . . 43 Summary of Velocity Measurements (with cover-plate) 44 Summary of Pressure Measurements (with cover-plate) . . . 145

Sample Computer Printout of Least-Squares Fit to Velocity 46 Profiles

Sample Computer Printout of Theoretical Pressure

Distributions

47

ILLUSTRATIONS

Figure

Sketch of Unsteady Flow over a Small Amplitude Wave

Sketch of Steady Flow over a Small Amplitude Wave . . . . 48

Plot of B vs kyc (after Miles (1959)) 49

View of Belt Used as Moving Boundary 53

View of Beam, Sprocket and Guides 51

View of Beam with the Installed Belt 52

General View of the Channel 53

Trace of Plastic Particles Floating on the Water Surface,

from Zagustin, et al. (1966) 54

Trace of Plastic Particles Floating on the Water Surface,

from Zagustin, et al. (1966) 54

General Dimensions of Apparatus 55

View of General Probe Arrangement 56

Probe Location in Cross Section of the Channel 57

Pitot-Static Probe Dimensions 57

. 1. 3. 8« , . . . 13,

(7)

ILLUSTRATIONS (continued),

Figure Page

14.,

15.

Sample. Calibration Curves °.!

Schematic Diagram of Probe Systems

r

58

58

16. Velocity Profile Across the Channel V! LI. tot 59

17a. Velocity Profile at Station 0 . N. 60

17b. Experimental and Theoretical Pressure Distributions

60

18a. Velocity Profile at Station 4 . . .4 in 61

18b.. Experimental and Theoretical Pressure Distributions 61

19a. Velocity Profile at Station 8 .

62

19b

Experimental and Theoretical Pressure Distributions

62

20a

Velocity Profile at Station 12 . 63.

20b. 21a.

Experimental and Theoretical Pressure 'Distributions

Velocity Profile, at Station 16 . 4

63,

64-21b. Experimental and Theoretical Pressure Distributions

22. Static Pressure Differential Across, the Channel

65

23.

24.

General View of Cover-Plate ,

View of Plastic Test Section, . I., 4i: AI go 66 25a. Velocity Profile at Station 0 (with cover-plate)

6?

25b. Experimental and Theoretical Pressure Distributions 67

26a. Velocity Profile, at Station 4 (with cover-plate) . . 68

26b.. Experimental. and Theoretical Pressure DistributiOnS 68

27a, 27b.

Velocity Profile at Station 8 (with cover-plate) Experimental and Theoretical Pressure Distributions

69

69

28a,

2811.

Velocity

Profile

at StatiOn 12 (with. cover-plate) Experimental and Theoretical Pressure. Distributions

70

70 29a. Velocity Profile at Station. 16 (with. cover-plate) *

71

29b. Experimental and Theoretical Pressure Distributions

71

Profile

(8)

ILLUSTRATIONS(continued)

11

Figure

Variation of the Experimental Pressure Distribution as

a Function of Distance from Moving, .Boundary (with

11.18±.

30.

cover-plate) . :A -A' oi

72

31.

Variation of the Experimental Pressure Distribution as

a Function of Distance from Moving Boundary (without

Cover-plate) ,e A A is; 73.

32.

Velocity Probe Calibration . .

41'.

74

33, Experimental and Theoretical Pressure Distributions

for'

Average 1J1 and yc (without cover-plate) , .

1

75

34. Experimental and Theoretical Pressure Distributions for i

k

Average U1 and yc (with cover-plate) . . , . , . . 76,

II

35.

Streamline Pattern (without cover-plate), Vb = 3.92 ,

Q...1.3 ,. . . ep 4 4 v

.

.

,

77

i

r

36. Streamline Pattern (with cover-plate), Vb = 3.92 ,

Q= 1-3

-qv

.

. ..

.',....

! ' ,,.':! t' 4

4 4:4m4 4:;.

i

(9)

LIST OF SYMBOLS

Amplitude of the wave

Celerity of the wave

Acceleration due to gravity Wave number, 27/X

Discharge

Wind speed at infinity

U(y) Wind velocity

U velocity,

v/To/P

1 Reference velocity, 17,/k

Relative Velocity, U(y) - c

Vb Belt velocity

Vw Water velocity

Direction in which wave propagates

Distance from moving boundary

yc Critical layer thickness

yo Streamline boundary

Dimensionless distance

zc Value of z at the critical level a Dimensionless pressure coefficient

Dimensionless pressure coefficient

Ap Difference in static pressure

6 Phase angle

von Kaman universal constant

X Wave length

Friction

(10)

-LIST OF SYMBOLS (continued)

Kinematic viscosity

Density

(11)

ACKNOWLEDGMENTS

The authors are indebted to Mr. Theodor Mogel for his suggestions concerning instrumentation and to Messrs. John Austin and Tom Wagner for their assistance in the experiments and in preparing the figures. Credit

goes to Mr. E. John Finnemore for the computer program shown in the Appendix. We are grateful to Mrs. Kay Mac Laury and Mrs. Ashby Longwell

(12)

1. INTRODUCTION

In the last decade, theoretical analyses, e.g., Miles (1957, 1959,

1960, 1962, 1967), Phillips (1957), Benjamin (1959) and Lighthill (1962),

have carefully delineated the basic mechanisms responsible for the gener-ation of waves by the wind. Phillips (1966) has given a detailed and

incisive account of these efforts. It is now accepted that the resonance

mechanism proposed by Phillips (1957) and the shear-flow mechanism pro-posed by Miles (1957) provide a reasonable qualitative description of the

initiation and growth of wind-generated gravity waves. However, the

experiments of Shemdin and Hsu (1966), Bole and Hsu (1967), and Hidy and

Plate (1965) in laboratory wind-wave facilities and those of Cox and

Snyder (1966) in the open ocean have shown that the theory and experiments do not agree quantitatively.

To verify one particular aspect of the developments by Miles,

i.e., the results of his 1957 analysis of an inviscid shear-flow over a

wavy boundary, Zagustin, et al. (1966) studied the flow of water past a

moving, wavy, yet solid boundary. The purposes of their experiments were

specifically to generate a real fluid flow that closely approximated the

theoretical model proposed by Miles (1957) and to compare the measured

pressure distributions near the boundary with those computed in accordance with Miles' theory. To precisely simulate this theory, the wave form must

remain fixed in shape and magnitude. Accordingly, wind-wave flume

exper-iments were not appropriate. Similarly, Zagustin, et al. (1966) argued

that the fixed wavy boundary experiments of Motzfeld (1937), Stanton, et

al. (1932), Thijsse (1952), Bonchkovskaya (1955), and Larras and Claria

(1960), etc., were not a correct simulation of the Miles model because

they did not permit the development of the so-called critical layer in

the flow.

The experiments of Zagustin, et al. (1966) were incomplete; how-ever, the experimental pressure profiles agreed reasonably well with those

predicted by the theory. In addition, because the experiments were run in

an open channel with the moving, wavy boundary forming one vertical wall of the channel, they experienced certain difficulties in measuring small

(13)

velocities and pressures because of the disturbances present on the free surface of the flow. As a result, they suggested a further investigation

of the areas near the critical layer and refinement of the experimental procedures.

The present work is a continuation and extension of the work of Zagustin, et al. (1966). Our objectives were to re-run and improve on

the accuracy and scope of their experiments. We undertook an analysis of

the Zagustin, et al. results; and while utilizing essentially the same experimental apparatus, we made specific changes to the measurement system and made extensive measurements to obtain detailed pressure and velocity profiles throughout the test section. Finally, a cover-plate was added

to the open channel, thereby making it a closed conduit in the region of the test section. The cover-plate removed many of the free surface

dis-turbances and made the flow at mid-depth in the channel essentially two-dimensional. Extensive measurements were made on this configuration also.

In the following sections we first summarize the principles and

results of the theoretical model upon which the experiments are based and the results of Zagustin, et al. (1966). Then the experimental apparatus

is described. Finally, the new results are cited and compared to theory

(14)

2. REVIEW OF THE EXPERIMENTS OF ZAGUSTIN, ET AL.

2.1. THE STEADY-STATE CONCEPT

The flow of a sheared fluid moving over a small amplitude,

sinu-*

soidal wave is shown in Fig. 1 . In this sketch the fixed wave form, with

amplitude a and length X , moves to the right at a constant speed or

celerity c . The velocity U of the fluid flowing over the waves varies with distance above the wave, i.e., U = U(y) , but remains constant in

time and direction.

To reduce this flow from an unsteady to a steady state, a velocity which is equal but opposite in direction to the wave celerity is superposed on the flow system. Under the assumption that for small amplitude waves the surface particle velocity variation between the crest and the trough is negligible, the wave-surface boundary is now moving to the left at a constant speed which is equal to the wave celerity c . Far from this

boundary, the velocity will be U(y) - c , while at the level where

U(y) = c the relative velocity v , defined as U(y) - c , will be equal

to zero. This level of zero velocity is called the "critical level," and the region between the moving boundary and this critical level is the "critical layer" (Miles (1957)). The critical layer thickness yc is

the distance from the boundary to the critical level. A sketch showing

the above steady-state conditions is shown in Fig. 2.

2.2. THE SHEAR-FLOW MODEL

In formulating his inviscid shear-flow model, Miles (1957) assumed

an inviscid, incompressible, two-dimensional parallel shear-flow coupled to a prescribed two-dimensional, deep-water, gravity wave train. Viscosity

and turbulence are supposed to play no essential part in the development, except that they are assumed to create and maintain a logarithmic mean velocity profile in the shear-flow. This logarithmic profile is expressed

as

U(y) = U1 ln(y/yo) (1)

*

For convenience, all figures are presented at the end of the work.

(15)

-where yo is the aerodynamic roughness height determined from extension

of the log profile to zero velocity, U(y) is the parallel shear-flow,

and U1 is the reference velocity defined by U1 =

U*/K

. Here, U* is

Prandtl's shearing stress velocity and K is von Karman's universal

turbulence constant, usually taken as 0.4 . This profile is of particular significance in the energy transfer mechanism, for as Miles (1957) stated,

"... the rate at which energy is transferred to a wave of speed c is

proportional to the profile curvature

- U(y)

at the level where U = c ."

The perturbations in the velocity and pressure in the shear-flow and associated with the wave motion were assumed to be two-dimensional and sufficiently small to permit the linearization of the equations of motion and application of the interface boundary conditions at the mean wave surface where y = 0 . In addition, only the component of the aerodynamic force that is in phase with the wave slope was considered to be significant;

that is, normal pressures were assumed to transfer the energy from the wind to the wave.

Through the use of the equations of motion and an assumed form for the aerodynamic pressure, the perturbed motion in the air was described

in terms of the Orr-Sommerfeld equation. Introducing dimensionless variables

into the equation, Miles formulated an aerodynamical boundary value problem,

the solution of which

would produce the values for dimensionless pressure

coefficients a and B . Miles presented an approximate solution to this problem in his 1957 paper. Later, Conte and Miles (1959) presented a

solution obtained by numerical integration of the Orr-Sommerfeld equation.

In both papers, a dimensionless wind profile parameter 2 = g yo/lq was

introduced, where g is the acceleration due to gravity. A dimensionless

wave speed c/U, was also defined, where c2 = g/k is the wave celerity

for deep-water waves and the wave number k is defined as 27/X , X

being the wave length. In Conte and Miles (1959), values of a ,

3,

and yc were tabulated for each value of the dimensionless wave speed

c/U1 and corresponding wind profile parameter

Q .

For each value of Q and

c/U1 the aerodynamic pressure

distri-bution over the wavy surface can be determined from the expression ,

(16)

p = - p a "Lq k

(a2

(32)1/2 cos(kx - 0) , (2)

where

is

the phase shift of the pressure distribution relative to the wave form and in the direction of wave propagation. This phase angle can

be determined by the expression

tan 8 = - (6/a) . (3)

In a later paper, Miles (1959) extended his previous analysis by presenting results based on a more accurate solution of the Orr-Sommerfeld equation (Conte and Miles (1959)). He imposed the interface boundary

con-dition at the wave surface, rather than at the mean surface, and included the dominant viscous term in the complete Orr-Sommerfeld equation. He

found that imposing the boundary condition at this location had no effect

on the end results; also, the viscous effects in the air just above the water surface were found to be small compared with the effect due to normal pressures. He concluded that his model for energy-transfer from

a parallel shear-flow to deep-water gravity waves gave a total energy transfer in order-of-magnitude agreement with observation.

Miles (1959) also included other results which are significant.

Defining a dimensionless distance to the critical level

zc = kyc , he presented an expression which relates 0 to

zc

U1\2

zc =

QL-

exp(E-) .

ul

In addition, noting that 0 versus

zc is essentially independent of

Q for

zc < 2 , he presented a figure which gives the coefficient B

as a function of

zc . This figure is reproduced as Fig. 3 of this study.

In Appendix B of his paper, Miles (1959) showed that for a small z,

the coefficient a could be obtained directly from

a = -

[(/7z)

-

62]1/2 . (6)

(5)

.

(17)

-Since the papers cited above, Miles has presented, as have others,

improvements to his basic shear-flow model. Miles (1960) combines the

inviscid shear-flow model with Phillips' resonance model. This

combin-ation formed a stochastic theory which can be applied over a complete spectrum of waves. Miles (1962) investigated the viscous effects on

capillary and short gravity waves. Miles (1967) added the wave-induced

perturbations in the turbulent Reynolds stresses for momentum transfer to his model. At present, what can be considered his combined model

pro-vides a reasonable description of the physical phenomena between wind and small amplitude gravity water waves.

2.3. THE PREVIOUS EXPERIMENTS

Since the Miles-Benjamin theory suggests the importance of the existence of the critical layer in the energy transfer from wind to wave,

it is evident that the theoretical conditions leading to formation of a layer must be correctly reproduced in order to simulate Miles' (1957)

shear-flow model. Zagustin, et al. (1966) argued that the previous fixed

wave form models failed to reproduce correctly these conditions, in that a critical layer was not formed above the wave. Using the assumption by

Miles (1957) that the variation in particle velocity between the crest and trough of a small amplitude wave is small, Zagustin, et al. used a moving belt apparatus to establish the correct boundary conditions. In

their experiments a belt capable of moving at a constant speed and following the prescribed shape of a wave was placed vertically in a flume. Water

was used as the working fluid.

The belt was a U.S.S. Cyclone, "Flex-Deck"-type, flat wire mesh covered with a rubber sheet to provide the required smooth surface (Fig.

4). The rollers at each edge of the belt ran in guide tracks that set

the shape for the belt to follow, while cast iron sprockets transmitted the driving force to the belt from a variable speed electric motor. The

guides and sprockets were supported by a light I-beam (Fig. 5). In turn,

the I-beam was held by three metal supports which allowed for the leveling of the wave guides. The I-beam was placed in the depressed center section

(18)

in place, with the belt installed, is shown in Fig. 6. A Venturi meter

was used to monitor the discharge into the stilling basin at the upstream end of the flume, while a tailgate, located at the downstream end, pro-vided for regulation of water depth. A general view of the apparatus is

given in Fig. 7.

Sinusoidal waves with lengths of two and three feet and respective belt speeds of 3.2 and 3.92 fps were investigated. These speeds are the

celerities of deep-water gravity waves with the corresponding wave lengths.

Discharges of 0.9 and 1.3 cfs, which produced mean water velocities of 0.8 and 1.15 fps respectively, were used for each wave length. Pressure

measurements were taken every 1/16 of the wave length in each case. Two pitot-static probes were used to register the pressure dif-ference between the undisturbed pressure near the straight wall and the pressure near the moving wall. Vertical and horizontal pressure and

velocity profiles revealed that the most desirable location for the probes was 0.35 ft below the water surface, in a water depth of 0.75 ft. The

stationary probe was located 1.5 in. from the fixed wall, while the second probe was kept a constant 0.2 in. from the moving boundary for all pres-sure meapres-surements.

Velocity measurements were taken across the channel at the same depth as the static pressure measurements, and for each particular dis-charge these profiles were measured at the crest and trough of each wave.

Logarithmic profiles were fitted to these measured velocity distributions with a resulting large systematic deviation occurring in the vicinity of

the critical level where the mean velocity is zero. It was argued that

these large deviations were caused by the combination of extremely low velocities near the critical layer and the large shearing action in the critical layer. As a result of these effects, the logarithmic profiles

were fitted to the measured profiles only in that region above the critical level. Using these fitted profiles, values of kyc and S

(Miles (1959)) were obtained and the theoretical pressure distributions

calculated.

With regard to the experimental pressure distributions, Zagustin, et al. remarked that the surface effects in the open flow made it

(19)

extremely difficult to obtain good measurement resolution. Even though

screens were installed along the straight wall to damp out these effects,

it was noted that in some cases the dynamic fluctuations in the static

pressure readings were as large as 50 percent of the reading.

In spite of the stated limitations, Zagustin, et al. point out

that the magnitudes of the measured pressure profiles agree reasonably well with the theoretical distributions; however, the phase shift in most of the experimental distributions was less than the theory predicted.

They attribute this difference in phase shift to viscous effects not

accounted for in the Miles (1957) theory. This differential in shift does,

they argue, support Benjamin's (1959) conclusion that viscosity tends to retard the phase shift in the pressure distribution.

To achieve qualitative verification of the Lighthill (1959) pre-diction of a stationary vortex located over the wave crest, Zagustin, et al. also made flow visualization experiments by taking time exposures of plastic balls moving in the flow. The experiments were performed on the

sixth wave, of 3-ft length, in the wave train. The belt speed was reduced

to 50 percent of the value used in the pressure measurements described

above. In addition, to increase the thickness of the critical layer, the average velocity was reduced to 75 percent of that used for the pressure measurements. These reductions were justifiable because the mechanism proposed by Miles does not require that the moving boundary have the celerity of the equivalent deep-water wave. No pressure or velocity

measurements were taken at the above conditions; therefore, no detailed analysis of the patterns were attempted. However, it did appear that the

resulting flow patterns (Figs. 8 and 9) displayed the "cat's-eye flow pattern" hypothesized by Lighthill (1962) in his physical interpretation of Miles' energy transfer mechanism.

2.4. DISCUSSION OF EXPERIMENTS

Although the measured pressures seem to agree reasonably well with the theoretical predictions, the analysis by Zagustin, et al. (1966) to obtain

U1 and yc from measured velocity profiles suggests further

(20)

mentioned in Section 2.2., Miles (1957) points out that the rate of energy transfer is proportional to the profile curvature at this region.

By considering only the velocity measurements above the critical level,

they obtained lower yc and higher U1 values than those that would

have been obtained if the velocities around U = c were accounted for.

As a result, their theoretical maximum pressures were lower than they would have been if the total profile had been considered. In the present

study we investigated the region mentioned above and have determined U1

and yc in terms of the entire logarithmic-portion of the velocity

distribution.

The flow visualization experiments did show the presence of vortex patterns above the wave. However, since no detailed pressure or velocity

measurements were made at the reduced belt speed and mean flow velocity of the visualization experiments, it was difficult to obtain more than a qualitative analysis of these patterns. To investigate the flow pattern

in the present study, we derived streamline patterns by integrating the

velocity profiles used to obtain U1 and yc . The belt speed and mean flow velocity were the same as used for the pressure measurements.

(21)

3. SMALL AMPLITUDE WAVE EXPERIMENTS

3.1. INTRODUCTION

In Sections 2.1. and 2.2., it was emphasized that the

Miles-Benjamin theory confirms that the moving boundary conditions existing in the steady-state picture of wind-generated waves play a major role in the energy transfer from the wind to wave. In particular, the critical layer

thickness

yc has considerable effect on the pressure distribution along

the wave. In Section 2.3. the apparatus used by Zagustin, et al. to

simulate the moving boundary conditions was reviewed. Unlike the fixed

boundary models, this apparatus did generate a critical layer with a finite thickness.

The small amplitude wave experiments described below are a contin-uation of the Zagustin, et al. (1966) experiments with particular emphasis now being placed on the region around the critical layer. The apparatus

used in the present experiments is essentially the same as that described in Section 2.3. with the exception of the specific changes discussed below.

3.2. THE APPARATUS

The general dimensions of the facility, along with a detailed description of the test section, are shown in Fig. 10. The test section

Is the four-foot long area described in Detail A of the figure. It can

be seen in this detail that the wave used throughout the experiments had an amplitude a of 0.75 in. and a wave length X of 3.0 ft. Because of

the limitations of our variable speed motor, the maximum speed at which the belt could be driven was Vb = 3.92 fps . This corresponds to the celerity of a deep-water gravity wave with a length of 3.0 ft. A greater

belt speed would have been desirable because it would have produced a thicker critical layer. The amplitude a = 0.75 in. was selected because

it gave a ratio a/X = 0.0208 , i.e., the wave was a small amplitude wave.

This ratio, in turn, gave a reasonably large pressure variation along the wave with a minimum surface disturbance in the open channel for the given

flow. The discharge Q = 1.3 cfs was also held constant throughout the

(22)

without overflowing because of the downstream control.

In the test section, two pitot-static probes made by the United

Sensor Corporation were used to make all velocity and pressure measure-ments. One probe was fixed 1.5 in. from the straight boundary, while

the other could be moved to any distance from the moving boundary by a traversing mechanism. This motorized traversing and position indicator

mechanism could position and give the horizontal location of the pitot-static probe relative to the moving boundary to ± 0.01 in. A special

mechanism adapted to the device allowed for the rotation and alignment of the probe in the vertical axis. The traversing mechanism and fixed probe

were supported above the test section on an aluminum carriage which could be moved longitudinally along the full length of the test section. This

combination of the carriage and traversing mechanism allowed for the accurate placement of the movable probe at any location in the test section. The traversing mechanism and carriage are shown in Fig. 11.

The test section was located on the fifth wave in a series of seven waves

along the channel. The fifth wave was used because the flow was fully

developed before it reached that section of the channel. The test section

included the region from the upstream crest (Station 0) to downstream crest (Station 16) of the three foot wave (see Fig. 10).

3.3. INSTRUMENTATION

A cross section of the probe arrangement in the channel can be seen in Fig. 12. A typical probe is sketched in Fig. 13. These probes

were connected by a series of three-way stopcocks to a Pace model P90 differential pressure transducer, with a range of ± 0.037 psi. The

trans-ducer was coupled to a Sanborn Recording Oscillograph, model 650, with a carrier-amplifier system. The transducer was calibrated after each major

series of experiments by a Harrison micromanometer which has a stated resolution of ± 0.0001 in. of water. Calibration curves that relate

pres-sure to output voltage (or galvanometer deflection) are given in Fig. 14.

The calibration curves displayed no significant change in the time span of this study.

(23)

and/or static pressure from the movable probe and static pressure from the stationary probe. These combinations permitted velocity readings to

be taken from the movable probe or pressure differential readings between the static pressures at the straight wall and at the moving wall (see

Fig. 15).

3.4. SPECIAL PROBE ARRANGEMENT

Also tied into the probes by means of stopcocks was a small reser-voir that was kept at an elevation of about 15 ft above the apparatus.

This device, shown in the schematic diagram of the probe setup (Fig. 15), allowed for the flushing of air out of the probes, lines, and transducer.

The surge of water through the lines during flushing also cleaned out

foreign particles that may have deposited in the dynamic end of the movable probe. Before each run heated water was placed in the small reservoir,

and the hot water was washed through the lines, probes and transducer.

As pre-heating the water removed the dissolved air, the water was deficient in air when it cooled in the lines before the run. This deficit caused

any small air pockets that were present in the lines to be re-dissolved into the solution. This process proved very effective in removing air

bubbles from the lines.

3.5. EXPERIMENTS IN AN OPEN FLUME

3.5.1. Velocity Measurements

Velocity measurements were made at Stations 0, 4, 8, 12 and 16.

At these stations, velocity profiles were taken over y-distances ranging from 0.2 in. to at least six critical layer thicknesses from the moving boundary. To investigate the cross-channel velocity distribution,

measure-ments were taken completely across the channel at Stations 0, 8 and 16 and are shown in Fig. 16. Care was taken to insure that the probe was always

pointed in the direction of the flow. This required rotation of the probe

by 1800 at the critical level. In addition, the distance from the moving

boundary was checked after each rotation. All velocity measurements taken

(24)

3.5.2. Pressure Measurements

Pressure measurements were taken every 1/16 of the wave length.

At each station the movable probe was located 0.2 in. from the moving boundary. Care was taken to insure that the probe was always oriented

into the flow. This was accomplished by turning the probe in the flow

until the static pressure difference was a maximum between the fixed and moving probe. Each run consisted of measuring one pressure reading at

each of the 17 stations along the wave. Thus, during one run a complete

pressure distribution along the wave was obtained. Three of these profiles

were taken, and they are plotted on each of the Figs. 17b-21b. The range

of the values at a given point is an indication of the experimental errors present in the three otherwise identical runs.

To study the change in the pressure profile as a function of dis-tance from the moving boundary, pressure measurements were taken from 0.2 in. through the critical layer at Stations 0, 2, 4, 6, 8, 10, 12, 14 and

16. The measurements at Stations 0, 8 and 16 were continued across the

channel to the fixed probe, to study the pressure decay with distance from the wave. These cross-channel profiles are plotted in Fig. 22.

All the above pressure measurements made in the open flume are summarized in Table 2.

3.6. EXPERIMENTS IN A COVERED FLUME

3.6.1. Flume Cover-Plate

To suppress the surface disturbances encountered in the experiments in the open flume, a cover was built and installed. This cover-plate made

the channel a closed conduit in the region of the test section, thereby making the flow at mid-depth essentially two-dimensional. The plate

covered three wave lengths upstream and two and a half wave lengths down-stream of the test section (Fig. 10). The upstream length was sufficient

to allow full development of the flow in the region of the test section.

Half-inch marine plywood was cut so that it fit snugly between the straight wall and the moving boundary (Figs. 10 and 12). Special fittings were

(25)

between the cover and the channel bottom was set at 9.0 in. (Fig. 12).

A

round transition on the upstream end of the plate insured a smooth entrance

of the flow into the test section. A general view of the cover is shown

in Fig. 23.

The 4-ft test section in the cover-plate had the general

dimen-sions shown in Fig. 10. This section was constructed out of half-inch

plexiglass, which allowed for a visual check of the probes in the flow. Slots were cut in the plexiglass at each station to allow for transverse

movement of the probes. In addition, the plexiglass plate was also split

longitudinally. This allowed the two plexiglass halves

to be slid over one another, facilitating access to the probes. An aluminum T section was

countersunk in the split to add strength to the plexiglass sections.

Countersunk screws with wing nuts held the section in place. Pressure

sensitive tape was used to seal all the slots that the probes did not

occupy. This insured that the surface exposed to the flow was completely

smooth. Figure 24 shows the plexiglass section with all slots taped,

except Station 8 where the probe is in place.

During all experiments, a thin film of water was allowed to flow over the entire cover-plate. This insured that the underside of the plate

was under a slight pressure, which in turn reduced the effect of air pockets that may have formed under the plate. Also, before any

measure-ments were made, all significant air bubbles were removed from under the

plexiglass. In both the velocity and pressure measurements, the probes

were located 4.5 in. below the plate (Fig. 12). This placed the probes in

the region of maximum horizontal velocity in the conduit.

3.6.2. Velocity Measurements

Velocity profiles were taken at Stations 0, 4, 8, 12 and 16. The

measurements ranged from 0.2 to 4.5 in. from the moving boundary. Again,

care was taken to rotate the probe 180° at the critical level and to check

the distance from the wave. A summary of velocities with distance from the

moving boundary is presented in Table 3. The velocity profiles are plotted in Figs. 25a-29a.

(26)

3.6.3. Pressure Measurements

The pressure measurements with the cover-plate were made in the same manner as those made without it. That is, the movable probe was

positioned 0.2 in. from the moving boundary and measurement was made at each station along the wave. The fixed probe was maintained 1.5 in. from

the straight wall at each station (Fig. 11). Two complete sets of

measure-ments were taken along the wave and are presented in each of the Figs. 25b-29b. Again, the range of values at a given point is an indication of

the experimental error in the two otherwise identical runs. To investigate

the pressure distribution along the wave as a function of distance from the wave, pressure measurements were also made at various distances from the moving boundary. Profiles were taken from 0.2 to 4.5 in. from the

moving boundary at Stations 0, 4, 8, 12 and 16. At each of the remaining

stations, profiles were measured from 0.2 to 1.0 in. Pressure

distribu-tions along the wave are given as funcdistribu-tions of distance from the wave in Fig. 30. The comparable result for the open flume is shown in Fig. 31.

3.7. EXPERIMENTAL ERRORS

In the open flume experiments, the same types of errors arose that Zagustin, et al. (1966) experienced in their experiments. These errors

were caused primarily by surface disturbance effects on the pressure readings. The surface disturbances resulted from the belt motion, and

their relative effect on the pressure reading diminished as the flow-rate was increased. Recall that a flow of Q = 1.3 cfs was used throughout

all experiments because it was the maximum flow the facility could handle without modification to the downstream outlet works. Also, three layers

of screens were attached along the straight wall to help damp out reflec-tions of surface undulareflec-tions. Even with these precautions, surface

dis-turbances caused a noticeable effect on the pressure and velocity measure-ments. To filter out high frequency disturbances induced by mechanical

vibrations, a 500 mfd capacitor was attached in parallel with the galvano-meter in the Sanborn 650. Despite the measures taken above, the

fluctu-ations around the mean value of the reading were about 20 percent without the cover-plate. When the cover-plate was used to make the flume a closed

(27)

flow, the fluctuations in the pressure readings were reduced to half of

those without the plate. To insure a representative reading, pressure

fluctuations were registered for two minutes at each point. The mean

value of the reading was taken from these readings by drawing a best-fit line by eye on the recording paper.

To insure that a compatible relationship between the shape and phase of the measured pressure distributions along the moving boundary was obtained, the pressure probe location relative to the belt was

main-tained at 0.2 in. during measurements with and without the cover-plate.

This distance was more than twice the probe diameter and so was large enough to avoid effects due to the presence of the solid boundary. In

addition, this distance places the probe in a region where the pressure change with distance from the moving wall is small (see Fig. 22 for an indication of the pressure variation with distance from the wall in the open flume).

As mentioned in Section 3.4., errors from the formation of air bubbles in the probes and lines were kept at a minimum by periodic flushing

of the system. After each flushing, the system was given adequate time to

restabilize before a new reading was attempted.

The flow in the experiments had a finite depth and width and, in the case of the open flume, a free surface. As a result, the experimental

flow was neither infinite nor fully two-dimensional, in contrast to the theory of Miles (1957). However, all the velocity and pressure measure-ments were taken at one depth where the flowfield was expected to be most

nearly plane. In addition, the flow developed quickly over the moving boundary, which tends to make the effect of the finite length of the test

section negligible. The maximum variation in channel width was ± 0.75 in., which caused a small variation in the average flow-velocity along the wave.

This variation would be a maximum of 8 percent between the crest and

,

trough if the flume water surface remained plane. However, it can be seen

in Fig. 16 that the actual variation in velocity is small in our open-flume

experiments. This would indicate that the free surface does adjust its level to compensate for the variations in channel width. In the closed

(28)

it appears that the errors due to finite channel width and depth are relatively small.

Small errors in velocity measurements did occur in the region

close to the critical level. Two causes of these errors were the extremely

low velocities near the critical level and the large shearing action in the critical layer. These errors are discussed below in more detail. To

help alleviate these errors, two readings were taken at each point in the region of the critical level and longer time spans were used to determine the mean for each reading.

3.8. DATA REDUCTION METHODS

3.8.1. Least-Squares Fit of Velocity Profiles

The velocity was determined for each data point from the difference in dynamic and static pressure at the movable pitot-static probe. The

velocity profiles at each station were taken from 0.2 in. to at least 4.5 in. from the moving boundary. The portion of the profile that was

con-sidered logarithmic was from 0.2 in. to a distance where the measured

velocities decreased or remained constant with distance from the moving boundary (Fig. 16). In most cases, this distance was about 3.5 in. The

velocity values were tabulated, digitized and fed into a computer, and a

least-squares-fit straight line was calculated and plotted through these data points on semi-logarithmic plots (Figs. 17a-21a and 25a-29a). The

intersection of the least-squares-fit line with the relative velocity value v = 0 gave the value for yc . In most cases, this value agreed

very well with the measured critical layer thickness. The value of the

reference velocity U1 was derived from the slope of the least-squares

line by using Eq. (1). The values of U1 and yc are listed on the

above plots. A sample computer print-out of the least-squares-fit line,

yc and U1 is given in Table 5.

3.8.2. Calculation of Theoretical Pressure Distribution

Using the six equations of Section 2.2. and a computer program (given in the Appendix) developed from the Conte and Miles (1959) paper, we calculated the theoretical pressure distributions predicted by Miles.

(29)

These distributions were then compared to the measured pressure distribu-tions described in Secdistribu-tions 3.5.2. and 3.6.3.

Having obtained the values of U1 and yc for a particular

station by the least-squares method, we used Eqs. (4) and (5) to compute

zc and Q . For a given set of values of Q ,

zc and c/U, , we

cal-culated 8 using the program mentioned above. The range of 8 values for these experiments is depicted in Fig. 3. Since

zc is less than

unity for the range of these experiments, a can be evaluated directly

from Eq. (6). The phase shift angle was calculated from Eq. (3). It

follows that the theoretical pressure distribution over the full wave length is given by Eq. (2). The theoretical distributions are plotted in

Figs. 17b-21b for the open flume and in Figs. 25b-29b for the closed flume experiments. Note that each theoretical distribution is uniquely

deter-mined by the parameters of the velocity profile used. Accordingly, the

theoretical pressure distribution along the wave determined by the velocity profile at Station 0 (Figs. 17a and b) is different from that determined by the velocity profile at Station 4 (Figs. 18a and b). The experimental

(30)

4. DISCUSSION OF EXPERIMENTAL RESULTS

4.1. VELOCITY PROFILES

In the velocity profiles taken without the cover-plate, there is

a noticeable discontinuity at the critical level where U - c = 0 , i.e.,

v = 0 (cf., Figs. 17a-21a). This discontinuity seems to be caused by

fluctuations in the extremely small velocities near the critical level.

The velocity fluctuations then appear as fluctuations in the reading of the static-dynamic pressure difference obtained from the pitot-static

probe. In some instances these fluctuations were as large as the mean

difference between the static and dynamic pressures. However, one would

expect that the fluctuations would be averaged out by our data recording technique and that no discontinuity would appear in the velocity profile.

The velocity fluctuations arise from two sources. First, small

surface disturbances may induce subsurface fluctuations. Addition of the

cover-plate would reduce the size of this source of disturbance. Second,

through the viscous and turbulent transfer action in the critical layer,

there is a high shearing action in the layer. This action is particularly

evident near the critical level where the mean velocity is zero.

Accord-ingly, the flow direction can be expected to reverse itself randomly in the region of the critical level as a result of the shear- and surface-disturbance-induced fluctuations.

A flow reversal has a serious impact on the apparent reading of a pitot-static probe of the type used here. To investigate this impact, a

probe was placed in a known flow and aligned in the proper or positive direction. The difference between static and dynamic pressures was

re-corded, and then the probe was reversed in the flow. Again, the pressure

difference was recorded. By carrying out this procedure for a number of

flow-rates, we were able to plot a probe calibration curve (Fig. 32). It

is seen that for a given velocity (pressure) the recording galvanometer

deflection is relatively very large for the probe-into-the-flow position.

Accordingly, in a fluctuating velocity field, the reading will be zero on the average only when the fluctuations are primarily in the negative direction (against the back of the probe). As we approach the critical

(31)

level, where v = 0 , from either side, the probe (which has been oriented

into the dominant flow direction) will give undue weight to positive flow.

In particular at the point where v = 0 , the probe will indicate a

posi-tive velocity. As a result, a discontinuity appears in our measured

velocity profiles where v should be equal to zero.

With the cover-plate in place, the surface source of velocity fluctuations was minimized. This reduced, but did not remove, the profile

discontinuity, indicating that the shearing action near the critical level is a significant source of direction-reversing velocity fluctuations.

Since the pitot-static probe was reversed when it passed through the critical level, the measured points to each side of the level tend to have equal, but opposite-in-sign, errors induced by the velocity

fluctua-tions. In the process of data reduction (see Sec. 3.8.1.) least-squares-fit straight lines were calculated for the presumed-logarithmic portion of each velocity profile (which included the critical level). The

least-squares differences were associated with the velocities in a data set of velocities versus distance from the moving boundary. Because of the

dis-tribution of the probe error near the critical level, the least-squares fit tended to balance out the velocity profile errors near v = 0 by

weighing the errors on both sides of this region equally, thus, averaging out their effect. In addition, the majority of the points used to establish the logarithmic profile lie outside the region where the probe error due to fluctuations is significant. This tends to decrease the effect of the few

points near the critical level. Note, however, that we are using the entire

measured velocity profile in our present analysis in contrast to Zagustin, et al. (1966) who used only points outside the critical layer.

4.2. PRESSURE PROFILES

4.2.1. Comparison With Miles' Theoretical Distribution

The procedure used to calculate the theoretical pressure distri-butions was discussed in Section 3.8.2. Given the values of U/ and yc

as input data, a computer program (see the Appendix) calculated the

(32)

program showing the distribution is given in Table 6. The distributions

predicted on the basis of the local profile at various stations are shown

as

solid lines in Figs. 17b-21b for experiments in the open flume and in Figs. 25b-29b for experiments with the cover-plate in place.

In the experiments without the cover-plate, the theoretical dis-tributions predicted for the velocity profiles at the two crests are es-sentially the same; that is, they have the same magnitude and phase shift.

This is expected because the flow is fully developed before it reaches the test section. For example, we see in Fig. 16 that the cross-channel

velocity profiles at the two crests are very similar. The magnitude and

phase shift of the theoretical pressure distribution at the crests are larger than at any other location on the wave. This is a result of the

combination of a large U1 and yc at these stations. The theoretical

distributions at Stations 4 and 12 are also very similar to each other.

The combination of a small

U1 and a large yc at Station 4 seems to

balance with a large U1 and a small yc at Station 12 to give

essen-tially the same theoretical pressure distribution. A combination of small

U and a small

yc at Station 8 produced a pressure distribution with 1

the minimum magnitude.

With the cover-plate on, the experimental pressure displayed less scatter than in the open flume experiments. In addition, with the

cover-plate on, the pressure measurements on the leeward side of the wave are slightly lower than in the open flume experiments. The difference could

be explained by the fact that the pressure measurements with the cover-plate on were taken in the region where the flow was fully two-dimensional and in the region of maximum horizontal velocity. This insured the

maxi-mum negative pressure readings. However, in the case of the open flume,

the free surface made it difficult to precisely determine a fully two-dimensional flow region in which the flow velocity was a maximum. As a

result, the pressure measurements may not have been taken in the region of maximum horizontal velocity.

In most cases, the theoretical pressure distribution displays a larger phase shift in the downwind direction than the experimental dis-tributions. This difference is probably due to viscous effects not

(33)

accounted for in the theory. That is, as mentioned above, Miles assumes

the air flow to be inviscid, incompressible and two-dimensional. Viscosity

acts only to maintain the logarithmic mean-velocity profile. However,

this difference in phase shift does coincide with Benjamin's (1959) con-clusion that the general effect of viscosity would be to retard the phase shift in the pressure distribution (cf., Zagustin, et al. (1966)).

It is clear from the above that the velocity profile changes sub-stantially from station to station along the wave and that this in turn affects the theoretical pressure distribution predicted from one local profile. It would seem reasonable to remove the profile variation by

averaging and to predict a "best" pressure distribution from the average profile parameters.

Accordingly, to investigate this effect of the local velocity pro-files on the theoretical pressure distribution, the values of U1 and

yc were averaged for Stations 0 and 16. These new values were then

averaged with the U1 and yc values taken from Stations 4, 8 and 12.

The theoretical pressure distribution associated with these average values is plotted in Fig. 33, for velocity profiles taken without the cover-plate,

and in Fig. 34 for those taken with the cover-plate in place. In both

cases, these distributions lie somewhere between the theoretical distribu-tions of Stadistribu-tions 4 and 12. The phase shift and magnitudes of the

experi-mental profiles agree reasonably well with the theoretical profiles pre-dicted by Miles.

4.2.2. Pressure Decay

Miles (1957) and Lighthill (1962) point out that within the approx-imations of Prandtl's boundary layer theory, the pressure variations across

the critical layer region can be assumed small. This assumption is

veri-fied by the results shown in Figs. 22, 30 and 31.

In Figs. 30 and 31, pressure distributions were plotted for equal distances from the moving boundary. The dashed lines in Fig. 31 were used

when insufficient data was taken at that distance to warrant a precise profile. In both figures, it can be seen that the pressure distribution

remains essentially constant through the critical layer, which was about

(34)

one inch thick, and starts to die out or decay with greater distance from

the critical layer. In Fig. 22, this decay can be seen more clearly. In

these plots, it can be seen that the static pressure difference between

the fixed and moving probe decays in an exponential fashion with distance

from the moving boundary. This would coincide with the Lighthill (1959) argument that pressure gradients decrease above the wave in an exponential manner.

4.3. STREAMLINE PATTERNS

The streamline patterns for the experiments with and without the cover-plate are shown in Figs. 35 and 36, respectively. These patterns

were sketched in order to obtain a qualitative picture of the flow

char-acteristics in the logarithmic region of the velocity profiles. To obtain

these streamlines, the velocity profiles at Stations 0, 4, 8, 12 and 16

were integrated to produce a set of distance points between each of which

0.05 cfs passed. These points of equal flow were plotted at each of the

above stations. A distorted scale was used to facilitate examination of

the region near the critical layer. A dashed line was placed through the

points where v = 0 . Starting from the moving boundary, we connected

points of equal flow by smooth lines that were guided between points by the contour of the wave. It was assumed that Lighthill's "cat's-eye

patterns" did exist around the critical layer. The Zagustin, et al.

(1966) flow visualization experiments supported this assumption (Figs. 8

and 9). Using Lighthill's physical interpretation of Miles' theory, we knew that the low pressure region was located in the center of the "cat's-eye pattern," while a high pressure region was located in the region where

the vortices meet. Reading from the experimental pressure distributions for each flow, one can see that the high pressure area is located at

Station 10 and the low at Station 2. Using this feature of the theory and

experiment, the center of the vortex and the region where the vortices

meet were located. Also known was the fact that the vortices distribute themselves essentially symmetrically around the low pressure area. Using

these facts, the estimated points where the streamlines crossed the

(35)

The large critical layer at Station 4, as well as the small criti-cal layer at Station 12, can be seen in the sketched patterns. The

vari-ance in the thickness of the critical layer perhaps can be attributed to the effect of a pressure gradient on boundary layer growth. In Figs. 30

and 31, it can be seen that the pressure distribution at Station 4 has a positive gradient and at Station 12 it has a negative gradient. As a

result of the positive, or favorable gradient, the boundary layer is the thickest at Station 4. Since the critical layer is directly affected by

the boundary layer, it also is the thickest at Station 4. As a result of

the negative, or adverse gradient, the boundary layer is diminished and the critical layer becomes thinner at Station 12. The streamline patterns

at Station 0 and Station 16 appear to be essentially the same. This

con-sistency supports the argument that the flow is fully developed at the test section.

(36)

5. CONCLUSIONS

The primary objective of this experimental investigation was to

extend and to improve the experimental data obtained by Zagustin, et al.

(1966) on steady-state simulation (in an open flume) of small amplitude

wind-generated waves. Extensive velocity and pressure measurements in the flow-field, particularly in the region of "critical level," were obtained to gain further insight into Miles' inviscid mechanism of trans-fer of energy from wind to water waves. In addition, measurements of

velocities and pressures were repeated with a cover-plate installed to minimize free-surface disturbances.

It has been found that the addition of the cover-plate over the

open flume reduces the free-surface disturbances and improves the quality

of the data. Fluctuations in pressure readings with the cover-plate in place were about one half of those in an open flume. The data agree with

the previous data obtained by Zagustin, et al. It may be concluded that

the free surface had a negligible effect in so far as the steady-state

simulation of wind blowing over progressive, small amplitude waves is

concerned.

It was determined that the critical layer thickness varies along

the wave and that this variation is correlated with the local pressure

gradient. As a result, the values of the logarithmic velocity profile

parameters U1 and

yc vary along the wave. The calculated theoretical

pressure distributions, based on the measured local values of U1 and yc at the wave crest and wave trough, had the largest and the smallest

amplitudes, respectively. The values of U1 and

yc measured at

inter-mediate stations along the wave yield calculated pressure amplitudes

lying between those at the crest and trough. However, most significantly, the theoretical pressure distribution based on the average values of U1

and yc along the wave agrees reasonably well with the experimental dis-tribution for the case of the open flume, as well as for the flume with a cover-plate in place.

Finally, the singular behavior of the velocity distribution in the region of critical level deserves further experimental investigation.

(37)

It would be extremely desirable to investigate the structure of this

region by employing hot-wire anemometry in a steady-state simulation of wind waves such as that accomplished here.

(38)

REFERENCES

Benjamin, T. B. (1959): Shearing flow over a wavy boundary, J. Fluid Mech., 6, pp. 161-205.

Bole, J. B. and E. Y. Hsu (1967): Response of Gravity Water Waves to Wind Excitation, Dept. of Civil Eng. Tech. Rept. No. 79, Stanford Univ., Stanford, Calif.

Bonchkovskaya, T. V. (1955): Wind Flow Over Solid Wave Models, Akademia Nauk SSSR, Morskoi Gidrofizicheskil Institute, 6, pp. 98-106.

Conte, S. D. and J. W. Miles (1959): On the numerical integration at the Orr-Sommerfeld equation, J. Soc. Indust. Appl. Math., 7,

pp. 361-366.

Hidy, G. and E. Plate (1965): Wind Action on Water Standing in a Labor-atory Channel, NCAR Manuscript No. 66, Boulder, Colo.

Larras, J. and A. Claria (1960): Wind Tunnel Research on Relative Wave and Wind Action, La Houille Blanche, No. 6, pp. 647-677.

Lighthill, M. J. (1962): Physical interpretation of the mathematical theory of wave generation by wind, J. Fluid Mech., 14,

pp. 385-398.

Miles, J. W. (1957): On the generation of surface waves by shear flows, J. Fluid Mech., 3, pp. 185-204.

Miles, J. W. (1959): On the generation of surface waves by shear flows, J. Fluid Mech., 6, pp. 568-582.

Miles, J. W. (1960): On the generation of surface waves by turbulent shear flows, J. Fluid Mech., 7, pp. 469-478.

Miles, J. W. (1962): On the generation of surface waves by shear flows. Part 4, J. Fluid Mech., 13, pp. 433-448.

Miles, J. W. (1967): On the generation of surface waves by shear flows. Part 5, J. Fluid Mech., 2, p. 417.

Motzfeld, H. (1937): Die turbulente Stromung an welligen Wanden, Z. angew. Math. Mech., 17, pp. 193-212.

Phillips, O. M. (1957): On the generation of waves by turbulent wind, J. Fluid Mech., 2, pp. 417-445.

Phillips, O. M. (1966): Dynamics of the Upper Ocean, Cambridge Univer-sity Press.

(39)

Shemdin, 0. H. and E. Y. Hsu (1966): The Dynamics of Wind in the Vicinity of Progressive Water Waves, Dept. of Civil Eng. Tech. Rept. No. 66, Stanford Univ., Stanford, Calif.

Snyder, R. L. and C. S. Cox (1966): A field study of the wind generation of ocean waves. Sears Foundation, J. Marine Res., 24(2),

pp. 141-178.

Stanton, T. E., D. Marshall and R. Houghton (1932): The Growth of Waves on Water Due to the Action of the Wind, Proc. Roy. Soc., A, 137, pp. 283-293.

Thijsse, J. Th. (1952): Growth of Wind-Generated Waves and Energy Transfer, Gravity Waves, Nat. Bur. Stand., Circular No. 521, Washington, D. C., pp. 281-287.

Zagustin, K., E. Y. Hsu, R. L. Street and B. Perry (1966): Flow over a Moving Boundary in Relation to Wind-Generated Waves, Dept. of Civil Eng. Tech. Rept. No. 60, Stanford Univ., Stanford, Calif.

(40)

APPENDIX

COMPUTER PROGRAM TO CALCULATE THEORETICAL PRESSURE DISTRIBUTIONS

This appendix presents the essential parameters needed as input data to a computer program that uses numerical integration of the Orr-Sommerfeld equation to solve for B and a . The program is based on

Conte and Miles (1959) and was developed for the IBM model 7090 by E. John Finnemore at Stanford University. For this study we have

modi-fied the program for use on an IBM model 360/67 computer. In addition,

we have added routines that use the CALCOMP x-y plotter to plot directly the pressure distributions presented in this text.

The program, of which a listing is given below, is written in Fortran IV and requires approximately one minute to compile and execute for one typical run. It consists of a main program and one subprogram

INTGRT.

PROGRAM VARIABLES

Input Variables

EPSILON (always positive) = w is the inte-gration limit used to bound the numerical Integration calculation away from the singu-larity in the Orr-Sommerfeld equation at

w =0 (E=1x10-5 ).

RNVDW 1/Aw , w = integrating step ( Aw = 1/64 ).

IDWPL Awt/Aw (must be an even integer multiple of

1/Aw) = 4.0 .

Number of stations on pressure plot.

NUMPAT Number of data sets or number of pressure plots.

WA Wave amplitude (inches).

WL Wave length (feet).

WC Wave celerity (ft/sec).

QCFS Water discharge (cfs).

Ul The reference velocity obtained from the slope of the logarithmic velocity profile.

YC The critical layer thickness. . ...

(41)

Important results calculated in the program are listed in the

sample printout in Table 6.

The subroutine INTGRT performs the numerical integration by the method of Runge and Kutta and is accurate to the order of (406.

The basic wave data is read in; Q , WO ( WO = - (WC/U1) ) and

zc are calculated by methods outlined in Section 3.8.2. The program

then uses the methods of Conte and Miles (1959) to calculate 6 and a .

Using a and 6 , the program then calculates the theoretical pressure

distribution above the wave by the methods of Section 3.8.2.

Sample Q and WO values from the table in Conte and Miles

(1959) were read into the program and the resulting 6 and a values were checked against those listed in the paper. In all cases the 6

values checked to i 0.01. In addition, in the pressure distribution

calculations, the computed values of 6 were checked against the 6

versus

zc from Miles (1959) (Fig. 3).

They agreed very well. The

a values calculated by the program method generally agreed with those

listed in Conte and Miles (1959). However, in some cases they disagreed

by ± 0.1.

In this study zc < 1 . Therefore, a was obtained directly

(42)

CALL STRTP1(10) 122 READ(5,50) E. ANV1J6, 106,PL, N 5O FORMAT( 20X, E10.2, F10.1, (10, 15) RD6PL = IDWPL DwPLuS = PIMPL/RNVDw 144 961 = 1.0/RNVOw OH? = OW1 - E E2 E*F 13 L2*E PI = 3.14159 UNDFLO = 0.1**35 CVP(LC = 10.0**35

NAMELIST / (ASTI / F, oWl, DWPLUS

WRITE(6,LIST1) NUMCUT = 1

RFAC(5,52) NUmDAT

NEW PAGE

PROGRAM ,,HICH USES MILES THEORY TO CALCULATE THE PRESSURE DISTRIBUTF-IN ABOVE A MOVING WAVE. FORM, USING THE THICKNESS CF THE CRITICAL LAYER (YC) AND THF CHARACTERISTICS IF THE LCGAkITE1C PROFILE (U1).

NOTATIGN

HA HAVE AFRLITUDE (INCHES)

wL = WAVE LENGTH (FEET)

wC = wAVF CELERITY (ET/SEC)

Wk WAVE NUMRER

QCFS = WATER DISCHARGE ICES)

Dl FOUND FRCM SLOPE OF SEMILOG PLOT OF EXP VELOCITIES

YC = CRITICAL THICKNESS FOUND FROM SAME PLOT (INCHES)

LC = 1 AT SINGULARITY (6=0)

WC = 4 AT Z=0 (HERE REQ. IWOI.ET.12 - SEE S.103)

= EPSILON (ALWAYS PuSITIVE) (= WBAR)

Dv. = DELTA

DwPLUS = DELTA 6+

RNVDW = 1/0w

Eli = PH1(SUB-1)

FIIX = SEkIES EXPANSION FOR Eli AT w = +CR- E FlIv = PHI(SU0-1) PRIME (DERIVATIVE W.R.T. W)

= SERIES EXPANSICN FOR FII AT w = +E

FI1XN = SERIES EXPANSION FCR FII AT W = -E

IDwPL = D6PLUS/C6 (+VE INTEGER)

(60 = HO/OW (-VI INTEGER)

NUmBER OF STATIONS ON THE PRESSURE PLOT NuMDAT = NUmBER OF DATA SETS

NUM)'LT = NUMBER OF EXPERIMENTAL PRESS PLOTS IN DATA SET

V REPRESENTS PRIME (DERIVITIVE W.R.T. w)

IMPLICIT PEAL*R(A-H, u-Z),INTEGEk*,,(I-N)

R(-AL*4 P(15)),PS(150),01(19).DS1(19) ,D2(19),DS2(19). 03(19),DS3(19),D4(19),OS4(19)

COMMON 6, OH, IC?, Y1, Y3, F11, FI3 SUBR

C. C, =

-FIIXP -= =

(43)

52 FURMAT(12) 1000 RL4D(5,51) wA, WL, WC, CFS, Ul, YC 51 FORMAT( 6F10.4) YCFT = YC / 12.0 WN = 2.0 * 3.1416/WL LC = WN * YCFT w0

- (WC /

01) CMEGA = 70 * (W(')**2 * OEXP(w0) WO = WO * 64.0 (WC = 40 RwO = IWO

TEMP = DABS (0 - RWO)

IF (TF.P.(17.0.5) IWO = IWO + I RWO = IWO WO = RWO/RNVCW 188 702 = 1C*10 W03 = WO*4,40*WO WPITE(6,60) 60 FORmA7(11,1, 13IH

/

WRITE(6,61) WO, OMEGA, ZC, YCFT

61 FORMAT(IHO, 8HFUR WO =, F6.2, 10H OMEGA =, F7.4, I4X, 7H...ZC =, F9.5,7H YCFT =,F6.4 ///)

AT SINGULAPITY PLUS AND MINUS EPSILON

200 71 = ZC2*E2/6.0 12 = ZC2*E3*7.0/35.0 73 = ZC2*(31.0+Z02+2C2)*E2*E2/ 240.0 74 = 7C2*(333.0+101.)*ZC2)*E2*E3 /5400.0 211 FI1XP = F*(1.0

+ LI

+ 72 Z3 + 74)

FI1XN = -FIIXP 2.0*112 + 14)*E

FIIVXP = 1.0 + 3.04,71 + 4.0*72 + 5.0*73 + 6.0*74 = FIIVXP - 6.0*72 - 12.0*14 22? 75 = (1.0 - ZC2)*E/2.0 Z6 = (3.0- 20.0*ZC2)*E2/36.0 Z7 = (6.0 - 137.0*ZC2 - 18.0*ZC2 *702)*E3/432.0 Z8 (45.0/1896.0 + 179.0*ZC2/ 632.0 - LC2*2C2)*74.0*E2* E2/900.0 233 FI3XP = -1.0 + L*(Z5 + 76 r 77 Z8)

FI3XN = FI3X12 - 2.0*176 + Z8)*E

FI3VXP = 2.0*75 + 1.0*Z6 + 4.0*77 5.0*Z8 FI3VXN FI3VXP - 4.0*Z5 - 8.0*Z7 -- - - --- -FIIVXN - -= -+ -+ -

(44)

-NUMERICAL INTEGRATION FOR VI, Y3, Ni, FI3 FROM W=-E TO WO

Y1V = DYI/DW = YI-F11(1/w-ZC2*EXR(2W)) OE 1 (1)

FIIV = DFII/OW = YI OF 2 (1)

Y3V = DY3/DW = Y3-FI3(1/W-ZC2*EXPI2W1)

+F11(1+1/W)/W-2Y1/W OF I (2/

FI3V = DFI3/DW - Y3 DE 2 (2)

C

NOTE..

EXP(X), X -Gr;,. 68.029692 NOT ALLOWED

KCUNT1 = 0 300 KOUNT1 = KOUNI1 + 1 If-(KCUNTI .CT. 1) GO IC 333 311 W = -E Vi = FIIVXN Y3 = FI3VXN FII = FI1XN FI3 = F13XN 322 DW = -DW2 GO TO 344 333 OW = -DWI 344 CALL INTGRT 355 KTCHK1 = -WO*RNVOW*1.2 IF(KOUNT1 .GT. KTCHK1) GO TO 911

IF( W .LT. WO-OW/2.0 .AND. W .GT. wO+DW/2.0 I GO TG 400

GC TO 300 HENCE, AT WO 400 YIWO = Y1 FIlw0 =

FII

Y3w0 = Y3 FI3W0 = FI3 wATwO = W KNTIWC = KOUNT1 FOR FI2W0..

FI2 = FIl*LOGW + FI3

C AT WO, A IS -VE, SO FI2 = FI2R + H*F12I COMPLEX

WHERE FI2R = FI1*LOGI-W1 + FI3 REAL PRT

FI2I -P1*FII [MAO PRT

THIS IS 4CCCUNTE0 FOR IN THE SOLUTION (REF 900) FOR A AND

411 ROGWO = OLOC(-WO)

FI2Rw0 = FII*RCCWO + FI3 REAL PRT

NUMERICAL INTEGRATION FOR Y1,

Y3. Ni, FI3

FROM W=+E TO W+

C

=

Cytaty

Powiązane dokumenty

In the GOCE case the algorithm results in two separate cross-wind data sets, one derived from linear accelerations (force-derived), the other from angular

In Chapter 6, basic mechanical models of high-rise buildings are used to model high-rise buildings to study the damping contribution of different energy dissipation sources to the

Niedojrzałe postawy wobec rodzicielstwa częściej wiążą się z takimi właściwościami indywidualnymi młodych dorosłych, jak niski poziom oceny wartości religijnych

Wartości religijne stanowią podstawę holistycznego rozwoju osobowości dzieci, dlatego dorośli nie powinni przyjmować wobec nich postawy bierności, bądź wrogości, lecz

Al- though reservoir depth and large bodies inside the reservoir are detectable, small-scale resistivity structures due to steam injection that are estimated from reservoir

1 „Ona (imaginacja) stwarza monstra, których w naturze nie masz; przeciw niej uzbroić się nam należy. Jeżeli szczęśliw ość w doczesnem życiu znajdować

Assessment of benefits from the projects is based on energy production, the cost of MWh produced in comparison with the level of NER300 funding (public return cost), avoided CO

Wanneer je naar Amsterdam kijkt, zie je echter dat de licht- en donkergrijze staafjes bijna even hoog zijn; vraag (huishoudgroottes) en aanbod (woninggroottes) komen min of