A NEW PARADIGM FOR GENERATING SURFACE-NORMAL FORCES FOR HULL-CLEANING ROBOTS By Kristina Maria Kamensky A DISSERTATION Michigan State University in partial fulfillment of the requirements Submitted to for the degree of Mechanical Engineering – Doctor of Philosophy 2020 ABSTRACT A NEW PARADIGM FOR GENERATING SURFACE-NORMAL FORCES FOR HULL-CLEANING ROBOTS By Kristina Maria Kamensky A Bernoulli pad uses an axial jet to produce radial outflow between the pad and a proximally located parallel surface, which may be either a mobile workpiece or a fixed wall. The flow field produces a force between the surfaces which depends upon their spacing Ò. The direction of this force is repulsive for large and small h, but two equilibria exist between these limits. The nearer equilibrium point (Òeq) is stable, and this distance is dependent on the direction and magnitude of the force the pad is required to apply. Increasing the flow rate increases the strength of the contactless grip, subject to cavitation or compressibility constraints, depending on the working fluid. Industry has created devices of this type to grip and transport a variety of workpieces without contact. The present research is inspired by the need to keep a submerged ship hull free of biofouling organisms. Preventative maintenance during idle periods of operations can improve efficiency while prolonging the original surface properties of the hull. The Bernoulli pad for this application is significantly larger and uses the surrounding water as the working fluid. In the present work, the flow field was investigated computationally and experimentally. Field tests were also performed to determine the ability of the device to mitigate biofouling. The computational work, which was validated with experimental results found in literature, indicates that a power-law relationship exists between Òeq and the inlet fluid power required to sus- tain this equilibrium spacing when each is appropriately scaled. This scaling is derived principally from the wall shear; an additional term incorporating the inlet Reynolds number is used to account for the force applied to the system. The relationship is valid over a range of forces acting on the system, geometric, and material properties. Major and minor geometry alterations provide insight to customizing pressure or wall shear stress profiles. The biofouling removal ability of a shear-based device was field tested on two submerged surface types, Garolite G-10 and AkzoNobel’s Intersleek 1100SR. The latter is a fouling-release coating. Each surface was groomed at four frequencies along with a control group during a seven-week grooming study conducted in Narragansett Bay in Rhode Island. An image-processing algorithm was developed and used to assess the effectiveness of the various grooming protocols, along with direct measurements of chlorophyll a per surface area. The image-processing data showed that the grooming resulted in approximately 50% cleanliness on the Garolite at the end of the study whereas the Intersleek was continuously restored to nearly its initial clean state. Chlorophyll a data supported these overall conclusions. These results indicate that surface cleanliness can be maintained effectively on Intersleek using frequent shear-based grooming. The key to success is to match or supersede the critical wall shear stress of settled biofouling organisms whose adhesive strength is exponential in time. Particle Tracking Velocimetry (PTV) measurements were also taken on the flow field. This La- grangian measurement approach uses an iterative particle reconstruction technique in combination with high seeding density to reconstruct a 4D (x, y, z, t) flow field. This 4D reconstruction allows the pressure field to be reconstructed using the Navier-Stokes equations. Various experiments have been conducted on confined radial outflow but PTV measurements are presented here for the first time. The PTV measurements were compared with computational results and while there is reasonable agreement in the velocity field data, there are discrepancies is the pressure field data. Recommendations are provided for future work that can reconcile these differences. Copyright by KRISTINA MARIA KAMENSKY 2020 Dedicated to Michael Anthony Donegan, who never let my confidence run dry. v ACKNOWLEDGEMENTS I would like to express my gratitude and thanks to my advisor, Dr. Ranjan Mukherjee, for his guidance and patience during my career at Michigan State University. Together, we have worked on more than just the research presented in this paper. He has fostered an incredible lab where he encourages my lab mates and myself to push ourselves while we push the boundaries on scientific knowledge. I would also like to thank my Navy mentor, Dr. Aren Hellum, who has provided insight at every step. He also has been an exceptional role model, and I am proud of my career trajectory. I also want to thank Dr. John Foss for his earlier investigation into confined radial flow, which definitely was the seed of inspiration for this research. Thank you Dr. Ricardo Mejia for being on my committee and letting me use your fluid flow visualization setup. Your help when I would get stuck at processing and evaluating the data was also instrumental in my research. Thank you Richard Prevost, from LaVision, for taking the time to lend your expertise on operating the Particle Tracking Velocimetry setup efficiently to get results within the first few tries. Thank you Dr. James Klausner for being on my committee. Your passion for sustainable engineering helps to give me critical feedback on my own research. Thank you Office of Naval Research, and in particular Maria Medeiros, for funding and sup- porting me through the Naval Undersea Research Program (NURP). I benefited highly from such a unique graduate school experience. I pursued engineering in order to make a difference in the world, in particular to make systems more energy efficient in order to reduce environmental footprint. It is not a coincidence that this dissertation reflects that goal. The biggest thank you to my parents, Martin and Lydia Kamensky, who loved and nurtured the curious scientist in me from a very early age. My dad instilled gardening in me, as well as a passion for tinkering with tools to fix and build things. I have so many memories of books with my mom. Not only did she buy me books (science, insect, and animal books were my favorites), but she also took me on countless library visits. I love books and the pursuit of knowledge more than ever. Thank you to my brother, Martin Ernest Kamensky, who was born first. By choosing your vi college and major before me, I was able to learn from you what engineering was truly all about. A shout-out to all my aunts, uncles, and cousins for their love and support. I also want to thank my best friend and sister that I always wanted, Kimberly June "Bug" Moody. You have always encouraged me to be my weird self. I cannot thank my partner, Michael Anthony Donegan, enough. We met in Flint, Michigan when were were both pursuing our Bachelors of Science in Electrical (him) and Mechanical (me). I’ve been infatuated with him ever since. Michael has been an endless source of strength and support for me while I continued being a professional student. His immediate and extended family has also been a blessing. They have never stopped cheering me on. There are so many people at MSU that I want to thank. I am glad I somehow sat next to Andrew Hess at a campus-wide teaching assistant orientation. I thought I was getting too old to make new best friends, but he thankfully proved me wrong. I am grateful of my research lab family for being who they are; Mahmoud Abdullatif, Amer Allafi, Sanders Aspelund, Connor Boss, Sheryl Chau, and Nilay Kant. Pursuing a PhD is an individual journey, but it is less daunting with people like them. Many thanks to the various faculty that welcomed my questions when I knocked on their office door or emailed them, regardless if I was taking a course with them. A huge thanks to the Division of Engineering Computing Services (DECS), who without them I would not have the resources or support to be successful. I also can not thank the people who run the Department Mechanical Engineering enough. I want to thank Roy Bailiff and Mike Koschmider from the Manufacturing Teaching Laboratory for all the random help and 3D printing they have provided. A huge thanks to Dr. Katy Colbry for her help and words of encouragement. I have so much gratitude toward Dr. Percy Pierre and the Sloan Engineering Program that he created. I am also thankful for Julie Rojewski and MSU BEST (Broadening Experiences in Scientific Training) for empowering me to make the most of my time on campus. A huge thanks to the various people who helped me with my field test studies in Newport, Rhode Island. Thanks to the Center of Corrosion and Biofouling Control at the Florida Institute of Technology for inviting me to visit their group and providing the special paint. Thanks to Dr. vii Mark Menesses from Boston University who was a fellow graduate student in NURP. He shared his knowledge of biofouling and experience in field test studies. Thanks to Natasha Dickinson from NUWC for regularly visiting me at the dock and identifying the invertebrates that were living on my experimental setup. Thanks to Dr. Pia Moisander and Abhishek Naik from University of Massachusetts Dartmouth for your collaboration. Thanks to Dr. Christin Murphy for taking an interest in me and becoming one of my newest mentors. Thanks to everyone who gave me a chance prior to my time at MSU, especially Dr. Homayun Navaz. When I visited his office to talk about joining his research group as an undergraduate student at Kettering University in 2007, I never could imagine how far that interview would take me. Not only did I write my first thesis on evaporation and vapor transport models within porous materials with him as my advisor, but I stayed in Flint to pursue a Masters. Dr. Navaz and I started a company to do consulting on optimizing commercial refrigeration. I wrote another thesis based on our own research on quantifying and visualizing the infiltration/exfiltration process in walk-in coolers. Within a few weeks of having my Masters degree in hand, I was asked by Dr. Craig Hoff and Dr. Henry Kowalski if I wanted to teach an undergraduate mechanical engineering course. Thank you for the opportunity, because my love for learning and new found love for teaching led me to applying to MSU’s PhD program. There are so many more people from Kettering that played a critical role and I want to name a few more. Thanks to Dr. Patrick Atkinson, Dr. Theresa Atkinson, Professor Reginald Bell, Dr. Susan Farhat, Dr. Gianfranco DiGiuseppe, Dr. Mary Gilliam, Dr. Ruben Hayrapetyan, Nick Keehn, Brenda Lemke, Dr. Massoud Tavakoli, and Dr. Ali Zand. Flint, Michigan has been my year round home since 2010. I feel as though I have been adopted by various people in this tight knit community. I wholeheartedly thank Dawn Hibbard for befriend- ing me and introducing me to the Rotary Club of Flint. Not only did I get the chance to meet the leaders and business owners of this tenacioius city, but I was invited to join their diverse group. At the time of this writing, we are in the midst a modern pandemic, ecological turmoil, and civil unrest. A last thank you for everyone that helped me to keep moving forward and focused during what had already been the most stressful time in my life. viii TABLE OF CONTENTS LIST OF TABLES . LIST OF FIGURES . CHAPTER 1 . . . . INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi CHAPTER 2 POWER SCALING OF RADIAL OUTFLOW: BERNOULLI PADS IN . . . . . . . . Introduction . . 2.1 2.2 Approach . EQUILIBRIUM . . . . . 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 2.2.1 Description of computations . . . . . . . . . . . . . . . . . . . . . . . . . 8 2.2.2 Validation using published results . . . . . . . . . . . . . . . . . . . . . . Finding Òeq using interpolation . . . . . . . . . . . . . . . . . . . . . . . . 2.2.3 9 Fluid power measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.2.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 . 16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 Scaling based on wall shear . . . . . . . . . . . . . . . . . . . . . . . . . . 19 Scaling modified to account for a . . . . . . . . . . . . . . . . . . . . . . 21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.5.1 2.5.2 2.5.3 Corner Effects . . 2.3 Results . 2.4 Sample Calculations of Discretization Error 2.5 Discussion . . . . . . . . . . . . . . . . . . . . . 2.6 Concluding Remarks . . . . . . . . . . . . . . . . . . CHAPTER 3 UNDERWATER SHEAR-BASED GROOMING OF MARINE BIOFOUL- . . . . . . . . . . Introduction . 3.1 . 3.2 Materials and Methods 3.3 Experimental Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ING USING A NON-CONTACT BERNOULLI PAD . . . . . . . . . . . . 26 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 . 26 . 29 3.3.1 Test Site . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 3.3.2 Experimental Design - Year 1 . . . . . . . . . . . . . . . . . . . . . . . . 29 3.3.3 Experimental Design - Year 2 . . . . . . . . . . . . . . . . . . . . . . . . 31 3.3.4 Grooming Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 3.3.5 Data Acquisition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 3.3.6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 3.3.7 Chlorophyll a Processing . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 Preliminary Findings - Year 1 . . . . . . . . . . . . . . . . . . . . . . . . 39 Image Analysis - Year 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 . 44 3.5 Temperature, Light, and Salinity Data . . . . . . . . . . . . . . . . . . . . . . . . 48 3.6 Biofouling Organisms Observed . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 . 53 3.4.1 3.4.2 3.4.3 Chlorophyll a - Year 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Image Setup and Collection Image Processing . 3.4 Results . . 3.6.1 . . . . . . . . . ix . 3.7 Discussion . . 3.8 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 CHAPTER 4 . . . . . . . . . Imaging . Introduction . 4.1 . 4.2 Description of Experiment . . 3D PARTICLE TRACKING VELOCIMETRY OF RADIAL OUTFLOW BETWEEN TWO PARALLEL PLATES . . . . . . . . . . . . . . . . . . . 58 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 4.2.1 Bernoulli pad in water tank . . . . . . . . . . . . . . . . . . . . . . . . . . 58 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 4.2.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 4.2.3 Experimental Procedure . 63 . . . . . . . . . . . . . . . . . . . . . . . . 67 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 . 4.5.1 Ensemble Average . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 4.5.2 Azimuthal Average . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 Pressure Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 4.5.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Experimental Measurements 4.4 Computational Fluid Dynamics Model 4.5 Results . 4.6 Analysis . . 4.7 Concluding Remarks . . . . . . . . . . . . . . . . . CHAPTER 5 CONCLUSIONS . APPENDIX . . . . . BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 . 92 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x LIST OF TABLES Table 2.1: Sample discretization errors in pressure computations plotted in Figure 2.1 (a) for  = 4 mm,  = 40 mm, Ò = 0.4 mm and (cid:164) = 5.1 × 10−4 kg/s. . . . . . . . . 17 Table 2.2: Sample discretization errors in force computations plotted in Figure 2.1 (b) for  = 4 mm,  = 40 mm and (cid:164) = 5.1 × 10−4 kg/s. . . . . . . . . . . . . . . . . . 18 Table 2.3: Sample discretization errors in power computations at equilibrium plotted in Figure 2.8 for  = 4 mm and  = 40 mm. . . . . . . . . . . . . . . . . . . . . . 19 Table 3.1: Results of rmANOVA testing the effects of grooming on the concentration of chlorophyll a in fouling material growing on Garolite and Intersleek plates. This data and table was provided by Abhishek Naik. . . . . . . . . . . . . . . . 47 Table 3.2: Effects of grooming on chlorophyll a on fouled plates. Values represent Cohen’s d estimates of effect size, a standardized mean difference between chl a on control and groomed plates across 3 replicate plates each. Cohen’s d values < 0 imply decreased chl a on groomed plates. Cohen’s d > 0 implies increased chl a on groomed plates. This data and table was provided by Abhishek Naik. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 . . . . Table A.1: Detailed legend for the following tables displaying all photos used in image analysis. The field study was initiated on a Wednesday in order to consistently distribute the labor throughout the week. . . . . . . . . . . . . . . . . . . . . . . 93 . Table A.2: 3× Garolite. . Table A.3: 2× Garolite. . Table A.4: 1× Garolite. . . Table A.5: 0.5× Garolite. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 . Table A.6: Control Garolite. Table A.7: 3× Intersleek. . Table A.8: 2× Intersleek. . Table A.9: 1× Intersleek. . . Table A.10:0.5× Intersleek. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 xi Table A.11:Control Intersleek. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 xii LIST OF FIGURES Figure 1.1: (a) Parameters characterizing radial outflow between a Bernoulli pad and a proximal parallel surface. The embodiments (b)-(e) correspond to a work- piece being lifted, a workpiece being levitated, the pad being levitated, and the pad being lifted, respectively. The applied force  is the sum of all forces acting on the freely moving component (pad or surface) which are not produced by the fluid flow between the pad and surface. The convention used here is that  > 0 ( < 0) attempts to increase (decrease) the gap between the pad and the proximal surface. . . . . . . . . . . . . . . . . . . . . . . . . . 1 Figure 2.1: Comparison of CFD simulation results with experimental results in [1]: (a) pressure  as a function of radius  for  = 4 mm,  = 40 mm, Ò = 0.4 mm and (cid:164) = 5.1 × 10−4 kg/s, analytical results available for  = 0 mm [1] are also presented (b) fluid force on workpiece  as a function of gap height Ò for a mass flow rate of (cid:164) = 5.1 × 10−4 kg/s. Stable and unstable equilibrium points are found at Ò = 0.29 mm and Ò = 5.1 mm respectively. The insets are provided to indicate the magnitude of the error bars where it would not be visible otherwise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 Figure 2.2: Example of interpolation to find equilibrium gap height Òeq for zero and a non-zero value of the applied force a. . . . . . . . . . . . . . . . . . . . . . . 10 Figure 2.3: Fluid power (cid:164)m,in required to maintain an equilibrium height Òeq for a = 0. The nominal dataset is  = 4 mm,  = 40 mm, s = 0.04 mm, based on air as the working fluid; each other dataset varies one parameter from this baseline. The dotted line is the best fit to the nominal dataset. . . . . . . . . . . . . . . . 12 Figure 2.4: Ratio of inlet fluid pressure to inlet momentum as a function of equilibrium height Òeq for a = 0. The nominal dataset is  = 4 mm,  = 40 mm, s = 0.04 mm, based on air as the working fluid; each other dataset varies one parameter from this baseline. The dotted line is the best fit to the nominal dataset. 13 Figure 2.5: Non-dimensional fluid power (cid:164)∗ required to maintain a given non-dimensional equilibrium height Ò∗. The nominal dataset is  = 4 mm,  = 40 mm, s = 0.04 mm, based on air as the working fluid; each other dataset varies one parameter from this baseline. The dotted line is the best fit to all data. . . . 14 Figure 2.6: Non-dimensional fluid power (cid:164)∗ required to maintain a given non-dimensional equilibrium height Ò∗ for different values of ∗ when the force correction term in Eq.(2.5) is not used. The nominal dataset is  = 4 mm,  = 40 mm, s = 0.04 mm, using air as the working fluid. The dotted line is the best fit to the ∗ = 0 dataset. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 . xiii Figure 2.7: Non-dimensional fluid power (cid:164)∗corr required to maintain a given non-dimensional equilibrium height Ò∗ for different values of ∗ when the force correction term in Eq.(2.5) is used with  = 2/5. The nominal dataset is  = 4 mm,  = 40 mm, s = 0.04 mm, using air as the working fluid. The dotted line is the best fit to the ∗ = 0 dataset. Points associated with ∗ = +6 at Ò∗ > 0.25 are not visible in the present spread. . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 Figure 2.8: Fluid power (cid:164)m,in required to maintain an equilibrium height Òeq for the nominal data set:  = 4 mm,  = 40 mm, s = 0.04 mm, and air as the working fluid with a = 0. The GCI errors bars are included with sample calculations shown in Table 2.3. . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Figure 2.9: Illustration of corner geometries: (a)  = 0, (b)  > 0, (c)  < 0. Comparison of CFD simulation results with experimental results in [1]. . . . . (d) . 23 Figure 2.10: The pressure contour plots and pressure profile of a Bernoulli pad with a mass flow rate of 2 kg/s pulled 5 mm away from the wall; (a) shows the simple geometry with a 90 degree corner generating 21 N of suction and (b) shows a pintle physically turning the axial flow to radial flow generating 45 N of suction. 24 Figure 2.11: The velocity contour plots and resulting wall shear profiles of a Bernoulli pad with a mass flow rate of 2 kg/s pulled 5 mm away from the wall; (a) shows the simple geometry with a 90 degree corner generating a larger peak wall shear stress sooner and (b) shows a pintle generating a lower peak wall shear stress at a larger radial location. . . . . . . . . . . . . . . . . . . . . . . . . . . 24 Figure 3.1: (a) Half of the cross-sectional view of radial outflow between a Bernoulli pad and a proximal parallel surface such as a hull with a magnified gap, (b) and (c) normal and tangential stresses on the vessel hull generated by a pad of dimensions  = 200 mm and  = 31.75 mm, operating at Òeq = 0.91 mm resulting from an inlet jet velocity of 1.9 m/s. . . . . . . . . . . . . . . . . . . 27 Figure 3.2: The shear stress range generated by the Bernoulli pad as implemented in this study, shown in the context of literature data. Critical shear stress (Pa) required for biofouling removal are shown as a function of time for biofilm or biofouling development prior to exposure to shear. With the flow rate generating a maximum wall shear stress of 5000 Pa and a minimum of 0 Pa on clean plates, significant biofouling mitigation is expected between the minimum and maximum (2 days to 2 weeks) time elapsed between all frequencies of grooming. This is shown with a dark grey box overlay. Previously published and adapted with permission from both the author and publisher (Menesses et al. 2017). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 Figure 3.3: An aerial view of the Naval Undersea Warfare Center (NUWC) Division Newport where both Year 1 and Year 2 studies were conducted. . . . . . . . . . 30 xiv Figure 3.4: (a) Pier at Stillwater Basin showing the suspended frames with plates sub- merged under water. (b) A frame hauled up to remove a plate scheduled for grooming. (c) Using a tote filled partially with ocean water to keep plate submerged while in transit to and from the grooming tank. (d) Suspended test design used in Year 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 Figure 3.5: Suspended test design used in Year 2, shown here in two rows for ease of illustration. The water level remained at least 0.3 meters above the top of the grooming frame during the lowest tide. . . . . . . . . . . . . . . . . . . . . . . 32 Figure 3.6: Grooming operation on a set of 4 Intersleek plates flush-mounted in the Garolite grooming frame; the clear acrylic Bernoulli pad is difficult to discern and is therefore outlined. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 Figure 3.7: Picture of the sheltered space taken with the Canon EOS Digital Rebel XT used for data acquisition, which is why there is no camera fixed into the mount. The photographing platform for the assembled grooming frame is directly underneath the point of view of the camera. The middle cutout was to provide addition contact prevention on the plates. The accent light is currently directed to the tank during the water exchange before scheduled grooming. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 . . . . . . Figure 3.8: The sampling device was sanitized in between plate sampling with isopropyl alcohol. (a) A plate slides into a slot designed into the bottom piece. Contact on the front plate was avoided by holding only the permanently installed tab. The top piece was carefully set down on the plate with the well containing the single o-ring targeted on the area not yet sampled. Engaging the clamp compresses the o-ring to create a leak-proof seal. The plastic top piece was constrained from ever coming into contact with the plate by resting completely on the flat face of the bottom piece of the sampling device. (b) 1 mL of 0.2 µm-filtered seawater was added to the well. (c) An overhead view in to show the water sitting in the sealed well. (d) The biofouling in the well was scraped off with a sterile plastic pestle before collected with a pipette. . . . . . . . . . . 36 Figure 3.9: The top row shows raw images of Garolite and Intersleek plates before and after grooming. The middle row shows computer generated images of Garolite and Intersleek plates before and after grooming. The pixels of the computer- generated images have colors that correspond to the entries of +; the thin black contour indicates entries of + = 4.5. The bottom row shows the measure of cleanliness () for different values of ; as expected, the plates . . . . . . . . . . . . . . . . . . . . . . . . . . . are cleaner after grooming. . 39 xv Figure 3.10: (a) The view from above during static grooming levitating the Garolite plate. (b) Select results of static grooming of Garolite plates in Year 1. The plates are shown before and after grooming in weeks 3 and 5 after they were submerged, for four different grooming frequencies. . . . . . . . . . . . . . . . . . . . . . . 40 Figure 3.11: Select results of scanned grooming of Garolite plates in Year 1. The plates are shown before and after grooming in weeks 1, 3 and 5 after they were submerged, for four different grooming frequencies. . . . . . . . . . . . . . . . 41 Figure 3.12: “Cleanliness” measured using () with  = 4.5 for (a) Garolite and (b) Intersleek plates for four different grooming frequencies over a period of seven weeks. The terms “before” and “after” indicate the pair of results obtained before and after grooming on a particular day. The points linked by the dotted lines in (a) correspond to the Garolite plates shown in Figure 3.9 (a). Similarly, the points linked by the dotted lines in (b) correspond to the Intersleek plates shown in Figure 3.9 (b). . . . . . . . . . . . . . . . . . . . . . 43 Figure 3.13: Chlorophyll a concentrations (µg cm−2) at 5 time points post-deployment. (a) Garolite plates; (b) Intersleek plates. For the Garolite 0.5×/wk groomed plates, samples were not collected on day 37 due to hard fouling growth. This data and figure was provided by Abhishek Naik. . . . . . . . . . . . . . . . . . 45 Figure 3.14: Effects of grooming on Garolite and Intersleek plates. Each point is a Cohen’s d estimate representing a standardized mean difference between chl a on control and groomed plates across 3 replicate plates each. Each estimate summarizes the effect size for one plate type, grooming frequency, and day. A Cohen’s d < 0 implies decreased chl a on groomed plates. A Cohen’s d > 0 implies increased chl a on groomed plates. For the 0.5×/wk groomed plates, samples were not collected on day 37 due to hard fouling growth. For Intersleek plates, Cohen’s d values could not be calculated for the day 9 time points due to negligible averages and zero standard deviations among chl a concentrations of groomed plates and control plates. The red lines indicate large Cohen’s d effect size of 0.8. This data and figure was provided by Abhishek Naik. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 Figure 3.15: The temperature and light intensity was logged in 15 minute intervals locally with a HOBO pendant sensor that was mounted on one of the submersible frames. The raw data points for temperature were plotted individually, with a moving mean overlay. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 Figure 3.16: The water quality stations from the T-Warf located at the southern point of Prudence Island represent the open bay water near the field test site. Water salinity data was collected at both surface and bottom stations every 15 minutes using YSI 6600 V2 and EXO2 data loggers. Each data point is plotted for the duration of the study along with a moving mean. . . . . . . . . . 49 xvi Figure 3.17: The front (top) and back (bottom) of the Garolite control plates at day 43 of the field test study. The box shows a close up picture of the area when submerged to better show the mussel that is attached. The top also shows circular areas that were used for sampling for chl a. Since the front of the plates had a significant amount of biomass by day 35, the sampling device could not maintain a seal for the circular well even if there were no hard fouling in the way. The samples for that day was taken with a sterilized cell lifter, the most noticeable resulting patch being on the front of the second plate from the left. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 Figure 3.18: The back of the control set of Garolite plates installed in its submersible frame is shown, revealing a variety of macrofouling at day 51 which is after the end of the 7 week study. The last extensive scraping of the back happened 2 weeks prior. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 . . . . . . . . Figure 3.19: Photos taken of (a) the back of the grooming frame after a group of 4 duplicate plates are installed, and (b) the front. A paper note was used to document the details and time of what was photographed. . . . . . . . . . . . . . . . . . . . . 54 Figure 4.1: The Bernoulli pad assembled with a plenum and flow straighteners. . . . . . . . 59 Figure 4.2: The BPA is mounted on a beam with roller wheels that allow the pad to be placed in close proximity to the tank wall. Double arrows indicate the direc- tion of movement of the BPA permitted by the roller wheels; this movement allows the pad to find its equilibrium configuration. . . . . . . . . . . . . . . . 59 Figure 4.3: A top view of the experimental setup showing the imaging hardware com- prised of four ultra high-speed cameras and a dual-head pulsing laser. The cameras are placed in an arc-like arrangement and focused on a volume that lies between the pad and the tank wall. The dual-head laser is located in the middle of the four cameras. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 Figure 4.4: (a) CAD model illustrating where the laser illumination is directed to on the BPA. Defined edges indicate where the image was masked in the flow visualization software. (b) The Bernoulli pad showing the matte surface finish and three equally spaced nylon-tipped set screws installed. These set screws help avoiding perturbations of the uniform equilibrium gap between the pad and the tank wall. (c) Side view of the laser illuminating the area of interest in the gap and partially into the jet outlet. . . . . . . . . . . . . . . . . 62 Figure 4.5: A single representative image at time T (taken from the set of 100,000 images) from the four cameras is shown. . . . . . . . . . . . . . . . . . . . . . . . . . . 64 Figure 4.6: Overlay of all 100,000 images are shown for each camera. These images (left to right) correspond to the cameras (left to right) shown in figure 4.3. . . . . . . 64 xvii Figure 4.7: a) Straight-on view of measurement volume that includes a small portion of the jet. b) and c) Tracks of the particles (Lagrangian) visible in the four images in figure 2.6 over 8 additional frames (equivalent to 400s). d) Vector fields (Eulerian) of the particles visible in the four images in Figure 2.6. . . . . 66 Figure 4.8: The force generated at different gap heights and four different mass flow rates (left). The force versus mass flow rate at a gap height of 1.25 mm with equilibrium shown with a dashed line (right). The maximum velocity generated at this equilibrium point in the fluid flow field at 1.25 mm is 8.4 m/s, which is below the limits of the hardware. . . . . . . . . . . . . . . . . . . 67 Figure 4.9: Portion of the inlet is visible on the bottom left, and the intermediate outlet of the experiment window is at the top of these plots. (a) Raw z-component of velocity, (b) Raw theta-component of velocity, (c) Raw radial-component of velocity, (d) refined radial-component of velocity, e) radial-component of velocity from CFD run at the same boundary conditions. . . . . . . . . . . . . . 69 Figure 4.10: (a) Raw pressure, (b) refined pressure, c) Pressure from CFD run at the same boundary conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.11: An illustration to show the boundaries of the data after completing the az- . 71 . 72 . 72 imuthal average. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.12: Mass flow rate calculated for every radial location using raw and refined radial velocity data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.13: The experimental radial velocity at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized PTV data is overlaid on the original unstructured PTV data. . . . . . . . . . . . . . . . . . . 73 Figure 4.14: The CFD radial velocity at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized CFD is overlaid on the original unstructured PTV data. . . . . . . . . . . . . . . . . . . . . . . . . 74 Figure 4.15: The experimental axial velocity at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized PTV data is overlaid on the original unstructured PTV data. . . . . . . . . . . . . . . . . . . 75 Figure 4.16: The CFD axial velocity at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized CFD is overlaid on the original unstructured PTV data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 xviii Figure 4.17: The experimental discretized and tangential velocity at different axial loca- tions within the fluid gap (/Òeq), where zero is at the glass wall. The dis- cretized PTV data is overlaid on the original unstructured PTV undiscretized data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 . . . . . . . . . Figure 4.18: The experimental pressure at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized PTV data is overlaid on the original unstructured PTV data. . . . . . . . . . . . . . . . . . . . . . . 78 Figure 4.19: The CFD pressure at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized CFD is overlaid on the original unstructured PTV data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 Figure 4.20: The discretized pressure curve along the bottom of the glass overlaid over all the unstructured pressure data in the entire height of the gap. The vertical dashed lines indicate the original data region from 10.8 mm and 64.4 mm. . . . 81 Figure 4.21: The discretized pressure curve along the bottom of the glass with an offset of +0.82 kPa overlaid over all the unstructured pressure data in the entire height of the gap. The vertical dashed lines indicate the original data region from 10.8 mm and 64.4 mm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 Figure 4.22: The discretized pressure curve along the bottom of the glass with an overshoot of +0.82 kPa overlaid over all the unstructured pressure data in the entire height of the gap. The vertical dashed lines indicate the original data region from 10.8 mm and 64.4 mm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 Figure 4.23: The cumulative summation of the force along the glass. Pressure at a radial location was multiplied with the annular surface area. The vertical dashed lines indicate the original data region from 10.8 mm and 64.4 mm. . . . . . . . 82 Figure 4.24: Normalized Radial velocity for multiple flow rates of water with the pad dimensions used in the PTV experiment. The velocity near the glass wall (top) shows no negative flow due to being between the separation bubble and glass, where the velocity near the pad (bottom) does capture the recirculation that occurs in the separation bubble. . . . . . . . . . . . . . . . . . . . . . . . 84 Figure 4.25: Normalized pressure for both PTV and CFD data of the water Bernoulli pad based on equation 4.5. The air case is included with the previously published experimental data measured with a traversing pressure tap. . . . . . . . . . . . 85 xix Figure 5.1: Modeling a water Bernoulli pad ( = 31.75 mm,  = 200 mm) with a mass flow rate of 1.5 kg/s resulting in an equilibrium gap height of 0.9 mm. Only a portion of geometry is shown from  ≈ 10 mm to 50 mm. The top contour plot shows the radial velocity in a torrid-shaped fluid space, if the flow did not have axial flow to turn the corner. The bottom contour plot shows how the recirculation bubble (indicated by the region with negative radial velocity) constricts the turning flow to achieve a higher values of velocities near the corner. 90 Figure A.1: Illustration of the location of each of the 4 duplicate plates and how the group looked before and after scheduled grooming on that day in the proceeding tables. 94 Figure A.2: Picture taken of Garolite plates after grooming, with points hand selected to analyze the pixels within the resulting polygon. . . . . . . . . . . . . . . . . . . 95 Figure A.3: A close up of the generated polygon of the plate that was never subject to biological sampling. Drop down menus for grooming details were available on the top left. The coating type can be selected on the top middle. The "continue" button on the top right proceeds to the next full image to repeat the point selection process. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 xx CHAPTER 1 INTRODUCTION A confined radial outflow can be produced by forcing a jet of fluid out from the center of a pad towards a proximally located parallel surface - see Figure 1.1 (a). Per Livesey [2], interest in this flow field grew out of application to lubricating films [3] away from the Re (cid:28) 1 limit, as inertial effects become equal in importance to viscous terms. Despite the apparent simplicity of the flow, it is not amenable to exact solution of the Navier-Stokes equations. A perturbation expansion which achieves reasonable agreement sufficiently far from the jet has been found [4], but the development region near  = /2 precludes a solution over the entire footprint of the pad. This lack of tractability engendered a strong tradition of experimental work on the topic [5, 6], which has continued to the present day for classical [1, 7] and perturbed [8, 9] geometries. Because the flow geometry is relatively simple, simulations have also been used to effectively model the system [10, 11]. Figure 1.1: (a) Parameters characterizing radial outflow between a Bernoulli pad and a proximal parallel surface. The embodiments (b)-(e) correspond to a workpiece being lifted, a workpiece being levitated, the pad being levitated, and the pad being lifted, respectively. The applied force  is the sum of all forces acting on the freely moving component (pad or surface) which are not produced by the fluid flow between the pad and surface. The convention used here is that  > 0 ( < 0) attempts to increase (decrease) the gap between the pad and the proximal surface. 1 Devices exhibiting this flow field see widespread use in the semiconductor industry [12], where their ability to manipulate delicate components without contact is valuable. A rich patent literature focused on this application has also been developed [13, 14, 15]. Because these devices are able to exert force to grip soft and pliable workpieces, they have also been explored in medical [16], apparel [8], and meat processing [17] applications. The magnitude and direction of the normal force produced are strongly dependent upon the distance between the pad and the surface. The reason this force changes direction can be intuitively understood as follows. In the limit where the pad contacts the surface, the repulsive force is equal to the product of the jet’s area and feed pressure. In the limit of very large distances, the jet is unconfined, and the repulsive force is proportional to the jet momentum. An attractive force is produced at intermediate distances, which can be understood in the context of the Bernoulli equation. Although this framework is not quantitatively accurate owing to the presence of large shear forces, it is useful for illustration. In combination with mass conservation, the result (cid:34) 2in (cid:35)2 () = (0) −  2 8Ò (1.1) is readily found, where Ò represents the gap between the pad and the surface. For small values of the product Ò, a net attractive force is produced. Because the force is repulsive in the limits of large and small Ò, with attractive forces produced in the middle, there exist two equilibrium points where the net force is zero; one stable and the other unstable [1]. The fact that equilibrium points exist is intuitively obvious, but this appears to have been dis- cussed for the first time in [12] and has since received limited attention [1]. In various embodiments of the system (see Figure 1.1), this equilibrium height is produced by the balance between the ap- plied force and the force generated by the pressure field. The applied force a is the sum of all forces acting on the freely moving component (pad or surface) which are not produced by the fluid flow between the pad and the surface. It is assumed to be positive (negative) when it attempts to increase (decrease) the gap between the pad and the proximal parallel surface. The equilibrium height is important in a variety of applications. The second chapter of this dissertation utilizes a parametric study to find equilibrium points in order to find a scaling relationship of fluid power. 2 For example, the equilibrium height is important in levitated mobile robot applications [18] where a sufficiently prominent surface feature could make undesired contact with the robot. In material handling applications such as the one described in [19], point contacts with a levitated workpiece prevent the workpiece from inadvertent lateral sliding motion; a sufficiently small equilibrium height is therefore required to engage these contacts. This work is motivated by an application in which a mobile robot moves around the submerged hull of a stationary marine vessel to groom in order to maintain peak efficiency when underway. Microorganisms, plants, and animals responsible for biofouling are ubiquitous in all bodies of water and have impacted human activities as long as maritime and aquatic industries have existed. If an initially clean surface submerged in the marine environment is left undisturbed, bacteria readily form a biofilm on it, followed by other attaching microorganisms, with the various subsequently colonizing species contributing to formation of a complex biofouling community [20]. A vessel hull that is continuously immersed in seawater experiences biofouling pressure that depends on the season, location, and the duration for which it remains idle [21, 22]. The degree to which the hull becomes fouled from this pressure is also dependent upon its surface characteristics and maintenance protocols. In the absence of a strategy that prevents fouling, the hull experiences the full extent of the fouling pressure. The resulting biofouling that accumulates on the hull produces significantly more drag, increasing fuel consumption by as much as 50% and reducing the vessel’s top speed by as much as 10% [23, 24]. Universally deployed and efficient anti-fouling protection would save the world’s merchant fleet an estimated $180 billion in 2020 [21]. For the destroyer class ships of the US Navy alone, the overall cost associated with combating biofouling is around $56 million per year [25, 26]. Apart from the economic losses, biofouling also results in a higher carbon footprint [27] and facilitates the spread of invasive species [28, 29, 30, 31]. These impacts continue to drive interest in the development of effective biofouling removal technology and management plans. Surface coatings have been developed to prevent or reduce biofouling and the most effective coatings are biocidal. Such coatings based on tributyltin (TBT) or other organotin compounds 3 have been banned [32, 33], while coatings based on copper have been scrutinized [34, 35, 20] because of their environmental impact. With more than 1.1 million 2 of the US Navy’s hull area is coated with self-polishing copper coatings, it is important to consider both the positive and negative long term effects of a hull husbandry strategy [36]. To avoid harmful environmental impacts from leaching metal, non-biocidal anti-fouling coatings have come into use as a common way of mitigating the fouling pressure on vessel hulls. So-called fouling-release and slime-release coatings are currently the most commercially available option [37, 38]. These coatings are ideal for fast vessels that are frequently underway, which assists with the release of biofouling organisms that accumulated while the vessel was stationary [21]. Biofouling is not completely preventable using currently available surface treatments; therefore, supplementary cleaning or treatment is common even on coated surfaces. There are two different approaches to either treatments or cleaning, proactive and reactive [39]. The more established the biofouling, the more aggressive the reactive response needs to be, since the adhesive strength of biofouling organisms is exponential in nature as it physically localizes onto a surface [40] using different adhesion phases and a variety of adhesion mechanisms [41]. The heavier the fouling pres- sure, the more aggressive the proactive response needs to be to impede the dynamics of biofouling colonization. Fouling pressure often has a seasonality depending on location, therefore requiring different levels of responses at different times of the year. The following methods lie somewhere on the proactive and reactive spectrum: mechanical grooming [42, 43], continuous aeration curtains [44, 40], ultraviolet (UV) light [45], lasers [46], steam [47], and ultrasonic sound waves that create resonance on the hull [48, 49, 50, 51]. Even if a method is successful at combating biofouling, there is the potential to degrade the efficacy of the anti-fouling properties of the surface coating over extended periods of time, depending on the frequency and aggressiveness of the treatment. Less aggressive and less abrasive cleaning, even if conducted more frequently, reduces damage to surface coatings [42, 52]. This gentler approach, required to prolong the life of current anti-fouling coatings, is typically referred to as “grooming” instead of “cleaning”. Another consequence of biofouling is the corrosion products created in a biofilm, which can accelerate the corrosion already 4 occurring or even induce biocorrosion in materials and surfaces that are typically expected to not experience corrosion such as stainless steel [53]. Technical improvements to the grooming process can also be made to further reduce damage to the coating. A recently-commercialized system [54, 55] for improving coating longevity uses a rotating brush co-developed with a self-polishing coating. A one-year study performed with immersed water jets shows that contactless cleanings done at a frequency of 0.5 × /month and 1 × /month was sufficient to maintain both anti-fouling and foul-release coatings without damage [56]. The grooming method described in the present work is produced entirely without surface contact. It is anticipated to improve hull performance while extending the life cycle of the coating relative to more abrasive cleaning methods. The focus of third chapter is to present field test study results of a biofouling mitigation device that produces both wall shear and contactless grip using a unique radial jet flow. The potential of the method was recently demonstrated computationally in the second chapter. The wall shear is used to dislodge nascent fouling communities, while the contactless grip keeps the device levitating near the hull. A test protocol is presented that was used to establish the efficacy of the grooming device, along with site details, testing procedures, image acquisition and biological surveys per- formed. The results of these field tests from two field seasons are presented and discussed, along with an image-processing algorithm used to objectively discern the level of cleanliness of surfaces subjected to biofouling pressure. Comparison between Intersleek®-coated and uncoated Garolite plates using both image-analysis and chlorophyll a measurements bounds the grooming frequency requirements. Finally, the potential of the described grooming approach and some directions for future study/development are presented. For the fourth chapter, the pad dimensions from the field test study in operation with the sub- mersible pump is the basis of an experiment to capture the unique fluid flow field in equilibrium. Using the latest hardware and software, along with techniques and algorithms in fluid flow field visualization, the Lagrangian tracks captured by using tracer particles can be converted to a vec- tor field on a regular grid for more convenient analyses [57]. Shake-the-Box is a time-resolved 5 Particle Tracking Velocimetry (4D PTV) that maps more densely seeded flows when compared to Particle Image Velocimetry (PIV) by incorporating as much as possible the temporal and spacial information available [58]. Multiple cameras synced to a laser volume can accurately achieve a 3D representation of the measured data after successful volume self-calibration (VSC) [59, 60, 61] enhanced by an Optical Transfer Function (OTF) to account for voxel pulse spreading to reduce the occurrence of ghost particles [60, 58]. STB leverages Iterative particle reconstruction (IPR) to find and triangulate a particle in 3D space by ’shaking’ the particle in space, where in every iteration the particle track is linearly extrapolated outwards to predict particle positions. From densely seeded flow, only some of the particles will be successfully tracked in the first iteration, leaving residual particles found from back-projection of the already tracked particles to repeat the ’shaking’ and extrapolation [62]. Processing the particle data using FlowFit optimizes the velocity field and consequently the pressure field, while Vortex-in-Cell-plus (VIC+) optimizes the vorticity field, velocity boundary, and accelaration boundary. The VIC+ method further achieves an accurate and ghostless reconstruction, but is more computationally intensive when compared to the approach of tomographic PIV [63]. Increasing the spatial resolution by a factor of 2 also doubles the grid nodes in each direction (23 = 8), thereby the total computation time increases in proportion. To optimize the processing while avoiding error at the computation boundary, VIC+ utilizes a multi-grid ap- proximation that is popular in the PIV community by providing an approximate solution quickly with a coarse grid before converging with every sequentially finer grid [63]. Since microfouling and macrofouling can occur on any submerged surface, the broader appli- cations of a contactless grooming device are discussed in the fifth and concluding chapter. The highest environmental and economic impacts from biofouling occur to the maritime vessels; which includes industry, private, and military. The proposed device can work in both saltwater and freshwater to maintain an initially clean and smooth surface of a vessel hull by utlizing the shear forces created by the confined radial outflow. Future work in terms of experiments and prototypes to increase the peak wall shear stress, and therefore the cleaning ability, is also discussed in this final chapter. 6 POWER SCALING OF RADIAL OUTFLOW: BERNOULLI PADS IN EQUILIBRIUM CHAPTER 2 2.1 Introduction This chapter is in large part from what became an article published in Journal of Fluids [64]. For a mobile robot-based grooming strategy to work, the equilibrium distance Òeq must be large enough to avoid interference with surface features on the hull, such as rivets. To remain at this equilibrium distance, some amount of power must be expended to accelerate ambient fluid to the jet velocity in and to overcome losses in the system. The present computational study indicates that Òeq and this required power (cid:164)m,in are closely related, with the latter varying over several orders of magnitude. When both quantities are appropriately scaled, a power-law describes the relationship between (cid:164)m,in and Òeq. A derivation for this power law is provided, which indicates that the work added to the system is primarily needed to overcome shear at the wall, and that the force applied to the system can be accounted for using a term incorporating the inlet Reynolds number. This relationship is valid over a range of applied forces, geometric and material parameters. 2.2 Approach 2.2.1 Description of computations Computations validated by published experimental results have been employed in the present work. This approach was taken because of the large parameter space being investigated. The parameter space produces more than 200 equilibrium points, each of which is obtained by the interpolation of multiple non-equilibrium configurations. Analytical models are not reliable near the jet exit [4], making them a poor choice for comparing different geometries. The numerical results in this chapter were computed using ANSYS-Fluent. To reduce run time, a 2D axisymmetric geometry was used. The flow conditions were steady-state. A four-equation Transition-SST model, based 7 on the k- transport equations coupled with intermittency and transition transport equations, was used to model turbulence. Pressure-velocity coupling was performed via the SIMPLE algorithm. Second-order spatial discretization was used for pressure, density, momentum, energy, dissipation rate and intermittency. The turbulent intensity at the inlet was fixed at 5%. The turbulent viscosity ratio was limited to 10 during solution. The parameters in the validation dataset [1] were chosen to avoid local compressibility effects anywhere in the flow field. Incompressible and compressible solvers had similar solutions to that set of parameters. For all of the data points presented in the present work, a compressible solver was employed to preserve a consistent working method. Supersonic and transonic velocities were not observed in any of the data presented in the present work. A roughness value characteristic of commercial steel (p = 0.04 mm) was used for the pad throughout this study. This roughness value was also used for the surface (s = 0.04 mm) for most of the data sets. Two sets were run using roughness values characteristic of smooth plastic (s = 0.007 mm) and galvanized iron (s = 0.15 mm) to determine the effect of surface roughness. Air was used as the working fluid for each case, save for a single set where 2 was used to confirm an interesting non-dimensional feature of the system. 2.2.2 Validation using published results A comparison between measurements of surface pressure and forces due to fluid flow from a set of experiments in [1, 7] and our CFD calculations is presented in Figure 2.1. The experimental apparatus in [1, 7] used a sliding tap to acquire high-resolution pressure data on a plate opposite the Bernoulli pad. The radial pressure distribution and total force were measured over gap heights ranging from 0.21 mm to 1.40 mm. The flow velocities used in the experiment were selected to avoid compressibility effects within the domain. The set of parameters being used in both the experiment and simulation are  = 4 mm,  = 40 mm, p = s = 0.04 mm, and air as the working fluid. These values represent the nominal parameter set in the present work. The variation of surface pressure on the workpiece with radial distance is compared in Figure 2.1 (a) for a fixed gap 8 height. The variation of fluid force on the workpiece as a function of gap height is compared in Figure 2.1 (b) for a constant mass flow rate. Both figures show error bars calculated from the fine grid convergence index (GCI) [65]; sample discretization error calculations are provided in Tables 2.1 and 2.2. As shown in Figures 2.1 (a) and (b), the simulations show a reasonable agreement with experiment with a small grid convergence error. The discrepancy with experimental data may be attributable to small uncatalogued differences in inflow development or geometry [66]. Figure 2.1 (a) also presents the single data point for the pressure at  = 0 mm that was derived analytically [1]; the analytical approach oversimplifies the complex flow behavior and under-predicts the pressure, making it an inappropriate choice for predicting the fluid power. Figure 2.1: Comparison of CFD simulation results with experimental results in [1]: (a) pressure  as a function of radius  for  = 4 mm,  = 40 mm, Ò = 0.4 mm and (cid:164) = 5.1×10−4 kg/s, analytical results available for  = 0 mm [1] are also presented (b) fluid force on workpiece  as a function of gap height Ò for a mass flow rate of (cid:164) = 5.1 × 10−4 kg/s. Stable and unstable equilibrium points are found at Ò = 0.29 mm and Ò = 5.1 mm respectively. The insets are provided to indicate the magnitude of the error bars where it would not be visible otherwise. 2.2.3 Finding Òeq using interpolation As shown in Figure 2.1 (b), and consistent with the literature [6, 12, 7], the force produced by a Bernoulli pad for a given mass flow is dictated largely by the gap height Ò. In the present study, this dimension is considered to be a dependent parameter, rather than an independent one, as is typically 9 done. This is consistent with the contactless nature of these devices, as commonly deployed; a high-pressure source is used to provide the fluid power, and the height of the gap is determined by the applied force a. To determine the equilibrium point, a geometry with a specific gap height Ò was drawn and meshed, then multiple mass flow rates were simulated. This gap height will be the equilibrium gap height Òeq for some value of the inlet mass flow rate (cid:164), at which the total force is equal to zero, such that a +  = 0,  = 2 (0, )  Þ /2 0 Each computational geometry (including the gap height Ò) is an equilibrium configuration for some value of inlet mass flow rate (cid:164), which is found by survey and interpolation as illustrated in Figure 2.2. This permits the equilibrium configurations to be found with much less user input and repeated meshing than would be required by the alternative, in which the inlet conditions are fixed and Ò is varied until Òeq is found. Figure 2.2: Example of interpolation to find equilibrium gap height Òeq for zero and a non-zero value of the applied force a. The error associated with this interpolation process was estimated as follows. The interpolation process at a given gap height Ò predicts a mass flow rate (cid:164) at which Ò = Òeq. The simulations 10 are then re-run at that predicted (cid:164) and the actual force produced by the pad is compared to the predicted value  = 0. For the nominal parameter set, the maximum normalized error associated with interpolation of the non-dimensional force ∗ =  (1/2) in2 in (2)/4 was 2.32 × 10−3; this is three orders of magnitude smaller than the values of the non-dimensional force investigated. 2.2.4 Fluid power measurement It is common in treatments of this radial flow field to produce the relationship between force generated by the flow in the gap and the gap height as a figure of merit. A different metric is required at equilibrium because the total force is zero. A practical choice is power, which is easy to measure and is relevant for engineering applications. The fluid power delivered to the inlet based on mass, has been employed instead of making assumptions about a prime mover or internal energy. For the geometry presented in Figure 1.1, the fluid power at the defined control surface, is given by the expression (cid:164)m,in = 2 ()  (2.1) Þ /2 0 (cid:20)1 2  [()]2 + () (cid:21) This fluid power is either lost or is convected out of the domain. The “stem” of the device, from the inlet to the corner, has been kept short to minimize shear losses in that region. In Eq.(2.1), the ratio of the two bracketed terms forms the following non-dimensional “fluid power ratio”:  = 2 2 (2.2) This is the ratio of fluid power needed to overcome losses and that needed to add kinetic energy to the fluid. This quantity also appears in [6], less the constant multiplier, where it is referred to as a non-dimensional plenum pressure. 11 2.3 Results Figure 2.3 provides the relationship between the equilibrium gap height Òeq and the required inlet power (cid:164)m,in for a = 0, as calculated in Eq.(2.1). The data for a ࣔ 0 will be discussed separately for better visualization. The nominal set of parameters is:  = 4 mm,  = 40 mm, p = s = 0.04 mm, and air as the working fluid. These parameters are the same as those used for validation in section 2.2.2. Each other dataset varies these parameters one at a time. A linear least squares method was used to find the best fit to the nominal dataset. The variation in required fluid power is observed to be approximately six orders of magnitude over the entire corpus of data. An estimate of the grid convergence error in power is provided for the nominal geometry in Table 2.3. Figure 2.3: Fluid power (cid:164)m,in required to maintain an equilibrium height Òeq for a = 0. The nominal dataset is  = 4 mm,  = 40 mm, s = 0.04 mm, based on air as the working fluid; each other dataset varies one parameter from this baseline. The dotted line is the best fit to the nominal dataset. It is immediately clear that the power required at a given Òeq is most dramatically dependent upon the pad diameter, such that increasing  by a factor of four increases the required power by ≈ 3.5 orders of magnitude. In contrast, the inner diameter  and surface roughness s are only 12 marginally influential. The relative importance of  indicates that the power requirements are dominated by the need to overcome losses ( in Eq.(2.1)), rather than by the need to add kinetic energy to the fluid (2/2 in Eq.(2.1)). This may be visualized in Figure 2.4, in which the ratio of these terms ( in Eq.(2.2)) is displayed as a function of Òeq. Figure 2.4: Ratio of inlet fluid pressure to inlet momentum as a function of equilibrium height Òeq for a = 0. The nominal dataset is  = 4 mm,  = 40 mm, s = 0.04 mm, based on air as the working fluid; each other dataset varies one parameter from this baseline. The dotted line is the best fit to the nominal dataset. The nominal dataset in Figure 2.4 is the same as that in Figure 2.3 ( = 4 mm,  = 40 mm, s = 0.04 mm, air), and the displayed dotted line is the best linear least squares fit to this dataset. The relative contribution of pressure drop to the power is large at small Òeq, but declines to parity with the kinetic energy at large Òeq. In contrast with Figure 2.3 (where (cid:164)m,in depends primarily on ), the power ratio  is primarily dependent upon the jet diameter  at a given Òeq. The variation in  at a fixed Òeq is considerably less than the variation in power (cid:164)m,in. Although not described as a ratio of relative power expenditures, the ratio  was also shown to have a strong dependence on  in [6]. There is also a set of scaling relationships which causes the (cid:164)m,in versus Òeq data to collapse to 13 a single trend. In Figure 2.5, the nondimensional power (cid:164)∗ is plotted against the non-dimensional equilibrium height Ò∗, where (cid:34) (cid:35) (cid:164)∗ = (cid:164)m,in in(/)  2 , Ò∗ = Òeq  (2.3) For the data presented in Figure 2.5, the applied force  = 0. The dotted line displayed in Figure 2.5 is the best linear least squares fit to all data. A derivation for this non-dimensional group, representing the measured fluid power at the inlet, is provided in section 2.5.1, with associated discussion. Figure 2.5: Non-dimensional fluid power (cid:164)∗ required to maintain a given non-dimensional equi- librium height Ò∗. The nominal dataset is  = 4 mm,  = 40 mm, s = 0.04 mm, based on air as the working fluid; each other dataset varies one parameter from this baseline. The dotted line is the best fit to all data. Figure 2.6 provides the relationship between Ò∗ and (cid:164)∗ for seven values of the non-dimensional external force ∗, defined below, when  ࣔ 0. For each case, the nominal geometry was employed. The non-dimensional force is defined as the ratio of the applied force and the inlet momentum, i.e. ∗ =  (1/2) in 2 in (2/4) 14 (2.4) Figure 2.6: Non-dimensional fluid power (cid:164)∗ required to maintain a given non-dimensional equi- librium height Ò∗ for different values of ∗ when the force correction term in Eq.(2.5) is not used. The nominal dataset is  = 4 mm,  = 40 mm, s = 0.04 mm, using air as the working fluid. The dotted line is the best fit to the ∗ = 0 dataset. It is clear from Figure 2.6 that the scaling relationship which collapses the data for the unloaded system does not do so when the system is loaded ( ࣔ 0). In particular, a dramatic spread is visible in the data for ∗ < 0 at large Ò∗. To account for the effect of the applied force, the non-dimensional fluid power is redefined as (cid:34) (cid:164)∗corr = (cid:164)m,in in(/) 8 ∗Rein  2 = (cid:164)∗ −  (cid:35)(cid:20)inin  (cid:21)(cid:34)2 (cid:35)  2  in(2/4) (1/2) in2 (cid:34) (cid:35) (cid:34)2 −  8 (cid:35)  2  (2.5) When the force correction is applied, the data collapse is largely recovered, with some spread for Ò∗ > 0.25 - see Figure 2.7. The empirical constant  = 2/5 was used to provide this fit. A derivation for this force correction, including and a proposed reason for the scaling breakdown at ∗ < 0 above Ò∗ ≈ 0.25, is provided in section 2.5.2. 15 (cid:164)∗corr required to maintain a given non-dimensional Figure 2.7: Non-dimensional fluid power equilibrium height Ò∗ for different values of ∗ when the force correction term in Eq.(2.5) is used with  = 2/5. The nominal dataset is  = 4 mm,  = 40 mm, s = 0.04 mm, using air as the working fluid. The dotted line is the best fit to the ∗ = 0 dataset. Points associated with ∗ = +6 at Ò∗ > 0.25 are not visible in the present spread. 2.4 Sample Calculations of Discretization Error Ansys Fluent Workbench 17.1 was utilized for all steps of the CFD modeling process. For determining grid independence, three different mesh sizes were used with high orthogonal quality and low orthogonal skew. The ratios between the fine and medium mesh elements and the medium and coarse mesh elements are denoted by 21 and 32. Mesh inflation was selected at the walls to capture the boundary effects. For the residual errors, the condition for iterative convergence was set to 1.0 × 10−5. The discretization error was computed using the Richardson extrapolation method described in [65]; the tables provided in this section are based on the nomenclature used in [65]. Tables 2.1 and 2.2 provide sample discretization errors for pressure and force computations used for comparison with experimental data - see section 2.2.2. For the pressure profile at a single gap height, 0.35 % of the total number of 570 points indicated 16 Table 2.1: Sample discretization errors in pressure computations plotted in Figure 2.1 (a) for  = 4 mm,  = 40 mm, Ò = 0.4 mm and (cid:164) = 5.1 × 10−4 kg/s.  (mm) 21 32 1 (Pa) 2 (Pa) 3 (Pa)  21 e (Pa) 21  21 e  21 f 10.101 1.34503 1.33196 0.03449 0.03423 0.03447 0.186 0.03232 5.068 1.34503 1.33196 -0.38839 -0.39636 -0.42685 4.735 -0.39403 19.852 0.000 1.34503 1.34503 1.33196 1.33196 0.00061 5.64860 0.00059 5.76250 0.00061 5.81960 0.168 2.177 0.00453 5.64884 2.016% 2.052% 0.753% 0.481% 3.641% 0.004% 1.431% 6.722% 1.496% 86.437% 2.17% 3.91% 15.134 1.34503 1.33196 0.02486 0.02474 0.02466 1.188 0.02524 2.21% 0.81% 0.52% Figure 2.8: Fluid power (cid:164)m,in required to maintain an equilibrium height Òeq for the nominal data set:  = 4 mm,  = 40 mm, s = 0.04 mm, and air as the working fluid with a = 0. The GCI errors bars are included with sample calculations shown in Table 2.3. 17 Table 2.2: Sample discretization errors in force computations plotted in Figure 2.1 (b) for  = 4 mm,  = 40 mm and (cid:164) = 5.1 × 10−4 kg/s. Ò (mm) 21 32 1 (N) 2 (N) 3 (N)  21 e (N) 21 a 21 e  21 f 0.24 0.27 0.4 1.32772 1.34405 0.03835 0.03369 0.03203 3.890 0.04066 1.34503 1.31687 1.33196 1.32476 -0.05752 0.11858 -0.05825 0.11200 -0.06103 0.11442 4.706 3.683 -0.05728 0.12232 5.55% 12.13% 1.27% 0.42% 3.06% 3.15% 0.64% 5.69% 6.56% 0.5 1.34745 1.34106 -0.06312 -0.06332 -0.06479 6.805 -0.06309 0.32% 0.05% 0.16% 0.9 1.35303 1.36832 -0.04372 -0.04376 -0.04405 6.531 -0.04372 0.08% 0.01% 0.04% 1.3 1.37792 1.37460 -0.03161 -0.03155 -0.03117 5.929 -0.03162 0.19% 0.03% 0.08% oscillatory convergence. The local order of accuracy  ranges from 0.013 to 19.7, with an average value of a = 2.6, which is used to calculate the GCI error bars in Figure 2.1 (a). In Table 2.1, the highest extrapolated error, 21 e, is relative to the pressure near the outer diameter, where the gage pressure goes to zero. Taking the integral of the pressure profile gives a value of the force generated by the fluid flow. When the mass flow rate is kept constant, the force is a function of gap height. The force plot had 38% of the 16 points exhibit oscillatory convergence. The local order of accuracy  ranges from 0.21 to 12.9, with an average value of a = 4.2, which is used to calculate the GCI error bars in Figure 2.1 (b). The main focus of the second chapter of this dissertation is to derive a scaling relationship for the fluid power and Table 2.3 provides sample discretization errors for power computations. For the nominal data set of the parametric study 65% of the 17 points indicated oscillatory convergence. The local order of accuracy  ranges from 0.60 to 18.9, with an average value of a = 8.3. There are no GCI error bars for the power plot in Figure 2.3 since the percent error is visibly negligible on the log scale used. The nominal data set in Figure 2.3, when replotted in Figure 2.8 using a regular scale, shows that the errors are minimal. 18 Table 2.3: Sample discretization errors in power computations at equilibrium plotted in Figure 2.8 for  = 4 mm and  = 40 mm. Ò (mm) 21 32 1 (W) 2 (W) 3 (W)  21 e (W) 21 a 21 e  21 f 0.24 1.31687 1.32476 13.4581 13.8643 14.5747 1.929 12.87804 3.02% 4.50% 0.50% 0.27 1.32772 1.34405 6.30099 6.07708 6.13503 4.940 6.37426 3.55% 1.15% 0.54% 0.4 1.34503 1.33196 0.75723 0.75714 0.73827 18.885 0.75723 0.01% 0.00% 0.00% 0.5 1.34745 1.34106 0.24350 0.23537 0.23296 3.990 0.24706 3.34% 1.44% 0.45% 0.9 1.35303 1.36832 0.00523 .00519 0.00461 8.112 0.00524 0.86% 0.08% 0.11% 1.3 1.37792 1.37460 0.00049 0.00049 0.00049 9.148 0.00049 0.04% 0.00% 0.00% 2.5 Discussion 2.5.1 Scaling based on wall shear Figure 2.4 indicates that the dominant contributor to the power required by the system at equilibrium is the pressure drop from the inlet to the outlet. The trends in Figure 2.3 show that the power required at a given Òeq is strongly dependent on . Therefore, a relationship based on the wall shear within the gap might reasonably be expected to collapse the data, and the non-dimensional groups in the first term of Eq.(2.3) can indeed be derived on this basis. Consider the expression below for (cid:164)s, the work expended to oppose the wall shear w: (cid:164)s = w  =   ∝   ࡁ  (2.6) where the terms ࡁ,  and  correspond to the characteristic length, area and velocity scales, respectively. Mass conservation between the inlet and the gap yields (cid:20)  (cid:21)  (cid:21)2 (cid:20)  2 2 Òeq  =  in (2.7) where  is a characteristic radius. Choosing  ∝  yields  ∝ (/Òeq)in. Since the total wall shear force should increase with decreasing gap height and increasing pad area, we choose ࡁ = Òeq 19 and  = 2. Substituting these characteristic scales into Eq.(2.6) yields (cid:34)  2 in2  (cid:35)(cid:20) Òeq (cid:21)−3  (cid:164)s ∝   in Ò2 eq 2  in Òeq ∝ (cid:20) Òeq (cid:21)−3  (2.8) ⇒ (cid:164)s in(/) ∝  2 These non-dimensional groups, found by a shear-based derivation are the same as those observed in Figure 2.5. Furthermore, the value of the derived exponent is a close approximation to the linear least squares best fit indicated in Figures 2.5-2.7. The range of the observed best fit exponent  was found to vary over the range 3.12 ≤  ≤ 3.24 depending on the subset of data being examined. The scaling relationship provided in Eq.(9) indicates that the scaling is dependent upon the dynamic viscosity . To test this, a set of simulations using 2 as the working fluid was performed. These data follow the same trend as the data using air when the power and gap height are made non-dimensional. A “cleaner" derivation may be apparent to the reader, in which the characteristic values of velocity, gradient, and area are instead taken at the generic location  such that  = in(/2)2 2 Òeq ,  ࡁ =  Òeq These may be integrated to form a characteristic shear work ,  =  (2.9) Þ 2 0  Òeq (cid:34)in(/2)2 2 Òeq (cid:35)2  (2.10) (cid:164)s = w  = Þ  (cid:20) Þ /2 (cid:21)−3 (cid:19)(cid:21)(cid:20) Òeq /2 (cid:18)  The result of this integration yields the characteristic work (cid:164)s ∝  2 in  ln   ⇒ (cid:164)s in  ln(/) ∝  2 (cid:20) Òeq (cid:21)−3  (2.11) Making the power data in Figure 2.3 non-dimensional using the non-dimensional groups in Eq.(2.11) does not collapse the data. A comparison between the two derivations makes it clear that the power required for a given Òeq is fixed by “worst-case” choices of characteristic parameters - largest velocity, largest gradient, and largest area - rather than by using averaged parameters or the values of the parameters at a fixed value of . 20 2.5.2 Scaling modified to account for a The scaling correction for non-zero force shown in Eq.(2.3) can be derived as follows. Consider the expression (cid:164)∗corr = (cid:164)m,in (2/) − 2  in(2/) 2 (2.12) (cid:34) where  is a characteristic velocity. The sign convention used in this work, where  < 0 attempts to decrease the gap between the pad and the proximal parallel surface, produces the sign of the force correction term  in Eq.(2.12). This can be understood physically as follows. If an applied force is helping to lift or levitate the free component at a given gap height Òeq, the flow field in the gap does not need to produce as much force to bring the free component into equilibrium. This smaller force is achieved by reducing the velocity of the working fluid, which reduces the fluid power required. Both the inlet kinetic energy and pressure are reduced, such that the dependence of  on ∗ is minimal relative to the overall trend shown in Figure 2.4. Because the fluid power required at equilibrium is less than that required by an unassisted workpiece, the force correction is positive for  < 0. The converse of these arguments holds for  > 0, where the applied force is attempting to increase the gap between the pad and the parallel surface. The force correction term in Eq.(2.12) can be manipulated to form four non-dimensional groups corresponding to a non-dimensional force, the inlet Reynolds number, an area ratio, and a velocity ratio: (cid:35)(cid:20)inin  (cid:21)(cid:34)2 (cid:35)(cid:20)c (cid:21)  2 in (2.13)  in(2/) = 2  8  in(2/4) (1/2) in2 In Figure 2.7, the characteristic velocity ratio c =  in where  = 2/5 was found to give a satisfactory fit to the data. It is not obvious a priori that this form of c is most appropriate - in particular, if c is a velocity within the gap, c ∝ in/Ò∗ rather than the form observed. Velocity ratios in this form did not successfully scale the data. It is notable that the scaling is less successful for ∗ > 0 above Ò∗ ≈ 0.25. Comparing the area in the gap at the corner point  = /2 to the inlet area yields the ratio g  in = Ò 2/4 = 4Ò∗ 21 meaning that for Ò∗ < 0.25, the flow geometry contracts at the corner, whereas at Ò∗ > 0.25, the geometry expands at the corner. This has several consequences on the ability of the pad to support or lift a load. First, because a supporting force is provided primarily by the central pressure near  = 0, the ∗ < 0 points are less problematic at large values of Ò∗. Second, because the area is expanding at the corner for Ò∗ > 0.25, the pressure gradient predicted by an inviscid analysis is positive at the corner; the recirculation bubble [66] prevents this from being strictly true, but it suffices to say that the first-order analysis of the problem no longer favors suction. This being the case, very large velocities (inlet Mach number greater than 0.5) are required to produce large amounts of lifting force at these values of Ò∗. As such, some of the deviation from the predicted scaling may be related to compressibility effects. 2.5.3 Corner Effects The pressure field near the centerline of the pad is quite sensitive to the geometry near the stem-pad corner, so I have provided several corner geometries for context in 2.9. Surface pressures produced by the nominal square corner geometry, a geometry with a radius at the corner ( > 0), and a reentrant geometry ( < 0) are provided. Industry favors the pintle design for the increase in load capacity [67] with no drawback to the reduced wall shear stress since cleaning is not a need. In regards to underwater grooming, the most effective grooming happens from the directly impinging flow as in Figures 2.10(a) and 2.11(a). Since the jet never directly impinges on the hull, there is no positive pressure generated as seen in Figures 2.10(b), but because of the distance it takes to turn the flow while keeping the initial radius of the stem the same as in the simple geometry causes the maximum velocity to happen at a larger radial location in the gap as seen in 2.11(b). 2.6 Concluding Remarks A computational study of Bernoulli pads has been performed, with the intention of determining the input power required to maintain stable equilibrium. At the stable equilibrium, some gap exists 22 Figure 2.9: Illustration of corner geometries: (a)  = 0, (b)  > 0, (c)  < 0. (d) Comparison of CFD simulation results with experimental results in [1]. between the pad and the workpiece; the size of this gap depends on the magnitude and direction of the applied forces as well as the pressure and mass flow at the inlet. The applied forces and inlet flow characteristics determine the inlet power required to operate the pad at equilibrium. The current work presents a set of scaling relationships that collapses the power requirements of a large number of pad geometries and operating conditions onto a single curve. This set of scaling relationships is derived initially for the case of zero applied force and extended to the case of non-zero applied force. The present work is the first to treat the inlet power as a parameter of interest, and the scaling relationships derived have not appeared in the literature. Over 200 realizations of equilibrium were used to derive these scaling relationships. This 23 Figure 2.10: The pressure contour plots and pressure profile of a Bernoulli pad with a mass flow rate of 2 kg/s pulled 5 mm away from the wall; (a) shows the simple geometry with a 90 degree corner generating 21 N of suction and (b) shows a pintle physically turning the axial flow to radial flow generating 45 N of suction. Figure 2.11: The velocity contour plots and resulting wall shear profiles of a Bernoulli pad with a mass flow rate of 2 kg/s pulled 5 mm away from the wall; (a) shows the simple geometry with a 90 degree corner generating a larger peak wall shear stress sooner and (b) shows a pintle generating a lower peak wall shear stress at a larger radial location. 24 large number or working geometries and conditions, coupled with the relatively simple geometry, makes it an ideal problem to study computationally rather than experimentally. Analytical models described in the literature do not accurately predict the inlet conditions, and therefore cannot be trusted to provide a good estimate of the inlet power. Two recommendations can be made to researchers wishing to perform computational studies of Bernoulli pads at equilibrium. First, because local velocities near the jet corner can be a large multiple of the inlet velocity, particularly for small gap heights, we recommend that compressible codes be used throughout any survey. Second, for a given pad geometry, it is better to create a single mesh for each Òeq and adjust the inlet conditions until equilibrium is reached. The alternate method, in which the inlet mass flow is fixed and the gap height is varied, requires a new mesh to be created for every gap height. The scheme we recommend significantly reduces the time required as it eliminates the need to repeatedly change the computational grid. The scaling relationships show that the primary contributor to the power requirement is related to the fluid losses in the system rather than the kinetic energy. Therefore, significant power reduction may be possible by reducing these losses. An analogy to pipe flow may be instructive, wherein the shear associated with the confined flow in the gap is equivalent to the “major losses”, and losses near the corner are the “minor losses”. Future work is envisioned to further investigate each of these types of loss in the flow field. Work is underway to more closely characterize the operation of the pad in equilibrium using water as the working fluid in the envisioned hull grooming application. 25 CHAPTER 3 UNDERWATER SHEAR-BASED GROOMING OF MARINE BIOFOULING USING A NON-CONTACT BERNOULLI PAD 3.1 Introduction The following chapter is the result of two summer internships through the Naval Research Enterprise Internship Program (NREIP) that allowed field tests to be conducted off the coast of Rhode Island. While biofilms and biofouling occurs from ubiquitous microorganisms in any body of water resulting in mitigation and repair costs to various industries, biofouling mitigation in an ocean environment was the focus in this research. 3.2 Materials and Methods The basis of the grooming device is a Bernoulli pad that uses an axial jet at its inlet to produce radial outflow between the pad and a proximally located parallel surface such as a hull. One-half of the cross-sectional area of this arrangement is shown in Figure 3.1a. As the flow moves from the jet region to the exit, the increasing area at each radius r causes the velocity to decrease. This flow field produces a force between the surfaces, which depends upon the gap height Ò. The total force can be calculated by taking the integral of the pressure curve shown in Figure 3.1b. The direction of this force changes sign, such that a stable equilibrium point exists for some spacing Ò = Òeq [1, 64]. The wall shear also changes radially, with the peak wall shear occurring beyond the stagnation area, when the pure axial flow turns into radial flow. A similar trend is also observed in unconfined radial outflow generated by an immersed water jet for hull-grooming [68]. The flow field in this work has potential use for hull-grooming applications for several reasons. First, the pad will remain levitating parallel to the hull without additional forces, given a consistent source of fluid power (Kamensky et al. 2019). Second, because the radial outflow is confined to a thin gap, it will reliably produce shear forces on the hull surface, which will remove biofouling 26 Figure 3.1: (a) Half of the cross-sectional view of radial outflow between a Bernoulli pad and a proximal parallel surface such as a hull with a magnified gap, (b) and (c) normal and tangential stresses on the vessel hull generated by a pad of dimensions  = 200 mm and  = 31.75 mm, operating at Òeq = 0.91 mm resulting from an inlet jet velocity of 1.9 m/s. organisms (Menesses et al. 2017). Finally, grooming performed using fluid shear does not require contact with the hull, alleviating potential damage to the coating. Figures 3.1b and 3.1c show the predicted normal and shear stresses on the vessel hull generated by the Bernoulli pad in Figure 3.1 of outer diameter  = 200 mm and inner diameter  = 31.75, operating at its equilibrium gap height of Òeq = 0.91 mm, achieved with an inlet jet velocity of 1.9 m/s [64]. Because the pad is in equilibrium, the integral of the normal stress over the total area covered by the pad is zero. The peak shear stress occurs near the jet exit and reaches a magnitude of 5 kPa; this is expected to be sufficient to remove recently attached bacteria, diatoms, and larvae using a reasonable grooming 27 frequency to prevent any organism from reaching maturity - see Figure 3.2 used with permission by Menesses et al. [40]. The wall shear stress is zero at the center of the jet and near the outer diameter of the pad. The shear forces produced by the Bernoulli pad result brackets a range of expected effectiveness when projected on natural marine biofouling organisms in Figure 3.2, based Figure 3.2: The shear stress range generated by the Bernoulli pad as implemented in this study, shown in the context of literature data. Critical shear stress (Pa) required for biofouling removal are shown as a function of time for biofilm or biofouling development prior to exposure to shear. With the flow rate generating a maximum wall shear stress of 5000 Pa and a minimum of 0 Pa on clean plates, significant biofouling mitigation is expected between the minimum and maximum (2 days to 2 weeks) time elapsed between all frequencies of grooming. This is shown with a dark grey box overlay. Previously published and adapted with permission from both the author and publisher (Menesses et al. 2017). 28 on published data on the organisms’ adhesion strength on various surface types shows the trend in adhesive strength of biofouling tends to be exponential in time. The expected produced shear force at the beginning of the study could not be maintained with the electrical input to the pump held constant, because the peak shear in the system will decrease with the increase in both the gap height and the thickness of biofouling. Other data on the adhesion strength of microfouling specific to foul-release coatings show that a shear stress of at least 300 Pa from an immersed water jet is sufficient to maintain a clean surface [41]. 3.3 Experimental Design 3.3.1 Test Site Field tests were conducted at the same location and time of the year as Menesses et al. 2017 [40]; off a pier at Stillwater Basin (latitude 41°32(cid:48)06.8(cid:48)(cid:48) N and longitude 71°18(cid:48)49.8(cid:48)(cid:48) W) in Narragansett Bay, Rhode Island, USA (see Figure 3.4). Both studies were completed within the most active season of biofouling (approximately from May to August). Bordered by the Atlantic Ocean, the study area is characterized by a moist continental climate. A stone breakwater shelters the pier where the study was conducted. The deployed plates were suspended such that they remained completely submerged even during low tide (Figures 3.4 and 3.5). For the second year, the average water temperature logged by a HOBO MX2202 pendant attached to one of the submersible frames during the study period (June 12 to July 29) was 211.7°C. For the same time period, the average salinity range from a nearby surface water quality monitor station (latitude 41°3442.10 N and longitude 71°1916.05 W) was 29.20.5 parts per thousand (ppt) [69]. More detail on the logged temperature, light intensity, and salinity data can be found in section 3.4. 3.3.2 Experimental Design - Year 1 During the summer of 2018, five plates (30.5 × 30.5 cm) of fiber-reinforced plastic, commercially known as Garolite, were fixed to PVC frames (Figure 3.4). When submerged, the PVC frames flooded with water and held the panels vertically facing the sun (Figure 3.4a). Four panels were 29 Figure 3.3: An aerial view of the Naval Undersea Warfare Center (NUWC) Division Newport where both Year 1 and Year 2 studies were conducted. subjected to different grooming frequencies established by the Center for Corrosion and Biofouling Control (CCBC) at Florida Institute of Technology (FIT): 3× per week, 2× per week, 1× per week and 0.5×/wk [43]. The fifth panel was the control, which was not groomed for the entire duration of the study. Based on the grooming schedule, individual plates were removed from their frames and transported to an acrylic tank in a laboratory for grooming; the acrylic tank was filled with 300 liters of dechlorinated tap water mixed with 10.9 kg of Instant Ocean® for a final salinity of approximately 30 parts per thousand (ppt). For an unobstructed view of the contactless grooming process, the Bernoulli pad was machined out of clear acrylic and fixed 2 cm below the water surface. The pad was connected to a pump with a maximum flow rate of 150 liters per minute (LPM). Each plate was held proximally below the pad and the pump was turned on. The fluid flow field then generated normal forces to levitate the plate and shear forces to groom without contact. The pump was operated for approximately one minute, resulting in “static grooming” with no relative motion between the plate and the pad. Before and after grooming, the plates were 30 Figure 3.4: (a) Pier at Stillwater Basin showing the suspended frames with plates submerged under water. (b) A frame hauled up to remove a plate scheduled for grooming. (c) Using a tote filled partially with ocean water to keep plate submerged while in transit to and from the grooming tank. (d) Suspended test design used in Year 1. transported between the pier and the acrylic tank in a plastic tub filled with ocean water. The results from this static grooming experiment were used to inform a second test during Year 1, in which the pad was "scanned" over the surface to be cleaned. These preliminary results of Year 1 test informed the design and test protocol of the Year 2 experiment. 3.3.3 Experimental Design - Year 2 In this study, we investigated the efficacy of combining our grooming method with a fouling-release paint treatment on the surfaces, similar to the protocol described in [42, 43] that were using different grooming technologies. To this end, two types of plate surfaces were compared: bare Garolite® 31 G101 plates in factory condition and Garolite plates painted on one side with Intersleek® 1100SR2 coating. These two surface types will be hereafter simply referred to as Garolite and Intersleek. Garolite was chosen for its resemblance to a smooth epoxy/fiberglass surface, and is expected to handle biofouling pressure worse than the Intersleek surface. A sponge and dish soap was used to eliminate any initial residue on the Garolite that occurred from the manufacturing process. Four plates of each surface type were exposed. A single plate was reserved for image analysis, and three dedicated replicate plates were included for biological sampling for every combination of surface type (Garolite and Intersleek) and grooming frequency (3×/wk, 2×/wk, 1×/wk, 0.5×/wk, and control, bringing the total number of plates that needed to be submerged to 40 (Figure 3.5). The Figure 3.5: Suspended test design used in Year 2, shown here in two rows for ease of illustration. The water level remained at least 0.3 meters above the top of the grooming frame during the lowest tide. 1This is an epoxy-grade industrial laminate and phenolic material manufactured by Accurate Plastics, Inc in Yonkers, NY. dam, Netherlands. 2This is a patented slime release fluoropolymer technology from AkzoNobel based in Amster- 32 grooming frequency protocol developed by CCBC [37] and previously used for year 1, was used again for year 2. Our 0.32 cm thick plates were chosen to have the dimensions of 10.2 by 10.2 cm to fit a custom biofouling sampling device previously used by University of Massachusetts Dartmouth (UMass-D). To mitigate the possibility of a shading bias from the pilings and pier structure, the locations of every set of plates were shifted every grooming day during the entire duration of the field test study. Shading was the most likely reason for less biofouling on the lowest grooming frequency plates as seen in Figure 3.10 from static grooming in Year 1. Since the back and edges of each plate did not experience any contactless grooming, biofouling did build up significantly to warrant removal every biological sampling day starting at week 3 (See Figures 3.15 and 3.16). 3.3.4 Grooming Procedure Grooming was carried out on the pier in a tank freshly filled with ocean water. The process of removal of the plates from the submersible rig to replacement in the rig did not exceed twenty minutes and the plates were never held out of ocean water for more than 30 seconds. The plates were transported between the rig and the grooming tank in a tote filled with clean ocean water, which was at a sufficient level to submerge the plates during transport. To simultaneously groom each set of 4 plates (that had the same grooming frequency and surface type), a neutrally buoyant Garolite frame was fabricated to flush-mount the plates (Figure 3.6). Only the tabs, edges and backsides experienced contact during the swap from the submersible frame to the grooming frame and back. The Bernoulli pad was coupled to a submersible pump with a maximum flow rate of 150 Liters/minute (LPM) and a self-excited flowmeter was placed in-line to measure the flow rate during operation. For an unobstructed view of the contactless grooming process, the Bernoulli pad was machined out of clear rigid acrylic with the dimensions previously modelled computationally (Figure 3.1). The water level was maintained to be 2 cm above the pad. When the pump was operated, the Garolite frame with a set of 4 plates automatically finds its equilibrium position under the pad and was manually moved laterally for scanned grooming for approximately one minute. The flow rate was chosen at the beginning of the study to be 33 the maximum with which no cavitation occurred with the plates in their pristine condition; this corresponded to a no-load flow rate of 84 LPM using a pulse width modulation (PWM) circuit to allow some form of flow rate adjustment of the pump. All grooming operations were performed by keeping the duty cycle of the PWM circuit fixed. As the study progressed this resulted in a gradual increase in the flow rate during grooming as there was an increase in the level of biofouling growth on the surfaces; this is due to the fact that heavier biofouling increases the drag and in turn the equilibrium gap height. The combination of the drag and larger gap height also decreases the maximum velocity, which correlates directly with maximum wall shear. The key to this proactive grooming is to stay ahead of biofouling pressure in order to maintain the peak wall shear stress. Consistency of the act of grooming was more crucial than the duration of shear exposure. Prolonging the scan by either multiple passes or slowing the lateral scanning movement is not expected to affect the amount of biofouling removed since the critical wall shear to biofouling removal is the threshold that is or is not reached with the constant electrical power supplied to the pump. Figure 3.6: Grooming operation on a set of 4 Intersleek plates flush-mounted in the Garolite grooming frame; the clear acrylic Bernoulli pad is difficult to discern and is therefore outlined. 34 3.3.5 Data Acquisition Pictures of the plates were taken before and after grooming using a Canon EOS Digital Rebel XT for the first year. The second year the Canon Rebel was mounted and accompanied with an Ikan Piatto PL90 Accent Light as seen in Figure 3.7. The mount provided more consistent straight on shots of the set of plates. The time that elapsed between pictures before and after grooming was no more than two minutes. At each grooming frequency, one plate was never sampled biologically. Pictures of this plate were fed into an image-processing algorithm, described in subsection 3.3.2. Chlorophyll a (chl a) concentration per unit surface area was used as a biological measure for grooming effect. Samples for chl a were collected once a week on the same day of the week for the duration of the experiment, on a day with no scheduled grooming. Triplicate plates were sampled per grooming frequency and plate type, along with triplicate non-groomed control plates. A custom Figure 3.7: Picture of the sheltered space taken with the Canon EOS Digital Rebel XT used for data acquisition, which is why there is no camera fixed into the mount. The photographing platform for the assembled grooming frame is directly underneath the point of view of the camera. The middle cutout was to provide addition contact prevention on the plates. The accent light is currently directed to the tank during the water exchange before scheduled grooming. 35 3D printed sampling device was used to collect chl a samples from the plate surfaces and can be seen in Figure 3.8. The complete assembly included two adjustable steel “Hold-Down Toggle Clamps” installed on the bottom piece and a single “Oil-Resistant Buna-N O-Rings” press fitted into one of the eight shallow cylindrical cavities. No additional o-rings were installed to prevent contact on any area that is not scheduled to be sampled. Only the edges and back experienced contact during the handling and sampling procedure. The device created leak-proof 16 mm diameter wells on the plate surface that could be quantitatively sampled without spill-over to surroundings. To sample for chl a, a plate was affixed into the sampling device (Figure 3.8a-b), 1 mL of 0.2 µm-filtered Figure 3.8: The sampling device was sanitized in between plate sampling with isopropyl alcohol. (a) A plate slides into a slot designed into the bottom piece. Contact on the front plate was avoided by holding only the permanently installed tab. The top piece was carefully set down on the plate with the well containing the single o-ring targeted on the area not yet sampled. Engaging the clamp compresses the o-ring to create a leak-proof seal. The plastic top piece was constrained from ever coming into contact with the plate by resting completely on the flat face of the bottom piece of the sampling device. (b) 1 mL of 0.2 µm-filtered seawater was added to the well. (c) An overhead view in to show the water sitting in the sealed well. (d) The biofouling in the well was scraped off with a sterile plastic pestle before collected with a pipette. 36 seawater was added to the well (Figure 3.8b-c), and the biofouling in the well scraped off with a sterile plastic pestle (Figure 3.8d). On day 35, plates groomed 0.5×/week could not be sampled due to hard fouling that prevented leak-proof attachment of the sampling device. From the suspended biofilm, 200 µL was collected into a microcentrifuge tube, centrifuged, the supernatant discarded, and the pellets stored at −20°C until analysis. Chlorophyll a was extracted from the thawed pellets with 90% acetone overnight, and concentrations measured using a non-acidification method and narrow band-pass filters (Welschmeyer 1994) with a Turner-10AU fluorometer calibrated with pure chlorophyll a standards and stability checked daily with a solid standard (Turner Designs, San Jose, CA). 3.3.6 Image Processing Although the human eye can qualitatively determine whether a surface is covered with biofouling, we developed an image-processing tool for year 2 in order to affix a quantitative value to the amount of visible fouling. This tool is amenable to integration with an automated grooming system. Each pixel within a selected section of an image has a value for its Red, Green, and Blue (RGB) channels. For an image section with N pixels, this yields the matrix:   1 2 · · ·    1 2 · · ·   P = , ,  1 2 · · ·    (cid:110) (cid:174), = (cid:111) (cid:174), (cid:174) (3.1)  (cid:110) (cid:110) Note that the selected portion of each image may contain a different number of pixels. Using a section of a clean plate (that has not been fouled), we first compute the mean  and standard deviation  of the RGB channels: (cid:111) (cid:111) −→  = −→  = , ,  , ,    37 (3.2) (3.3) These vectors are representative of the plate material to which we can compare the plates, which are submerged. A separate “clean” vector-pair is computed for each coating type. For a raw image of a submerged plate (top row of Figure 3.9), the vector  + = 1 3 |1−| |2−|   + |1−| + |2−|  + |1−| + |2−|    · · ·  |−|  + | −| + |−|   (3.4) is a measure of the difference between each pixel’s color and the mean value for that channel observed on the clean plate, normalized by the standard deviation associated with that channel. This is illustrated by the colored contour maps (computer generated images) shown in the middle row of Figure 3.9. In each image, some number of the pixels fall under the threshold + < . These pixels are considered “clean”, and the fraction of clean pixels is given by the value of the cumulative density function at , (). By analogy, a pixel is fouled above + > , and the fraction of fouled pixels is given by 1 − (). The threshold  = 4.5 was trained using user input on a subset of the images. In the bottom row of Figure 3, the effect of  on () is provided for the sample image. 3.3.7 Chlorophyll a Processing During year 2, Chlorophyll a (chl a) concentrations in controls and groomed plates over time were compared using a two-way repeated measures ANOVA (rmANOVA) in SPSS (v26.0.0). Cohen’s d effect sizes were calculated for each grooming frequency and time point to estimate how chl a concentration differed on the surfaces of control and groomed plates [70]. Cohen’s d values of 0.2, 0.5, and 0.8 were interpreted as small, medium, and large effect sizes, respectively. Graphs were created using RStudio (v1.1.456) using ggplot2 (v3.2.1). Sampling and processing the data was completed by Abhishek Naik from University of Massachusetts - Dartmouth. More details the chl a can be found in [71], which was the result of the collaboration between MSU, UMass-D, and 38 Figure 3.9: The top row shows raw images of Garolite and Intersleek plates before and after grooming. The middle row shows computer generated images of Garolite and Intersleek plates before and after grooming. The pixels of the computer-generated images have colors that correspond to the entries of +; the thin black contour indicates entries of + = 4.5. The bottom row shows the measure of cleanliness () for different values of ; as expected, the plates are cleaner after grooming. NUWC. 3.4 Results 3.4.1 Preliminary Findings - Year 1 Results from static grooming confirmed that the maximum shear forces are generated near the jet outlet on the plate. Although a significant portion of the biofouling was removed in the initial weeks, some organisms remained attached to the plates in the region below the outer diameter of 39 the pad. This remnant biofouling increased the surface roughness, causing reduction in the efficacy of the grooming operation. Over time, the Bernoulli pad failed to generate sufficient forces for both levitation and grooming. Some results from static grooming are shown in Figure 3.10. It is clear from these figures that grooming was effective in week 3 but less effective in week 5. Also, grooming was most effective in the central region where the shear forces were the highest. Another observation was that the sun exposure was non-uniform on the submerged plates, which is likely to have resulted in less biofouling of the plates groomed with the lowest frequency. To mitigate the shading bias, the locations of the plates were rotated every grooming day during the second experiment in Year 1. In addition, the mode of grooming was changed to “scanned grooming” in the second experiment to alleviate the problems associated with static grooming. In scanned grooming, which was also performed for one minute, the plate was moved laterally relative to the pad such that the entire surface of the plate was subjected to the maximum shear forces. To facilitate scanned grooming, the outer diameter of the pad was reduced to  = 200 mm from its original size of  = 300 mm. This reduction in size was performed after simulations confirmed Figure 3.10: (a) The view from above during static grooming levitating the Garolite plate. (b) Select results of static grooming of Garolite plates in Year 1. The plates are shown before and after grooming in weeks 3 and 5 after they were submerged, for four different grooming frequencies. 40 that the peak shear stress and the pressure profiles were marginally affected. Although scanned grooming was more effective than static grooming, it was not possible to completely dislodge all organisms. Scanned grooming kept the plate surfaces cleaner for a longer duration of time but was eventually rendered ineffective similar to static grooming. Some results from scanned grooming are shown in Figure 3.11. Similar to the results obtained with static grooming, these results also indicate that the effectiveness of grooming reduces over time. Scanned grooming, however, removed organisms more uniformly from the entire surface area of the plates. Also, as evident from the results of week 3 before grooming, uniform sun exposure through rotation of the plates achieved the expected difference in biofouling for different grooming frequencies. After both rounds of tests were completed, it was concluded that it is difficult to quantify the level of biofouling and hence objectively evaluate the effectiveness of grooming. This indicated the need for normalized image-based analysis and more detailed biological testing. During Year 1 it was also determined that the time taken to transport the plates between the pier and the laboratory was Figure 3.11: Select results of scanned grooming of Garolite plates in Year 1. The plates are shown before and after grooming in weeks 1, 3 and 5 after they were submerged, for four different grooming frequencies. 41 not optimal. The grooming was done directly on the pier during Year 2 in order to reduce the potential biases to the experiment resulting from extended time the plates were out of the ocean environment. 3.4.2 Image Analysis - Year 2 The image-processing method was used to assess the effectiveness of grooming undertaken at four different frequencies compared to controls, over a period of seven weeks. The processed “cleanliness” data are shown in Figure 3.12a and b for the Garolite and Intersleek plates, respectively. There is no statistical testing of these data since the image analysis was limited to a single plate for every grooming frequency and surface type. For Garolite, at the beginning of the study (first two weeks), () 1 for plates both before and after grooming. This was true regardless of the grooming frequency. After this initial period, during which the plates became colonized by biofouling organisms, the trend in () before grooming essentially tracked that of the Garolite control, and () after grooming was (with isolated exceptions) higher than that observed before grooming. While this was expected, it was not anticipated that the grooming effect would be larger on the 2 × / and 1 × / plates than on the 3×/ plates. More frequent grooming seemed to promote a robust monoculture of a tenacious biofilm [37], and it is hypothesized that more frequent grooming facilitates the growth of either a more tenacious species or promotes an alternative compact growth habit that can better tolerate the grooming frequency. A similar trend has been observed in industrial water systems, where turbulent flows promote homogeneous and slimy biofilms which are significantly more robust and harder to eradicate than biofilms that develop in laminar flow [53]. The data for the Intersleek plates indicate that up to day 46, grooming frequencies of 1× and more frequent grooming per week were capable of maintaining the plates to the clean state, such that ()1 even for plates before grooming (Figure 3.12). On the plate which was groomed 0.5×/wk, the () for the plate before grooming began to track that of the control plate on day 29, although it appears that grooming was able to reduce the plate back to the clean state. A similar result was 42 observed when the control plate was groomed for the first time on day 48, which was the last day of the study. The before and after photographs of this control plate grooming are shown at the top right of Figure 3.11 and the complete set of four plates are shown at the bottom of Table A.11 in the Appendix. It is also notable that the () for the 0.5×/wk groomed Intersleek plate, prior to grooming, was decreasing over the test. This may have been caused by a rise in temperature over Figure 3.12: “Cleanliness” measured using () with  = 4.5 for (a) Garolite and (b) Intersleek plates for four different grooming frequencies over a period of seven weeks. The terms “before” and “after” indicate the pair of results obtained before and after grooming on a particular day. The points linked by the dotted lines in (a) correspond to the Garolite plates shown in Figure 3.9 (a). Similarly, the points linked by the dotted lines in (b) correspond to the Intersleek plates shown in Figure 3.9 (b). 43 the duration of the experiment, given that 0.5×/wk grooming was apparently effective in removing the biofouling of the Intersleek plates. It is also apparent that the biofouling prevention characteristics of the Intersleek and Garolite control plates were not substantially different when judged on the basis of the image data. This appears to confirm the proposed working method of the Intersleek formulation. As designed, the surface does not simply passively prevent the growth of macroscopic biofouling, but acts as a fouling-release coating only under the action of fluid shear, such as that produced by a moving vessel. 3.4.3 Chlorophyll a - Year 2 Increase in chl a over time was observed on all plate types, although the degree of growth varied across plate types and grooming treatments (Figures 3.13 and 3.14). Overall, chl a concentrations were always substantially lower on the Intersleek vs. Garolite plates. This is a different result from that obtained via image processing. In particular, image-processing data indicated that the passive fouling prevention characteristics on the control plates were not sensitive to plate surface type. On the Garolite plates, the chl a concentration of control plates increased from 0.03 ± 0.001 µg cm−2 at 9 days post-deployment to 3.61.27 µg cm−2 at 37 days post-deployment (Figure 3.13a). Changes in chl a concentration were dependent on grooming frequency. Cohen’s d estimates suggest that grooming of bare plates was associated with reduction of chl a until day 30 for plates groomed 0.5×/wk, 1×/wk and 3×/wk (dashed lines of Figure 3.14, Table 3.2). However, compared to control plates, there were no statistically significant reductions in chlorophyll a on groomed Garolite plates when considering the entire duration of growth (37 days) (Table 3.2). Among the 0.5×/wk-groomed Garolite plates, there was a statistically significant reduction in chl a concentra- tions compared to control plates only for the first 30 days, after which they could not be sampled due to hard fouling growth (rmANOVA; F = 15.34, p < 0.05). While 0.5×/wk grooming visibly removed some soft fouling from Garolite plates (Table A.5 in Appendix). However, the 0.5×/wk groomed plates could not be sampled at day 37 due to hard fouling growth. These results suggest 44 that 0.5×/wk grooming may remove soft fouling, it may have simultaneously allowed hard fouling to accumulate. For Garolite plates groomed 2×/wk and 1×/wk, there was some soft fouling removal (Tables A.4 and A.3 in the Appendix), but no statistically significant difference in chl a concen- trations between control and groomed plates (Figure 3.13a). At 3×/wk grooming frequency, there was a significant difference in chl a concentrations between control and groomed bare plates only until day 30 (F = 26.255, p < 0.05), but not when considering all 37 days (F = 0.706, p = 0.448). On day 37, chl a concentrations were higher on groomed plates compared to control plates (positive effect size, Figure 3.14, bottom right). Hence, despite low post-grooming cleanliness based on image analysis, 3×/wk grooming may have removed chlorophyll-rich biomass from bare plates, but Figure 3.13: Chlorophyll a concentrations (µg cm−2) at 5 time points post-deployment. (a) Garolite plates; (b) Intersleek plates. For the Garolite 0.5×/wk groomed plates, samples were not collected on day 37 due to hard fouling growth. This data and figure was provided by Abhishek Naik. 45 Figure 3.14: Effects of grooming on Garolite and Intersleek plates. Each point is a Cohen’s d estimate representing a standardized mean difference between chl a on control and groomed plates across 3 replicate plates each. Each estimate summarizes the effect size for one plate type, grooming frequency, and day. A Cohen’s d < 0 implies decreased chl a on groomed plates. A Cohen’s d > 0 implies increased chl a on groomed plates. For the 0.5×/wk groomed plates, samples were not collected on day 37 due to hard fouling growth. For Intersleek plates, Cohen’s d values could not be calculated for the day 9 time points due to negligible averages and zero standard deviations among chl a concentrations of groomed plates and control plates. The red lines indicate large Cohen’s d effect size of 0.8. This data and figure was provided by Abhishek Naik. only until day 30. However, there was a grooming frequency-dependent increase in light green and apparently shear-tolerant macroalgae (Tables A.2 and A.3 in the Appendix). The visually observed high abundance of this community appears to be correlated with lower post-grooming cleanliness at 2×/wk and 3×/wk grooming on bare plates (Figure 3.12a). The presence of this community would likely also have influenced measured chl a concentrations. On Intersleek control plates, chl a concentrations were from 0 µg cm−2 at 9 days post- 46 deployment, increased to 0.82 ± 0.19 µg cm−2 on day 30, and decreased to 0.266 ± 0.16 µg cm−2 at day 37 (Figure 3.14b). Fouling growth on Intersleek control plates was loosely attached, hence the lower final concentration may be a result of natural fouling removal by in situ shear stresses. There was a significant reduction of chl a concentration due to grooming at all frequencies on Intersleek plates, relative to the control plates (all rmANOVA tests p < 0.02). Grooming was consistently associated with large reduction in chl a concentrations on Intersleek plates (negative Cohen’s d, solid lines of Figure 3.14 and Table 3.2). However, the effects were generally not con- sistent across time, with effect size decreasing between day 30 and 37 due to the low concentrations of chl a concentrations on the control plate at 37 days post-deployment. Table 3.1: Results of rmANOVA testing the effects of grooming on the concentration of chlorophyll a in fouling material growing on Garolite and Intersleek plates. This data and table was provided by Abhishek Naik. 47 Table 3.2: Effects of grooming on chlorophyll a on fouled plates. Values represent Cohen’s d estimates of effect size, a standardized mean difference between chl a on control and groomed plates across 3 replicate plates each. Cohen’s d values < 0 imply decreased chl a on groomed plates. Cohen’s d > 0 implies increased chl a on groomed plates. This data and table was provided by Abhishek Naik. 3.5 Temperature, Light, and Salinity Data A HOBO MX2202 pendant to log temperature and light intensity every 15 minutes was attached to one of the submersible frames 10 days into the field test study. For the complete data logged shown in Figure 3.15, the average was 21 ±1.7 °C. The salinity from a water quality monitor station 5 km from the submersible rig was acquired from the National Estuarine Research Reserve System (NERRS) in partnership with the National Oceanic and Atmospheric Administration (NOAA). The complete data logged for the time period of the field test study is shown in Figure 3.16. The average surface salinity range was 29.2 ±0.5 parts per thousand (ppt) is representative to what the plates experienced. 48 Figure 3.15: The temperature and light intensity was logged in 15 minute intervals locally with a HOBO pendant sensor that was mounted on one of the submersible frames. The raw data points for temperature were plotted individually, with a moving mean overlay. Figure 3.16: The water quality stations from the T-Warf located at the southern point of Prudence Island represent the open bay water near the field test site. Water salinity data was collected at both surface and bottom stations every 15 minutes using YSI 6600 V2 and EXO2 data loggers. Each data point is plotted for the duration of the study along with a moving mean. 49 3.6 Biofouling Organisms Observed The scheduled grooming occurring only on the front, which by orientation from being suspended off the pier and restricted from rotating resulted in more direct sun exposure. Due to the shading effect, the back and edges of all plates experienced the biofouling pressure differently from that of the front. The combination of the shade and no grooming promoted a completely different array of organisms, to the point where the back and edges of all plates had to be scraped weekly once significant biofouling accumulated. On the sun-side of Garolite plates that were subject to grooming, sampling, and imaging, the fouling community was soft for the first 5 weeks; consisting of green and brown algae, biofilm, and intermittent arborescent bryozoans. After 5 weeks, hard fouling appeared in the form of tubeworms and slipper snails. The sun-side of the Garolite control plates experienced the highest level of hard fouling in the form of bryozoans, slipper snails, tubeworms and mussels. Some of these organisms can be seen in Figure 3.17 of the set of Garolite control plates in the grooming frame. Figure 3.18 shows the same set installed in the submersible frame where the plates are kept physically separated. Keeping the plates underwater allowed us to gently move the buoyant soft fouling aside to better investigate the community. On the sun-side of all Intersleek plates, the fouling community was soft; consisting of green and brown algae, and noticeable biofilm where there was no algae attached. The biofouling on Intersleek also was significantly less dense when compared to the Garolite counterparts of the same grooming frequency, demonstrating the effectiveness of the slime release coating. The Intersleek control did have most biofouling coverage compared to any of the groomed Intersleek but was still significantly less fouled than the Garolite control. To mitigate the possibility of a shading bias on the front of plates from the pilings and pier structure itself, the locations of every set of plates were shifted away from the coast during every grooming day (M, W, F) since the beginning of the field study. This was conveniently and quickly accomplished using threaded connecting links made from stainless steel installed on the ropes to detach and re-secure the individual submersible frames. All of these types of interactions with the plates were completed as much as possible in submerged conditions with the help of multiple 50 ocean water filled totes. The grooming test tank was set up in a shack at the deployment dock that provided protection from the elements, such as sun and rain. The observed organisms that established themselves on all the exposed surfaces of the plate Figure 3.17: The front (top) and back (bottom) of the Garolite control plates at day 43 of the field test study. The box shows a close up picture of the area when submerged to better show the mussel that is attached. The top also shows circular areas that were used for sampling for chl a. Since the front of the plates had a significant amount of biomass by day 35, the sampling device could not maintain a seal for the circular well even if there were no hard fouling in the way. The samples for that day was taken with a sterilized cell lifter, the most noticeable resulting patch being on the front of the second plate from the left. 51 Figure 3.18: The back of the control set of Garolite plates installed in its submersible frame is shown, revealing a variety of macrofouling at day 51 which is after the end of the 7 week study. The last extensive scraping of the back happened 2 weeks prior. (which comprises of the sun-side, edges, and back) is in agreement to what Menesees et al. [40] observed on Garolite plates that were free of any coating. The plates in that study were 15.2× 22.9 cm and suspended at a 22.5° angle from the vertical to both mimic the non-vertical portion of a hull and ensure continuously generated wall shear stress created by the bubbles as they rose along the Garolite surface. The single stream of bubbles created a distinctly clean region. Moving away from the bubble stream, the fouling growth gradually increases. Since the back and edges of each plate in the presented study did not experience any contactless grooming, substantial biofouling did build up, warranting removal during every biological sampling day (Thursday) starting at week 3. The method of carefully scraping with a small putty knife was accomplished by holding the plate only by the permanently mounted tab to maintain no physical 52 contact to the front side. The scraping allowed the plates to fit into the chlorophyll a sampling device while maintaining no direct contact with the groomed-side other than the o-ring that created a seal for the sampling well as described in the earlier section regarding data acquisition. The scraping also prevented any biofouling that started on the edges or back of the plate from influencing or encroaching onto the sun side. Based on visual observations that occurred while handling plates, the same variety and rate of biofouling was observed on the back of all plates, which was expected since the Intersleek plates had the slime release coating only on the front side, leaving the back surface in original Garolite condition. Thus, the edges and back of all 40 plates in this study had the same surface type and same level of light exposure, and consisted of mainly colonial tunicates, arborescent bryozoans and encrusting bryozoans (as previously seen in Figures 3.17 and 3.18). Image Setup and Collection 3.6.1 Four 10.2× 10.2 cm plates were lined up in order and secured flush to a grooming frame. To begin assembly, the grooming frame was positioned upside down on spacers in the corners to prevent the sample plates from coming into contact with anything on the sun-side. Touching was strictly on the tabs and the back or shade-side if needed to ensure flushness. A set of four duplicate plates was labeled on the back with frequency and duplicate number. They were consistently placed in decreasing order (Figure 3.19) so that when the grooming frame was flipped, they were in increasing order. Figure 3.19b shows the flush side that was exposed to contactless grooming via shear stress. The pink foam visible in Figure 3.19a was attached to achieve neutral buoyancy of the assembled grooming frame when submerged in the grooming tank. The fasteners were permanently mounted to the frame. On the other side of the grooming frame, the screws acted like a fixture to mount the plates consistently. 53 Figure 3.19: Photos taken of (a) the back of the grooming frame after a group of 4 duplicate plates are installed, and (b) the front. A paper note was used to document the details and time of what was photographed. 3.7 Discussion The biofouling mitigation effectiveness of a Bernoulli pad grooming device was tested on a Garolite and Intersleek surfaces. The tests took place over a 48-day period at four different grooming 54 frequencies, during which image and chl a data were acquired. An image-processing algorithm was developed to produce a quantitative measure of cleanliness. For the Garolite plates, the image analysis indicates that grooming rendered the plate cleaner than the control plate at the end of the test for all grooming frequencies, but ultimately did not restore the plates to a nearly clean state after 20 days. The chl a data support this observation, and indicates that grooming was minimally effective for all grooming frequencies. The chl a data also suggests only 0.5× and 3×/wk grooming had less biofouling than the control up to day 28. Both visual and biological methods suggest that the higher grooming frequencies removed the less tenacious fouling organisms. Note the before and after grooming difference in C() is consistently smaller for the 3×/wk compared to the other grooming frequencies indicating a the peak wall shear was below critical threshold for a significant portion of the biofouling present, which is why both the 0.5× and 1×/wk appear to be cleanest on day 43. This is consistent with previous observations that stress on the organism produced by turbulent flow or grooming influences the structure of biofouling community [72, 53, 37]. To the extent that there are differences between the methods’ findings, we offer the following possible explanations. First, the image-processing threshold  was calibrated visually, such that the notably clean and fouled areas were consistent between the raw and processed images, as shown in 3.12. In contrast, the chl a data records only photosynthetic biofouling, including that found within the growing layer. A plate with significant thickness of fouling will produce elevated levels in the chl a survey compared to the image analysis. Second, the image-processing algorithm considers the plate to be either clean or dirty at a given location. This means that the "dirtiness" of the plate at that location is saturated. This is not the case for the chl a data, where the thickness of fouling buildup can yield additional chl a. Finally, not all biofouling organisms produce chl a, meaning that the approaches will diverge when the surface is under biofouling pressure from non-photosynthetic organisms. On the Intersleek plates, grooming was able to restore the plate to a nearly-clean state at all treatment frequencies, as measured by both the image-processing and chl a methods. By day 42, 55 some fouling was visible on the 0.5×/wk plate prior to the grooming step, indicating that whatever residual organisms not dislodged by the previous grooming action were able to recover to a visually measurable level in the interim two weeks. The results demonstrate that it is considerably easier to remove biofouling from the Intersleek plates than the Garolite surfaces using the fluid shear grooming protocol. It is also notable that the Intersleek control plate was groomed to a nearly clean state at the end of the study, indicating that a 48-day or longer grooming frequency may be appropriate for this surface type under the in situ conditions prevailing during this study. The Bernoulli pad uses fluid shear stress as a cleaning mechanism. The images therefore fully corroborate fouling-release as Intersleek’s mode of action from wall shear stress generated by an underway vessel, whereas the chl a data suggest that this coating also limits attachment and/or buildup in a to-be-determined manner. There is a range of other foul-release coatings from a variety of manufacturers that rely on the shortest possible idle periods to prevent the required critical shear stress for removal to surpass the generated fluid shear stress [21]. The most successful and thereby most popular foul-release coatings on the market all use surface properties to make the surface either more hydrophilic to increase resistance to cell adhesion and protein adsorption [73] or hydrophobic to lower the surface energy that prevents an organism from forming a strong adhesive bond [74]. When polymers are formulated to be simultaneously both hydrophilic and hydrophobic, the result is a coating such as the Intersleek 1100SR used in this study [75]. We therefore expect that the main results in the present work will be generalizable to other fouling release coatings. For future field test studies of this method, both the gap height and input electrical power to the pump should be measured and logged in real-time. The input electrical power should be varied to allow the device to compensate for events where biofouling has become sufficiently abundant to cause drag in the fluid flow field. Using the gap height and the flow rate data, the maximum velocity in the gap can be calculated, and we can estimate the generated wall shear stress. This information can be incorporated into a feedback loop so that the flow rate can be adjusted to generate sufficient force to match the adhesion strength of biofouling present. This approach of force matching will 56 ensure optimal grooming while extending the life of the anti-fouling or foul-release coating. The electrical input can also be used to determine how restricted the flow is during grooming. A duplicate contactless grooming device machined from clear acrylic could be dedicated to study the cavitation bubbles generated when the flow rate is high enough to generate extreme low pressure region immediately after the corner of the stem-pad interface where the axial flow from the jet begins to turn into radial flow. Such an investigation will determine the location and extent of cavitation erosion (potentially on both pad and contactlessly groomed surface) as well as quantify the improvement on biofouling mitigation when the maximum possible wall shear stress is generated consistently. 3.8 Concluding Remarks A Bernoulli-based grooming device is capable of effectively removing fouling from an Intersleek surface without damaging or abrading the Intersleek surface material. The same cannot be said of the Garolite surface. The method of cleaning is fluid shear stress, which is the same mechanism used to produce fouling release when a vessel is underway [21] and can help maintain a vessel hull that experiences biofouling pressure when idle, by preventing biofouling before it gets too strongly attached (Figure 3.2). One shortfall of relying on the speed of the vessel to shed biofouling is that the generated wall shear stress will vary depending on the location on the hull. This grooming effect, produced entirely without surface contact, is anticipated to improve hull performance while extending the life cycle of the coating relative to more abrasive cleaning methods. The image-processing algorithm described here offers a way to achieve immediate feedback in an automated grooming system. Furthermore, the radial outflow supplied by the Bernoulli pad enforces a self-equilibrating gap between the pad and the hull. This removes any need to sense contact pressure on the hull, which might ordinarily be required to remain attached while not creating excessive shear force on the hull surface. We therefore believe that this type of grooming system is well-suited to incorporation in an automated or remotely operated system, and we plan to explore this in future work. 57 CHAPTER 4 3D PARTICLE TRACKING VELOCIMETRY OF RADIAL OUTFLOW BETWEEN TWO PARALLEL PLATES 4.1 Introduction For the Bernoulli pad to be effective as a cleaning device, there needs to be sufficient wall shear stress generated by the radial fluid flow to remove organisms that settle on the surface of a vessel hull. Since the flow rate and equilibrium height have an inverse relationship; with increasing the flow rate directly increases the generated suction force, which decreases the equilibrium gap height. The CFD model has been validated with previously published experimental data using air as the working fluid as previously presented in Chapter 2, but only in terms of radial pressure profiles and the suction force generated at different gap heights. There was no information available on experimentally measured velocity field and the derived pressure to compare to CFD. With water as a working fluid, an experiment was conducted in a glass tank with a Bernoulli pad fabricated with the dimensions used previously in the field test study as discussed in Chapter 3. 4.2 Description of Experiment 4.2.1 Bernoulli pad in water tank The Bernoulli pad used in our experiments had an inner diameter of 31.75 mm and an outer diameter of 200 mm. The pad was made of opaque matte black acrylic to block the view of particles on the other side of the pad and reduce glare within the measurement volume. The pad was connected to a flow conditioning system comprised of a plenum chamber, a honeycomb, a series of meshes (coarse mesh with 46% opening area; medium mesh with 36% opening area, fine mesh with 31% opening area), and a converging nozzle [76]. This flow conditioning system produced a very close to homogeneous velocity distribution at the inlet of the Bernoulli pad. Figure 4.1 shows the Bernoulli pad together with the flow conditioning unit in exploded and assembled views. This assembly will 58 Figure 4.1: The Bernoulli pad assembled with a plenum and flow straighteners. Figure 4.2: The BPA is mounted on a beam with roller wheels that allow the pad to be placed in close proximity to the tank wall. Double arrows indicate the direction of movement of the BPA permitted by the roller wheels; this movement allows the pad to find its equilibrium configuration. henceforth be called the Bernoulli pad assembly (BPA). Our experiment was conducted in a glass water tank of dimensions 1.22 m× 0.305 m× 0.762 m. The water tank was placed on an optical table. A modular frame with tracks was secured to the optical table to mount the BPA on rollers and with its pad parallel to one of the tank walls. When water flows radially outward between the pad and the wall, the BPA would move freely toward or away from the wall to find the equilibrium gap between the wall and the pad. As shown in figure 59 wire screen mesheshoneycomb structureconverging nozzleBernoulli padacrylic chamberend capinlet frompump(a) exploded view(b) assembled view(a) top view(b) side viewBernoulli padglass water tankmodular framebeam with roller wheelspumproller wheelsbeam 4.2, a submersible pump (1/3 HP) is connected to the Bernoulli pad assembly to drive the flow. This setup is shown in figure 4.2. 4.2.2 Imaging For imaging, we used four Phantom v2512 cameras (maximum frame rate at full resolution: 25,700 fps at 1280 pixel x 800 pixel); they were placed in an arc-like pattern (see figure 2.2) such that the optical axes of their lenses were coplanar and converged to a single point located in the gap between the BPA and the tank wall. Each camera utilized a Scheimpflug mount to keep the entire image plane in focus. A dual head Photonics Industries class IV laser, DM30-527DH, (maximum energy: 70 mJ/pulse at 527 nm, maximum pulsing rate: 10 kHz/head) was used to illuminate a volume encompassing the focal point. A programmable timing unit (PTU) was used to synchronize the cameras with the laser for data acquisition. This imaging setup was implemented with DaVis software by LaVision, Inc. We used Time-resolved 3D Particle Tracking Velocimetry (4D-PTV) for flow characterization. This method is based on syncing multiple cameras to look at a laser-illuminated volume from different angles simultaneously, to track the motion of individual particles suspended in the fluid. It uses an Iterative Particle Reconstruction (IPR) technique in combination with a 4D algorithm using time-information of the particles for track reconstruction (assuming that tracer particles track fluid particles sufficiently well, a track can be defined as a surrogate path-line). To increase our signal-to-noise ratio, premixed fluorescent-coated microspheres (38 − 45 m), which fluoresce at a wavelength of 606 nm peak emission upon excitation with green light (527 nm in our case) were used to seed the water in the tank. Because the light fluoresced off the particles was of a different frequency, any stray laser light was filterd out by notch filters centered at the fluorescent wavelength. With 50 mm lenses and a 1.5 magnification factor, particle images were typically around 2 pixels in diameter. To minimize peak-locking, diffuser optical filters were in- stalled in the Scheimpflug mounts to broaden the point-spread function of particle images. 60 Figure 4.3: A top view of the experimental setup showing the imaging hardware comprised of four ultra high-speed cameras and a dual-head pulsing laser. The cameras are placed in an arc-like arrangement and focused on a volume that lies between the pad and the tank wall. The dual-head laser is located in the middle of the four cameras. Spatial calibration was performed with a dual-plane target with a distance of 1 mm between planes. Once the laser volume and cameras were aligned with the target, the camera view was masked to the area of interest in order to reduce memory requirements and disk space. The target was used to make the cameras look approximately at the same field of view and to obtain an initial coarse spatial calibration. Since the depth of the volume of interest was ∼ 1 mm, it was unnecessary to traverse the target along the depth to construct the initial spatial calibration. The spatial calibration was then refined with a self-calibration algorithm applied to particle images [?]. After volumetric self-calibration and refinement were performed to reduce disparities, the optic transfer function was generated. The resulting fit model was a third-order polynomial with an RMS of fit of 0.0013 pixel and a scale factor of 14.11 pixel/mm. The final dewarped image was 227 × 851 pixels in size. 61 4.2.3 Experimental Procedure After removing the calibration target, the resulting laser volume between the wall and the BPA was as shown in figure 4.4. This laser volume captured a portion of the incoming axial flow and a radial strip that extends beyond half the radius of the pad, where the average flow has become purely radial. Given the arc-like camera distribution, and to achieve optimal focus by using the smallest possible angle between the cameras, the orientation of the laser volume was chosen to be vertical. Figure 4.4: (a) CAD model illustrating where the laser illumination is directed to on the BPA. Defined edges indicate where the image was masked in the flow visualization software. (b) The Bernoulli pad showing the matte surface finish and three equally spaced nylon-tipped set screws installed. These set screws help avoiding perturbations of the uniform equilibrium gap between the pad and the tank wall. (c) Side view of the laser illuminating the area of interest in the gap and partially into the jet outlet. Each laser head was run at 10 kHz, and the two were combined in cavity stacking mode to achieve a laser pulse frequency of 20 kHz. A total of 100,000 sets of four images each were acquired over a time interval of 5 s. With the purpose of reducing loss of pairs and peak locking due to exceedingly large or small displacements respectively, flow conditions were adjusted at the maximum acquisition rate to obtain particle displacements no greater than 10 pixels between frames [77]. This was achieved by finding an optimal combination of equilibrium height Òeq and . For the present experiment, a configuration based on simulations indicated Òeq ≈ 1.25 mm,  ≈ 0.77 m/s, resulting in max ≈ 8.4 m/s. These simulation-derived parameters were found to satisfy the above-mentioned 62 requirement for pixel displacement, and the process is discussed more in Section 4.4. Higher values of max were avoided because they would induce cavitation and would also result in very small equilibrium gaps. The equilibrium gap was maintained by first adjusting the nylon-tipped set screws (see Figure 4.4b-c) to provide the nominally correct gap as defined by our simulation. The electrical power delivered to the pump was then increased until the support force provided by the screws was approximately zero. The screws were then retracted slightly to confirm that the pad was indeed in equilibrium, then replaced to make gentle contact with the tank wall. The aforementioned simulations indicated that the mean radial velocity ¯() = 2 m/s at the outermost radial location within the measurement volume. The total time t = 5 s duration of data acquisition corresponds to t ¯()/ = 4.2 times the radial distance of the measurement volume. This duration of data acquisition was deemed likely to provide a sufficient number of statistically independent data samples to ensure convergence. 4.3 Experimental Measurements In this section, the raw images from the four cameras to obtain the velocity vector field in the volume of measurement are outlined. This experiment was carried out with a relatively low seeding density, requiring a large number of instantaneous realizations (100,000) to ensure convergence of flow statistics. For illustration, a single particle image taken from the set of 100,000 is shown in Figure 4.5, which we will refer subsequently as time  = . Figure 4.6 shows an overlay of all raw images. This composite particle field was used to detect if there were any large regions of the volume of interest consistently devoid of tracers in the course of our experiment. A region like that would be unable to yield statistically significant data. While there are a few narrow gaps that were not visited by tracers, their azimuthal extent is small enough to be taken care by data interpolation after converting the track field into a structured-grid velocity field. Each set of four 2D particle images, like the set shown in figure 4.5, is used to determine the in- stantaneous location of each particle in the 3D volume of measurement. The particle distribution of 63 Figure 4.5: A single representative image at time T (taken from the set of 100,000 images) from the four cameras is shown. Figure 4.6: Overlay of all 100,000 images are shown for each camera. These images (left to right) correspond to the cameras (left to right) shown in figure 4.3. 64 the next time step is then predicted based on the current particle distribution. The errors introduced with the prediction are then corrected with image matching before repeating the prediction for the following time step. This algorithm also avoids the problem of creating ghost particles. Following this prediction-matching approach, the motion of each individual particle is tracked using subse- quent instantaneous realizations until the particle exits the field of view. The time delay and particle displacement between instantaneous realizations are used to determine instantaneous velocity. The set of all instantaneous velocities and positions of a single particle define particle tracks (path-line surrogates). When the entire set of 100,000 images is processed, the particle-tracking algorithm produces a large set of particle tracks that define a flow field based on a Lagrangian description. This PTV technique was implemented using LaVision-DaVis. Reference [57] presents a detailed description of this algorithm. The particle-tracking algorithm was performed to the entire illumi- nated volume multiple times for refinement, where the total volume consisted of the gap and the bottom portion of the jet flow 15 mm into the stem. To keep both the processing time and file size reasonable, the final volume of interest was defined to be 13 mm azimuthally by 53 mm radially, by 1.5 mm axially from the glass tank wall. Transforming the data from the Langrangian specification of the field (Figure 4.7b) to the Eulerian specification (Figure 4.7c) requires converting the tracks to a vector field on a structured grid. The PTV data had vector fields computed for every time step with the following specifications: collect tracks inside sub-volumes sized 4 voxels with 75% overlap resulting in a grid spacing of 1 voxel (which was equivalent to a cube of size 0.07 mm in our experimental setup) using a third-order spatial polynomial. The collected tracks are then converted using a filter length of 5 time steps and a second-order polynomial fit. This interpolation resulted in 100,000 sparse velocity fields that were ensemble-averaged to obtain the average velocity field. Given their sparsity, grid points with no velocity information were excluded from the calculation. Ensemble averaging and any further calculation on the vector fields were implemented with MATLAB as explained in the following sections. 65 Figure 4.7: a) Straight-on view of measurement volume that includes a small portion of the jet. b) and c) Tracks of the particles (Lagrangian) visible in the four images in figure 2.6 over 8 additional frames (equivalent to 400s). d) Vector fields (Eulerian) of the particles visible in the four images in Figure 2.6. 66 4.4 Computational Fluid Dynamics Model To help with both the setup and analysis, a 2D axisymmetric CFD model was created and run in ANSYS-Fluent with steady-state flow conditions. A four-equation Transition-SST model, based on the k- transport equations coupled with intermittency and transition transport equations, was used to model turbulence. For more detail on the setup of the CFD, please refer to section 2.2.1. The pad outer diameter and jet diameter corresponded to both the device used in the field test study from Chapter 3, as well as the black acrylic pad machined specifically for the PTV experiment in this Chapter. The 31.75 mm stem diameter in the CFD model and experimental both had a length of 25 mm, where a MFR boundary condition in CFD produced top hat velocity profile to match the top hat velocity profile of water created by the flow straightener entering the clear acrylic stem of the BPA. In both cases, the flow was allowed to develop past the initial entrance of the stem. With the limit of the hardware being 10 m/s in order to prevent a seed particle from moving more than 10 pixels in a single time step, the gap height had to be set to an approximate gap height and flow rate. Otherwise, significant error will be introduced during the cross-correlation to achieve particle track reconstruction. To find a suitable pair of parameters, a parametric sweep was done with 4 different mass flow rates and 8 gap heights as seen in Figure 4.8. The MFR needed to Figure 4.8: The force generated at different gap heights and four different mass flow rates (left). The force versus mass flow rate at a gap height of 1.25 mm with equilibrium shown with a dashed line (right). The maximum velocity generated at this equilibrium point in the fluid flow field at 1.25 mm is 8.4 m/s, which is below the limits of the hardware. 67 be significant enough to generate enough suction (negative force) with a steep F/h slope to allow the suspended BPA to overcome rolling friction on the rails. Based on the preliminary results, The equilibrium gap height needed to be larger than the experimental study, which a smaller flow rate would achieve. Since the experiment will have outliers in regards to the maximum velocity, the CFD model indicates that a mass flow rate of 0.605 kg/s will result in an equilibrium gap height of 1.25 mm with a peak velocity magnitude of 8.4 m/s. 4.5 Results Radial outflow has a unique fluid flow field, where the initial axial flow transitions into radial flow. This involves separation of the flow at the corner of the stem/pad interface before reattaching on the pad-side further downstream. The DaVis Software has the ability of converting the particle tracks into a vector field and calculating the pressure field. 4.5.1 Ensemble Average The above mentioned conversion from particle tracks to vector fields in an unstructured mesh, and then to vector fields in a structured rectangular mesh, produced a total of  = 100, 000 3- dimensional vector fields of grid size  = 171 ×  = 756 ×  = 22; with velocity components at each grid point (, , ) given by (, , , , , , , ,). The value at each grid point of the ensemble average of each velocity component (, , and ), is defined as the average velocity value across all  instantaneous realizations at each grid point. That is:   =1 =1 , , =   =1 =1 , , () () , , , , , = , , () () , , , , , = , , () () , , (4.1)   =1 =1 Where () , , = 1 if the (, , )−th grid point of the −th instantaneous realization has data, and () , , = 0 if it does not. Equations (4.1) were used by DaVis software to generate an average efficiently without having to store the individual vector fields for every time step. 68 4.5.2 Azimuthal Average Utilizing the nature of the axis-symmetric flow, both the Cartesian coordinate system and the corresponding data were transformed onto a cylindrical frame of reference and collapsed onto the rz-plane. To focus on the longest particle tracks that originally generated the data, a wedge that starts at the origin of the jet and intersects with the bottom of the experimental window was defined from 4.64 to 4.78 radians, where 3/2 radians is the center of the wedge that was illuminated at the bottom of the BPA. The transition from the 3D (, , ) space to a 2D (, ) plane resulted in the data remaining structured in the z-direction, but became unstructured in the r-direction. Bins of the size 0.07 mm were used to preserve the resolution, and were used to sort and average the number of instances that occurred in each bin. The resulting fluid flow field on a 2D plane is shown in Figure 4.9a-c. Figure 4.9: Portion of the inlet is visible on the bottom left, and the intermediate outlet of the experiment window is at the top of these plots. (a) Raw z-component of velocity, (b) Raw theta- component of velocity, (c) Raw radial-component of velocity, (d) refined radial-component of velocity, e) radial-component of velocity from CFD run at the same boundary conditions. With the nature of flow visualization using seeding particles, there were instances in which no 69 seed passed through a voxel in space nearest the walls, even after binning. The velocity was forced to be zero m/s at the walls, and a shape-preserving piecewise cubic spline interpolation algorithm in MATLAB called ’pchip’ was used to fill in the gaps near the wall as shown in Figure 4.9d. With an estimated 0.59 kg/s mass flow rate in CFD that will be discussed in more detail the next section, there is good agreement of the radial velocity contour plots between PTV and CFD (Figure 4.9d and e). 4.5.3 Pressure Distribution From the LaVision DaVis software that was used to conduct the experiment and process the captured images, the pressure was directly computed using LaVision Pressure from PIV, a 4D pressure solver based on the method of computing the pressure gradient from the velocity field and using Poisson’s equation to spatially integrate [78]. Using the velocity field generated from the average of all 100k images as described previously, the time-averaged pressure was calculated from the following parameters; working fluid was defined as water at room temperature with the boundary condition set using gauge pressure equal to 0  at point  = 60,  = 0.637, which this solver treats as a Dirichlet boundary condition. The 4D pressure solver developed in part by NIOPLEX (Non-Intrusive Optical Pressure and Loads Extraction for Aerodynamic Analysis) [79] where the 4D spatio-temporal blocks consist of multiple 3D domains as well as their Lagrangian connections in the time domain. Boundary conditions are defined by the user to the first time step, and the calculated pressure is used in the next time step. The resulting pressure field from PTV is shown in Figure 4.10, with the main discrepancy being the pressure generated in the central region. 70 Figure 4.10: (a) Raw pressure, (b) refined pressure, c) Pressure from CFD run at the same boundary conditions. 4.6 Analysis Once the PTV data had been collapsed onto the rz-plane, the mass flow rate was estimated at every radial location using the following summation 22 =1 (cid:164)() = 2 (, ) (4.2) where L is the grid spacing between the 19 points within the gap in the axial direction. There were 3 additional points that include data captured in the stem, where  > 1.275 as seen in Figure 4.11, and were not included in this calculation. This summation is shown in Figure 4.12 was repeated with the refined PTV data, and shows that the radial velocity near the wall can be better estimated. The focus of this chapter is the velocity and pressure field that was well defined by enough instances of particles in the 100k sets of images. The estimation of a MFR of 0.7 kg/s was used in CFD. 71 Figure 4.11: An illustration to show the boundaries of the data after completing the azimuthal average. Figure 4.12: Mass flow rate calculated for every radial location using raw and refined radial velocity data. To better visualize the different components of velocity and the pressure field in the gap, scatter plots with each z-location was color coded. The axial velocities for PTV and CFD are shown in Figures 4.13 and 4.14 respectively. The axial velocity before the corner of the stem/pad interface has good agreement. The axial velocity does achieve a larger negative values after the corner when compared to CFD. The flow that occurred near the glass wall experienced only positive radial velocity and near 0 m/s axial flow. At the corner of the pad/stem interface, the flow separation 72 Figure 4.13: The experimental radial velocity at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized PTV data is overlaid on the original unstructured PTV data. 73 Figure 4.14: The CFD radial velocity at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized CFD is overlaid on the original unstructured PTV data. 74 Figure 4.15: The experimental axial velocity at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized PTV data is overlaid on the original unstructured PTV data. 75 Figure 4.16: The CFD axial velocity at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized CFD is overlaid on the original unstructured PTV data. 76 Figure 4.17: The experimental discretized and tangential velocity at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized PTV data is overlaid on the original unstructured PTV undiscretized data. 77 Figure 4.18: The experimental pressure at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized PTV data is overlaid on the original unstructured PTV data. 78 Figure 4.19: The CFD pressure at different axial locations within the fluid gap (/Òeq), where zero is at the glass wall. The discretized CFD is overlaid on the original unstructured PTV data. 79 results in a region of recirculation. This seperation bubble appears to be larger in the PTV visualization as evidenced by the locations 0.67 < /Ò < 0.94 to reach negative radial velocity values whereas only the locations 0.83 < /Ò < 0.94 in the CFD reach negative radial velocity values as seen in Figures 4.15 and 4.16 respectively. This recirculation region also where the highest tangential velocities occur as seen in Figure 4.17. The mean of the discretized points is -0.04 m/s. CFD results of the tangential velocity is omitted due to being zero everywhere in the gap, even with axis-symmetric swirl allowed. After the negative pressure peak is generated, the pressure asymptotically goes to zero gauge pressure by the end of the experimental window as seen in Figure 4.18. While there is discrepancy in peak positive pressure between the PTV and CFD Figures 4.18 and 4.19 respectively, both reach near zero gauge pressure by D/2 = 50 mm. This confirms the initial boundary condition used in the DaVis pressure solver was a reasonable one. Further investigation to the pressure at this point is discussed at the end of this section. Since only a portion of the radial flow was captured, the pressure gradient beyond 64.4 mm is unknown. Based on the tail spread of available undiscretized data seen, the shape of the remaining pressure curve is unknown. Integrating the pressure curve at the glass, over the radially growing surface area, the force can be estimated if the data is extended from 0 mm to 10.8 mm and from 64.4 mm to 100 mm as seen in Figure 4.20. From past CFD cases as well as the experimental data set from Li and Kagawa [1] used in the CFD model validation in Chapter 2, the asymptotic portion of the pressure curve has been observed to slightly overshoot the ambient pressure before the outlet. Since the DaVis pressure solver calculates the pressure gradient, the earlier estimation of 0 kPa at 60 mm can be modified in such a way to achieve F = 0 N when the entire pressure curve is integrated as seen in Figure 4.21. With 0 kPa at 100 mm, the data points in between 64.4 and 100 mm was populated with data using a shape-preserving piecewise cubic spline interpolation algorithm. Plotting the cumulative summation of force along the radial length does confirm that the pressure gradient still had an upwards trajectory as seen at the second vertical dashed line in Figure 4.23, but even modifying how the pressure profile extends by folowing the trend before 64.4 80 mm would not achieve a total force of 0 N even with an overshoot to 0.82 kPa as seen in 4.22. The offset of 0.82 kPa does preserve the pressure gradient and makes a reasonable estimation of the shape of both the pressure curve and the force summation after the end of the experimental window. The offset also falls within the spread of the tail end of the unstructured data. Figure 4.20: The discretized pressure curve along the bottom of the glass overlaid over all the unstructured pressure data in the entire height of the gap. The vertical dashed lines indicate the original data region from 10.8 mm and 64.4 mm. Figure 4.21: The discretized pressure curve along the bottom of the glass with an offset of +0.82 kPa overlaid over all the unstructured pressure data in the entire height of the gap. The vertical dashed lines indicate the original data region from 10.8 mm and 64.4 mm. 81 Figure 4.22: The discretized pressure curve along the bottom of the glass with an overshoot of +0.82 kPa overlaid over all the unstructured pressure data in the entire height of the gap. The vertical dashed lines indicate the original data region from 10.8 mm and 64.4 mm. Figure 4.23: The cumulative summation of the force along the glass. Pressure at a radial location was multiplied with the annular surface area. The vertical dashed lines indicate the original data region from 10.8 mm and 64.4 mm. 82 To normalize the captured and processed PTV data, the maximum radial velocity,  is unique to the fluid field flow since keeping the velocity at the inlet constant will generate different maximum radial velocity at every different gap height. The radial velocity is normalized as follows: ∗  = (, )  with the radial location normalized with respect to the radius of the jet outlet. where the normalized pressure, ∗, is based on the pressure forces and inertial forces. ∗ =    ∗ = (, ) 2  (4.3) (4.4) (4.5) Each data set was defined in non-dimensional terms of its own Euler and Reynold’s numbers:  = (, 0) 2   =     (4.6) (4.7) The near the walls of the gap show reasonable agreement of the normalized radial velocities in Figure 4.24. The case previously validated in Chapter 2 where the air was the working fluid and the pad was  = 4,  = 40 is also shown in red. The discrepancy between the pressure near the centerline of the jet still exists after normalization as seen in Figure 4.25, but there is good agreement between the three CFD cases of water (green, grey, and cyan) and the CFD of air (red). The experimental data with air (magenta) measured a higher pressure than what was predicted with CFD, further confirming that the flow is complex, and any change to the size in th seperation bubble impacts the pressure field. 83 Figure 4.24: Normalized Radial velocity for multiple flow rates of water with the pad dimensions used in the PTV experiment. The velocity near the glass wall (top) shows no negative flow due to being between the separation bubble and glass, where the velocity near the pad (bottom) does capture the recirculation that occurs in the separation bubble. 4.7 Concluding Remarks Confined radial outflow has never been imaged before using PTV, but the current hardware and software made it possible to capture at an acceptable level of accuracy and provide insight to the 84 Figure 4.25: Normalized pressure for both PTV and CFD data of the water Bernoulli pad based on equation 4.5. The air case is included with the previously published experimental data measured with a traversing pressure tap. region of separation and reattachment. The annular recirculation bubble is advantageous in our application of a contactless grooming device because of the direct correlation to the maximum velocity that is induced between the bubble and the glass wall. After processing the data and converting to grid, the gap height was measured to be 1.275 mm, providing supporting evidence that the pad was indeed levitating without contact when the pump was in operation. The setup can be improved by incorporating a mass flow meter in line at the inlet and several pressure taps embedded in the pad; such as after the corner and before the outer diameter. Rolling friction in the suspended rail configuration was could prevent the pad from finding the exact point of equilibrium. As seen in Figure 4.8, the slope of the force curve is steep near the stable equilibrium point, and would still indicate the results presented in this Chapter to be near equilibrium. The set screws were mainly to provide stability and prevent any oscillations of the levitating pad to keep a fixed gap as the reattachment point has been observed to fluctuate [66] which in turn would 85 slightly fluctuate the force gnerated by the fluid flow field. Another major improvement would be to add a precision laser sensor to measure the gap height. Ideally in multiple locations to ensure the pad is indeed parallel to the glass wall. The seeding density was at the minimum desirable level and should be increased in the future. Once the experimental setup is robust, multiple data sets at different flow rates and different gap heights can be analyzed to see the development of the turbulent and transition region. The stable equilibrium point with no applied force,  = 0 will continue to be the main aspect of intereset due to the final application of the contactless hull grooming device described in Chapter 3. Another shortcoming of the research presented here is that only a portion of the radial length was captured. The relative pressure at the outlet in the experiment could not be confirmed. It was assumed that the pressure had reached 0 gauge pressure at the outlet, but looking at the raw PTV data before descretization in Figure 4.18 a range of pressures measured in the experimental window contracts from 20 mm to 55 mm before spreading out at the end of the experimental window. Discretization did result in an average pressure with minimal fluctuation in the last portion of the radial outflow. Having a complete pressure profile curve would allow for the force to be accurately calculated through integration. Reducing the size of the pad is not recommended since a shallower experimental window would occur at equilibrium with no applied Force, . Holding the gap with some  would still be of some interest. Keeping the current size of the Bernoulli pad, the fluid flow along entire diameter can be captured by stitching data from multiple data sets from several different experimental windows. This can be accomplished without moving the cameras or laser volume if the radial outflow in the horizontal is the focus, and the BPA can still move laterally as in the original experimental setup as seen in Figure 4.2. The resulting pressure versus radial location can be integrated to determine the total force generated by the flow field at specific parameters and operating conditions. With the recommended improvements to the PTV setup, the larger variety of acquired data will serve as a better benchmark for refining the CFD model. With only the velocity field and pressure gradient from the experiment, there was no efficient way to investigate the Reynolds-averaged 86 Navier-Stokes (RANS) equations and how they impact the region of separation and reattachment. Another turbulent flow simulation method should also be investigated in future studies, such as Scale Resolving Simulations (SRS). The main disadvantage of SRS is how computationally ex- pensive the modeling is since both high grid resolution and small times steps are required for this inherently unsteady method. 87 CHAPTER 5 CONCLUSIONS The oceans cover 71% of the Earth’s surface which the international seaborne trade navigated to transport 11 billion tons of cargo in 2018 [80]. As stated in the introduction, optimal fuel efficiency can be achieved with a clean hull of transport vessels. This results in a cost savings and a reduced carbon footprint. The other important aspect to consider is that every vessel has the potential to transport invasive species, which is why it is essential to prevent biofouling in order to stop these biological hitchhikers from spreading further. The lessons learned from this research is not just for the ocean environment and the US Navy. Increasing the efficiency of any system that is submerged in salt or fresh water is significant. Even stationary systems, such as a power plant, highly benifits from biofouling prevention. For mobile surfaces that experience submersion, such as a boat or vessel, cleaning reduces or eliminates the unintentional transportation of invasive species to another waterway. The state of Michigan has been dealing with the consequences of the non-native zebra mussel (Dreissena polymorpha) that have been first detected in Lake St. Clair in 1988 and found to be established in all of the Great Lakes by 1990 [81]. The zebra mussel was indigenous to the lakes of southern Russia and Ukraine, but has spread globally as a result of discharged transoceanic ship ballast water contaminated with mussels [81]. Juveniles tend to settle on surfaces that experience flow rates less than 1.5 m/sec [82]. Once established, they can grow so densely that they block pipelines and screens; resulting in clogged water intakes of hydroelectric companies, municipal water plants, and other industries. Zebra mussels causes accelerated corrosion of concrete and steel, which impacts structural integrity [81]. For a vessel of any size, zebra mussels can encrust the hull which results in significant drag or restrict the passages of the engine cooling system which results in overheating [81]. The national economical impact has been estimated to be $1 billion per year [83], with the cost of damage and management of zebra mussels in the Great Lakes alone exceeds $500 million per year [84, 85]. Another side effect of invasive species is how they impact ecosystems by disrupting the food chain 88 and out-compete native species. Since the gap between prevention and removal of biofouling organisms is continuous, the key is early and continuous attention. A custom solution by combining several or more biofouling prevention tactics holds the most promise long term. As part of the maintenance and repair schedule, there are three categories: passive (such as the fouling-release paint that was used in Chapter 2), active (such as the contactless grooming proposed by this dissertation operated by a SCUBA diver), and active but autonomous (such as developing the contactless grooming device into a autonomous robot). The biofouling mitigation schedule is recommended to utilize more active tools and approaches during the peak of the local biofouling season. Investigating larger sizes of Bernoulli pads with flexible material or an array of Bernoulli pads to conform to underwater surfaces with noticeable curvature would increase the power and versatility of the non-contact grooming device. The geometry of the stem and pad holds the most potential for creating the optimal fluid flow field, but from early studies there are often trade-offs. Both the force of suction and the size of the levitation window can be increased by using a pintle to help turn the flow from axial to radial, but at the cost of reducing the peak wall shear as seen in Figure 2.11. Directly impinging flow with out being confined has demonstrated successful removal of bio- fouling while the adhesion strength is still under the threshold of the peak wall shear stress generated by the flow [41, 68, 56]. When compared to the confined radial outflow approach in this research, fluid power of the unconfined radial outflow is reduced due to entrainment of the ambient water. The other benefit the Bernoulli pad has is that the device will automatically self equilibriate the gap as it is moved across a surface. The main limitation to the Bernoulli pad approach are surface features that protrude more than the gap height or inaccessible areas of a ship such as the side thrusters. The bulbous bow and stern have the most extreme variances in geometry. Creating an array of Bernoulli pad devices would increase the effective surface area cleaned, and would allow for more physical flexibility even with rigid pads while providing stability when cleaning portions of the hull with extreme variances. If one pad of the array reaches the unstable equilibrium point, the pads that remain levitating can help re-orientate the dislodged pad to find the stable equilibrium 89 point and continue grooming operations. An aspect of operation that has potential to be beneficial or detrimental to the groomed surface is when cavitation is induced with higher flow rates. Inves- tigating caviation with another field test study would provide insight to the durability of both the device and the groomed surface. Physically constraining an impinging jet to a confined gap has several interesting attributes. The corner geometry determines how quickly the axial flow is turned into radial flow. There also is a flow separation induced at this corner, which does reattach downstream based on the main operating conditions of flow rate and gap height. The larger the outer diameter, the more room there is for reattachment. With this in mind the gap can be made larger before levitation ceases when the unstable equilibrium point is reached. In between the separation and reattachment point on the surface of the pad, lies the annular recirculation bubble. This region . Not only is the cross sectional area reduced from the jet diameter ((/2)2) to the first instance of the cross-sectional area in the gap (Ò/2), but the presence of the recirculation bubble artificially reduces the cross- sectional area that the majority of the flow passes though after the corner as seen in Figure 5.1. The maximum velocity in the entire fluid flow filed can be found in this specific region. The maximum wall shear stress on the surface parallel to the pad correlates directly to this maximum velocity. The jet still generates a stagnation region where the maximum pressure and minimum velocity of zero occur. After the axial flow turns into radial flow, the velocity also quickly diminishes with respect Figure 5.1: Modeling a water Bernoulli pad ( = 31.75 mm,  = 200 mm) with a mass flow rate of 1.5 kg/s resulting in an equilibrium gap height of 0.9 mm. Only a portion of geometry is shown from  ≈ 10 mm to 50 mm. The top contour plot shows the radial velocity in a torrid-shaped fluid space, if the flow did not have axial flow to turn the corner. The bottom contour plot shows how the recirculation bubble (indicated by the region with negative radial velocity) constricts the turning flow to achieve a higher values of velocities near the corner. 90 to  due to increasing cross-sectional area in the gap. The flow does recover after the separation bubble, and becomes purely radial as seen in Figure 5.1. Since wall shear stress is the critical feature of this Bernoulli pad device for contactless grooming applications, additional experiments using a hot film sensor to find the wall shear profile generated along . The hot film probe can be mounted from the back-side of a test plate with the surface flush to the front-side that would be parallel to the pad. Including a force torque sensor to ensure equilibrium, where  = 0, and a laser measurement device to determine the gap height at equilibrium conditions. A flow rate sensor can be incorporated to measure fluid power as defined in Chapter 2. Electrical power to the pump should also be monitored and recorded in real-time. There is potential to reverse-engineer a wall shear stress profile, by creating the desired wall shear stress profile first to steer the design and operating conditions. Only the center of a pad can be designed this way since there is no effective way to counter the increasing cross-sectional area in the radial direction without affecting the desired clearance of the pad relative to a surface. 91 APPENDIX 92 APPENDIX YEAR 2 FIELD TEST STUDY DETAILS AND COMPLETE COLLECTION OF RAW IMAGES USED FOR IMAGE ANALYSIS Table A.1 is a legend for the tables A.2 to A.11. The images are arranged in pairs where the top picture is the ‘before’, while the bottom is the ‘after’ as seen in Figure A.1. In the subsequent pages, note that position 4 was left untouched for image processing, while positions 1,2,3 are the sampled plates where little circles free from biofouling were the result of sample collection as described earlier, in Chapter 3. Table A.1: Detailed legend for the following tables displaying all photos used in image analysis. The field study was initiated on a Wednesday in order to consistently distribute the labor throughout the week. The Garolite is originally slightly transparent, as seen in the first few weeks after the initial submersion. The vertical grey portion in the middle of the plates are the mounting tabs that fit in both the submersible rig and the grooming frame. The vertical red potion is an intermediate gasket material between the back of the plate and additional standalone tabs as seen on the top of Figure 3.19. The tab that was physically and permanently mounted onto the back of the plates using stainless steel hex drive flat head screws (82 Degree Countersink Angle, 2-56 Thread Size) flush with the front surface and stainless steel locknuts. The diameter of the screw heads was 5 and were not included in the image analysis as seen in Figures A.2 and A.3 generated by the code written and run in MATLAB. If there was glare from light or overhanging algae that grew from the head of the screw, this code had the flexibility to make a polygon with more than 4 points to omit 93 Figure A.1: Illustration of the location of each of the 4 duplicate plates and how the group looked before and after scheduled grooming on that day in the proceeding tables. areas that would skew the + number, also referred to as the measure of cleanliness. The original pictures were 1728 by 1152 pixels. The polygon made manually on the fourth plate ended up being only 3 to 6% of the original amount of pixels depending on how much the mounted camera was zoomed in. 94 Figure A.2: Picture taken of Garolite plates after grooming, with points hand selected to analyze the pixels within the resulting polygon. Figure A.3: A close up of the generated polygon of the plate that was never subject to biological sampling. Drop down menus for grooming details were available on the top left. The coating type can be selected on the top middle. The "continue" button on the top right proceeds to the next full image to repeat the point selection process. 95 Table A.2: 3× Garolite. 96 Table A.2 (cont’d) 97 Table A.2 (cont’d) 98 Table A.2 (cont’d) 99 Table A.2 (cont’d) 100 Table A.2 (cont’d) 101 Table A.2 (cont’d) 102 Table A.3: 2× Garolite. 103 Table A.3 (cont’d) 104 Table A.3 (cont’d) 105 Table A.3 (cont’d) 106 Table A.3 (cont’d) 107 Table A.3 (cont’d) 108 Table A.3 (cont’d) 109 Table A.4: 1× Garolite. 110 Table A.4 (cont’d) 111 Table A.5: 0.5× Garolite. 112 Table A.6: Control Garolite. 113 Table A.7: 3× Intersleek. 114 Table A.7 (cont’d) 115 Table A.7 (cont’d) 116 Table A.7 (cont’d) 117 Table A.7 (cont’d) 118 Table A.7 (cont’d) 119 Table A.7 (cont’d) 120 Table A.8: 2× Intersleek. 121 Table A.8 (cont’d) 122 Table A.8 (cont’d) 123 Table A.8 (cont’d) 124 Table A.8 (cont’d) 125 Table A.8 (cont’d) 126 Table A.8 (cont’d) 127 Table A.9: 1× Intersleek. 128 Table A.9 (cont’d) 129 Table A.10: 0.5× Intersleek. 130 Table A.11: Control Intersleek. 131 BIBLIOGRAPHY 132 BIBLIOGRAPHY [1] X. Li and T. Kagawa, “Theoretical and experimental study of factors affecting the suction force of a Bernoulli gripper,” Journal of Engineering Mechanics, vol. 140, no. 9, p. 04014066, 2014. J. Livesey, “Inertia effects in viscous flows,” International Journal of Mechanical Sciences, vol. 1, no. 1, pp. 84–88, 1960. [2] [3] R. Brand, “Inertia forces in lubricating films,” Journal of Applied Mechanics, vol. 77, pp. 363– 364, 1955. [4] S. Savage, “Laminar radial flow between parallel plates,” Journal of Applied Mechanics, vol. 31, no. 4, pp. 594–596, 1964. [5] P. Moller, “Radial flow without swirl between parallel discs,” The Aeronautical Quarterly, vol. 14, no. 2, pp. 163–186, 1963. [6] C. Wark and J. Foss, “Forces caused by the radial out-flow between parallel disks,” Journal of Fluids Engineering, vol. 106, no. 3, pp. 292–297, 1984. [7] K. Shi and X. Li, “Optimization of outer diameter of Bernoulli gripper,” Experimental Thermal and Fluid Science, vol. 77, pp. 284–294, 2016. [8] G. Fantoni et al., “Grasping devices and methods in automated production processes,” CIRP Annals-Manufacturing Technology, vol. 63, no. 2, pp. 679–701, 2014. [9] D. Eisenberg and C. Deschamps, “Experimental investigation of pressure distribution in turbulent flow between parallel and inclined disks,” Journal of Fluids Engineering, vol. 137, no. 11, p. 114501, 2015. [10] X. Brun and S. Melkote, “Modeling and prediction of the flow, pressure, and holding force gen- erated by a Bernoulli handling device,” Journal of Manufacturing Science and Engineering, vol. 131, no. 3, p. 031018, 2009. [11] F. Possamai, R. Ferreira, and A. Prata, “Pressure distribution in laminar radial flow through inclined disks,” International journal of heat and fluid flow, vol. 22, no. 4, pp. 440–449, 2001. [12] J. Paivanas and J. Hassan, “Attraction force characteristics engendered by bounded, radially diverging air flow,” IBM Journal of Research and Development, vol. 25, no. 3, pp. 176–186, 1981. [13] J. Logue, “Apparatus for supporting articles without structural contact and for positioning the supported articles,” Aug. 11 1970. US Patent 3,523,706. [14] H. Frey, “Tool for the contact-free support of plate-like substrates,” Oct. 19 1999. US Patent 5,967,578. 133 [15] L. McIlwraith and A. Christie, “Contactless handling of objects,” Aug. 5 2003. US Patent 6,601,888. [16] M. Trommelen, “Development of a medical Bernoulli gripper,” Master’s thesis, TU Delft, 2011. [17] E. Misimi, E. Oye, A. Eilertsen, J. Mathiassen, O. Asebo, T. Gjerstad, J. Buljo, and O. Skotheim, “Gribbot-robotic 3d vision-guided harvesting of chicken fillets,” Computers and Electronics in Agriculture, vol. 121, pp. 84–100, 2016. [18] M. Wagner, X. Chen, M. Nayyerloo, W. Wang, and J. G. Chase, “A novel wall climbing robot based on Bernoulli effect,” in 2008 IEEE/ASME International Conference on Mechtronic and Embedded Systems and Applications, pp. 210–215, Oct 2008. [19] R. G. Olsson and E. C. Williams, “Contactless lifter,” April 1969. US Patent 3,438,668. [20] S. Cao, J. Wang, H. Chen, and D. Chen, “Progress of marine biofouling and antifouling technologies,” Chinese Science Bulletin, vol. 56, no. 7, pp. 598–612, 2011. [21] C. Hellio and D. Yebra, Advances in marine antifouling coatings and technologies. Elsevier, 2009. [22] K. Z. Hunsucker, G. J. Vora, J. T. Hunsucker, H. Gardner, D. H. Leary, S. Kim, B. Lin, and G. Swain, “Biofilm community structure and the associated drag penalties of a groomed fouling release ship hull coating,” Biofouling, vol. 34, no. 2, pp. 162–172, 2018. [23] M. E. Callow and J. A. Callow, “Marine biofouling: a sticky problem,” Biologist, vol. 49, no. 1, pp. 1–5, 2002. [24] P. Viett, “New hull coatings cut fuel use, protect environment.,” Currents, pp. 36–38, 06 2009. [25] J. Bendick, M. Schultz, E. Holm, and W. Hertel, “Economic analysis of underwater hull coatings and maintenance,” tech. rep., NSWCCD-63-TR-2010/17. Bethesda (MD): Naval Surface Warfare Center Carderock . . . , 2010. [26] M. Schultz, J. Bendick, E. Holm, and W. Hertel, “Economic impact of biofouling on a naval surface ship,” Biofouling, vol. 27, pp. 87–98, 01 2011. [27] S. Dobretsov, J. Thomason, D. N. Williams, et al., Biofouling methods. Wiley Online Library, 2014. [28] A. D. Coutts and M. D. Taylor, “A preliminary investigation of biosecurity risks associated with biofouling on merchant vessels in new zealand,” New Zealand Journal of Marine and Freshwater Research, vol. 38, no. 2, pp. 215–229, 2004. [29] I. Davidson, C. Scianni, C. Hewitt, R. Everett, E. Holm, M. Tamburri, and G. Ruiz, “Mini- review: Assessing the drivers of ship biofouling management–aligning industry and biosecu- rity goals,” Biofouling, vol. 32, no. 4, pp. 411–428, 2016. 134 [30] E. Georgiades, A. Growcott, and D. Kluza, “Technical guidance on biofouling management for vessels arriving to new zealand,” Ministry for Primary Industries, Wellington, ISBN, no. 978-1, pp. 77665–793, 2018. [31] K. Z. Hunsucker, E. Ralston, H. Gardner, and G. Swain, “Specialized grooming as a mechan- ical method to prevent marine invasive species recruitment and transport on ship hulls,” in Impacts of Invasive Species on Coastal Environments, pp. 247–265, Springer, 2019. [32] L. Gipperth, “The legal design of the international and european union ban on tributyltin antifouling paint: direct and indirect effects,” Journal of environmental management, vol. 90, pp. S86–S95, 2009. [33] IMO, “International convention on the control of harmful anti-fouling systems on ships,” 2001. [34] S. Nehring, “After the tbt era: alternative anti-fouling paints and their ecological risks,” Senckenbergiana maritima, vol. 31, no. 2, p. 341, 2001. [35] J. A. Callow and M. E. Callow, “Trends in the development of environmentally friendly fouling-resistant marine coatings,” Nature communications, vol. 2, p. 244, 2011. [36] M. Tribou and G. Swain, “The effects of grooming on a copper ablative coating: a six year study,” Biofouling, vol. 33, no. 6, pp. 494–504, 2017. [37] J. Hearin, K. Z. Hunsucker, G. Swain, A. Stephens, H. Gardner, K. Lieberman, and M. Harper, “Analysis of long-term mechanical grooming on large-scale test panels coated with an an- tifouling and a fouling-release coating,” Biofouling, vol. 31, no. 8, pp. 625–638, 2015. [38] J. Breglia, “Hull coatings, maintenance, and repair,” Encyclopedia of Maritime and Offshore Engineering, pp. 1–12, 2017. [39] C. Scianni and E. Georgiades, “Vessel in-water cleaning or treatment: identification of environmental risks and science needs for evidence-based decision making,” Frontiers in Marine Science, vol. 6, p. 467, 2019. [40] M. Menesses, J. Belden, N. Dickenson, and J. Bird, “Measuring a critical stress for continuous prevention of marine biofouling accumulation with aeration,” Biofouling, vol. 33, no. 9, pp. 703–711, 2017. [41] D. Oliveira and L. Granhag, “Matching forces applied in underwater hull cleaning with adhesion strength of marine organisms,” Journal of Marine Science and Engineering, vol. 4, no. 4, p. 66, 2016. [42] M. Tribou and G. Swain, “Grooming using rotating brushes as a proactive method to control ship hull fouling,” Biofouling, vol. 31, no. 4, pp. 309–319, 2015. [43] J. Hearin, K. Z. Hunsucker, G. Swain, H. Gardner, A. Stephens, and K. Lieberman, “Analysis of mechanical grooming at various frequencies on a large scale test panel coated with a fouling-release coating,” Biofouling, vol. 32, no. 5, pp. 561–569, 2016. 135 [44] N. C. Dickenson, J. S. Krumholz, K. Z. Hunsucker, and M. Radicone, “Iodine-infused aeration for hull fouling prevention: a vessel-scale study,” Biofouling, vol. 33, no. 10, pp. 955–969, 2017. [45] K. Hunsucker, C. Braga, H. Gardner, M. Jongerius, R. Hietbrink, B. Salters, and G. Swain, “Using ultraviolet light for improved antifouling performance on ship hull coatings,” Biofoul- ing, pp. 1–11, 08 2019. [46] V. Kostenko, A. Y. Bykanova, and A. Y. Tolstonogov, “Underwater robotics complex for inspection and laser cleaning of ships from biofouling,” in IOP Conference Series: Earth and Environmental Science, vol. 272, p. 022103, IOP Publishing, 2019. [47] P. W. Joyce, R. N. Cuthbert, L. Kregting, K. Crane, G. Y. Vong, E. M. Cunningham, J. T. Dick, and N. E. Coughlan, “Stay clean: direct steam exposure to manage biofouling risks,” Marine pollution bulletin, vol. 142, pp. 465–469, 2019. [48] L. Heng, N. Jun, H. Wen-jie, and L. Guibai, “Algae removal by ultrasonic irradiation– coagulation,” Desalination, vol. 239, no. 1-3, pp. 191–197, 2009. [49] S. Guo, H. P. Lee, S. L. M. Teo, and B. C. Khoo, “Inhibition of barnacle cyprid settlement using low frequency and intensity ultrasound,” Biofouling, vol. 28, no. 2, pp. 131–141, 2012. [50] A. D. Trigiani, “Ultrasonic algae control,” 2017. [51] “Algae and biofilm control by mimicking turbulence.” [52] O. P. Abioye, C. Loto, and O. Fayomi, “Evaluation of anti-biofouling progresses in marine application,” Journal of Bio-and Tribo-Corrosion, vol. 5, no. 1, p. 22, 2019. [53] S. Coetser and T. E. Cloete, “Biofouling and biocorrosion in industrial water systems,” Critical reviews in microbiology, vol. 31, no. 4, pp. 213–232, 2005. [54] T. Fjeldaas and R. Pedersen, “Robot with magnetic wheels for cleaning ship hulls,” September 2019. [55] B. Coxworth, “Aquatic robot crawls hulls to brush off barnacles,” Mar 2020. [56] D. R. Oliveira and L. Granhag, “Ship hull in-water cleaning and its effects on fouling-control coatings,” Biofouling, pp. 1–19, 2020. [57] D. Schanz, S. Gesemann, and A. Schröder, “Shake-the-box: Lagrangian particle tracking at high particle image densities,” Experiments in Fluids, vol. 57, p. 70, Apr 2016. [58] D. Schanz, S. Gesemann, and A. Schröder, “Shake-the-box: Lagrangian particle tracking at high particle image densities,” Experiments in fluids, vol. 57, no. 5, p. 70, 2016. [59] B. Wieneke, “Volume self-calibration for 3d particle image velocimetry,” Experiments in fluids, vol. 45, no. 4, pp. 549–556, 2008. 136 [60] D. Schanz, S. Gesemann, A. Schröder, B. Wieneke, and M. Novara, “Non-uniform optical transfer functions in particle imaging: calibration and application to tomographic reconstruc- tion,” Measurement Science and Technology, vol. 24, no. 2, p. 024009, 2012. [61] B. Wieneke, “Improvements for volume self-calibration,” Measurement Science and Technol- ogy, vol. 29, no. 8, p. 084002, 2018. [62] B. Wieneke, “Iterative reconstruction of volumetric particle distribution,” Measurement Sci- ence and Technology, vol. 24, no. 2, p. 024008, 2012. [63] Y. Jeon, J. Schneiders, M. Müller, D. Michaelis, and B. Wieneke, “4d flow field reconstruction from particle tracks by vic+ with additional constraints and multigrid approximation,” in Proceedings 18th International Symposium on Flow Visualization, ETH Zurich, 2018. [64] K. M. Kamensky, A. M. Hellum, and R. Mukherjee, “Power scaling of radial outflow: Bernoulli pads in equilibrium,” Journal of Fluids Engineering, vol. 141, no. 10, p. 101201, 2019. [65] I. B. Celik, U. Ghia, P. J. Roache, et al., “Procedure for estimation and reporting of uncertainty due to discretization in {CFD} applications,” {ASME} Journal of Fluids Engineering, vol. 130, no. 7, 2008. [66] K. Nakabayashi, T. Ichikawa, and Y. Morinishi, “Size of annular separation bubble around the inlet corner and viscous flow structure between two parallel disks,” Experiments in Fluids, vol. 32, no. 4, pp. 425–433, 2002. [67] V. Savkiv, R. Mykhailyshyn, P. Maruschak, L. Chovanec, E. Prada, I. Virgala, and O. Pren- tkovskis, “Optimization of design parameters of bernoulli gripper with an annular nozzle,” in Transport Means 2019: Sustainability: Research and Solutions: Proceedings of the 23rd International Scientific Conference, pp. 2–4, 2019. [68] D. R. Oliveira, L. Larsson, and L. Granhag, “Towards an absolute scale for adhesion strength of ship hull microfouling,” Biofouling, vol. 35, no. 2, pp. 244–258, 2019. [69] N. N. E. R. R. S. (NERRS)., “System-wide monitoring program. data accessed from the noaa nerrs centralized data management office website.” accessed 12 October 2012. [70] J. Cohen, “Statistical power analysis for the social sciences,” 1988. [71] K. Kamensky, A. Hellum, R. Mukherjee, A. Naik, and P. Moisander, “Underwater shear-based grooming of marine biofouling using a non-contact bernoulli pad device,” Biofouling, 2020. [72] P. Stoodley, J. Boyle, A. B. Cunningham, I. Dodds, H. Lappin-Scott, and Z. Lewandowski, “Biofilm structure and influence on biofouling under laminar and turbulent flows,” Special Publication-Royal Society of Chemistry, vol. 242, pp. 13–24, 1999. [73] S. Krishnan, C. J. Weinman, and C. K. Ober, “Advances in polymers for anti-biofouling surfaces,” Journal of Materials Chemistry, vol. 18, no. 29, pp. 3405–3413, 2008. 137 [74] P. J. Molino and R. Wetherbee, “The biology of biofouling diatoms and their role in the development of microbial slimes,” Biofouling, vol. 24, no. 5, pp. 365–379, 2008. [75] T. P. Galhenage, D. C. Webster, A. M. Moreira, R. J. Burgett, S. J. Stafslien, L. Vanderwal, J. A. Finlay, S. C. Franco, and A. S. Clare, “Poly (ethylene) glycol-modified, amphiphilic, siloxane–polyurethane coatings and their performance as fouling-release surfaces,” Journal of Coatings Technology and Research, vol. 14, no. 2, pp. 307–322, 2017. [76] A. K. M. F. Hussain and V. Ramjee, “Effects of the Axisymmetric Contraction Shape on Incompressible Turbulent Flow,” Journal of Fluids Engineering, vol. 98, pp. 58–68, 03 1976. [77] C. Cierpka, B. Lütke, and C. J. Kähler, “Higher order multi-frame particle tracking velocime- try,” Experiments in Fluids, vol. 54, p. 1533, May 2013. [78] B. W. van Oudheusden, “PIV-based pressure measurement,” Measurement Science and Tech- nology, vol. 24, p. 032001, Jan. 2013. [79] P. Van Gent, D. Michaelis, B. Van Oudheusden, P.-É. Weiss, R. de Kat, A. Laskari, Y. J. Jeon, L. David, D. Schanz, F. Huhn, et al., “Comparative assessment of pressure field reconstructions from particle image velocimetry measurements and lagrangian particle tracking,” Experiments in Fluids, vol. 58, no. 4, p. 33, 2017. [80] UNCTAD, “Review of maritime transport 2019.” [81] T. Nalepa and D. Schloesser, Quagga and Zebra Mussels: Biology, Impacts, and Control, Second Edition. CRC Press, 2013. [82] S. L. Sprecher and K. D. Getsinger, “Zebra mussel chemical control guide.,” tech. rep., U.S. Army Engineer Waterways Experiment Station, Vicksburg, MS, 2000. [83] D. Pimentel, R. Zuniga, and D. Morrison, “Update on the environmental and economic costs associated with alien-invasive species in the united states,” Ecological economics, vol. 52, no. 3, pp. 273–288, 2005. [84] R. De Leon, “The silent invasion: Finding solutions to minimize the impacts of invasive quagga mussels on water rates, water infrastructure and the environment,” 2008. [85] C. for Invasive Species Research, “Quagga zebra mussels,” 2020. 138