ABSTRACT Title of dissertation: AIR ENTRAINMENT IN THE TURBULENT SHIP HULL BOUNDARY LAYER Nathan Washuta, Doctor of Philosophy, 2016 Dissertation directed by: Professor James Duncan Department of Mechanical Engineering Turbulent fluctuations in the vicinity of the water free surface along a flat, ver- tically oriented surface-piercing plate are studied experimentally using a laboratory- scale experiment. In this experiment, a meter-wide stainless steel belt travels hori- zontally in a loop around two rollers with vertically oriented axes, which are sepa- rated by 7.5 meters. This belt device is mounted inside a large water tank with the water level set just below the top edge of the belt. The belt, rollers, and supporting frame are contained within a sheet metal box to keep the device dry except for one 6-meter-long straight test section between rollers. The belt is launched from rest with an acceleration of up to 3-g in order to quickly reach steady state velocity. This creates a temporally evolving boundary layer analogous to the spatially evolv- ing boundary layer created along a flat-sided ship moving at the same velocity, with a length equivalent to the length of belt that has passed the measurement region since the belt motion began. Surface profile measurements in planes normal to the belt surface are con- ducted using cinematic Laser Induced Fluorescence and quantitative surface pro- files are extracted at each instant in time. Using these measurements, free surface fluctuations are examined and the propagation behavior of these free surface ripples is studied. It is found that free surface fluctuations are generated in a region close to the belt surface, where sub-surface velocity fluctuations influence the behavior of these free surface features. These rapidly-changing surface features close to the belt appear to lead to the generation of freely-propagating waves far from the belt, outside the influence of the boundary layer. Sub-surface PIV measurements are performed in order to study the modifi- cation of the boundary layer flow field due to the effects of the water free surface. Cinematic planar PIV measurements are performed in horizontal planes parallel to the free surface by imaging the flow from underneath the tank, providing streamwise and wall-normal velocity fields. Additional planar PIV experiments are performed in vertical planes parallel to the belt surface in order to study the bahvior of stream- wise and vertical velocity fields. It is found that the boundary layer grows rapidly near the free surface, leading to an overall thicker boundary layer close to the sur- face. This rapid boundary layer growth appears to be linked to a process of free surface bursting, the sudden onset of free surface fluctuations. Cinematic white light movies are recorded from beneath the water surface in order to determine the onset location of air entrainment. In addition, qualitative observations of these processes are made in order to determine the mechanisms leading to air entrainment present in this flow. AIR ENTRAINMENT IN THE TURBULENT SHIP HULL BOUNDARY LAYER by Nathan Washuta Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2016 Advisory Committee: Dr. James H. Duncan, Chair/Advisor Dr. James Baeder, Dean’s Representative Dr. Kenneth Kiger Dr. Johan Larsson Dr. Amir Riaz c© Copyright by Nathan Washuta 2016 Acknowledgments First and foremost, I would like to thank my advisor, Dr. James Duncan, for all that he has taught me over the years. Without his advice and guidance these past six years, I would not have become half the student, engineer, or researcher I am today. I will always strive to match the example he sets. I would also like to thank my labmates, without whom I would never have been able to accomplish anything worthwhile in graduate school. To Naeem Masnadi, An Wang, Ren Liu, Martin Erinin, Christine Ikeda, Kyle Corfman, Jayson Geiser, Rahul Mulinti, Dana Ehyaei, and so many others, for being so smart, hard-working, generous, and kind. I wish them all the success they so clearly deserve. I also want to thank Dr. Ken Kiger, Dr. Amir Riaz, Dr. Johan Larsson, and Dr. James Baeder for generously donating their valuable time and agreeing to serve on my dissertation committee and for the many valuable and productive discus- sions. Special thanks go to Dr. Kiger for giving me my first crack at research as an undergraduate. My deepest thanks go to Jennifer Dahne, who endured my forever open-ended quest for a Ph.D. Without her love and support, I would never have dared to attempt anything so great. Thank you again to Kyle Corfman and to Rosie Myers, for taking me in and for being great friends. And last, but certainly not least, I want to thank my family. The love and caring they have always given to me inspire me every day to want to be better. ii Contents List of Figures iv 1 Introduction 1 1.1 Background and Motivation . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Prior Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.3 Research Considerations for a Laboratory-Scale Experiment . . . . . 12 2 Experimental Details 15 2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 2.2 Experimental Facilities . . . . . . . . . . . . . . . . . . . . . . . . . . 16 2.2.1 Water Tank . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 2.2.2 Ship Boundary Layer Device . . . . . . . . . . . . . . . . . . . 20 2.3 Experimental Techniques . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.3.1 Surface Profile Measurements . . . . . . . . . . . . . . . . . . 32 2.3.2 Fluid Velocity Measurements . . . . . . . . . . . . . . . . . . 38 2.3.3 Air Entrainment Measurements . . . . . . . . . . . . . . . . . 59 3 Results and Discussion 63 3.1 Surface Profile Measurements . . . . . . . . . . . . . . . . . . . . . . 63 3.1.1 Wave Propagation . . . . . . . . . . . . . . . . . . . . . . . . 68 3.1.2 Free Surface Bursting . . . . . . . . . . . . . . . . . . . . . . . 76 3.1.3 Surface Fluctuations . . . . . . . . . . . . . . . . . . . . . . . 77 3.2 Flow Field Measurements . . . . . . . . . . . . . . . . . . . . . . . . 79 3.2.1 PIV measurements with a horizontal light sheet . . . . . . . . 79 3.2.2 PIV Measurements with a vertical light sheet . . . . . . . . . 99 3.3 Air Entrainment Events . . . . . . . . . . . . . . . . . . . . . . . . . 110 3.3.1 Event Timing . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 3.4 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . 122 3.5 Directions for Future Work . . . . . . . . . . . . . . . . . . . . . . . . 125 iii List of Figures 1.1 Photograph of naval combatant ship showing zone of white water next to the hull. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.2 Drawing from Stern (1986) defining the various flow regions in a ship hull boundary layer. . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.3 Drawing from Grega et al. (1995) hypothesizing the secondary flow motions that exist in the mixed boundary corner. . . . . . . . . . . . 5 1.4 A series of plots from Sreedhar and Stern (1998) comparing the sec- ondary flow motions found under a variety of boundary conditions. . 7 1.5 Regions of various types of surface motions for free surface turbu- lence with velocity fluctuation magnitude q (vertical axis) and length scale L (horizontal axis), from Brocchini and Peregrine (2001). Air entrainment and spray production occur in the upper region, above the uppermost curved line. The three data points are values obtained for the turbulent boundary layer on a flat plate with q taken as the rms of the spanwise (which is vertical for the boundary layer along a ship hull) velocity fluctuation (w′) and L taken as the boundary layer thickness (δ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 1.6 Two plots from Kim et al. (2014), showing wall-normal distributions of turbulent velocity fluctuations in the (a) streamwise and (b) ver- tical (−) and wall-normal (−·) directions. . . . . . . . . . . . . . . . . 12 1.7 Two figures from Kim et al. (2013), showing a visualization of air entrainment in a turbulent Couette flow from (a) a side view and (b) an end view. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 2.1 A photo of the open-surface water tank, showing the transparent acrylic walls and steel support frame. . . . . . . . . . . . . . . . . . . 16 2.2 A top view photo of one of the seams between acrylic panels that form the walls of the water tank. Each gap between panels is first filled with flexible Tygon tubing to reduce the thickness of the silicone caulk layer, allowing for a fully-cured caulk layer. . . . . . . . . . . . 17 2.3 A photo of the skimmer at one end of the tank. Its height is adjustable so that it can be set just below the desired water level in the tank. . . 18 2.4 Top view schematic of the SBL device and the tank where experiments are performed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 2.5 A photo of one of the rollers during construction of the SBL. . . . . . 21 iv 2.6 A photo of the belt construction in progress, showing major com- ponents of the device, including the rollers, steel support frame, hy- draulic motors, and stainless steel belt. . . . . . . . . . . . . . . . . . 22 2.7 A photo of one end of the SBL, showing the stainless steel sheet metal dry box. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 2.8 A photo of the overflow drain piping that allows the SBL dry box to drain in the event of a sump pump failure. . . . . . . . . . . . . . . . 24 2.9 A top view schematic of the entrance fairing, designed to create a smooth transition over the backwards-facing step at the location where the belt exits the SBL dry box. . . . . . . . . . . . . . . . . . . 25 2.10 Two photos of the hydraulic system showing (a) the Hydraulic Power Unit (HPU) and (b) hydraulic fluid accumulators. . . . . . . . . . . . 26 2.11 Two photos of the crane system used to move the SBL device in and out of the tank. Image (a) shows one of the cranes that is used and (b) shows a photo of the process of lifting the device. . . . . . . . . . 27 2.12 Measurements of belt trajectory for 5 repeated runs at each belt speed. The left column shows the belt velocity, U , versus time rel- ative to the trigger signal that comes from the belt control system. The right column shows the amount of belt travel, x, versus time relative to trigger. Belt speeds of 3 m/s are shown in (a) and (d), 4 m/s are shown in (b) and (e), and 5 m/s are shown in (c) and (f). . 29 2.13 A comparison of the average belt launch trajectory for each belt speed. Plot (a) shows the belt velocity U versus time, while plot (b) shows the belt velocity U versus x, the amount of belt travel. . . 30 2.14 Schematic drawing showing the set up for the cinematic LIF mea- surements of the free surface shape from (a) a side view and (b) an end view perspective. . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.15 One sample frame from an LIF movie during a belt launch to 5 m/s. Image (a) shows a raw image from the movie camera and image (b) shows that same image with the result of the image processing plotted on top. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 2.16 Sample images from intermediate steps of the processing alogirthm used to extract instantaneous surface profiles from LIF images . . . . 35 2.17 A side view schematic of the horizontal PIV setup. . . . . . . . . . . 39 2.18 A photo of the horizontal PIV setup made necessary by space con- straints. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 2.19 A photo of the Type 309-15 calibration target used to correct for optical distortions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 2.20 Two images of the periscope used to create a light sheet below the wa- ter surface showing (a) an overall schematic and (b) the configuration used in the horizontal light sheet experiments. . . . . . . . . . . . . . 42 2.21 Two images of the vertical PIV setup showing (a) a side-view schematic and (b) a photo of the actual setup through the wall of the tank, showing the orientation of the light sheet. . . . . . . . . . . . . . . . 44 2.22 Particle frequency response for a range of particle diameters. . . . . . 49 v 2.23 Four photos of the particle seeding system. Image (a) shows the basin in which high-concentration particle solution is mixed. Image (b) shows the pump that draws from this basin and pumps to (c) the particle seeder. Image (d) is a photo from beneath the water of an initial test run of the particle seeder using fluorescent dye rather than particles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 2.24 Two images of the results from the PIV experiments with a horizontal sheet far from the free surface during a launch to 4 m/s at x = 4.2 m. Image (a) shows a sample raw PIV image and image (b) shows the processed vector field with the vector spacing reduced by 1/2 and the vector length increased by 10 times for clarity. The background is colored by vector length with a scale running from 0 to 2.5 m/s. . . . 54 2.25 Two images of the results from the PIV experiments with a vertical sheet close to the belt surface during a launch to 3 m/s at x = 5.9 m. Image (a) shows a sample raw PIV image and image (b) shows the processed vector field with the vector spacing reduced by 1/2 and the vector length increased by 10 times for clarity. The background is colored by vector length with a scale running from 0 to 1.0 m/s. . . . 55 2.26 A plot showing the viscous sublayer thickness based on the Schultz- Grunow formula for skin friction, taken from Schlichting (1979) . . . 57 2.27 Measurements of the y-direction belt position throughout each run with each plot showing the results of 20 repeated runs at (a) 3 m/s, (b) 4 m/s, and (c) 5 m/s. . . . . . . . . . . . . . . . . . . . . . . . . . 58 2.28 Measurements of the RMS velocity fluctuations of the belt, normal- ized by belt speed throughout each run. . . . . . . . . . . . . . . . . . 59 2.29 A photo of the sub-surface planar bubble measurement system, illu- minated using two 650 watt flood lamps. . . . . . . . . . . . . . . . . 60 2.30 A schematic of the underwater stereo bubble measurement system . . 61 3.1 A sequence of five images from a high speed movie of the free surface during a belt launch to 5 m/s. These images are taken at equivalent belt lengths of (a) 0 m (b) 1 m (c) 2 m (d) 3 m (e) 4 m and (f) 5 m from the bow of the ship. The high reflectivity of the stainless steel belt makes it appear as a symmetry plane on the left side of the images. The horizontal field of view for these images is approximately 31 cm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 3.2 A sequence of five images from a high speed movie of the free surface during a belt launch to 5 m/s. These images are taken at equivalent belt lengths of (a) 0 m (b) 5 m (c) 10 m (d) 15 m and (e) 20 m from the bow of the ship. The high reflectivity of the stainless steel belt makes it appear as a symmetry plane on the left side of the images. The horizontal field of view for these images is approximately 31 cm. 65 vi 3.3 Sequence of profiles of the water surface during belt launch to 3 m/s starting from x = 21 m. The time between profiles is 4 ms and each profile is shifted up 2 mm from the previous profile to reduce overlap and show propagation of surface features throughout time. The belt is positioned at the left edge of the image (y = 0). . . . . . . . . . . . 67 3.4 Cross correlation maps for profiles from x = 0 to 5 m. The left column shows cross correlation maps averaged over a region close to the belt and those in the right column are averaged over a region far from the belt. The three rows from top to bottom contain plots from U = 3, 4, and 5 m/s, respectively. The black line in each plot indicates the ∆x location of the maximum correlation for each ∆t slice. . . . . . . 69 3.5 Cross correlation maps for profiles from a launch to 3 m/s. Images (a) and (c) show cross correlation maps averaged over a region close to the belt and (b) and (d) are averaged over a region far from the belt. Images (a) and (b) are averaged over x positions from 5.85 m to 17.85 m and images (c) and (d) are averaged over x positions from 17.85 m to 29.85 m. The black line in each plot indicates the ∆x location of the maximum correlation for each ∆t slice. . . . . . . . . . 71 3.6 Cross correlation maps for profiles from a launch to 4 m/s. Images (a) and (c) show cross correlation maps averaged over a region close to the belt and (b) and (d) are averaged over a region far from the belt. Images (a) and (b) are averaged over x positions from 5.85 m to 17.85 m and images (c) and (d) are averaged over x positions from 17.85 m to 29.85 m. The black line in each plot indicates the ∆x location of the maximum correlation for each ∆t slice. . . . . . . . . . 72 3.7 Cross correlation maps for profiles from a launch to 5 m/s. Images (a) and (c) show cross correlation maps averaged over a region close to the belt and (b) and (d) are averaged over a region far from the belt. Images (a) and (b) are averaged over x positions from 5.85 m to 17.85 m and images (c) and (d) are averaged over x positions from 17.85 m to 29.85 m. The black line in each plot indicates the ∆x location of the maximum correlation for each ∆t slice. . . . . . . . . . 73 3.8 Wave propagation speed far from the belt averaged over 5.85 m < x <17.85 m versus belt speed. . . . . . . . . . . . . . . . . . . . . . . 74 3.9 A sequence of six images from a high speed movie of the free surface during a belt launch to 3 m/s. These images are cropped to have a horizontal field of view of approximately 7 cm. Each image from (a) to (f) is spaced by 20 ms, corresponding to 6 cm of belt travel between images.The first image occurs at x = 2.45 m . . . . . . . . . 75 3.10 Two plots showing (a) the average x location of the onset of free surface bursting versus belt speed and (b) the associated Reynolds number based on x. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 3.11 RMS surface height fluctuation averaged over y and x versus belt speed for an equivalent ship length of 30 m. . . . . . . . . . . . . . . 77 vii 3.12 RMS surface height averaged over repeated experimental runs versus distance from the belt for three belt speeds. . . . . . . . . . . . . . . 79 3.13 Mean streamwise velocity profiles at a belt speed of U = 3 m/s for (a),(c) D = 14 cm and (b),(d) D = 2.5 cm. Images (a) and (b) each plot a profile at each 0.5 m of belt travel from x = 0 to 5 m, while (c) and (d) each plot a profile at each 5 m of belt travel from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 3.14 Mean streamwise velocity profiles at a belt speed of U = 4 m/s for (a),(c) D = 14 cm and (b),(d) D = 2.5 cm. Images (a) and (b) each plot a profile at each 0.5 m of belt travel from x = 0 to 5 m, while (c) and (d) each plot a profile at each 5 m of belt travel from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 3.15 Mean streamwise velocity profiles at a belt speed of U = 5 m/s for (a),(c) D = 14 cm and (b),(d) D = 2.5 cm. Images (a) and (b) each plot a profile at each 0.5 m of belt travel from x = 0 to 5 m, while (c) and (d) each plot a profile at each 5 m of belt travel from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 3.16 Plots showing (a-c) momentum thickness and (d-f) displacement thick- ness versus x calculated from measured boundary layer profiles for belt speeds of (a, d) 3 m/s, (b, e) 4 m/s, and (c, f) 5 m/s. . . . . . . 84 3.17 Plots showing shape factor, H, as a function of x for belt speeds of (a) 3 m/s, (b) 4 m/s, and (c) 5 m/s. . . . . . . . . . . . . . . . . . . 86 3.18 Plots showing different scaling parameters for bursting onset, cal- culated using measured velocity profile data. In all plots, blue dots indicate velocity from D = 14 cm and red dots indicate velocity closer to the surface at D = 2.5 cm. . . . . . . . . . . . . . . . . . . . . . . 88 3.19 Streamwise rms velocity profiles at a belt speed of U = 3 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. These profiles are plotted in (a) and (b) at each 1 m of belt travel from x = 1 to 5 m and in (c) and (d) at each 5 m of belt travel from x = 5 to 30 m. . . . . . . 90 3.20 Streamwise rms velocity profiles at a belt speed of U = 4 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. These profiles are plotted in (a) and (b) at each 1 m of belt travel from x = 1 to 5 m and in (c) and (d) at each 5 m of belt travel from x = 5 to 30 m. . . . . . . 91 3.21 Streamwise rms velocity profiles at a belt speed of U = 5 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. These profiles are plotted in (a) and (b) at each 1 m of belt travel from x = 1 to 5 m and in (c) and (d) at each 5 m of belt travel from x = 5 to 30 m. . . . . . . 92 3.22 Streamwise profiles comparing the velocity profiles at different dis- tances from the free surface. The left column shows plots of U and the right column shows plots of urms. Each column contains plots at each 1 m from x = 1 to 5 m. The top-most plots correspond to x = 1 m, while the bottom plots correspond to x = 5 m. . . . . . . . . . . 93 viii 3.23 Plot of x locations of four different events that indicate transition to turbulence. Points in blue represent data from measurements at D = 14 cm and points in red represent data from measurements at D = 2.5 cm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 3.24 Wall-normal rms velocity profiles at a belt speed of U = 3 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. In plots (a) and (b), profiles are plotted at each 0.5 m of belt travel from x = 0 to 5 m, while in plots (c) and (d), profiles are plotted at each 5 m of belt travel from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . 95 3.25 Wall-normal rms velocity profiles at a belt speed of U = 4 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. In plots (a) and (b), profiles are plotted at each 0.5 m of belt travel from x = 0 to 5 m, while in plots (c) and (d), profiles are plotted at each 5 m of belt travel from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . 96 3.26 Wall-normal rms velocity profiles at a belt speed of U = 5 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. In plots (a) and (b), profiles are plotted at each 0.5 m of belt travel from x = 0 to 5 m, while in plots (c) and (d), profiles are plotted at each 5 m of belt travel from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . 97 3.27 Wall-normal rms velocity profiles for belt speeds of (a) 3, (b) 4, and (c) 5 m/s at D = 14 cm. . . . . . . . . . . . . . . . . . . . . . . . . . 98 3.28 Streamwise mean velocity profiles at a belt speed of U = 3 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . 100 3.29 Streamwise mean velocity profiles at a belt speed of U = 4 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . 101 3.30 Streamwise mean velocity profiles at a belt speed of U = 5 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . 102 3.31 Streamwise rms velocity profiles at a belt speed of U = 3 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 ix 3.32 Streamwise rms velocity profiles at a belt speed of U = 4 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 3.33 Streamwise rms velocity profiles at a belt speed of U = 5 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 3.34 Vertical rms velocity profiles at a belt speed of U = 3 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 3.35 Vertical rms velocity profiles at a belt speed of U = 4 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 3.36 Vertical rms velocity profiles at a belt speed of U = 5 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 3.37 Three sequences of images from underwater white light movies at three different belt speeds. The three columns contain frames from belt launches to 3, 4, and 5 m/s respectively. Each row contains frames from each 5 m of belt travel from x = 0 to 30 m. . . . . . . . 112 3.38 Two sequences of images from LIF movies depicting potential air entraining events. Images in the left column depict a trench closure event and images in the right column depict a wave breaking event. Both sets of images are from a belt launch to 5.0 m/s. . . . . . . . . . 113 3.39 Two sequences of images from underwater white light movies of the same air entrainment event. The belt is moving from left to right. Images from the left column were captured with the left camera and images from the right camera were captured with the right camera. This event occured during a launch to 4.0 m/s. Because of the lighting and imaging configuration shown above in Figure 2.30, each image contains both the bubble of interest and its shadow on the belt. . . . 116 3.40 Two sequences of images from underwater white light movies of the same air entrainment event. The belt is moving from left to right. Images from the left column were captured with the left camera and images from the right camera were captured with the right camera. This event occured during a launch to 5.0 m/s. . . . . . . . . . . . . 117 x 3.41 A set of timelines depicting the x location of various events described throughout this dissertation. . . . . . . . . . . . . . . . . . . . . . . . 119 3.42 Plots showing (a) θ vs x and (b) urms vs x for the initial 5 m of belt travel at each belt speed. . . . . . . . . . . . . . . . . . . . . . . . . . 120 3.43 Plots showing Weber number vs Froude number, calculated using θ as a length scale and a velocity scale of (a) U or (b) urms at y = 3.04 mm, the location of the secondary peak in streamwise velocity fluctuations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 3.44 A figure from Brocchini and Peregrine (2001) depicting critical regions for air entrainment based on length and velocity scales, with recorded data of θ and urms plotted on top. . . . . . . . . . . . . . . . . . . . . 122 xi Chapter 1: Introduction 1.1 Background and Motivation It has long been observed that a layer of white water is present along the hulls of large naval ships. This white water indicates that air entrainment is occurring, but it is not clear whether these bubbles are entrapped by breaking bow waves and swept downstream into the boundary layer or if turbulent fluctuations in the boundary layer are strong enough to entrain air. Although the mechanism of this entrainment is still unclear, the result is that bubbles formed during this process are swept into the wake of the ship, where they can persist for over a mile. This bubbly wake can be detected by sonar, thus enhancing the remotely-detected signature of naval combatant ships. While the air entrainment caused by breaking bow waves has been studied previously, the possibility of bubble generation due to near-surface velocity fluctuations in the turbulent boundary layer has received little attention. 1.2 Prior Research Research into boundary layer turbulence in the vicinity of a free surface first began in the 1980s when Hotta and Hatano (1983) considered the turbulence quan- 1 Figure 1.1: Photograph of naval combatant ship showing zone of white water next to the hull. tities in the near wake of a double tanker ship hull model, which is symmetric about a horizontal center plane. In one condition, this model is fully submerged and in another case, the model is set to pierce the free surface where the air-water interface coincides with the model symmetry plane. In this way, the authors make compar- isons to the flow around the ship hull model with and without a free surface. These experiments were performed in a circulating water channel and velocity was mea- sured in both cases using a hot wire probe at depths of 0.4d, 0.58d, and d, where d is the distance from the horizontal symmetry plane to the bottom of the model. They found that in the case where a free surface is introduced, the turbulence intensities associated with vertical velocity fluctuations are greatly reduced at shallow depths. 2 In an attempt to isolate the turbulent boundary layer flow in the vicinity of a free surface from the effects of hull shape, theoretical and numerical work by Stern (1986) and the corresponding experimental research by Stern et al. (1989) looked at the influence of a small-amplitude, two-dimensional Stokes wave on the boundary layer of a surface-piercing flat plate. The first of these studies focused on using scaling arguments to determine important flow field scaling parameters for regions influenced by the solid wall and free surface boundary conditions, as shown in Figure 1.2. The authors focused on flow in the region where the free surface boundary condition had the largest influence (Region II in Figure 1.2), the region where effect of the ship hull boundary layer was dominant (Region III), and the region where both of these boundary conditions had a significant effect (Region IV). It was found that wave steepness was the primary parameter that had a significant effect on the boundary layer near the free surface, resulting in large variations in boundary layer thickness. The wave effects present near the free surface in Regions II and IV were found to have a significant influence at depths up to half of the wavelength of the imposed surface wave. Following this research, two primary groups of researchers began attempting to isolate the effects of boundary layer turbulence from the flow produced by the shape of the ship hull by studying what was referred to as a ”mixed-boundary corner flow.” This type of flow is characterized by flow in the vicinity of a corner formed by a vertical, stationary wall and a horizontal, flat, free-shear boundary. The first of these studies, performed by Grega et al. (1995), consisted of an experi- mental and numerical investigation of flow near a surface-piercing flat plate. In this 3 Figure 1.2: Drawing from Stern (1986) defining the various flow regions in a ship hull boundary layer. work, turbulent structures and free-surface effects were studied. Experiments were performed along a 4.5 meter-long flat plate that pierced the free surface in an open- surface water tunnel. In these experiments the Froude number, which describes the ratio of turbulent kinetic energy to gravitational potential energy, was limited to 0.003, effectively eliminating the effects of waves and surface deformations from this flow. The numerical study emulates this condition using a flat, free-shear boundary condition at the free surface. The experiments were performed using both Laser Induced Fluorescence (LIF) and shadowgraph for qualitative flow visualization and Laser Doppler Velocimetry (LDV) to record single-component point velocity mea- surements. Experiments were performed with free-stream velocities up to 21 cm/s. From flow visualization studies, the authors found that vortex lines that connected to the free surface caused localized depressions in the free surface, which act as a sig- nificant source of transport for surfactant from the free surface into the bulk. Both numerical and experimental results indicated the presence of a streamwise vortex 4 Figure 1.3: Drawing from Grega et al. (1995) hypothesizing the secondary flow motions that exist in the mixed boundary corner. near the intersection of the flat plate and the free surface. Additional secondary flow features were found, but these results differed between experiments and simulations. The authors hypothesized a secondary flow motion shown in Figure 1.3. Longo et al. (1998) performed an experimental investigation of a towed, vertically- oriented, surface-piercing flat plate. In this study, the flat plate was towed at low speed (corresponding to a low Froude number) to minimize free surface disturbances. To measure the flow in the vicinity of the plate, three-component point-velocity measurements beneath the free surface were performed using a Laser Doppler Ve- locimetry (LDV) system. It was found that a streamwise vortex was formed near the solid/free-surface junction, while a second streamwise vortex rotating in the op- posite direction was formed beneath the free surface in the outer boundary layer. 5 These vortices were found to have the effect of increasing turbulent mixing between the boundary layer and outer flow region, leading to a thickening of the flat plate boundary layer near the free surface. Sreedhar and Stern (1998) performed a similar numerical study using Large Eddy Simulation (LES) on a temporally evolving junction flow, incorporating a rigid lid as an analog for the free surface. The rigid lid boundary condition is cited as an approximation for a flat free surface, where vertical velocity as well as gradients of the horizontal velocity components in the vertical direction are set to zero. A comparison was made between this solid/rigid-lid and a solid/solid corner flow. The mean cross-stream profiles appeared to be quite similar between these two cases, forming an inner streamwise vortex with an outer flow that moved up along the vertical wall and out along the horizontal boundary. The presence of streamwise vortices at the solid/rigid-lid junction and reduction in vertical velocity fluctuations found by the authors was consistent with prior research studies and their results are shown in Figure 1.4. Additionally, it was found that despite a reduction in vertical velocity fluctuations, an overall increase in turbulent kinetic energy near the free surface was observed due to redistribution of vertical velocity fluctuations to the two horizontal components. Hsu et al. (2000) performed three-component LDV and horizontal planar Par- ticle Image Velocimetry (PIV) measurments near a surface-piercing wall in an open- surface water tunnel, extending the work from Grega et al. (1995). Measurements were again performed at low speed in order to minimize free surface deformations. In this research, the measurements of Longo et al. (1998) were called into question, 6 Figure 1.4: A series of plots from Sreedhar and Stern (1998) comparing the sec- ondary flow motions found under a variety of boundary conditions. as the flow was thought to be transient at the time of measurment. Additionally, the computations of the solid-solid corner flow presented in Sreedhar and Stern (1998) were questioned due to the lack of symmetry along a line that bisects the corner. The subgrid turbulence model was also called into question, as it assumes isotropy of turbulent fluctuations at small scales, which is considered by the authors to be an invalid assumption. In their own study, the authors found that near the solid boundary, streamwise velocity fluctuations are suppressed as the free surface is approached, while far from the solid wall, distance from the free surface has little effect on this quantity. Vertical velocity fluctuations are seen to decrease near the free surface at almost all wall-normal distances measured. A monotonic decrease in horizontal, wall-normal velocity fluctuations as the free-surface is approached is seen, although far from the belt, wall-normal fluctuations peak immediately at the 7 free surface. Additionally, it is found that turbulent kinetic energy production and dissipation are greatly reduced near the free surface, resulting in a redistribution of turbulent fluctuations to surface-parallel components. Grega et al. (2002) performed PIV in a cross-stream plane in order to con- clusively determine the nature of secondary flows in the mixed boundary corner, finding a slowly rotating streamwise vortex and wall-normal outflow from the wall near the free surface., similar to the hypothesized flow shown in Figure 1.3. Broglia et al. (2003) performed LES of an open-surface duct flow, utilizing LES to simulate this flow at three different Reynolds numbers. This work attempts to resolve discrepencies between prior research studies of a mixed boundary corner. The authors find similar weak secondary flow motions as have been seen in previous research, but also indicate the presence of a second outer vortex below the inner streamwise vortex, which is assumed to be caused by the presence of the solid- solid corner boundary at the bottom of the duct. The authors also examine the budgets of the streamwise vorticity equation, examining the contribution of each term to the mean secondary motion. These secondary motions are found to be caused by anisotropy of Reynolds stresses, with the production and pressure-strain redistribution terms having the largest effect. To this point, boundary layer turbulence in the vicinity of a free surface has only been studied at relatively low ship speed conditions where the free surface is approximately flat or disturbed only by imposed gravity waves. In order to study the physics of how turbulence interacts with the air-water interface, Shen and Yue (2001) performed DNS of a turbulent shear flow in the presence of a deformable free 8 surface. It was found that the energy cascade to small scales desreases near the free surface, while the stress becomes highly anisotropic due to the free-shear boundary condition. Additionally, it was found that vortex lines connect to the free surface, leading to the creation of regions where fluid is convected towards and away from the free surface, referred to as ’splats’ and ’antisplats’ respectively. The authors utilize these findings to develop more phyically-relevant sub-grid scale models for LES simulations. Teixeira and Belcher (2006) performed a theoretical study of the process by which pressure fluctuations generated in a turbulent shear flow drive free surface height fluctuations. It is found that surface fluctuations grow rapidly during the initial development of the boundary layer, but this growth slows considerably as time progresses. This is due to the development of long streaks in the flow, which work to shift the turbulent forcing to larger scales. It is also found that under certain conditions waves tend to grow through resonance, where the length and velocity scales of the turbulent forcing matches that of the free surface waves. This tends to cause preferential growth for freely-propagating waves. Additionally, this resonant forcing corresponds to a time scale, which affects the decorrelation time of the surface features. The flow conditions leading to air entrainment have long been an object of interest in a wide variety of flows. Chanson (1996) reviews a number of different types of high velocity, air-entraining flows and discusses the commonalities between them. The author finds that air entrainment in high-velocity flows can be caused by breaking waves, droplets or jets impinging onto the free surface, by the action of 9 vorticies near the air-water interface, or by free-surface instabilities and turbulent fluctuations near the free-surface. In reviewing previous studies, the author finds a wide range of onset conditions for air entrainment specific to each flow geometry. The author states that this variety of air entrainment processes are each dependent in some way on gravity effects, surface tension effects, or viscous effects, which are characterized by the Froude, Weber, and Reynolds numbers. Therefore, the author argues, experimental investigations of these phenomena should only be performed at full scale in order to retain the pyhsics of air entrainment that occur at full scale. In order to better understand the conditions under which air entrainment oc- curs, Brocchini and Peregrine (2001) utilize scaling arguments for critical Froude and Weber numbers in order to determine the dimensional regimes for turbulent velocity (q) and length (L) scales that correspond to air entrainment. The authors review a wide range of flows in which turbulence interacts with a free surface and combine this with scaling arguments to characterize surface features for different regions of the q-L plane. A plot from this paper showing different free surface behaviors in the q-L plane is shown in Figure 1.5, in which the upper region corresponds to the conditions necessary for air entrinament. Three data points have been included in this graph as estimates of q versus L in the ship boundary layer, based on classical correlations for flat plate boundary layers for a plate (ship) speed of 15 m/s at three streamise (X) locations. Because these estimates lie in the uppermost region of the plot, air entrainment is predicted. Only in recent years has air entrainment in wall-bounded turbulent flows be- come the focus of some research studies. In research presented by Kim et al. (2013), 10 X = 0.5 m X = 10.0 m X = 100.0 m Ballistic Figure 1.5: Regions of various types of surface motions for free surface turbulence with velocity fluctuation magnitude q (vertical axis) and length scale L (horizontal axis), from Brocchini and Peregrine (2001). Air entrainment and spray production occur in the upper region, above the uppermost curved line. The three data points are values obtained for the turbulent boundary layer on a flat plate with q taken as the rms of the spanwise (which is vertical for the boundary layer along a ship hull) velocity fluctuation (w′) and L taken as the boundary layer thickness (δ). direct numerical simulations (DNS) were performed of a turbulent two-phase couette flow. In this study, two parallel, vertical walls move in opposite directions to create a shear flow in the vicinity of a free surface. It is found that turbulent fluctuations reach a maximum at a small distance from each sidewall, as shown in Figure 1.6. Additionally, small air bubbles tend to accumulate near the sidewalls, while larger bubbles are more highly concentrated near the center of the channel. A visualization of the computed domain is shown in Figure 1.7. The very large concentration of bubbles reported in this study seems unphysical and appears to be contradicted by 11 Figure 1.6: Two plots from Kim et al. (2014), showing wall-normal distributions of turbulent velocity fluctuations in the (a) streamwise and (b) vertical (−) and wall-normal (−·) directions. experiments performed by Washuta et al. (2014), although the periodic nature of the computational domain would limit the outflow of bubbles to only occur at the free surface. This would essentially create a couette flow channel of infinite length, as opposed to the finite length of the experimental domain. Kim et al. (2014) later indicated that increasing the depth of the computational domain had a large effect of reducing air entrainment, which could also contribute to the differences seen here. 1.3 Research Considerations for a Laboratory-Scale Experiment In the proposed study, the turbulent free-surface boundary layer adjacent to a ship hull is explored via a novel laboratory-based experimental setup. The diffi- culty of designing laboratory scale experiments on ship boundary layers, including 12 Figure 1.7: Two figures from Kim et al. (2013), showing a visualization of air entrainment in a turbulent Couette flow from (a) a side view and (b) an end view. air entrainment due to the near-surface turbulence, centers around overcoming the effects of gravity and surface tension. Consider that in the free surface boundary layer, the air entrainment process is controlled by the ratios of the turbulent ki- netic energy to the gravitational potential energy and the turbulent kinetic energy to the surface tension eenergy. These ratios are given by the square of the Froude number (Fr2 = q 2 gL ) and the Weber number (We = ρq 2L σ ), where g is the gravita- tional acceleration, ρ is the density of water, σ is the surface tension of water, q is the characteristic magnitude of the turbulent velocity fluctuations and L is the length scale of this turbulence. In order to retain the wave and bubble behavior that control the air entrainment process at full scale, the Froude and Weber num- bers must be matched in model scale. Because these experiments will be performed on earth and in water, the gravity, density, and surface tension remain constant be- tween scales. Therefore, in order to retain both Froude and Weber number scaling, the full-scale length and velocity scales of naval combatant ships must be retained. All previous studies of this penomenon have been performed near the limit of zero 13 Froude number with a nearly flat free surface, leading to a large gap in the present knowledge. While some previous research has attempted to create a general frame- work for understanding air entrainment processes in turbulent flow, these criteria and mechanisms have yet to be applied to this problem. Therefore, this proposed research will attempt to study the mechanisms by which turbulent boundary layer fluctuations are capable of entraining air as is seen in real world flows by performing experiments at full length and velocity scales of large naval ships. In order to achieve this, these experiments will be performed using a flat-sided, surface-piercing belt which will launch impulsively from rest and travel along its length until it quickly reaches a constant speed. The development of the boundary layer along the belt in time will be equivalent to the boundary layer development along the length of a ship, where the total distance traveled by the belt is equivalent to the length along the hull of a flat-sided ship moving at constant velocity, making no bow waves. By using this impulsively started belt rather than a towed flat plate, there is no leading edge, meaning that the air entrainment produced along the flat belt can be isolated from other flow effects seen along the flat mid-hull section of a ship. In the following section, further details of the experimental apparatus will be presented, along with techniques for measuring water surface profiles, sub-surface velocity fields, and mechanisms for air entrainment. The final section will contain re- sults of these measurements and a discussion of the findings, followed by conclusions and directions for future work. 14 Chapter 2: Experimental Details 2.1 Overview The present research study is an experimental investigation of the boundary layer along a surface-piercing flat wall moving horizontally at full-scale ship speeds. To create this moving wall in a laboratory setting, a stainless steel belt is driven in a loop around two vertically-oriented rollers. The belt is started suddently from rest to create a temporally-evolving boundary layer analagous to the spatially-evolving boundary layer developed along a flat ship hull. The interaction of the boundary layer turbulence and the free surface is investigated using a range of experimental techniques. Water surface profiles are recorded in order to characterize height fluc- tuations and wave propagation behavior at the free surface, while particle image velocimetry measurements are performed beneath the free surface to determine the interaction of turbulence with the free surface under a range of conditions. Addi- tionally, white light movies from beneath the water surface are used to characterize the air entrainment behavior at each flow condition. 15 2.2 Experimental Facilities 2.2.1 Water Tank Experiments are performed in the large, open-surface water tank in the Hy- drodynamics Laboratory at the University of Maryland, as shown in the photo in Figure 2.1. This tank is approximately 13.4 m long, 2.4 m wide, and 1.3 m deep. The walls and floor of the tank are made of 3.18 cm thick transparent, abrasion- resistant acrylic sheets to allow for optical access from the bottom and on all four sides. The supporting structure of the tank is composed of steel beams that support Figure 2.1: A photo of the open-surface water tank, showing the transparent acrylic walls and steel support frame. 16 the acrylic walls with minimal deflection under the pressure of the water. By mini- mizing the deflection of the tank walls, the optical distortion when looking through the polycarbonate is limited. Figure 2.2: A top view photo of one of the seams between acrylic panels that form the walls of the water tank. Each gap between panels is first filled with flexible Tygon tubing to reduce the thickness of the silicone caulk layer, allowing for a fully-cured caulk layer. Because this tank is made of 25 separate acrylic panels, the seams between panels had to be sealed with silicone caulk to prevent leaks. This silicone caulk needs to be exposed to air in order to cure and therefore can only be applied in layers up to 3/8” thick. Because the gaps between panels were as deep as the thickness of the panels (3.18 cm), filling the entire gap with silicone would prevent the entire depth 17 of the seam from fully curing. Therefore, each seam between panels was first filled with flexible Tygon tubing that was forced into the gaps, as shown in Figure 2.2, before being covered with silicone sealant. Figure 2.3: A photo of the skimmer at one end of the tank. Its height is adjustable so that it can be set just below the desired water level in the tank. A skimmer is fixed at one end of the tank, see Figure 2.3, with adjustable height so that it can be used for a range of tank water levels. This skimmer is made of 1 cm-thick acrylic and is sealed with silicone sealant to prevent water from leaking into the box. This skimmer box is designed to be set just below the desired water level of the tank so that the top layer of water pours over the edge of the box, forming a small waterfall that contains any debris or surfactants that have accumulated on the water surface. The skimmer is connected to the plumbing system of the tank through a hose that connects a bulkhead fitting at the bottom of the skimmer to one of two bulkhead fittings through the floor of the water tank. Using this connection, the water that enters the skimmer can be either sent directly to the drain or pumped 18 through a diatomaceous earth filter and back into the opposite end of the tank to keep the water level in the tank constant. Additionally, water from the bulk of the tank (rather than from the skimmer) may be circulated through the filter. The tank is typically filtered continuously when experiments are not being performed to ensure that the water remains clean of particulate matter. This circulation pump is turned off while experiments are performed to minimize disturbances in the tank. Additionally, before experiments are performed, the skimmer is opened to the drain and water is continuously added to the tank so that any surfactant or debris floating on the water surface is skimmed to the drain. Two stainless steel feet are set into the concrete floor of the lab and pass through the bottom of the tank. These feet act as a mounting surface to support the above-mentioned moving belt device, herein called the Ship Boundary Layer (SBL) device, which is discussed below. These feet were made to be adjustable in the horizontal plane before being permanently fixed so that the distance between them could be set to match the distance between the feet of the SBL device. These feet were positioned and fixed to the floor of the lab prior to the installation of the acrylic walls of the tank. When the SBL device is in the tank, two large threaded rods pass through the frame of the device and thread into these feet, securing the SBL in place. 19 2.2.2 Ship Boundary Layer Device In this experiment, a temporally-evolving boundary layer, which simulates the boundary layer on the flat portion of the hull of a ship, is created using a hydraulically-powered belt device. A top view schematic of the SBL device can be seen in Figure 2.4. This device is made up of two hydraulically-driven rollers with vertically-oriented axes that are separated by 7.5 m and that are attached to a steel support frame. A stainless steel belt is driven in a loop around these two rollers and one of the two straight sections of belt between rollers forms a vertical, surface-piercing wall that translates along its length. measurement region water entrance fairing 2.4 m 13.4 m 6 m 7.5 m Figure 2.4: Top view schematic of the SBL device and the tank where experiments are performed. One of these rollers that makes up this device can be seen laying on its side in Figure 2.5, a photo taken before the device was assembled. Each roller is 0.46 m in diamater with circumferential grooves designed to prevent hydroplaning, similar to those on a rain tire. On either end of the roller, the axle of the roller passes 20 through a roller bearing, which allows the roller to rotate about its axis. On the top of each roller, two hydraulic motors are arranged in opposite orientations, with each motor driving a pulley attached to the axis of the roller, providing torque to drive the roller. Figure 2.5: A photo of one of the rollers during construction of the SBL. To assemble the SBL device, the two rollers are attached at either end of a steel support frame, as shown in Figure 2.6, a photo of the construction of the belt device. One roller is directly fixed to one end of the SBL support frame and the second roller is attached to the other end of the SBL support frame at the top and bottom via two horizontally oriented hydraulic pistons. These pistons push in and out to provide tension to the belt and to tilt the second roller relative to the 21 first in order to dynamically maintain the vertical position of the belt. When fully assembled, a 0.8 mm-thick stainless steel belt is looped around these two rollers. Figure 2.6: A photo of the belt construction in progress, showing major compo- nents of the device, including the rollers, steel support frame, hydraulic motors, and stainless steel belt. The majority of the SBL device is enclosed in a stainless steel sheet metal box, as shown in Figure 2.7, in order to keep the support frame and driving mechanism dry and prevent the belt from hydroplaning off of the rollers. A schematic of this setup can be seen in Figure 2.4. One of the two straight sections of belt between the rollers exits the dry box and is exposed to the water in the tank for approximately 6 m before re-entering the box. The outer surface of this section of the belt forms the surface of the simulated ship hull. The water level in the tank is set to a height such that the belt pierces the free surface. 22 Figure 2.7: A photo of one end of the SBL, showing the stainless steel sheet metal dry box. At the locations where the belt enters and exits the sheet metal box, phenolic laminate slats press against the belt to act as scrapers in order to keep the air con- tained inside the box from leaking out into the test section and vice versa. Because these slats and various other seals are imperfect, two sump pumps are positioned at the bottom of the dry box at each end so that any water leaking into the box from the tank is returned into the tank to keep the overall tank water level constant. Each of these sump pumps is capable of pumping 4000 gallons per hour and has a built in float switch so that they do not remain on when the box is completely dry. The outlet of these pumps is positioned behind the SBL dry box so that this outflow 23 does not effect the flow field of the experiment. An overflow drain was plumbed into the side of the SBL box at a height of approximately 5 cm from the bottom of the sheet metal dry box, as shown in Figure 2.8. In the event that the pumps fail, this drain allows excess water in the dry box to flow directly to the drain instead of remaining in the dry box. Figure 2.8: A photo of the overflow drain piping that allows the SBL dry box to drain in the event of a sump pump failure. Because of the thickness of the laminate scrapers and their mounting hardware, the shape of the dry box at the location where the belt enters the water forms a backwards-facing step. In order to reduce the effect of the flow separation caused by this step, a stainless steel entrance fairing is attached to the dry box, as can be seen 24 in Figure 2.9. This fairing is designed with a length of 38 cm and a maximum angle of 15 degrees in order to turn the backwards-facing step into a smooth, gradual transition. The opposite problem occurs where the belt re-enters the box, so a simpler fairing is placed at this location to reduce the splash caused by this high- pressure region in front of the forward-facing step. This second fairing is located significantly downstream of the measurement location and has no noticeable effect on the experiment. 15.0° 38.00 Figure 2.9: A top view schematic of the entrance fairing, designed to create a smooth transition over the backwards-facing step at the location where the belt exits the SBL dry box. A stationary, vertically-oriented, acrylic wall is fixed in the tank parallel to the belt surface at a distance of approximately 1.17 m. The wall is 4.87 m long and begins approximately 0.5 m downstream of the fairing exit. This wall is far enough away from the belt to where it appears to have no effect on the flow field or waves produced during a run. The area between this stationary wall and the tank wall opposite the belt surface provides a region for return flow from the end of the tank. This wall is also able to be moved closer to the belt for couette flow experiments not discussed in this dissertation. The SBL device is powered by a 60 horsepower hydraulic power unit (HPU), as shown in Figure 2.10(a). This HPU contains a 250 gallon reservoir of hydraulic fluid 25 with two pumps that can achieve hydraulic fluid pressures up to 3000 psi. When launching the belt, there is a large demand for high-pressure hydraulic fluid. In order to meet this demand, the pressure line of the hydraulic system contains three acumulators, two of which are shown in Figure 2.10(b). Each of these accumulators is precharged with nitrogen gas to 1400 psi and contains a piston that separates this gas from the hydraluic fluid. When entering high-pressure mode, the HPU pumps hydraulic fluid up to 3000 psi, which compresses the nitrogen and fills the the accumulators with high-pressure hydraulic fluid. This large volume of high- pressure hydraulic fluid is then available to drive the hydraulic motors during the launch of the belt. (a) (b) Figure 2.10: Two photos of the hydraulic system showing (a) the Hydraulic Power Unit (HPU) and (b) hydraulic fluid accumulators. The hydraulic system is designed to be used to power both the hydraulic system as well as a high-speed towing carriage in the same water tank. The hydraulic lines connected to the SBL device utilize quick disconnect fittings so that, when 26 transitioning to another experiment, these hoses can be disconnected and the SBL device can be removed from the tank. Because this device weighs approximately 12,000 lbs, this is accomplished using two 3-ton cranes attached to horizontal ceiling beams that run over the tank, as shown in Figure 2.11(a). In order to transition between experiments, the hydraulic hoses are disconnected, the threaded rods that secure the SBL device to the floor of the tank are removed, and two crane operators simultaneously lift the device in the air, as shown in Figure 2.11(b). These cranes are then able to translate along the ceiling beams to which they are attached. When passing over the side wall of the tank, the SBL device has approximately 1 inch of clearance, making this process difficult to coordinate. The device is then lowered onto the floor outside of the tank, where it remains when not in operation. (a) (b) Figure 2.11: Two photos of the crane system used to move the SBL device in and out of the tank. Image (a) shows one of the cranes that is used and (b) shows a photo of the process of lifting the device. When performing experiments, the belt launches from rest and is capable of accelerating at up to 3 times gravitational acceleration until reaching constant belt speeds of up to 15 m/s. In order to attain this rapid acceleration, the system needs 27 to utilize the power of all four hydraulic motors during launch. This four-motor operation can only be sustained for very short periods of time before the demand for high-pressure hydraulic fluid exceeds the ability of the HPU and accumulators to supply this fluid. The hydraulic control system contains an option to switch from four-motor operation to single-motor operation in the middle of a run, as a single motor is capable of sustaining the belt motion at constant speed. This switch from four motors to one creates a difficult challenge from a controls perspective, as that single motor must supply four times the torque it had been supplying, while also suddenly receiving all of the hydraulic fluid that had been going to all four motors. Quite often, before the single motor can reach a new equilibrium, the velocity of the belt experiences a large peak in velocity just after the switch. At the lower belt speeds used in this set of experiments, this peak was found to reach approximately 1.5 m/s above the desired belt speed. To avoid this large discontinuity, it was determined that using a single motor throughout the entire run was more desirable. Due to this, the acceleration that the system is able to achieve is lower and the launch time is increased, but the belt motion is more repeatable and overshoot is greatly reduced. The belt speeds used in these experiments are 3, 4, and 5 m/s. These speeds seem to span a range of air entrainment conditions, where little to no air is entrained at 3 m/s, some air is entrained at 4 m/s, and a large amount of air is entrained at 5 m/s. Further details of this air entrainment will be discussed later. For each experimental setup, the belt launch reamins the same for a given belt speed. By taking white-light video of the belt and cross-correlating successive images, it was 28 Time relative to trigger (s) 0 0.5 1 1.5 U (m /s ) 0 0.5 1 1.5 2 2.5 3 3.5 run 1 run 2 run 3 run 4 run 5 Time relative to trigger (s) -0.5 0 0.5 1 1.5 x (m ) 0 1 2 3 4 5 run 1 run 2 run 3 run 4 run 5 (a) (d) Time relative to trigger (s) 0 0.5 1 1.5 U (m /s ) 0 1 2 3 4 5 run 1 run 2 run 3 run 4 run 5 Time relative to trigger (s) -0.5 0 0.5 1 1.5 x (m ) 0 1 2 3 4 5 6 7 run 1 run 2 run 3 run 4 run 5 (b) (e) Time relative to trigger (s) 0 0.5 1 1.5 U (m /s ) 0 1 2 3 4 5 6 run 1 run 2 run 3 run 4 run 5 Time relative to trigger (s) -0.5 0 0.5 1 1.5 x (m ) 0 1 2 3 4 5 6 7 run 1 run 2 run 3 run 4 run 5 (c) (f) Figure 2.12: Measurements of belt trajectory for 5 repeated runs at each belt speed. The left column shows the belt velocity, U , versus time relative to the trigger signal that comes from the belt control system. The right column shows the amount of belt travel, x, versus time relative to trigger. Belt speeds of 3 m/s are shown in (a) and (d), 4 m/s are shown in (b) and (e), and 5 m/s are shown in (c) and (f). 29 determined that the belt launch is repeatable to within a 5 to 10 cm of belt travel over the length of comparable runs, as can be seen in Figure 2.12. Time relative to trigger (s) 0 0.5 1 1.5 2 U (m /s ) 0 1 2 3 4 5 6 U = 3 m/s U = 4 m/s U = 5 m/s X (m) 0 1 2 3 4 5 U (m /s ) 0 1 2 3 4 5 6 V = 3 m/s V = 4 m/s V = 5 m/s (a) (b) Figure 2.13: A comparison of the average belt launch trajectory for each belt speed. Plot (a) shows the belt velocity U versus time, while plot (b) shows the belt velocity U versus x, the amount of belt travel. Figure 2.13 shows a comparison between the belt launch trajectory for each of the three belt speeds. As you can see in (a), each launch accelerates at approximately the same rate, so at higher belt speeds, the belt travels farther before reaching full speed. This effect is captured in (b), which shows the belt speed U versus the length of belt travel, x. From these plots, it is found that during belt launches to 3, 4, and 5 m/s, the belt travels 0.85 m, 1.45 m, and 2.29 m before reaching 95% of the final belt speed. These measurements of belt travel during launch will be disucssed later in relation to the timing of other interesting events in these experiments. When performing experiments, the high-speed cameras are triggered with a circuit closure signal which is output by the hydraulic control system. In this way, the timing of different runs can be matched to compare results to each other and to the timing of the launch process. For a majority of the experiments discussed 30 herein, the length of belt travel studied is limited by the memory capacity of the high-speed cameras. This typically limits the length of each run to ten seconds. Therefore, at a belt speed of 3 m/s, the length of belt travel would be approximately 30 m, while at 5 m/s the belt would travel approximately 50 m. In order to make equivalent comparisons between belt speeds, most of the results presented below are only discussed for 30 m of belt travel. It is assumed that the length of belt that travels past a fixed point can be considered equivalent to the distance along the length of a ship hull. Therefore, the temporally-evolving boundary layer created along the belt is considered comparable to the spatially-evolving boundary layer along a flat-sided ship that makes no bow waves. In a perfect experimental setup using this configuration, the rollers would be so far apart that any fluid particle to pass through the measurement region during a run would have experienced the same effect of the shear induced by the belt for the entire experiment up to that point. In considering the effect of the finite length of the belt test section, this assumption holds true for the beginning part of each experiment in which the belt that passes through the measurement region has been in contact with water for the entire run to that point. Because the section of belt exposed to water at a given time is approximately 6 m, measurements are performed 4.75 m downstream of the end of the fairing and approximately 1 m upstream of the downstream fairing. In this way, the measurement region is located as far downstream as possible to maintain this assumption while still remaining outside of the region of influence of the downstream fairing. Therefore, for the first 4.75 m of each experiment, this assumption of fluid particles feeling the influence of the 31 belt for the entire experiment is valid. After this point, some fluid particles in the flow are entrained from the region outside the influence of the belt and must be accelerated into the boundary layer and it is unclear what influence this has on the flow field. 2.3 Experimental Techniques 2.3.1 Surface Profile Measurements Belt motion Free surface Light sheet from Argon CW laser Digital movie camera 300 pps Transparent tank bottom Fluorescing dye Water mixed with uorescein dye Side View Belt moving normal to page Light sheet for LIF Argon CW laser Stainless steel belt Water surface End View (a) (b) Figure 2.14: Schematic drawing showing the set up for the cinematic LIF measure- ments of the free surface shape from (a) a side view and (b) an end view perspective. To study the water surface motion, a cinematic Laser Induced Fluorescence (LIF) technique is utilized, as shown in Figure 2.14. In this technique, a continuous- wave Argon Ion laser beam is converted to a thin sheet using a system of spherical and cylindrical lenses. This sheet is projected vertically down onto the water surface with the plane of the light sheet oriented normal to the plane of the belt. This laser emits light primarily at wavelengths of 488 nm and 512 nm. During these 32 experiments, the water in the tank is mixed with fluorescein dye at a concentration of about 5 ppm and dye within the light sheet is excited by the laser light, fluorescing at a longer wavelength than the incident laser light. Two cameras view the intersection of the light sheet and the water surface from both upstream and downstream with viewing angles of approximately 20 degrees from horizontal. The cameras (Phantom V641 by Vision Research, Inc.) capture 4-Mpixel, 12-bit black-and-whtie images at frame rates of 1000 Hz. Utilizing 200 mm Nikon lenses, the images captured by the cameras are each 2560 x 600 pixels in size and have a resolution of 7.96 pixels/mm. This reduced vertical frame size allows the camera to store over twice as many images in memory as with the full frame size, allowing the camera to record at this high frame rate for the entire duration of a run. A long-wavelength-pass 550 nm optical filter is placed in front of each camera lens. These filters block out direct reflections of the laser light and transmit the light from the fluorescing dye, thus preventing specular reflections of the laser light from the water surface or belt surface from entering the camera lenses. The use of two cameras allows for more accurate surface measurement in the event that the intersection of the light sheet and the water surface is blocked by large deformations of the free surface between the plane of the light sheet and either camera, although this was unnecessary for the conditions presented in this research. A sample LIF image from a belt launch to 5 m/s is shown in Figure 2.15(a). The sharp boundary between the upper dark and lower bright region of each image is the intersection of the light sheet and the water surface. The lower portion of each image contains light that has been refracted when the light sheet passes 33 (a) (b) Figure 2.15: One sample frame from an LIF movie during a belt launch to 5 m/s. Image (a) shows a raw image from the movie camera and image (b) shows that same image with the result of the image processing plotted on top. through the free surface and again when the fluorescing light passes through the surface towards the camera, and is therefore not quantitative. The high-contrast boundary where the light sheet intersects the free surface is analyzed quantitatively to extract instantaneous surface profiles. Because these images were recorded from the downstream camera, the belt surface appears on the right side of each image and the belt moves out of the page. Because of the reflectivity of the belt, the intensity pattern to the right of the belt location is a reflection of the light pattern on the left due to the high reflectivity of the smooth surface of the belt. This line of symmetry gives a good indication of the position of the belt in each image. In order to extract these surface profiles, a novel image processing algorithm was developed and applied to each frame of the LIF movies. In the first step of this algorithm, this image is first broken into 64-pixel-wide windows. Each of these win- 34 (a) (b) (c) (d) (e) Figure 2.16: Sample images from intermediate steps of the processing alogirthm used to extract instantaneous surface profiles from LIF images 35 dows is then thresholded and recombined into one image, as shown in Figure 2.16(a), in order to account for horizontal variations in light sheet intensity. In this image, regions of connected white pixels above the free surface represent illuminated fea- tures that are behind the plane of the light sheet. Any such regions not connected to the surface are eliminated, as shown in image (b). The top-most white pixel in each column of this image is used as the initial guess for the profile extraction algorithm. This algorithm, discussed in more detail below, requires the input of a matrix that contains a ridge of high pixel intensities that can be traced to form the surface. Due to the many unique characteristics of the LIF images, a combination of a number of different image transformations is used to create this matrix. The first operation performed on the raw image is to determine its gradient in the horizontal and verti- cal directions. Due to the nature of the free surface shape, most surface features can be captured simply by using the gradient in the vertical direction. However, when surface features develop a large slope, as is often the case at the meniscus, where the water contacts the belt surface, this vertical gradient fails. Therefore, these individ- ual gradients in each direction are combined into the total gradient image shown in Figure 2.16(c), which is capable of capturing vertical or nearly vertical free surface shapes. While the total gradient image shows many small edges, particularly below the free surface, the intersection of the water free surface and the light sheet results in a large band of high contrast. In order to emphasize this band, a separate op- eration is performed to calculate the large-scale vertical intensity gradient. In this operation, for each pixel in the image, the average of the intensity of the 5 pixels above this location is subtracted from the average intensity of 5 pixels below, as 36 shown in (d). These images are then combined using a weighted average and the resulting intensity map is again broken into 64-pixel-wide windows to normalize the intensity of each part of the image. In order to extract the free surface shape from this intensity image, a version of the snake method, presented by Kaas et al. (1988), using dynamic programing, as discussed by Amini et al. (1990) is utilized. In this method, the free surface is defined as a snake, which is as an energy-minimizing spline that is guided by internal spline forces and external image forces, defined below as: Esnake = ∫ 1 0 Eint(v(s)) + Eimage(v(s))ds Eint = (α(s)|vs(s)|2 + β(s)|vss(s)|2)/2 The first equation defines the energy of the snake as the sum of Eint, the in- ternal energy of the snake, and Eimage, the energy of the image, which is defined by the intensity map discussed in the previous section. The second equation defines the internal spline energy, where v(s) is the position of the snake and α and β are ad- justable parameters used to weight the relative importance of these two terms. The first term in the internal spline energy equation is designed to penalize sharp spatial changes in the shape of the snake, allowing it to act like a membrane, while the second term penalizes the curvature of the snake, allowing it to act like a thin plate. Therefore, the snake is attracted to high intensity regions of the prescribed intensity map, Eimage, but penalized for discontinuous or non-smooth shapes. Through trial 37 and error, it was determined that values of α and β of 0.02 and 0, respectively, gave good performance of this algorithm. Because of the rapid changes in curvature of the free surface, use of a non-zero coefficient for the curvature term gave poor per- formance. Typically this algorithm is capable of achieving consistent results within ± 1 pixel of the surface location judged by eye. Because of the complex optical phenomena often present on the free surface, occasional errors were caused by the presence of bubbles or by refracted light patterns below the light sheet intersection line. These profiles were verified by eye and, if necessary, corrected using simpler thresholding or sobel edge detection algorithms. The horizontal location of the belt surface in each image was determined after calculating these surface profiles. Because of the reflectivity of the belt, the image processing algorithm was able to extract the surface shape even in the reflection of the belt. Therefore, located towards the end of each profile, the portion of the profile next to the belt surface was reflected as a mirror image of itself. To determine the belt location, a one-dimensional cross-correlation was performed between a section of the extracted profile and it’s mirror image. Because of this mirroring effect, the location of the highest correlation value was taken to be the location of the belt surface. 2.3.2 Fluid Velocity Measurements In order to fully examine the complex interaction between boundary layer tur- bulence and the free surface, it is desirable to study the turbulent velocity field 38 beneath the free surface. To accomplish this, two configurations of cinematic planar Particle Image Velocimetry (PIV) measurements are utilized. In this technique, a high-speed digital movie camera (Phantom V641, by Vision Research, Inc.) records image pairs of tracer particles, illuminated by a laser light sheet, as they are con- vected by the flow. These image pairs are recorded at a frame rate of 250 Hz, with a frame size of 2560 x 1600 pixels. By cross-correlating the intensity map of small interrogation windows between the frames of each image pair, a velocity vector field is derived. Using this technique, the first set of measurements is performed by form- ing a horizontal light sheet at varying distances from the free surface. In this way, it is possible to study the impact of the free surface boundary condition on streamwise and wall-normal velocity components. Belt Surface Water Free Surface Horizontal Light Sheet Transparent Tank Floor High-Speed Movie Camera Figure 2.17: A side view schematic of the horizontal PIV setup. 39 Figure 2.18: A photo of the horizontal PIV setup made necessary by space con- straints. A schematic of the setup used for the horizontal PIV experiment is shown in Figure 2.17. In this setup, a high-speed movie camera beneath the water tank views a horizontally oriented light sheet located just beneath the free surface. Because of limitations on space, the camera faces horizontally, with a large, flat mirror posi- tioned at a 45 degree angle in front of it, as can be seen in the photo in Figure 2.18. This mirror directs the camera’s view upward so that the illuminated horizontal plane can be viewed. Using this orientation, the camera can look normally through the floor of the tank, eliminating any optical aberrations that are caused by oblique viewing. 40 Figure 2.19: A photo of the Type 309-15 calibration target used to correct for optical distortions. Calibration of the setup is performed prior to experiments using LaVision’s DaVis imaging software and a Type 309-15 calibration target provided by LaVision, as shown in Figure 2.19. This target is 309 mm by 309 mm with 3 mm diameter white dots on a black background that are each spaced 15 mm apart. This calibration target is used to transform the coordinates of the raw camera images into real-world coordinates using a 3rd-order polynomial. This calibration procedure corrects any distortions in the optical setup. A horizontal light sheet is created using a high-repetition rate, double-pulsed Nd:YLF laser which is directed above the tank using a series of laser-line mirrors, and a set of two spherical convex lenses are used to prevent the beam from expanding 41 (a) (b) Figure 2.20: Two images of the periscope used to create a light sheet below the water surface showing (a) an overall schematic and (b) the configuration used in the horizontal light sheet experiments. as it travels. The effective focal length of a two-lens system is given by the equation: 1 f = 1 f1 + 1 f2 − d f1f2 In this equation, f is the effective focal length of the set of two lenses, f1 and f2 are the focal lengths of each individual lens, and d is the distance between these lenses. When d = f1 + f2, the terms on the right hand side cancel out and the effective focal length goes to infinity. Assuming the incoming beam is collimated, the beam that exits this two-lens system would also be collimated. Next, consider the case when d = f1 + f2 - , where  is some small distance. In this case, the effective focal 42 length of the system would become: f = f1f2  Again, if  is small, this would result in the lens system having a long focal length, potentially much longer than either of the individual lenses, and with a focal point tunable by changing the distance between lenses. Therefore, when setting up the optical system to create a light sheet, two spherical lenses with focal lengths of 10 and 15 cm were chosen simply because they were surplus equipment from a previous experiment and the approximate combined distance between them of 25 cm was convenient for the geometry of the optical system. First, these lenses were aligned with the path of the beam that was set to travel up to the ceiling and above the tank and then the distance between them was tuned so that the beam focused to a point at the approximate distance from the laser that the light sheet will be located. This is used as a sort of coarse focus for the light sheet. From above the tank, this thin beam is directed downward and passed through a piece of equipment similar to a periscope, as shown in Figure 2.20. This periscope is made up of a tube with focusing optics and a pill-shaped head that contains a set of laser-line mirrors and a cylindrical lens. This periscope assembly is air-tight so that the pill-shaped head can be submerged in water while the optics remain dry. Using this periscope, a laser beam is introduced from above the water surface into the open top end of the tube and a light sheet is projected from the device beneath the water surface. Because of this, the water free surface does not refract the beam 43 as it passes through the surface. In the horizontal PIV setup, the periscope head is positioned approximately 80 cm from the belt surface and the light sheet is projected directly towards the belt surface. In doing this, the distance of particle-filled water through which the beam passes is reduced when compared to introducing the light sheet from the far side of the tank. This is preferrable because the scattering of light by particles outside of the measurement region reduces the intensity of light that would eventually reach the region of interest. This periscope also contains a set of focusing optics to allow for fine tuning of the light sheet thickness in the measurement region. Stainless steel belt Moving normal to page Laser light sheet High-speed movie camera Water surface (a) (b) Figure 2.21: Two images of the vertical PIV setup showing (a) a side-view schematic and (b) a photo of the actual setup through the wall of the tank, showing the orientation of the light sheet. The second set of measurements is performed using a vertical light sheet par- allel to the belt surface. This configuration provides a more detailed picture of the variation of the streamwise and vertical velocity components as the free surface is ap- proached. A schematic of this second PIV configuration is shown in Figure 2.21(a). In this setup, a vertical light sheet is projected parallel to the belt surface using the same periscope device discussed in the previous section. This periscope is positioned 44 downstream of the measurement region with the pill-shaped head reoriented so that the light sheet is projected upstream and vertically up at an angle, as can be seen in the photo in Figure 2.21(b). By positioning the head of the periscope downstream and below the edge of the belt, its effect on the upstream flow is minimized. A high- speed movie camera is mounted outside of the sidewall of the tank and views the vertically-oriented light sheet straight on, without the plate mirror that is needed in the horizontal PIV configuration. A similar calibration procedure as previously mentioned is used to correct for optical abberations that are introduced by looking through the tank wall. When passing the laser beam into the top of the periscope tube, the periscope is initially moved to its highest point to ensure that the beam is centered with respect to the tube. Next, the tube is lowered to its eventual location and alignment with the periscope tube is performed. Using this procedure, the beam is aligned with the periscope tube so that there are no reflections from the reflective interior of the metal tube. Positioning of the light sheet for each set of experiments is performed using a ruler with a resolution of 1 mm. For the horizontal sheet, the water level is first set by positioning the skimmer at the desired water level in the tank and a small, continuous stream of water is added to the tank so that the skimmer continuously overflows to the drain. This keeps the tank water level constant. Next, the light sheet is focused using the initial position of the optics and periscope so that a thin light sheet is formed near the belt surface. A clear ruler is then positioned in front of the light sheet just at the location where the light sheet exits the pill-shaped head of the periscope in order to give an initial measurement of the distance of the light 45 sheet from the free surface. Then, a measurement is performed near the belt surface, far from the periscope. This should be the same as the previous measurement in order to ensure that the light sheet is perfectly horizontal. The angle of the periscope head is adjusted and these measurements are repeated iteratively until the distance of the light sheet from the free surface remains constant over its entire path toward the belt. Next, the vertical position of the light sheet is adjusted by translating the periscope tube vertically until the desired light sheet position with respect to the free surface is reached. When aligning the vertical light sheet, a similar procedure is followed, with ruler measurements performed with respect to the belt surface rather than the free surface. It should be noted that the focusing optics in the periscope tube can cause some eccentric movement of the light sheet when adjusted, changing the distance of the light sheet from the free surface or belt surface, so this is performed prior to any light sheet alignment. The flow is seeded with fluorescent tracer particles which are selected in or- der to provide a sufficiently small settling velocity and sufficiently high frequency response to faithfully follow the flow. As with the LIF experiments discussed above, an optical long-wavelength-pass filter is placed in front of the camera in order to block direct scattering of laser light by bubbles, the free surface, and the belt surface and allow only fluorescent light from the particles to pass through to the camera sen- sor. These particles are manufactured in the laboratory using the process described by Pedocchi et al. (2008). In this process, a mixture of Rhodamine WT in epoxy is created by first mixing 200 mL of MAS epoxy resin with 100 mL of slow-hardening MAS epoxy hardener in a shallow pot or cake pan. Due to the exothermic curing 46 of the epoxy, these epoxy disks should remain 1 to 1.5 cm thick to ensure full cure and this amount of epoxy is typically suitable for a 9-inch diameter pan. These two components of epoxy should be mixed thoroughly for approximately 90 seconds to ensure that the a proper homogeneous mixture of resin to hardener is main- tained. After this mixing, 5 mL of Keyacid Rhodamine WT Liquid (703-010-27) from Dyes.com is added to the epoxy mixture and is further mixed for 90 seconds, or until the mixture appears to become completely homogeneous. This epoxy mixture should then be allowed to cure for 24 hours to ensure that the mixture has fully cured to become a hardened disk. This disk is then sanded iwith a belt sander to produce a wide range of polydispersed particles. A sanding belt with 220 grit is typically suitable to obtain a range of particles in desirable size ranges. This sand- ing is typically performed in a ventilation hood with the operator of the belt sander completely protected from inhaling these particles by using a respirator with N100 or P100 cartridges, which are rated to filter over 99.97% of particles. The operator should also wear a disposable painter suit, rubber gloves that cover the wrists, and goggles in order to prevent particles from getting onto their skin or clothes. This sanding should be performed within a box in the vent hood in order to contain the majority of particles that are created during this process. These particles can then be easily collected using a dedicated vacuum with a 3-layer pleated filter, rated to capture 99.5% of particles that are 0.5 microns or larger. By sifting these particles through sieves of different mesh sizes, specific size ranges of particles can be created. The vast majority of particles created by this process lie in the size ranges of 25 µm to 53 µm and 53 µm to 106 µm. 47 In order to be considered suitable as tracers, these particles must have a set- tling velocity that is negligible in comparison to flow velocities of interest. These particles also should have a small enough settling velocity that the particles do not settle out of the region of interest prior to or during the experiment. Using Stokes’ Law as a simplified analysis of the particle settling velocity, it is found that the smaller of these size ranges has a settling velocity between 0.044 mm/s and 0.20 mm/s, while the larger size range contains settling velocities up to 0.80 mm/s. An additional consideration is that the tracer particles must have a sufficient fre- quency response in order to represent the frequencies of fluid motion in the flow field. Following the work of Mei (1996), an analysis of particle frequency response can be performed using the following relation: vˆp(ω)e iωt = Hp(ω) · u(ω)eiωt In this equation, u(ω)eiωt is the equation for a sinusiodal oscillation of a fluid particle, vˆp(ω)e iωt is the particle response function, and Hp(ω) represents the transfer function between them. Via the analysis in Mei (1996), an equation for the square of this transfer function, called the energy transfer function can be devised using asymptotic analysis, taking the following form: |Hp(ω)|2 = |Hp()|2 = (1 + ) 2 + (+ 2 3 2)2 (1 + )2 + [+ 2 3 2 + 4 9 (ρ− 1)2]2 48 In this equation, ρ is the density ratio of the particle to the fluid and  is a Stokes number, defined as:  = √ ωa2/2ν In this equation, ω is the frequency of fluid oscillation in the flow, a is the particle radius, and ν is the kinematic viscocity of the fluid. By substituting the particle sizes in the ranges previously discussed, as well as the viscocity of water and the density ratio of the particles of 1.13, according to information from the epoxy company, this energy transfer function can be calculated for each particle size over a large range of frequencies, as shown in Figure 2.22. Frequency (Hz) 100 101 102 103 104 105 106 107 108 |H p |2 0.84 0.86 0.88 0.9 0.92 0.94 0.96 0.98 1 25 µm particles 53 µm particles 106 µm particles Figure 2.22: Particle frequency response for a range of particle diameters. It can be seen from this analysis that any of the particles in these size ranges retain over 84 percent of the energy of the forcing fluid flow at all frequencies. Additionally, if you were to define an acceptable energy level for this energy transfer function, say 90%, and evaluate each particle size at that energy level, the largest 49 particles would reach approximately 1.5 kHz, the mid-sized particles would reach approximately 6 kHz, and the smallest particles would reach approximately 30 kHz. Therefore, the smaller size range provides a considerably higher frequency range over which the majority of energy in the flow is well-represented by the particle response. Finally, to be suitable for use as PIV tracers, images of these particles must be a sufficient size and brightness and with a high enough seeding density in order to perform proper cross correlation between image pairs. By performing preliminary PIV tests with both size ranges of particles, it was determined that the particles with diameters between 25 µm and 53 µm were the most appropriate for these experiments. While the larger particles appeared brighter, the smaller particles provided a better settling velocity and higher frequency response while maintaining an acceptable level of brightness. To introduce these particles into the region of interest of the flow, a local seeding system is used rather than filling the entire tank with a homogeneous mixture of particles. The concentration of particles that is required in the flow for properly- seeded PIV images is high enough that a homogeneous mixture would cause the water to appear cloudy, greatly reducing visibility as the camera has to look through a large distance of water before seeing the region illuminated by the laser light sheet. In addition, the amount of particles required to properly seed the approximately 8000 gallons of water in the tank would be prohibitive. Additionally, these particles would settle out of the flow between runs and remixing the entire tank would be difficult. 50 (a) (b) (c) (d) Figure 2.23: Four photos of the particle seeding system. Image (a) shows the basin in which high-concentration particle solution is mixed. Image (b) shows the pump that draws from this basin and pumps to (c) the particle seeder. Image (d) is a photo from beneath the water of an initial test run of the particle seeder using fluorescent dye rather than particles. 51 To create a local seeding system, a separate basin of approximately 85 gallons is filled with a 50 mg/L mixture of particles in water, as shown in figure 2.23(a). This mixture is then pumped from a bulkhead at the bottom of the basin into the flow upstream of the measurement region using a 1/2 horsepower pump, shown in image (b). To introduce these particles into the correct region of the flow, a length of cylindrical PVC pipe with a long, narrow slot was created to evenly discharge the particles along the length of the pipe, shown in images (c) and (d). The pump feeds directly from the seeding basin into this pipe, which is positioned either horizontally or vertically, depending on the orientation of the measurement plane in the PIV setup. This pipe is positioned approximately 5 meters upstream and displaced significantly from the measurement plane to avoid any disturbances in the wake of the pipe from influencing the flow. Before performing experiments, the pump is started and the belt is turned on at low speed (typically 1 m/s) for 30 seconds to introduce some particles into the measurement region and then the belt is turned off for 60 seconds to allow the water to calm down before an experimental run begins. The particle seeder pumps out particles with a velocity of approximately 10 cm/s and does not appear to have a significant effect on the flow. Planar velocity fields consisting of streamwise and wall-normal velocity compo- nents were measured in horizontal planes oriented perpendicular to the belt surface and aligned with the streamwise flow direction. In these experiments, two separate measurement planes were considered at distances below the undisturbed free sur- face, D, of 14 cm and 2.5 cm. The location of the first of these planes, positioned at a depth of 14 cm, was chosen to be far enough below the free surface to not be 52 influenced by any free surface waves or by the effect of the free surface boundary condition. The second measurement location, positioned at a depth of 2.5 cm, was chosen in order to be below the majority of bubbles and free surface depressions that would disrupt the measurement technique, while still capturing effects of the free surface on the boundary layer flow. These measurement locations allow for comparison of the flow fields in order to study the modified physics in the vicinity of the free surface. Separate experiments were performed in which planar velocity fields consist- ing of streamwise and vertical velocity components were measured in vertical planes parallel to the belt surface. In these experiments, two separate planes were consid- ered at distances from the belt surface, y, of 1.9 cm and 3.8 cm. These measurement locations were chosen in order to study the primary region in which wave breaking appears to occur. In all experiments, images were recorded with a high-speed digital movie cam- era (Phantom V641, by Vision Research, Inc.) using a Canon 180 mm lens with remote focus and aperature control, allowing for adjustments to be made while the camera was positioned under the tank. Image pairs were recorded at a frame rate of 250 Hz, with a frame size of 2560 x 1600 pixels and a long-wavelength-pass 550 nm optical filter was placed in front of the lens to filter out any specular reflections from the belt surface or bubbles in the flow. The images collected for the 14-cm-deep PIV experiments have a resolution of 23.4 pixels/mm, with a corresponding field of view of 10.9 cm x 6.8 cm. In moving the second sheet closer to the free surface and farther from the camera for the D = 2.5 cm case, the field of view was increased to 53 (a) (b) Figure 2.24: Two images of the results from the PIV experiments with a horizontal sheet far from the free surface during a launch to 4 m/s at x = 4.2 m. Image (a) shows a sample raw PIV image and image (b) shows the processed vector field with the vector spacing reduced by 1/2 and the vector length increased by 10 times for clarity. The background is colored by vector length with a scale running from 0 to 2.5 m/s. 54 (a) (b) Figure 2.25: Two images of the results from the PIV experiments with a vertical sheet close to the belt surface during a launch to 3 m/s at x = 5.9 m. Image (a) shows a sample raw PIV image and image (b) shows the processed vector field with the vector spacing reduced by 1/2 and the vector length increased by 10 times for clarity. The background is colored by vector length with a scale running from 0 to 1.0 m/s. 55 12.1 cm x 7.6 cm, with a corresponding pixel resolution of 21.1 pixels/mm. In the vertical-plane PIV experiments, the images of the light sheet closer to the belt have a resolution of 14.6 pixels/mm and a field of view of 17.8 cm x 11.1 cm, while the y = 3.8 cm condition yields images with a resolution of 14.8 pixels/mm and a field of view of 17.2 cm x 10.8 cm. These PIV images were collected using a double-frame mode, resulting in image pairs, which were processed using LaVision’s DaVis imag- ing software. To process these images, a multi-pass PIV processing algorithm with decreasing window size was used with a final interrogation window size of 32 x 32 pixels, using a 2 x 1 elliptical weighting function and a 50% overlap. The result- ing vector fields were post-processed using median filtering and bad vectors were replaced with values interpolated from surrounding vectors. Some sample images from both the horizontal and vertical PIV setups can be seen in Figures 2.24 and 2.25. Each figure shows a typical raw PIV image on the left and its corresponding velocity vector field on the right using this processing algorithm. This PIV process- ing resulted in a vector spacing of 0.68 mm and 0.76 mm for the horizontal PIV measurements at D = 14 cm and 2.5 cm, respectively. By calculating the viscous sublayer thickness using the Schultz-Grunow formula from Schlichting (1979), as shown in Figure 2.26, this vector spacing results in the first vector laying well out- side of the viscous sublayer at around 60-120 viscous wall units. While it would be preferrable to resolve this region, LIF movies show that air entrainment tends to occur at larger scales. Vector spacings of 1.09 mm and 1.08 mm were attained for the vertical-plane PIV measurements at y = 1.9 cm and 3.8 cm, respectively. 56 x (m) 0 10 20 30 40 50 60 δ ν (μ m ) 0 10 20 30 40 50 60 U = 3 m/s U = 4 m/s U = 5 m/s Figure 2.26: A plot showing the viscous sublayer thickness based on the Schultz- Grunow formula for skin friction, taken from Schlichting (1979) Because there is water on both sides of the portion of belt that is exposed to water, the belt is able to move in the wall-normal (y) direction under the influence of the water. When designing the SBL device, it was assumed that the large surface area of the belt would result in a large added mass, preventing it from moving significantly. When performing PIV measurements, the belt was observed to move in the y direction throughout each run. Therefore, the belt was tracked from frame to frame using the raw PIV images in each movie. In this method, 25 pixel by 100 pixel interrogation windows in the vicinity of the belt surface were cross correlated with their mirror images in the y direction. This utilized the reflectivity of the belt surface to determine the belt position in each frame and this was used as the reference from which the PIV data was measured. The results of these measurements are shown in Figure 2.27, which compares the belt position for 20 runs at each speed. It can be seen that the belt moves outward by approximately 1 cm throughout each run with 57 x (m) 0 5 10 15 20 25 30 y (m m ) 0 2 4 6 8 10 12 (a) x (m) 0 5 10 15 20 25 30 y (m m ) 0 2 4 6 8 10 12 (b) x (m) 0 5 10 15 20 25 30 y (m m ) 0 2 4 6 8 10 12 (c) Figure 2.27: Measurements of the y-direction belt position throughout each run with each plot showing the results of 20 repeated runs at (a) 3 m/s, (b) 4 m/s, and (c) 5 m/s. 58 x (m) 0 5 10 15 20 25 30 v r m s U ×10-3 0 1 2 3 4 5 U = 3 m/s U = 4 m/s U = 5 m/s Figure 2.28: Measurements of the RMS velocity fluctuations of the belt, normal- ized by belt speed throughout each run. a fairly repeatable motion. Using central difference to calculate the y-direction belt velocity from this position data, the RMS velocity fluctuation of the belt can be determined for each belt speed, as shown in Figure 2.28. It is unclear whether the high-frequency content of these measurements is caused by actual belt vibrations or by measurement error in using this technique. Due to the large surface area of the belt, high-frequency vibrations seem to be a less likely cause of these fluctuations. In either case, it is clear that these wall-normal belt fluctuations are very small compared to the belt speed. As will be shown later, these values are about a factor of 20 smaller than the wall-normal velocity fluctuations of the fluid near the wall. 2.3.3 Air Entrainment Measurements In addition to the above-described surface profile and velocity field measure- ments, observations of the onset of air entrainment are also desirable. While some 59 Figure 2.29: A photo of the sub-surface planar bubble measurement system, illu- minated using two 650 watt flood lamps. indications of surface breaking that might typically be associated with air entrain- ment can be seen from surface profile measurements, this above-water perspective cannot be used to definitively say whether a given event leads to air entrainment. Therefore, to record these events, cinematic white light movies are recorded beneath the water free surface, shown in Figure 2.29. The camera view used in this exper- imental setup is similar to that described above in the vertical-sheet PIV section. In these experiments, a high-speed camera is positioned to look directly at the belt surface from outside the side wall of the tank. This region is illuminated using two 650 watt flood lamps. Because the belt surface is partially reflective, this ar- rangement results in any surface depressions or bubbles appearing as dark features on a bright background. The resulting movies can be examined qualitatively to determine the time at which air entrainment occurs for a given belt speed. 60 Belt Surface Flood light High-speed movie camera Belt Motion Water WaterWater Acrylic tank wall (3.18 cm thick) 1. 83 m Glass walls Figure 2.30: A schematic of the underwater stereo bubble measurement system A second configuration for this experiment is shown in Figure 2.30. This configuration uses a cinematic stereo camera system in which two cameras view the flow just under the water free surface adjacent to the belt. The cameras are set up along the side wall of the tank that is parallel to the belt with horizontal viewing directions and lines of sight rotated ±30◦ from a vertical plane that is normal to the belt surface. A photo flood light with a translucent screen in front of it is placed next to each camera. Since the belt surface is moderately reflective, each light provides the illumination for the opposite camera. The cameras (lights) view (illuminate) the scene through water-filled prism boxes that are set to create a straight line of sight that is perpendicular to the side wall of the prism tank. Each camera is also equipped with a Scheimpflug lens mount, which inclines the sensor plane relative to the lens plane so that each camera may focus on a plane parallel to the belt despite 61 viewing from a displaced angle. Using two cameras at offset angles allows for a more accurate characterization of the non-axisymmetric free surface features found during entrainment events. This gives a better insight into the physical processes of air entrainment that are occurring. 62 Chapter 3: Results and Discussion 3.1 Surface Profile Measurements Surface profile measurements were performed at belt speeds, U , of 3, 4, and 5 m/s. Through initial trials, it was determined that a frame rate of 1000 fps was necessary to provide a sufficient temporal resolution so that surface features could be identified and tracked smoothly in successive frames. LIF images of the water free surface next to the belt for an experimental run with U = 5 m/s are shown in Figures 3.1 and 3.2. Though the camera recorded at a frame rate of 1,000 frames per second, the five images in the figure are spaced out equally by distance of belt travel, with the first image (a) in each figure taken at 0.0 s, the time when the belt first starts to move. The instantaneous belt speed from the beginning of belt motion through the acceleration portion until the belt reaches constant speed has been measured separately and is used to correlate the time of each frame to the belt travel distance. Here and in the following, rather than refer to images and data by the time after the belt started moving, we refer to them by the distance, x, from the leading edge of an equivalent flat plate, which is also the distance that the belt has traveled. Thus, the images in Figure 3.1, showing the beginning portion of the run, were captured at (a) 0 s, (b) 0.5 s, (c) 0.76 s, (d) 0.96 s, (e) 1.16 s, and (f) 1.35 s, 63 Belt Surface Contact point (a) (b) (c) (d) (e) (f) Figure 3.1: A sequence of five images from a high speed movie of the free surface during a belt launch to 5 m/s. These images are taken at equivalent belt lengths of (a) 0 m (b) 1 m (c) 2 m (d) 3 m (e) 4 m and (f) 5 m from the bow of the ship. The high reflectivity of the stainless steel belt makes it appear as a symmetry plane on the left side of the images. The horizontal field of view for these images is approximately 31 cm. 64 Belt Surface Contact point (a) (b) (c) (d) (e) Figure 3.2: A sequence of five images from a high speed movie of the free surface during a belt launch to 5 m/s. These images are taken at equivalent belt lengths of (a) 0 m (b) 5 m (c) 10 m (d) 15 m and (e) 20 m from the bow of the ship. The high reflectivity of the stainless steel belt makes it appear as a symmetry plane on the left side of the images. The horizontal field of view for these images is approximately 31 cm. 65 corresponding to x = 0.0, 1.0, 2.0, 3.0, 4.0, and 5.0 m, respectively. Similarly, images from Figure 3.2 depict the later portion of the run, with images (a), (b), (c), (d) and (e) captured at 0 s, 1.35 s, 2.35 s, 3.35 s, and 4.35 s, respectively, corresponding to x = 0.0, 5.0, 10.0, 15.0 and 20.0 m. As discussed in the previous section, the plane of the vertical light sheet is oriented normal to the belt surface and the cameras look parallel to the belt surface and down at the water surface at a small angle from horizontal. The images in Figure 3.1 are from the downstream camera, and these images have been flipped horizontally for convenie so that the belt is near the left side of each image and is moving out of the page, which will match the coordinate system of later plots. The position of the belt is marked on the left side of image (a) and the intensity pattern to the left of this location is a reflection of the light pattern on the right due to the high reflectivity of the smooth surface of the belt. This line of symmetry gives a good indication of the position of the belt in each image. The sharp boundary between the upper dark and lower bright region of each image is the intersection of the light sheet and the water surface. The upper regions of the later images contain light scattered from roughness features on the water surface behind the light sheet. These roughness features include bubbles that appear to be floating on the water surface and moving primarily in the direction of the belt motion. An example series of water surface profiles from a run with U = 3.0 m/s over a range of x from 21 m to 21.3 m is shown in Figure 3.3. The horizontal axis in the plot is horizontal distance, y, from the belt surface and the vertical axis is water surface height above the mean water level. The profiles are equally spaced in x by 1.2 cm and each successive profile is plotted 1.5 mm above the previous 66 0 50 100 150 200 250 300 0 50 100 150 y (mm) z (m m ) Figure 3.3: Sequence of profiles of the water surface during belt launch to 3 m/s starting from x = 21 m. The time between profiles is 4 ms and each profile is shifted up 2 mm from the previous profile to reduce overlap and show propagation of surface features throughout time. The belt is positioned at the left edge of the image (y = 0). profile so that overlap is reduced and the evolution of surface features can be seen. Surface features like ripple crests can be tracked over a number of successive profiles and the slopes of imaginary lines connecting these features are an indication of their horizontal speed away from the belt surface. It is clear that close to the belt surface, the ripple features last for only a few profiles and their paths do not contain a single dominant propagation speed. This region appears to be more chaotic than the region to the right. In this outer region, surface features remain visible over many frames giving support to the idea that they are freely propagating ripples, and their velocity remains fairly constant. The red line in the figure, which was drawn by 67 eye to approximate that slope of the imaginary lines connection the ripple crests in the outer region, corresponds to a velocity of about 34 cm/s. It should be kept in mind that this is only the y-component of the ripple phase speed. Information about the component of the phase speed in the x direction is unavailable from these experiments, but could be obtained from experiments with the LIF system oriented to record water surface profiles in planes parallel to the belt surface. 3.1.1 Wave Propagation In order to study the differences in the propagation of free surface ripples in a more quantitative way, we can perform cross correlation between profiles both close to the belt (0.6 cm > y > 6.9 cm) and far away (21.4 cm > y > 27.6 cm), as shown in Figures 3.4, 3.5, 3.6, and 3.7. The cross correlation function R is defined as: R(∆y,∆t) = ∑ (Z1(y, t)− Z¯1)(Z2(y + ∆y, t+ ∆t)− Z¯2)√∑ (Z1(y, t)− Z¯1)2 √∑ (Z2(y + ∆y, t+ ∆t)− Z¯2)2 In this case, y is the horizontal, wall-normal coordinate, t is the time at which each profile is measured, ∆y and ∆t are defined as the spatial and temporal shift between correlated profiles, and Z1 and Z2 are the two profiles which are being correlated. This cross correlation is performed for all profiles over lengths of belt travel from x = 0 to 30 m, and the resulting correlation maps are summed into three groups from x = 0 to 5 m, 5.85 m to 17.85 m, and 17.85 m to 29.85 m. This procedure is repeated for belt speeds of 3, 4, and 5 m/s. 68 (a) (b) (c) (d) (e) (f) Figure 3.4: Cross correlation maps for profiles from x = 0 to 5 m. The left column shows cross correlation maps averaged over a region close to the belt and those in the right column are averaged over a region far from the belt. The three rows from top to bottom contain plots from U = 3, 4, and 5 m/s, respectively. The black line in each plot indicates the ∆x location of the maximum correlation for each ∆t slice. 69 The first set of correlation maps in Figure 3.4 contains plots from early in each run. The left column contains cross correlation maps from close to the belt and the right column contains cross correlation maps far from the belt. The three rows from top to bottom contain results from belt speeds of 3, 4, and 5 m/s, respectively. It is clear that at this early stage, no waves have reached the region far from the belt, so only a flat surface is being correlated. Close to the belt, a well-defined ridge is beginning to develop, with correlations lasting for fairly long ∆t. Still, the ridge in each correlation map is fairly broad, indicating that free surface features are moving with a fairly wide range of speeds. It is clear when comparing the results from the mid- and late- run cross corre- lation maps that the correlation drops off very quickly close to the belt and that it is difficult to pick out a single dominant velocity component. Far from the belt, how- ever, the correlation lasts for long ∆t and has a clearly defined ridge, indicating a single dominant velocity, particularly in the range of x from 5.85 m to 17.85 m. This behavior could indicate that ripples close to the belt are heavily influenced by tur- bulent velocity fluctuations beneath the surface and change in form rapidly. These nonpropagating surface fluctuations are most likely the generators for the freely propagating waves in the outer region, which retain their form for longer periods of time. Alternatively, this could simply be an indication that these surface features are primarily convected in the x direction and have a short enough wavelength in that direction to become quickly uncorrelated. In addition, the correlation maps far from the belt appear to weaken later in each run, which could be an indication that 70 (a) (b) (c) (d) Figure 3.5: Cross correlation maps for profiles from a launch to 3 m/s. Images (a) and (c) show cross correlation maps averaged over a region close to the belt and (b) and (d) are averaged over a region far from the belt. Images (a) and (b) are averaged over x positions from 5.85 m to 17.85 m and images (c) and (d) are averaged over x positions from 17.85 m to 29.85 m. The black line in each plot indicates the ∆x location of the maximum correlation for each ∆t slice. 71 (a) (b) (c) (d) Figure 3.6: Cross correlation maps for profiles from a launch to 4 m/s. Images (a) and (c) show cross correlation maps averaged over a region close to the belt and (b) and (d) are averaged over a region far from the belt. Images (a) and (b) are averaged over x positions from 5.85 m to 17.85 m and images (c) and (d) are averaged over x positions from 17.85 m to 29.85 m. The black line in each plot indicates the ∆x location of the maximum correlation for each ∆t slice. 72 (a) (b) (c) (d) Figure 3.7: Cross correlation maps for profiles from a launch to 5 m/s. Images (a) and (c) show cross correlation maps averaged over a region close to the belt and (b) and (d) are averaged over a region far from the belt. Images (a) and (b) are averaged over x positions from 5.85 m to 17.85 m and images (c) and (d) are averaged over x positions from 17.85 m to 29.85 m. The black line in each plot indicates the ∆x location of the maximum correlation for each ∆t slice. 73 the growing boundary layer has a larger effect on modifying the free surface features farther from the belt. U (m/s) 2.5 3 3.5 4 4.5 5 5.5 c p y (c m / s) 28 30 32 34 36 Figure 3.8: Wave propagation speed far from the belt averaged over 5.85 m < x <17.85 m versus belt speed. Looking at the correlation maps far from the belt during the middle portion of each run, as can be seen from Figures 3.5(b), 3.6(b), and 3.7(b), the slope of the ridge in the cross correlation maps appears to increase as the belt speed increases. The black line in each image shows the ∆x location of the maximum correlation for each ∆t slice. Using this, the slope of the ridge can be tracked, showing the propagation speed of surface features, shown in Figure 3.8. This propagation speed in the wall-normal direction increases with belt speed. Further information about the propagation in the x direction is needed to more fully explore the significance of this behavior. 74 (a) (d) (b) (e) (c) (f) Figure 3.9: A sequence of six images from a high speed movie of the free surface during a belt launch to 3 m/s. These images are cropped to have a horizontal field of view of approximately 7 cm. Each image from (a) to (f) is spaced by 20 ms, corresponding to 6 cm of belt travel between images.The first image occurs at x = 2.45 m 75 3.1.2 Free Surface Bursting Another interesting phenomenon that occurs during the launch of the belt that can be seen from these LIF images is that the surface appears to burst with activity shortly after reaching constant velocity. Initially during launch, the belt appears to have no noticeable effect on the free surface and appears to simply slip past without creating waves. Shortly after reaching constant belt speed, the water free surface near the belt appears to burst very suddenly, as you can see from the sequence of images in Figure 3.9. Before this bursting occurs, the surface looks very similar to image (a) for the entire acceleration period. After this point, free surface fluctuations are continually generated close to the belt and this generation region grows in time. This appears to be an indication of transition to turbulence, but this transition point could be a moderated turbulence signal that has been filtered by surface tension and gravity, which would act to smooth out free surface motions. U (m/s) 2.5 3 3.5 4 4.5 5 5.5 x on se t (m ) 0 0.5 1 1.5 2 2.5 U (m/s) 2.5 3 3.5 4 4.5 5 5.5 R e x ×106 0 2 4 6 8 10 (a) (b) Figure 3.10: Two plots showing (a) the average x location of the onset of free surface bursting versus belt speed and (b) the associated Reynolds number based on x. 76 Therefore, the turbulence intensities may have to reach a critical threshhold in order to mobilize the air-water interface. This will be further discussed in the following sections. The x location of the onset of bursting in each run can be easily seen from the LIF movies and additional repeated runs were performed to increase the statistical significance of this measurement. Figure 3.10(a) shows that the average onset location decreases with increasing belt speed, averaged over 20 runs at each belt speed. The average Reynolds number based on x location of onset, shown in (b) appears fairly consistent, but more belt speeds would be required to see if this trend continues to hold true. Other potential scaling parameters for this bursting onset will be discussed in more detail below. 3.1.3 Surface Fluctuations 2.5 3 3.5 4 4.5 5 5.5 0 0.5 1 1.5 2 2.5 3 3.5 4 U (m/s) H r m s (m m ) Figure 3.11: RMS surface height fluctuation averaged over y and x versus belt speed for an equivalent ship length of 30 m. 77 In order to characterize these free surface motions, the instantaneous surface profiles extracted from the LIF movies are analyzed to determine RMS surface height fluctuations, Hrms. A mean profile is first determined by averaging the time series of surface height data for each wall-normal position throughout a given run. This average surface height profile is then subtracted from each individual profile and RMS fluctuations about the mean are calculated from the resulting data. This data is averaged over five repeated runs for belt speeds of 3 and 4 m/s, while ten repeated runs are utilized for the belt speed of 5 m/s due to its increased surface fluctuation and wave breaking behavior. The RMS surface height is first averaged across both wall-normal position and equivalent ship length to quantify the overall mean fluctuation amplitude for each belt velocity, as shown in Figure 3.11. Because these profiles are averaged across every surface height measured in both x and y for each speed, the RMS values are the most statistically significant of the quantities presented here. The RMS height fluctuations increase monotonically with increasing belt speed, but a larger range of belt speeds would be required to gain an insight into whether this trend is linear or if it follows some higher order behavior. More detail about the RMS height fluctuations can be found by examining their distribution in y. By averaging the RMS height fluctuation over the time series of points for each wall-normal position, the average RMS surface height fluctuation versus distance from the belt can be determined, as shown in Figure 3.12. These profiles for each speed are averaged over the same equivalent ship length of 30 m. While the surface fluctuations increase steadily from one speed to the next at all values of y, some differences can be seen in the shape of the distributions. Each 78 y (cm) 0 5 10 15 20 25 30 H r m s (m m ) 0 1 2 3 4 5 6 U = 3 m/s U = 4 m/s U = 5 m/s Figure 3.12: RMS surface height averaged over repeated experimental runs versus distance from the belt for three belt speeds. surface height fluctuation profile reaches a maximum at the belt, but in the highest speed case the values decrease more gradually with increasing distance from the belt than in the lower speed cases. 3.2 Flow Field Measurements 3.2.1 PIV measurements with a horizontal light sheet Because the boundary layer evolves temporally along the entire belt at the same time, the processed PIV vector fields were able to be averaged in the streamwsie direction for greater statistical convergence. Experiments at each light sheet location and belt speed were also repeated 20 times for the purpose of ensemble averaging. In addition, some averaging in x by ± 8 frames was performed for each condition, which corresponds to 9.6 cm, 12.8 cm, and 16 cm of belt travel for belt speeds of 3, 4, and 5 m/s, respectively. Velocity profiles, u(y), averaged in this way for belt speeds 79 U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.13: Mean streamwise velocity profiles at a belt speed of U = 3 m/s for (a),(c) D = 14 cm and (b),(d) D = 2.5 cm. Images (a) and (b) each plot a profile at each 0.5 m of belt travel from x = 0 to 5 m, while (c) and (d) each plot a profile at each 5 m of belt travel from x = 5 to 30 m. 80 U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.14: Mean streamwise velocity profiles at a belt speed of U = 4 m/s for (a),(c) D = 14 cm and (b),(d) D = 2.5 cm. Images (a) and (b) each plot a profile at each 0.5 m of belt travel from x = 0 to 5 m, while (c) and (d) each plot a profile at each 5 m of belt travel from x = 5 to 30 m. 81 U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.15: Mean streamwise velocity profiles at a belt speed of U = 5 m/s for (a),(c) D = 14 cm and (b),(d) D = 2.5 cm. Images (a) and (b) each plot a profile at each 0.5 m of belt travel from x = 0 to 5 m, while (c) and (d) each plot a profile at each 5 m of belt travel from x = 5 to 30 m. 82 of 3, 4, and 5 m/s can be seen in Figures 3.13, 3.14, and 3.15, respectively. The plots in the left column of each figure contain profiles from the D = 14 cm condition, while the plots in the right-hand column contain profiles from the D = 2.5 cm condition. The plots in the top row depict profiles from each 0.5 m of belt travel from x = 0 to 5 m, while the bottom row contains profiles at each 5 m from x = 5 to 30 m. It can be seen by comparing the plots in the top row that the boundary layer growth appears quite different between these different conditions. Far from the surface, the profiles from x = 0 to 3 m are very closely clustered together before the growth rate suddenly accelerates after 2 to 3 m of travel. Near the surface, this burst happens much more quickly, with the growth rate accelerating after only about 1 m of belt travel. This burst appears to be an indication of transition to turbulence and the onset location of this mean burst will be discussed later in relation to other events that occur in the flow. These mean velocity profiles can be used to calculate both the displacement thickness and the momentum thickness up until the x location where the bound- ary layer grows beyond the frame of the measurement region. The displacement thickness is the distance that the surface would have to be displaced in order for a uniform free stream velocity profile to retain the same flow rate and is defined using the following equation: δ∗ = ∫ ∞ 0 (1− u(y) U )dy The momentum thickness is the distance that the surface would have to be displaced in order for a uniform free stream velocity profile to retain the same total momentum 83 x (m) 0.25 1 2 3 4 5 θ (m m ) 0 1 2 3 4 D = 14 cm D = 2.5 cm x (m) 0.25 1 2 3 4 5 δ ∗ (m m ) 0 1 2 3 4 5 D = 14 cm D = 2.5 cm (a) (d) x (m) 0.25 1 2 3 4 5 θ (m m ) 0 1 2 3 4 D = 14 cm D = 2.5 cm x (m) 0.25 1 2 3 4 5 δ ∗ (m m ) 0 1 2 3 4 5 D = 14 cm D = 2.5 cm (b) (e) x (m) 0.25 1 2 3 4 5 θ (m m ) 0 1 2 3 4 D = 14 cm D = 2.5 cm x (m) 0.25 1 2 3 4 5 δ ∗ (m m ) 0 1 2 3 4 5 D = 14 cm D = 2.5 cm (c) (f) Figure 3.16: Plots showing (a-c) momentum thickness and (d-f) displacement thickness versus x calculated from measured boundary layer profiles for belt speeds of (a, d) 3 m/s, (b, e) 4 m/s, and (c, f) 5 m/s. 84 and is defined using the following equation: θ = ∫ ∞ 0 u(y) U (1− u(y) U )dy These relations can be applied to the velocity profiles at each x location for both horizontal planes, as shown in Figure 3.16. These plots confirm that the boundary layer grows faster closer to the surface, with both momentum thickness and dis- placement thickness both indicating a thicker boundary layer at every x location. Additionally, it can be seen in each momentum thickness profile that after an initial period of slow boundary layer growth, a kink in the profile occurs, followed by a higher boundary layer growth rate. This kink is an indicator of transition to turbu- lence and again appears to occur earlier in the plane closer to the surface. As with the x location of bursting seen in mean velocity profiles, the x location of this kink can be measured from these plots and will be discussed in more detail later. From the calculations of displacement thickness and momentum thickness, the shape factor H can be defined as the ratio of displacement thickness to momentum thickness. For a laminar Blasius boundary layer, this typically has a value of 2.59 and for a turbulent boundary layer, a value between 1.3 and 1.4 is typical. Figure 3.17 shows the shape factor as a function of x for each belt speed and light sheet location. These plots show a value of H between 2 and 2.25 that gradually increases during the initial portion of each launch, except for the plot for U = 3 m/s at the deeper measurement location. At this speed, in the time between the belt reaching full speed (around 0.85 m) and transition to turbulence, the shape factor grows to a 85 x (m) 0.25 1 2 3 4 5 H 1 1.5 2 2.5 3 D = 14 cm D = 2.5 cm (a) x (m) 0.25 1 2 3 4 5 H 1 1.5 2 2.5 3 D = 14 cm D = 2.5 cm (b) x (m) 0.25 1 2 3 4 5 H 1 1.5 2 2.5 3 D = 14 cm D = 2.5 cm (c) Figure 3.17: Plots showing shape factor, H, as a function of x for belt speeds of (a) 3 m/s, (b) 4 m/s, and (c) 5 m/s. 86 value of around 2.9. Because the belt reaches full speed in this condition well before transition, this could be an indication that the acceleration portion of the belt launch is initially suppressing the shape factor. Alternatively, this higher peak could be due to the relatively higher level of overshoot that occurs at this belt speed, as shown in Figure 2.13, which could lead to an adverse pressure gradient. In each plot, the initial value of shape factor forms a kink and transitions toward a value more typical of a turbulent boundary layer, just below 1.4. As with momentum thickness, this kink produces another indication of transition to turbulence. These numerous indications of transition will be discussed in further detail below. These momentum thickness calculations can also be performed at the x loca- tions of free surface bursting presented earlier, as shown in Figure 3.18 (a). These plots use blue dots to indicate velocity measurements from the deeper horizontal plane, D = 14 cm and red dots indicate velocity measurements from the shallow plane, D = 2.5 cm. Plot (a) shows the same plot for the x location of onset shown in Figure 3.10(a). Figure 3.18 (b) shows the displacement thickness calculated at those locations, while plot (c) shows the momentum thickness at each of those locations. Plots (d), (e), and (f) use these calculated momentum thickness values as a length scale in order to calculate Reynolds number, Froude number, and Weber number, respectively. It can be seen in these plots that both the displacement thickness and momentum thickness are greater closer to the surface, which is in agreement with the idea that the boundary layer grows faster near the free surface. The Froude and Weber number at bursting onset both appear to show a monotonic increase 87 U (m/s) 2.5 3 3.5 4 4.5 5 5.5 x on se t (m ) 0 0.5 1 1.5 2 2.5 U (m/s) 3 4 5 R e θ 0 1000 2000 3000 4000 D = 14 cm D = 2.5 cm (a) (d) U (m/s) 3 4 5 δ ∗ on se t (m m ) 0 0.5 1 1.5 2 2.5 D = 14 cm D = 2.5 cm U (m/s) 3 4 5 F r θ 0 20 40 60 80 100 120 D = 14 cm D = 2.5 cm (b) (e) U (m/s) 3 4 5 θ on se t (m m ) 0 0.2 0.4 0.6 0.8 1 1.2 D = 14 cm D = 2.5 cm U (m/s) 3 4 5 W e θ 0 50 100 150 200 250 D = 14 cm D = 2.5 cm (c) (f) Figure 3.18: Plots showing different scaling parameters for bursting onset, calcu- lated using measured velocity profile data. In all plots, blue dots indicate velocity from D = 14 cm and red dots indicate velocity closer to the surface at D = 2.5 cm. 88 with belt speed, while Reynolds number based on momentum thickness appears to provide the most consistent indicator of bursting onset. In addition to mean velocity profiles, additional information can be gained from looking at RMS velocity fluctuations. The RMS velocity at each location is defined as the RMS fluctuation about a mean velocity profile at that x location. At belt speeds of 3, 4, and 5 m/s, Figures 3.19, 3.20, and 3.21 each show a range of streamwise RMS velocity fluctuations at a range of x values. In each plot, the top row of plots contains profiles from every 1 m of belt travel from x = 1 to 5 m and the bottom row contains profiles at each 5 m of belt travel from x = 5 to 30 m. In each figure, the left column of plots again shows profiles for the condition D = 14 cm, while the right column of plots shows profiles for D = 2.5 cm. Far from the surface, early profiles have a sharp peak near the belt and decrease away from the belt. After about 2 to 3 m of belt travel, a second peak in this distribution begins to grow around y = 3 to 4 mm. Close to the surface, a similar sudden burst of activity occurs after only about 1 m of belt travel. The appearance of this second peak in the urms distribution again seems to coincide with transition to turbulence. After 5 m of belt travel, this second peak flattens out and spreads quickly away from the belt surface. In order to better compare the streamwise flow at these two measurement depths, Figure 3.22 shows plots of profiles at the same condition for both light sheet depths, D = 2.5 and 14 cm. In these plots, mean velocity profiles are shown in the left columnn and streamwise RMS velocity profiles in the right column. Each row contains plots at each 1 m of belt travel from x = 1 to 5 m. These plots clearly 89 urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 1.0 m X = 2.0 m X = 3.0 m X = 4.0 m X = 5.0 m urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 1.0 m X = 2.0 m X = 3.0 m X = 4.0 m X = 5.0 m (a) (b) urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.19: Streamwise rms velocity profiles at a belt speed of U = 3 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. These profiles are plotted in (a) and (b) at each 1 m of belt travel from x = 1 to 5 m and in (c) and (d) at each 5 m of belt travel from x = 5 to 30 m. 90 urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 1.0 m X = 2.0 m X = 3.0 m X = 4.0 m X = 5.0 m urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 1.0 m X = 2.0 m X = 3.0 m X = 4.0 m X = 5.0 m (a) (b) urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.20: Streamwise rms velocity profiles at a belt speed of U = 4 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. These profiles are plotted in (a) and (b) at each 1 m of belt travel from x = 1 to 5 m and in (c) and (d) at each 5 m of belt travel from x = 5 to 30 m. 91 urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 1.0 m X = 2.0 m X = 3.0 m X = 4.0 m X = 5.0 m urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 1.0 m X = 2.0 m X = 3.0 m X = 4.0 m X = 5.0 m (a) (b) urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m urms(y) U 0 0.1 0.2 0.3 0.4 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.21: Streamwise rms velocity profiles at a belt speed of U = 5 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. These profiles are plotted in (a) and (b) at each 1 m of belt travel from x = 1 to 5 m and in (c) and (d) at each 5 m of belt travel from x = 5 to 30 m. 92 U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 D = 14 cm D = 2.5 cm urms(y) U 0 0.5 1 y (m m ) 0 10 20 30 D = 14 cm D = 2.5 cm U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 D = 14 cm D = 2.5 cm urms(y) U 0 0.5 1 y (m m ) 0 10 20 30 D = 14 cm D = 2.5 cm U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 D = 14 cm D = 2.5 cm urms(y) U 0 0.5 1 y (m m ) 0 10 20 30 D = 14 cm D = 2.5 cm U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 D = 14 cm D = 2.5 cm urms(y) U 0 0.5 1 y (m m ) 0 10 20 30 D = 14 cm D = 2.5 cm U−u(y) U 0 0.5 1 y (m m ) 0 10 20 30 D = 14 cm D = 2.5 cm urms(y) U 0 0.5 1 y (m m ) 0 10 20 30 D = 14 cm D = 2.5 cm Figure 3.22: Streamwise profiles comparing the velocity profiles at different dis- tances from the free surface. The left column shows plots of U and the right column shows plots of urms. Each column contains plots at each 1 m from x = 1 to 5 m. The top-most plots correspond to x = 1 m, while the bottom plots correspond to x = 5 m. 93 depict the burst process observed in the earlier plots, as well as an overall faster growth of the boundary layer near the free surface. The numerous indications of transition to turbulence discussed within this section can be compared, as shown in Figure 3.23. It is clear that each of these events can be used as an indication of transition to turbulence and this plot gives an indication of the types of events that this transition sets into motion. This x location of transition will later be used as a reference point in discussing the timing of other events recorded throughout these experiments. U (m/s) 3 4 5 x tr a n si ti on (m ) 0 0.5 1 1.5 2 2.5 3 3.5 4 Momentum Thickness Kink RMS Knee Mean Burst Shape Factor Kink Momentum Thickness Kink RMS Knee Shape Factor Kink Shape Factor Kink Figure 3.23: Plot of x locations of four different events that indicate transition to turbulence. Points in blue represent data from measurements at D = 14 cm and points in red represent data from measurements at D = 2.5 cm. In addition to the streamwise velocity component, it is also useful to examine the behavior of wall-normal velocity fluctuations, as shown in Figures 3.24, 3.25, and 3.26. Each figure contains plots of profiles at each 0.5 m from x = 0 to 5 m in 94 vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.24: Wall-normal rms velocity profiles at a belt speed of U = 3 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. In plots (a) and (b), profiles are plotted at each 0.5 m of belt travel from x = 0 to 5 m, while in plots (c) and (d), profiles are plotted at each 5 m of belt travel from x = 5 to 30 m. 95 vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.25: Wall-normal rms velocity profiles at a belt speed of U = 4 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. In plots (a) and (b), profiles are plotted at each 0.5 m of belt travel from x = 0 to 5 m, while in plots (c) and (d), profiles are plotted at each 5 m of belt travel from x = 5 to 30 m. 96 vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.26: Wall-normal rms velocity profiles at a belt speed of U = 5 m/s for (a), (c) D = 14 cm and (b), (d) D = 2.5 cm. In plots (a) and (b), profiles are plotted at each 0.5 m of belt travel from x = 0 to 5 m, while in plots (c) and (d), profiles are plotted at each 5 m of belt travel from x = 5 to 30 m. 97 vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m vrms(y) U 0 0.05 0.1 0.15 y (m m ) 0 10 20 30 40 50 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) (c) Figure 3.27: Wall-normal rms velocity profiles for belt speeds of (a) 3, (b) 4, and (c) 5 m/s at D = 14 cm. the top row and plots at each 5 m of belt travel from x = 5 to 30 m in the bottom row. Plots in the left column are far from the free surface and plots in the right column are close to the free surface. In each figure, plot (a) shows that far from the surface, these wall-normal velocity fluctuations contain a peak slightly away from the wall that grows and moves farther from the wall with increasing x. Close to the surface, however, these vrms profiles contain a sharp peak close to the belt surface with a plateau farther from the belt. These profiles also appear to show overall larger fluctuations in wall-normal velocity at all wall-normal positions. In comparing these profiles for the three belt speeds in the D = 14 cm case, see Figure 3.27, it can be seen that this near-wall peak seen in the near-surface conditions also appears to grow with increasing belt speed in the deep-water condition. Plot (a) shows a peak in the range of y = 3 to 5 mm for the lowest-speed case, while (b) shows a plateau of this quanitity near the wall, and (c) shows a large, sharp peak near the wall. It is possible 98 that this region exists at low speed, with a layer that is simply too thin to resolve with the present resolution. Additionally, this near-wall peak could be influenced slightly by the y direction belt fluctuations, as shown earlier in Figure 2.27. 3.2.2 PIV Measurements with a vertical light sheet While the velocity profiles extracted from the horizontal velocity field measure- ments show that the flow is altered in the vicinity of the free surface, vertical velocity fields are necessary to quantify this trend. In these vertical velocity fields, averaging at each belt speed is cut off near the surface at a level below the deepest free sur- face depressions, as determined from raw images. In this way, any cross-correlations performed in regions without particles that may bias statistics are avoided. This distance was found to be 8 mm, 17 mm, and 21 mm for belt speeds of 3, 4, and 5 m/s, respectively. Figures 3.28, 3.29, and 3.30 show streamwise mean velocity profiles plotted at a range of x locations, with the top row of plots in each figure containing profiles from every 0.5 m of belt travel from x = 0 to 5 m and the bottom row containing profiles at every 5 m of belt travel from x = 5 to 30 m. The plots in the left column contain profiles from close to the belt (y = 1.9 cm) and plots in the right column contain profiles farther from the belt (y = 3.8 cm). It is clear in all cases that these velocity profiles increase significantly as the free surface is approached, particularly at low speed where statistics are valid for shallower depths and at greater x locations when the boundary layer has grown significantly. This behavior confirms the trends seen in the horizontal PIV measurements that indicate 99 u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.28: Streamwise mean velocity profiles at a belt speed of U = 3 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. 100 u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.29: Streamwise mean velocity profiles at a belt speed of U = 4 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. 101 u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m u(y) U 0 0.2 0.4 0.6 0.8 1 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.30: Streamwise mean velocity profiles at a belt speed of U = 5 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. 102 a thicker boundary layer near the free surface. This depth of influence of the free surface appears to become thicker at higher belt speeds, reaching depths of up to 5 cm at U = 5 m/s. Additional information about the effect of the free surface boundary condi- tion can be gained from examining streamwise velocity fluctuations, as shown in Figures 3.31, 3.32, and 3.33, which each contain urms profiles plotted at a range of x locations, with the top row of plots in each figure containing profiles from every 0.5 m of belt travel from x = 0 to 5 m and the bottom row containing profiles at every 5 m of belt travel from x = 5 to 30 m. The plots in the left column contain profiles from close to the belt (y = 1.9 cm) and plots in the right column contain profiles farther from the belt (y = 3.8 cm). It can be seen in Figure 3.31, and particularly in plots (a) and (b) from early in the run, that these streamwise rms velocity profiles increase close to the surface, with the strongest effect appearing to be limited to within 20 cm of the surface. In plot (a) of each of these figures, a slight inclination of these profiles also appears to exist during the development of the boundary layer close to the belt. Overall, the level of noise present in these profiles prevents any conclusive statements about the behavior of the streamwise rms velocity fluctuations from being made. In addition to streamwise velocity profiles, vertical velocity fluctuations can provide a valuable insight into the effects of the free surface boundary condition. Figures 3.34, 3.35, and 3.36 each contain wrms profiles plotted at a range of x locations, with the top row of plots in each figure containing profiles from every 0.5 m of belt travel from x = 0 to 5 m and the bottom row containing profiles 103 urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.31: Streamwise rms velocity profiles at a belt speed of U = 3 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. 104 urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.32: Streamwise rms velocity profiles at a belt speed of U = 4 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. 105 urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m urms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.33: Streamwise rms velocity profiles at a belt speed of U = 5 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. 106 wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.34: Vertical rms velocity profiles at a belt speed of U = 3 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. 107 wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.35: Vertical rms velocity profiles at a belt speed of U = 4 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. 108 wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 0.0 m X = 0.5 m X = 1.0 m X = 1.5 m X = 2.0 m X = 2.5 m X = 3.0 m X = 3.5 m X = 4.0 m X = 4.5 m X = 5.0 m (a) (b) wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m wrms(y) U 0 0.05 0.1 0.15 0.2 z (m m ) 0 20 40 60 80 X = 5.0 m X = 10.0 m X = 15.0 m X = 20.0 m X = 25.0 m X = 30.0 m (c) (d) Figure 3.36: Vertical rms velocity profiles at a belt speed of U = 5 m/s for (a), (c) y = 1.9 cm and (b), (d) y = 3.8 cm. Plots (a) and (b) are from early in the run, with profiles plotted every 0.5 m of belt travel from x = 0 to 5 m. Plots (c) and (d) contain profiles plotted every 5 m from x = 5 to 30 m. 109 at every 5 m of belt travel from x = 5 to 30 m. The plots in the left column contain profiles from close to the belt (y = 1.9 cm) and plots in the right column contain profiles farther from the belt (y = 3.8 cm). In examining these profiles, it can be seen that, as with vertical profiles of urms, the largest effect is seen within 20 mm of the undisturbed free surface at U = 3 m/s where data from this region is considered valid. These plots appear to show an increase in vertical velocity fluctuations close to the surface early in the run. Close to the belt, in plot (a) of each figure, a slight, gradual increase in vertical velocity fluctuations can be seen as the free surface is approached. This trend appears to reverse later in each run, with a gradual decrease in vertical fluctuations close to the free surface as the boundary layer becomes thicker. This behavior seems to show that the thicker boundary layer leads to a quick increase of turbulent fluctuations close to the surface early on, but as the boundary layer develops, the constraining effect of the free surface has greater influence and works to reduce these fluctuations. As with the streamwise velocity fluctuations, the level of noise present in these profiles prohibits firm conclusions about these behaviors. 3.3 Air Entrainment Events In order to study the air entrainment behavior in this boundary layer flow, it is necessary to record movies from below the free surface. Using the same camera view as was used in the vertical plane PIV experiments, white light movies were captured in order to determine the onset location of air entrainment. Frames captured from 110 movies of belt launches to 3, 4, and 5 m/s at every 5 m of belt travel from x = 0 to 30 m are shown in Figure 3.37. It is clear from watching these movies that almost no air entrainment occurs at U = 3 m/s. While a few bubbles are occasionally entrained, this belt speed does not exhibit sustained air entrainment as the higher speeds do. It can also be seen that the volume and quantity of bubbles entrained at U = 5 m/s appear to be greater than during a launch to 4 m/s. From watching these movies, air entrainment can be observed and recorded in each movie. One method of judging the onset of air entrainment is to record the first instance of air being entrained within the field of view. This method yields x onset locations of 3.61 m and 2.66 m for belt speeds of 4 and 5 m/s, respectively. Another method is to record the location of the first bubble to enter the field of view of the camera. This yields air entrainment onset locations of x = 2.51 and 2.31 m for belt speeds of 4 and 5 m/s, respectively. While the first method gives an exact measurement of when air is entrained in a given field of view, this method appears to be somewhat more random and less consistent because it relies more on local flow conditions to generate air entrainment. In addition, it is clear that air entrainment often begins earlier than this recorded onset location because of the presence of bubbles passing into the field of view of the camera. In either case, no consistent air entrainment can be deduced for belt speeds of 3 m/s. These two measurements of air entrainment onset location will later be discussed in relation to the flow statistics presented previously. From both surface profile movies and underwater white light movies, a vari- ety of air entrainment events can be seen. Perhaps the most prominent of these 111 Figure 3.37: Three sequences of images from underwater white light movies at three different belt speeds. The three columns contain frames from belt launches to 3, 4, and 5 m/s respectively. Each row contains frames from each 5 m of belt travel from x = 0 to 30 m. 112 (a) (f) (b) (g) (c) (h) (d) (i) (e) (j) Figure 3.38: Two sequences of images from LIF movies depicting potential air entraining events. Images in the left column depict a trench closure event and images in the right column depict a wave breaking event. Both sets of images are from a belt launch to 5.0 m/s. 113 mechanisms begins with the development of deep narrow troughs that are elongated in a quasi-streamwise direction. Some such troughs become so deep that the sides collapse together. When the sides of the trough meet, air is often entrapped beneath the free surface and bubbles begin to pinch off and convect downstream. Another air entrainment mechanism is caused by ejections from the free surface. These ejec- tions can fold over the surface and trap a pocket of air, similar to the process of air entrainment in plunging breaking waves. From the LIF movies, some evidence of both types of air entrainment events can be seen. Figure 3.38 contains two sets of images to help illustrate the mechanisms for air entrainment as seen from above the free surface. In the left column, a trough collapse is depicted. In images (a) and (b), a deep narrow trough can be seen clearly toward the left side of the image. In these first images, the left side of the trough is beginning to turn over. In image (c), the two sides of the trough meet and a droplet is ejected into the air. Presumably, this process entraps a pocket of air beneath the free surface. In the right column, the second type of surface breaking event can be seen. In the first image, (f), a jet can be seen forming near the contact point of the free surface with the belt, toward the left side of the image. This jet ejects from the surface and becomes nearly horizontal by image (h). In image (j), the jet makes contact with the free surface, potentially trapping any air that remained below it during the formation of the jet. Since the camera is above the free surface and because of blockage of the camera line of sight by other free surface features between the camera and the light sheet, the LIF images do not give us a clear 114 indication of whether or not air was actually entrained by either of these particular events. In order to be able to determine if air is entrained by surface breaking events, another way to visualize the events at the air-water interface is by recording white light movies beneath the free surface. This arrangement reduces line of sight block- age by other surface features and allows for a more detailed study of how the air is entrained rather than just showing the surface features that are likely to produce bubbles. Figure 3.39 shows two sets of images from two views of the same event. Each image contains both the feature of interest and its shadow projected onto the belt by the light source for each camera. This can be seen especially well in im- age pair (d)-(j). In image (d), the camera is on the left side, with the light source coming from the right side, so that the actual trough marked as object 2, casts a shadow shown in object 1. In image (j), the camera is on the right side with the light source to the left, so the same trough is shown as object 4 with its shadow projected to the right, shown as object 5. The relative horizontal distance between and object and its shadow is an indication of the distance of the object from the belt, where the smaller the distance, the closer the object is to the belt surface. In the first three sets of images, the trough can be seen to become deep and narrow. In the subsequent images, the left side of the trough can be seen turning over and contacting the opposite side of the trough, as it pinches off and becomes a large bubble in the final image pair. These events, where one side turns over completely, seem to primarily form large bubbles when entraining pockets of air. 115 (a) (g) (b) (h) (c) (i) 1 2 3 5 64 (d) (j) (e) (k) (f) (l) Figure 3.39: Two sequences of images from underwater white light movies of the same air entrainment event. The belt is moving from left to right. Images from the left column were captured with the left camera and images from the right camera were captured with the right camera. This event occured during a launch to 4.0 m/s. Because of the lighting and imaging configuration shown above in Figure 2.30, each image contains both the bubble of interest and its shadow on the belt. 116 (a) (g) (b) (h) (c) (i) (d) (j) (e) (k) (f) (l) Figure 3.40: Two sequences of images from underwater white light movies of the same air entrainment event. The belt is moving from left to right. Images from the left column were captured with the left camera and images from the right camera were captured with the right camera. This event occured during a launch to 5.0 m/s. 117 A second type of air entrainment event can be seen in Figure 3.40. As with the previous Figure, this event is shown from two camera views in two columns. This event begins, as shown in image (g), with the two sides of the trough closest and farthest from the right camera meeting at a single point in the middle of the trough and forming a round hole in the trough on the top left of the image. Following this event, the hole spreads rapidly, propagating outward from the meeting point and forming a large ligament of air that can be seen in image (h). In the following images, this ligament can be seen retracting towards the surface as it deposits a series of small bubbles into the flow. 3.3.1 Event Timing In an attempt to study the influence of flow field parameters on free sur- face bursting and air entrainment, the timing of these events can be compared. Figure 3.41 shows a set of timelines that depict the timing of a variety of events dis- cussed previously at each belt speed, with flow field events taken from near-surface measurements. In these timelines, colored lines connect the same event between belt speeds. It an be seen that the numerous events indicating transition to turbulence, as discussed earlier and plotted in red, occur at a similar x location. The onset of free surface bursting occurs closer to this transition point with increasing belt speed. At 5 m/s, it appears that bursting occurs a short time after transition. In all cases, bursting occurs before the secondary peak in streamwise RMS reaches a maximum. The rise of this second peak away from the wall could contribute to the onset of 118 bursting and will be discussed in more detail below. Both measurements of air en- trainment, presented in green, occur well after bursting and after the streamwise velocity fluctuations away from the wall reach a maximum and begin to decrease. x (m) 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 U (m /s ) 3 4 5 Full Speed Mean Transition Momentum Transition Shape Factor Transition RMS Transition Max RMS Peak Free Surface Bursting First Bubble Air Entrainment Figure 3.41: A set of timelines depicting the x location of various events described throughout this dissertation. The time series of velocity data collected for each speed can be used to calculate the momentum thickness, θ, and urms at the location of the secondary peak (y = 3.04 mm) versus x throughout each run, as shown in Figure 3.42. Using these values as length and velocity scales, a variety of scaling parameters can be calculated for each x value in an attempt to determine the important physical processes influencing free surface fluctuations and air entrainment. While the length scale θ increases steadily throughout the run, this velocity scale reaches a peak and then decreases fairly steadily after that point. This could indicate that air entrainment processes are strongest in this region of rapid boundary layer growth and are reduced later in each run. From these velocity and length scales, the Weber and Froude numbers can be calculated at each x location and these parameters can be compared across belt 119 x (m) 1 2 3 4 5 θ (m m ) 0 1 2 3 4 U = 3 m/s U = 4 m/s U = 5 m/s x (m) 0 1 2 3 4 5 u rm s (m / s) 0 0.1 0.2 0.3 0.4 0.5 U = 3 m/s U = 4 m/s U = 5 m/s (a) (b) Figure 3.42: Plots showing (a) θ vs x and (b) urms vs x for the initial 5 m of belt travel at each belt speed. speeds for both bursting and air entrainment. Figure 3.43 shows plots of Weber number vs Froude number at each speed, with the values at bursting and air en- trainment marked with connected lines, as in Figure 3.41. Plot (a) shows Weber number and Froude number calculated using θ as a length scale and U as a velocity scale. Using this scaling, both bursting and air entrainment appear to depend on reaching a critical Weber number in order for onset to occur. While this appears to be a reasonable conclusion for free surface bursting, this may not be true for air entrainment. Because the momentum thicnkess length scale should continue to grow beyond what has been calculated here and the velocity scale remains constant, the Weber number at each belt speed will continue to grow. Therefore, even at a belt speed of 3 m/s, the critical value of Weber number for air entrainment at higher speeds will be surpassed. Therefore, perhaps a more appropriate velocity scale is the urms of the secondary peak, discussed above. A plot of Weber vs Froude number using both θ as a length scale and urms as a velocity scale are shown in Figure 3.43 (b). Using this scaling, it is unclear how free surface bursting depends on Weber or 120 Froude number, but air entrainment appears to scale better using these length and velocity scales. From both measurements of air entrainment, it appears that a criti- cal Weber number of approximately 3 to 3.5 must be reached before air entrainment occurs. While this threshhold is reached at 3 m/s, it appears to only occur briefly, while at higher speeds, the Weber number increases sharply after this point. Weθ,U 0 500 1000 1500 F r θ ,U 0 20 40 60 80 100 U = 3 m/s U = 4 m/s U = 5 m/s Bursting First Bubble Air Entrainment Weθ,urms 0 2 4 6 8 10 F r θ ,u rm s 0 1 2 3 4 5 U = 3 m/s U = 4 m/s U = 5 m/s Bursting First Bubble Air Entrainment (a) (b) Figure 3.43: Plots showing Weber number vs Froude number, calculated using θ as a length scale and a velocity scale of (a) U or (b) urms at y = 3.04 mm, the location of the secondary peak in streamwise velocity fluctuations. An alternative explanation to the idea of a single critical Weber or Froude number is that there is some combined effect of both scaling parameters. This idea is expressed in Brocchini and Peregrine (2001), as shown in Figure 1.5. The recorded data discussed above can be plotted on top of this figure, as shown in Figure 3.44. While the boundary for air entrainment onset does not appear to be appropriate for the length and velocity scales used here, the curves for higher belt speeds do appear to reach further into the upper air entrainment region of the plot, indicating that this figure could be describing the physics correctly. 121 Figure 3.44: A figure from Brocchini and Peregrine (2001) depicting critical regions for air entrainment based on length and velocity scales, with recorded data of θ and urms plotted on top. 3.4 Summary and Conclusions The entrainment of air due to turbulent fluctuations in the boundary layers of large naval ships is an important problem for the detectability of these ships. Due to the large range of scales important to the physics of this problem, research on this topic using traditional means, both experimental and computational, has been limited to this point. A review of the literature found only research that is tangentially related to this problem. In this dissertation, a novel laboratory- scale device was created in order to experimentally study the interaction of the turbulent boundary layer with a free surface. This device utilizes a stainless steel belt, driven by hydraulic motors around two rollers as a surface piercing vertical wall of infinite length. This belt accelerates in under 0.5 seconds to constant speed 122 in an effort to mimick the sudden passage of a flat-sided ship. Utilizing the full length and velocity scales of large naval ships, this device creates a temporally- evolving boundary layer analagous to the spatially-evolving boundary layer along the length of a ship. Water surface profiles were recorded with a cinematic LIF system to study the generation of surface height fluctuations by the sub-surface turbulence. Sub-surface velocity fields were recorded using a cinematic planar PIV system in order to study the modification of the flow field in the vicinity of the free surface. Underwater white-light movies were recorded to determine the onset location of air entrainment as well as mechanisms for this entrainment. It was found that the free surface remains calm during the acceleration portion of the belt launch before bursting with activity close to the belt. This burst location was seen to vary from 1 to 2 m, depending on belt speed. Reynolds number based on x appear to show a consistent indicator of bursting around Rex = 6 ×106. After this point, the free surface fluctuates and changes form rapidly near the belt surface where wave breaking begins to occur. These free surface height fluctuations are seen to peak directly at the belt surface and decrease gradually away from the belt. At all wall- normal positions, these fluctuations increase with belt speed. This Hrms peak near the belt reaches a value approximately 50% higher than in the far field. The free surface ripples created near the belt surface appear to lead to the generation of freely- propagating waves far from the belt surface. The speed of these waves normal to the belt surface increases with increasing belt speed, with waves traveling approximately 25% faster at U = 5 m/s when compared to those at U = 3 m/s. In studying sub-surface velocity fields, it was found that the boundary layer exhibits similar 123 bursting to what is seen in free surface profiles, with an accompanying secondary peak in streamwise velocity fluctuations. This second peak appears to occur at a y location of approximately 3 mm from the belt surface and evidence of this second peak is found in the literature for high Reynolds number boundary layer flows. The rapid boundary layer growth seen in these velocity profiles happens earlier in planes closer to the surface, leading to an overall thicker boundary layer in the vicinity of the free surface. By analyzing these mean streamwise velocity profiles and comparing the results to the x locations of free surface bursting, it is found that both displacement thickness and momentum thickness at the location of bursting decreases monotonically with belt speed, while Reynolds number based on momentum thickness provides a fairly consistent indicator of free surface bursting at Reθ of approximately 2300 based on deep velocity profiles and 4400 based on shallow velocity profiles. Additionally, an overall increase in wall-normal velocity fluctuations is seen close to the surface early in each run, with unusual behavior occurring near the belt surface. While the near-surface velocity profiles always show a near-wall peak in the vrms profiles, this peak is reduced at U = 4 m/s and completely eliminated at 3 m/s when far from the surface. In vertical velocity fields, boundary layer growth of up to 30% is found near the free surface and this effect is apparent for depths of up to 5 cm. Free-surface normal fluctuations appear to show a small increase toward the surface during the period of rapid boundary layer growth, which then reverses and becomes a small, gradual decrease as the free surface is approached later in the run. This again lends credence to the idea that the boundary layer grows faster near the surface, while later in the run, the 124 restraining effects of surface tension and gravity become more dominant forces. From underwater white-light movies, it is found that no sustained air entrainment occurs at U = 3 m/s, while some air entrainment occurs at 4 m/s, and a large amount of air entrainment occurs at 5 m/s. Additionally, different types of air entrainment events are detected at small distances away from the belt, which entrain a variety of bubble sizes. It is apparent that deep, quasi-streamwise troughs form close to the belt and may become unstable before collapsing. In separate events, plunging wave breaking appears to overturn the surface and entrap pockets of air. In all cases, these air entrainment events appear to occur a small distance away from the belt surface, appearing to coincide with location of the growing secondary peak in steamwise velocity fluctuations. Onset of both bursting and air entrainment appear to depend in some way on Weber number. Free surface bursting appears to scale well with Weber number of approximately 200 based on momentum thickness and belt velocity, while air entrainment appears to occur after reaching a Weber number of approximately 3 to 3.5 based on momentum thickness and streamwise velocity fluctuations. These scaling parameters appear to agree with the behavior of the scaling arguments presented in Brocchini and Peregrine (2001), if not the exact values for this air entrainment boundary. 3.5 Directions for Future Work In future work, further exploration of vertical velocity fluctuations near the surface could lead to a greater understanding of the coupling between sub-surface ve- 125 locity fluctuations and free surface deformations. An increase in repeated runs with a vertically-oriented light sheet in order to increase statistical significance would be the first step towards improving our understanding of this behavior. Alternatively, while the present study measured these separately, a measurement campaign that combined LIF with vertical-plane PIV could allow for measurements to more closely approach the free surface and give insight into the flow field near particular events. Additionally, LIF in planes parallel to the belt surface would lead to a greater un- derstanding of the complete wave propagation behavior, particularly when close to the belt. Finally, it would be desireable to study the mechanisms and location of air entrainment in more detail, perhaps by extracting quantitative information from stereo white light movies of bubbles, perhaps combining this with LIF in order to direcly match wave breaking to air entrainment. 126 Bibliography Amini, A., Weymouth, T., and Jain, R., “Snakes: Active contour models,” IEEE Transactions on Pattern Analysis and Machine Intelligence, Vol. 12, no. 9, 1990, pp. 321–331. Brocchini, M. and Peregrine, D., “The dynamics of strong turbulence at free sur- faces. part 1. description,” Journal of Fluid Mechanics, Vol. 449, no. 1, 2001, pp. 225–254. Broglia, R., Pascarelli, A., and Piomelli, U., “Large-eddy simulations of ducts with a free surface,” Journal of Fluid Mechanics, Vol. 484, 2003, pp. 223–253. Chanson, H., Air Bubble Entrainment in Free-Surface Turbulent Shear Flows, Aca- demic Press, 1996. Grega, L., Hsu, T., and Wei, T., “Vorticity transport in a corner formed by a solid wall and a free surface,” Journal of Fluid Mechanics, Vol. 465, 2002, pp. 331–352. Grega, L., Wei, T., Leighton, R., and Neves, J., “Turbulent mixed-boundary flow in a corner formed by a solid wall and a free surface,” Journal of Fluid Mechanics, Vol. 294, 1995, pp. 17–46. Hotta, T. and Hatano, S., “Turbulence measurements in the wake of a tanker model on and under the free surface,” Journal of the Society of Naval Architects of Japan, Vol. 154, 1983, pp. 261–270. Hsu, T., Grega, L., Leighton, R., and Wei, T., “Turbulent kinetic en- ergy transport in a corner formed by a solid wall and a free surface,” Journal of the Fluid Mechanics, Vol. 410, 2000, pp. 343–366. Kaas, M., Witken, A., and Terzopoulos, D., “Snakes: Active contour models,” International Journal of Computer Vision, Vol. 1, 1988, pp. 321–331. Kim, D., Mani, A., and Moin, P., “Numerical simulation of bubble formation by breaking waves in turbulent two-phase couette flow,” CTR Ann. Res. Brief, 2013, pp. 37–46. Kim, D., Mani, A., and Moin, P., “Investigation of bubble for- mation by breaking waves in turbulent two-phase couette flows,” Proceedings of the 30th Symposium on Naval Hydrodynamics, Hobart, Tas- mania, Australia, 2014. 127 Longo, J., Huang, H., and Stern, F., “Solid/free-surface juncture boundary layer and wake,” Experiments in Fluids, Vol. 25, no. 4, 1998, pp. 283–297. Mei, R., “Velocity fidelity of flow tracer particles,” Experiments in Fluids, Vol. 22, 1996, pp. 1–13. Pedocchi, F., Martin, J. E., and Garcia, M. H., “Inexpensive fluorescent particles for large-scale experiments using particle image velocimetry,” Experiments in Fluids, Vol. 45, 2008, pp. 183–186. Schlichting, H., Boundary Layer Theory, McGraw-Hill, New York, 1979. Shen, L. and Yue, D., “Large-eddy simulation of free-surface turbulence,” Journal of Fluid Mechanics, Vol. 440, 2001, pp. 75–116. Sreedhar, M. and Stern, F., “Prediction of solid/free-surface juncture bound- ary layer and wake of a surface-piercing flat plate at low Froude number,” Jounral of Fluids Engineering - Transactions of the ASME, Vol. 120, no. 2, 1998, pp. 354–362. Stern, F., “Effects of Waves on the Boundary Layer of a Surface-Piercing Body,” Journal of Ship Research, Vol. 30, no. 4, 1986, pp. 256–274. Stern, F., Hwang, W., and Jaw, S., “Effects of Waves on the Boundary Layer of a Surface-Piercing Flat Plate: Experiment and Theory,” Journal of Ship Research, Vol. 33, no. 1, 1989, pp. 63–80. Teixeira, M. and Belcher, S., “On the initiation of surface waves by turbulent shear flow,” Dynamics of Atmospheres and Oceans, Vol. 41, 2006, pp. 1–27. Washuta, N., Masnadi, N., and Duncan, J. H., “The turbulent boundary layer on a horizontally moving, partially submerged, surface-piercing verticalwall,” Proceedings of the 30th Symposium on Naval Hydrodynamics, Hobart, Tasma- nia, Australia, 2014. 128