The Effect of Towed Array Orientation on the 3D Acoustic Picture for Sound Sources and the Vertical Ambient Noise Profile by ARCHIVES kSSACI4UFTTS INSTrTUTE OF FECHNOLOLOY JUL 3 0 2015 Arthur Anderson LIBRARIES B.S., The Pennsylvania State University (2006) S.M., Massachusetts Institute of Technology (2013) Naval Engineer, Massachusetts Institute of Technology (2013) Submitted to the Department of Mechanical Engineering in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Mechanical Engineering at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY June 2015 Massachusetts Institute of Technology 2015. All rights reserved. A uthor................... Signature redacted Department of M Certified by.. Signature redacted al Engineering May 18, 2015 rof Henrik Schmidt Professor of Mechanical and Ocean Engineering Thesi Supervisor Accepted by........ Signature redacted.... David E. Hardt Chairman, Department Committee on Graduate Students 2 The Effect of Towed Array Orientation on the 3D Acoustic Picture for Sound Sources and the Vertical Ambient Noise Profile by Arthur Anderson Submitted to the Department of Mechanical Engineering on May 18, 2015, in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Mechanical Engineering Abstract The three dimensional (3D) acoustic arrival structure of the undersea ambient noise field is important for many reasons, and can give us significant insights into the Arctic environment. For example, the anthropomorphic sound of the ice cracking along the ice edge could be used to track the location of the ice edge as it advances and retracts throughout the seasons. The noise sources could also be used as a noise source to acoustically map the bathymetry of the largely unexplored Arctic seabed. In addition, vertical arrival structure of the ambient noise field could give hints and clues that allow for improvements in both acoustic communications and target tracking. In this research, we will examine the ability of an autonomous underwater vehicle (AUV) equipped with a towed array in a virtual environment to develop an accurate 3D acoustic picture of the undersea environment. While prior towed array experiments are generally limited to the arrays being towed in a horizontal manner, here a "yoyo" maneuver is introduced. In a yoyo maneuver, the vehicle moves up and down in the water column as it traverses in order to break up the ambiguity of the measured vertical arrival structure. This thesis presents a method to measure the "verticalness" introduced into the towed array by this maneuver, and quantifies how this improves the quality of the 3D arrival structure. The results conclude that within the vehicle maneuvering limits of a Bluefin-21 AUV, a fully pitched yoyo pattern vs. a constant depth pattern results in a relative increase in the maximum beam response of a source by approximately 6.5 dB, and also decreases the 3-dB down bandwidth in the vertical direction by approximately 120. This is done without any significant losses for the bandwidth in the horizontal direction. When using a towed array to characterize a horizontally isotropic noise field, we find that within the AUV's maneuvering limits, the 3D beam response patterns are not sufficient to produce an accurate acoustic picture. To measure these fields, a vertical array is the most appropriate. 3 Thesis Supervisor: Prof. Henrik Schmidt Title: Professor of Mechanical and Ocean Engineering 4 Acknowledgments This research has been the culmination of a lot of work form a lot of different people that I'd like to thank. First of all, I'd like to thank my family, who has been there to support me over and over again over the last 5 years I've spent here at MIT. An especially important thank you to my mom, who despite not having an engineering background, has read my thesis from cover to cover to help me edit out my spelling and grammar mistakes. Next, I'd like to thank all of the members of my thesis committee. The inputs I've received from each of you have been especially invaluable. To Michael Benjamin - your work in autonomy and support of me has really inspired me to pursue further work in the field. I've learned so much about autonomy and robots through the many experiences we've shared, and I will take that with me for the rest of my career. To Franz Hover - thank you for always keeping a critical eye to my work, and pointing out how I could take it to that next step, and make the product better. To John Leonard - you're dedication to helping students succeed, despite your very busy schedule, is unparalleled. Finally, a big thank you to my thesis adviser, Henrik Schmidt. Henrik, your guidance throughout this entire endeavor has gotten this work to where it is now. Your vision was the inspiration for this work, and your insights and experience in the acoustic and autonomous communities have been invaluable to me throughout this process. Thank you. A final thank you to all my LAMSS lab mates. A big thank you to Stephanie Fried, who introduced me into 3D acoustic processing and wrote many of the preliminary versions of the code I use. Thank you Ian Katz for being there at the beginning, when I was still trying to make sense of all the LAMSS repositories. To Alon - I've so much enjoyed spending time with you and tinkering with robots, particularly during RobotX. I always knew the longer I spent time with you the more knowledge I'd get. Also thank you to Thom, Tom, Erin, Sheida, Stephanie. I know I've gotten help in so many different ways from each of you throughout the years. Being with you all and a part of this distinguished lab has been an absolute pleasure. 5 6 Contents 17 Introduction Thesis Objectives and Original Contributions . . . . . . . . . . . . 19 1.2 Thesis Outline. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 . 1.1 . 1 2 Background The Arctic Environment . . . . . . . . . . . . . . . . . . . . . . . . 23 2.2 Studying the Arctic . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 2.3 The Arctic Acoustic Environment . . . . . . . . . . . . . . . . . . . 28 2.4 IC E X 2016 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 2.5 A Sample 3D Image . . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.6 Types of AUVs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 2.7 The Bluefin-21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 . . . . . . . 2.1 General Specifications . . . . . . . . . . . . . . . . . . . . . 36 2.7.2 Equipped Acoustic Array - DURIP . . . . . . . . . . . . . . 38 . . 2.7.1 41 3.2 . 41 41 Virtual Environment . . . . . . MOOS-IvP and LAMSS 3.1.2 Vehicle Behaviors . . . . 43 3.1.3 Building a Virtual Acoustic Environment 44 3.1.4 Modeling Array Shape . 47 3D Acoustic Pictures . . . . . . 49 3.2.1 49 . 3.1.1 . 3.1 . Overview of Methods . 3 23 2D Beamforming..... 7 Creating the 3D Beam Response . . . . . . . . . . . . 51 Beam Response Patterns . . . . . . . . . . . . . . . . . . . . . 53 3.3.1 Deconvolving the Noise Field . . . . . . . . . . . . . . 55 3.3.2 Modeling the Noise Field . . . . . . . . . . . . . . . . . 56 3.3.3 Iteratively Solving for the Noise Field . . . . . . . . . . 57 3.3.4 Statistical Convergence . . . . . . . . . . . . . . . . . . 59 . . . . . . . . . . . . . . . . 60 3.4.1 Instantaneous Verticalness . . . . . . . . . . . . . . . . 61 3.4.2 Tracking Verticalness over Time . . . . . . . . . . . . . 62 3.4.3 uVertScoreKeeper . . . . . . . . . . . . . . . . . . . . . 62 . . . . . . . 3.2.3 . . Measuring Vertical Directionality Noise Field with a Point Source . . . . . . . . . . . . . . 65 4.1.1 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . 65 4.1.2 Noise Field Results . . . . . . . . . . . . . . . . . . . . . . . 67 Experimental Testing on a Simple Source . . . . . . . . . . . . . . . 74 4.2.1 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . 74 4.2.2 Results and Discussion........ .. .. .. .. ... .. . 78 . . . Quantifying Resolution . . . . . . . . . . . . 4.2 65 . 4.1 Convergence of a Noise Field . . . . . . . . . . . . . . . . . . . . . . 80 4.4 Range Finding on a Source . . . . . . . . . . . . . . . . . . . . . . . 84 4.4.1 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . 84 4.4.2 Range Finding Calculation . . . . . . . . . . . . . . . . . . . 85 4.4.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . 86 . . . . . 4.3 Horizontally Isotropic Vertical Noise Fields 91 M otivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 5.2 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 5.3 Initial Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 5.4 Exploring a One-Sided Vertical Noise Profile . . . . . . . . . . . . . . 97 5.5 Revisiting the Notch Profile with Fixed Arrays . 5.1 . . 5 50 . 4 . . . . . . . . . . . 3.4 Extension to 3D Beamforming . . . . . 3.3 3.2.2 8 100 5.6 6 Beam Responses for a Vertical Notch . . . . . . . . . . . . . . . . .1 100 Future work and Conclusions 107 6.1 Real World Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 6.1.1 O bjectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 6.1.2 Identifying the Notch . . . . . . . . . . . . . . . . . . . . . . . 112 6.1.3 Proposed Experiments . . . . . . . . . . . . . . . . . . . . . . 112 6.2 Better Ranging Algorithms . . . . . . . . . . . . . . . . . . . . . . . . 114 6.3 Ice Cracking Research . . . . . . . . . . . . . . . . . . . . . . . . . . 116 6.4 Autonomy Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 6.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 A Modeling the Pressure Field A .1 119 R ay Tracing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 A.2 Finding the Phase Difference . . . . . . . . . . . . . . . . . . . . . . . 122 A.3 Calculating Ray Pressures . . . . . . . . . . . . . . . . . . . . . . . . 123 A.4 Finding the Total Pressure . . . . . . . . . . . . . . . . . . . . . . . . 125 B Bellhop Inputs 127 C Modeling of a Towed Array 129 C.1 Modeling of an Array . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 D Spherical to Conical Transformation 133 E 3D Acoustic Pictures for a Simple Source 137 F 3D Acoustic Pictures for Noise Notch 143 G Noise Code 151 9 10 List of Figures The Yoyo Pattern Producing a 3D Picture 18 2-1 Receding Sea Ice and Arctic Routes . . . . . . . . . . . . . 24 2-2 Arctic Sea Route Navigability . . . . . . . . . . . . . . . . 25 2-3 NOAA/NASA Pathfinder Satellite . . . . . . . . . . . . . 26 2-4 T-AGOS-20 USNS Able . . . . . . . . . . . . . . . . . . . 27 2-5 Ambient Noise Levels in the Arctic for Different Frequencies 29 2-6 Ice Noise Measured as Individual Hotspots . . . . . . . . 30 2-7 ICEX 2016 Graphic . . . . . . . . . . . . . . . . . . . . . 32 2-8 Example of a 3D Acoustic Noise Profile in Shallow Water 33 2-9 Bluefin G lider . . . . . . . . . . . . . . . . . . . . . . . . 35 2-10 B luefin 21 . . . . . . . . . . . . . . . . . . . . . . . . . . 37 2-11 DURIP Array Schematic . . . . . . . . . . . . . .. . . . . 39 2-12 DURIP Array Photograph . . . . . . . . . . . . . . . . . 39 . . . . . . . . . . . 1-1 MOOS Database Tree . . . . . . . . . . . . . . . . . . . 42 3-2 IvP Helm Structure . . . . . . . . . . . . . . . . . . . . . 43 3-3 The Simulation Autonomy Code Architecture . . . . . . 45 3-4 pMarineViewer from a MOOS-IvP Simulated Environment 47 3-5 2D Array Direction . . . . . . . . . . . . . . . . . . . . . 50 3-6 3D Conical Angle . . . . . . . . . . . . . . . . . . . . . . 51 3-7 2D vs. 3D Beamforming . . . . . . . . . . . . . . . . . . 52 3-8 Distance Vector . . . . . . . . . . . . . . . . . . . . . . . 53 3-9 3D Beam Examples . . . . . . . . . 3-1 . . . . . . . . . . . . . 11 54 3-10 Iterative Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 3-11 Iteration Process Diagram . . . . . . . . . . . . . . . . . . . . . . . . 58 3-12 Verticalness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 3-13 uVertScoreKeeper Output . . . . . . . . . . . . . . . . . . . . . . . . 63 4-1 Simple Source Experimental Setup . . . . . . . . . . . . . . . . . . . 67 4-2 Simple Source Results for Period 300 . . . . . . . . . . . . . . . . . . 68 4-3 Simple Source Results for Period 1000 . . . . . . . . . . . . . . . . . 70 4-4 Simple Source Results for a Constant Depth . . . . . . . . . . . . . . 70 4-5 Path Plots for Yoyo vs. Constant Depth Maneuvers . . . . . . . . . . 71 4-6 Vertical Noise Profiles at the Source Location for Multiple Levels of Array Verticalness 4-7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 Horizontal Noise Profiles for Multiple Levels of Array Verticalness Looking at a Simple Source . . . . . . . . . . . . . . . . . . . . . . . 73 4-8 Path Distance Geometry . . . . . . . . . . . . . . . . . . . . . . . . . 75 4-9 Relationships of Verticality to Acoustic Picture Resolution . . . . . . 78 4-10 Series of Snapshots of the Generated Noise Field with Increasing Amounts of D ata . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 4-11 Correlations of the Noise Field Generated as a Function of Time . . . 83 4-12 Range Finding Experimental Setup . . . . . .... . . . . . . . . . . . 84 4-13 Range Finding Geometry . . . . . . . . . . . . . . . . . . . . . . . . . 86 4-14 Range Finding Acoustic Picture - Yoyo Pattern . . . . . . . . . . . . 87 4-15 Range Finding Acoustic Picture - Constant Depth . . . . . . . . . . . 87 4-16 Range Estimation - Vertical Arrival Structure Comparison . . . . . . 88 4-17 Range Finding - Experimental Range Measurements and Uncertainty 89 5-1 Vertical Notch Graphic . . . . . . . . . . . . . . . . . . . . . . . . . . 92 5-2 Vertical Notch Graphic . . . . . . . . . . . . . . . . . . . . . . . . . . 93 5-3 Noise Notch as Measured by a Vertical Array 94 5-4 3D Acoustic Pictures by a Towed Array of the Vertical Notch . . . . 95 5-5 Vertical Notch as Measured by 3 Towed Array Experiments . . . . . . 96 12 . . . . . . . . . . . . . 5-6 Zoomed View of Vertical Profile for 3 Towed Array Experiments.. . 97 5-7 One Sided Ambient Noise Profile as Measured by a Vertical Array 98 5-8 One Sided Ambient Noise Profile as Measured by a Towed and Two . Fixed Diagonal Arrays . . . . . . . . . . . . . . . . . . . . . . . . . . 5-9 99 One Sided Ambient Noise Profile - Vertical Noise Picture Comparison for 3 Experim ents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 5-10 Towed Array and Fixed Diagonal Arrays in a One-Sided Horizontally Isotropic Noise Field . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 5-11 Beam Responses for a Horizontally Isotropic Vertical Noise Field at Different Array Tilts . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 5-12 Maximum Normalize Beam Response vs. Different Array Tilts for a Horizontally Isotropic Vertical Noise Field . . . . . . . . . . . . . . . 103 5-13 Conical Beams measuring a a Horizontally Isotropic Noise Field . . . 104 6-1 OASES Predicted Beam Pattern from Ice Cracking Noise . . . . . . . 110 6-2 Ranging Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . Il 6-3 Bottom Response Objective 6-4 Ranging Using Multiple Paths . . . . . . . . . . . . . . . . . . . . . . 115 6-5 Ice Cracking Patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-6 Example Arctic Autonomy Utility Function A-1 Snell's Law . . . . . . . . . . . . . . . . . . . . . . . 112 116 . . . . . . . . . . . . . . 117 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 A-2 Ray Tube . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 . . . . . . 131 C-1 Model of the Individual Forces on a Towed Array Element D-1 Transformation from Spherical to Conical Coordinates . . . . . . . . 134 E-1 Simple Source Results for Period 300 . . . . . . . . . . . . . . . . . . 138 E-2 Simple Source Results for Period 500 . . . . . . . . . . . . . . . . . . 138 E-3 Simple Source Results for Period 600 . . . . . . . . . . . . . . . . . . 139 E-4 Simple Source Results for Period 700 . . . . . . . . . . . . . . . . . . 139 E-5 Simple Source Results for Period 1000 13 . . . . . . . . . . . . . . . . . 140 . . . . . . . . . 140 E-7 Simple Source Results for Period 2000 . . . . . . . . . . 141 . 141 . E-6 Simple Source Results for Period 1500 E-8 Simple Source Results for Constant Depth . . . . . . . . . . . . . F-2 Vertical Ambient Noise Field Results for Period 300 . . . . . . . . . 144 F-3 Vertical Ambient Noise Field Results for Period 500 . . . . . . . . . F-4 Vertical Ambient Noise Field Results for Period 600 . . . . . . . . . 145 F-5 Vertical Ambient Noise Field Results for Period 700 . . . . . . . . . F-6 Vertical Ambient Noise Field Results for Period 1000 . . . . . . . . 146 F-7 Vertical Ambient Noise Field Results for Period 1500 . . . . . . . . 147 F-8 Vertical Ambient Noise Field Results for Period 2000 . . . . . . . . 147 . . . . . . . . F-1 Vertical Ambient Noise Field Results for a Vertical Array 144 145 146 F-9 Vertical Ambient Noise Field Results for 450 Tilted Array. . . . . . . 148 F-10 Vertical Ambient Noise Field Results for 600 Tilted Array. . . . . . . 148 F-11 Vertical Ambient Noise Field Results for 75' Tilted Array. 149 14 List of Tables A Brief Description of the Key Code Elements . . . . . . . . . . . . 46 4.1 Simple Source Experiments List . . . . . . . . . . . . . . . . . . . . 66 4.2 3D Acoustic Resolutions for Simple Source Experiments . . . . . . . 73 4.3 Experiments Conducted on Verticality . . . . . . . . . . . . . . . . 77 4.4 Full Verticality Experimental Results . . . . . . . . . . . . . . . . . 79 5.1 Summary of Towed Array Experiments for Measuring a Vertical Notch 94 6.1 Objectives for the ICEX Experiment . . . . . 3.1 . . . . . . . . . . C.1 Physical Variables used in Modeling the DURIP Array 15 . . . . . . . . 108 . . . . . . . . 132 16 Chapter 1 Introduction For many reasons, the three-dimensional (3D) arrival structure of the undersea ambient noise field is important for research and development in the acoustic community. Three dimensional arrival structures can be used to estimate the beam noise of an array. This is important when estimating the performance of that array as a scientific or Navy asset, and also there are clues in the vertical arrival structure that give hints and clues as to the propagation paths for the noise. While the ideal measurement tool to understand the 3D arrival structure would be a high-resolution volumetric array sonar system, these systems are prohibitively expensive and generally not available. While not ideal, towed arrays have been and can be used to estimate the 3D ambient noise field [38]. In typical 2D beamforming on a line array, the azimuthal direction of a noise source is by measuring the phase angle of the signal as it approaches each element. But this doesn't present the whole picture. Rather than at angles relative to the array on just the horizontal plane, the actual beam patterns are conically symmetric about the axis of the array. It has been shown that even with small array tilts, a measured 3D noise field can be established using an iterative algorithm which deconvolves the noise from the beam pattern. With the new technology of autonomous underwater vehicles (AUVs), large towed arrays no longer need to be deployed from a ship, and this gives the opportunity that towed array may be given verticalness, or array tilt, introduced by an AUV maneuvering up and down in the water column in a "yoyo" 17 3 'R 3D Noise Rosette *4,' *%* / 4, I * i I? a S I I b ~* Figure 1-1: A set of pictures that illustrates the new yoyo pattern taken by an AUV vs. the traditional studies done using a horizontal path. The yoyo pattern helps to rapidly break up the inherent ambiguity in the vertical arrival structure [32]. 18 pattern as it moves also on the horizontal plane. In theory, this should allow for rapidly breaking the vertical and horizontal angle ambiguity and present a shorter time scale for noise statistics, thus allowing for better tracking of temporal events [32]. In this thesis we will study this idea in a virtual acoustic environment, and examine the improvements to the 3D noise field that are introduced by these maneuvers. 1.1 Thesis Objectives and Original Contributions In this thesis I explore the 3D "Sound Scape" produced for various environments using a towed array that is towed not only in a circular pattern in the horizontal plane, but also up and down in the water column in what we term a "yoyo" pattern. This research is unique because while 3-dimensional beam forming has been performed on arrays in the past, those arrays were only towed horizontally, and not in any sort of vertical manner. This research is done in preparation for experiments planned in the Arctic for the Spring of 2016. I will present here a method I've developed for both quantifying the amount of "verticalness" in a towed array, as well as quantifying the improvement in the resulting noise as a function of that verticalness. I do this by leveraging existing acoustic and autonomy software to run experiments in a virtual environment in order to characterize how verticalness affects a 3D acoustic picture. In this thesis I make two other original contributions to the field. First, I demonstrate that using the vertical arrival structure of the noise field in a 3D acoustic picture as measured by a towed array, one can estimate the range to that source. Second, I examine in depth the ability of a towed array to measure a horizontally isotropic ambient noise field. The motivations for measuring these types of fields are related to the unique Arctic sound speed profile and discussed in Chapter 5. The original contributions are summarized as follows: Quantifying Verticalness vs. Noise Directionality Improvement. In the course of this research I've developed a method to quantify the "verticalness" in a towed array over time, and used this as a metric to predict the signal excess and resolution improvement of the resulting 3D noise directionality. 19 Ranging to a Source By using the vertical arrival structure of sound from a source in a 3D acoustic measurement, I demonstrate that it is feasible to estimate the range to a target using ray tracing techniques. Resolving Horizontally Isotropic Vertical Noise Fields In this research I conduct an in depth study into the inherent ability of a towed array, with and without verticality, to resolve a horizontally isotropic vertical noise field. 1.2 Thesis Outline Chapter 2 discusses the background of the project. This includes a background on the Arctic environment, possible methods for studying it, and a literature review on the Arctic's acoustic properties. In Chapter 2, I also discuss the upcoming experiment that was the inspiration for this research (ICEX 2016) and the particular AUV and array that will be used during this experiment. Chapter 3 is an overview of the methods and processes used, covering 3 main topics. The first of these is creation and structure of the virtual environment. This includes a description of the software architecture used, and as well a discussion on how we can simulate acoustic propagation. The second topic is how I define the verticalness of an array, and the third and final topic is a description of the mathematics involved in 3D beams and beam response patterns, and how the noise field can be deconvolved from those beam responses. Chapter 4 deals with experiments involving a simple source, and covers the first two of the three contributions listed in Section 1.1. It begins with a small series of experiments in which the verticalness of an array is compared against the resolution of the resulting noise field. Next, it expands into a larger design of experiments, where a full range of experiments 'with different vertical scores are tested, compared, and discussed. Also in this section, I show through experimentation the data requirements for picture convergence, and the feasibility of determining the range to a source using 20 the vertical noise arrival structure. Chapter 5 is the exploration into quantifying how a towed array's vertical score will affect the 3D noise picture picture when measuring horizontally isotropic vertical noise fields. This covers the third of the three contributions. At the beginning of the chapter, I discuss the motivation for measuring these types of fields, and then discuss a series of results of a number of experiments in measuring these types of ambient noise fields. Finally, I take an in-depth look into why towed arrays are inherently poor at resolving these types of fields by closely examining the beam response patterns. Chapter 6 consists of suggestions for future work and conclusions. Included in the future work is a list of objectives and experiments that can be used to test the results presented here with real world experiments, as well as several other suggestions for follow on work in different (but related) subject areas. 21 22 Chapter 2 Background 2.1 The Arctic Environment The Arctic Ocean is comprised of nearly 5.4 million square miles, which is roughly 1.5 times the size of the United States. It is well known that the extent of the polar sea ice in the arctic region is retreating from year to year. Over the coming decades, the extent of multi-year Arctic sea ice extent is expected to recede. The region will become increasingly more widely used, as commercial interests seek to use trade routes that were previously inaccessible, and nations seek to assert their rights over the region's abundant oil and gas resources. The increased activity will manifest itself in the form of resource extraction, fishing, tourism, shipping, and Naval activity [10]. Figures 2-1 and 2-2 show the effect of the retreating sea ice on the commercial routes through the region. The blue bars represent times for open water conditions, or less than 10% ice coverage, while the yellow bars represent the shoulder season, defined by less than 40% ice coverage. Currently, the Bering Strait only sees open water conditions for approximately 145 days with a shoulder season of 40 days each year. By the year 2030, that time is expected to increase to 190 open water days, with 70 days of shoulder. Other routes will experience similar effects. The Northern Sea Route, which currently only experiences 15 days of open water conditions and 40 days of shoulder, will see open water times on average of 50-60 days with 70 days of shoulder season by 2030. The Northwest Passage and the Transpolar routes, which 23 Figure 2-1: This figure shows the anticipated future transit routes superimposed over the expected sea ice yearly minimum (also called multi-year ice), projected to year 2030 as anticipated by the United States Navy [10]. 24 Arctic Sea Route Navigability Bering Strait (BS) Transp 440 Vesel zou m -aa..a Ro te(TPR) 0 Vesses 20M0 2020 Northwest Pas Northen Sea Route (NSR) (5.225 HIM) (4,7401NM) 44 Veselsesel 2012 2025 1 1 > Ves 202S5 A LM.0A44 LOU Oct IA$"A 201 (NWP) kMbdlee saftoo 40% sea ice Shoulder Season: 10-40% sea Ice Open Water: < 10% sea Ice Figure 2-2: Projected navigability of the Arctic transit routes projected to year 2030 as anticipated by the Office of Naval Intelligence [10]. as of 2012 do not even see shoulder seasons, could each experience on average 30-35 days of open water per year, both with a significant shoulder seasons. The Arctic region also holds a vast amount of mineral resources with significant wealth potential. Estimates for the hydrocarbon reserve exceed $1 trillion dollars. The Alaskan Arctic alone is estimated to hold 29.9 billion barrels of oil, 221 trillion cubic feet of natural gas, and 5.9 billion barrels of natural gas liquids [8]. In addition, the Arctic environment may hold significant amounts of mineral resources, including but not limited to iron ore, zinc, nickel, graphite, and palladium [10]. The new open transit routes, along with the new potential for the extraction of resources, means that the Arctic region has regained its importance since its decline at the end of the Cold War. 2.2 Studying the Arctic Because understanding the Arctic region is of critical importance, we must seek ways to understand it. Important areas for research include persistent ice edge surveillance 25 Figure 2-3: The Pathfinder satellite is a joint project between NOAA and NASA [24]. It is equipped with an Advanced Very High Resolution Radiometer (AVHRR), which allows it to measure sea surface temperatures and approximate sea ice thickness. for the reasons of safe navigability, seabed mapping, discovering tools for resource discovery, and fundamental climate change research. So the question arises, how do we explore these areas, and what platforms could we use? The first area, ice edge surveillance, is particularly important. Naval operations are likely to expand with newly discovered resource competition, and other oceangoing vessels will use the North East and North West passages as regular shipping lanes during more months of the year, increasing the likelihood of loss of life due to collisions with floating icebergs and ice floes [29]. Persistent ice edge tracking, however, proves a difficult problem. Satellite or aerial vehicle imagery would seem to be a good solution, however this proves difficult in practice. Persistent monitoring aircraft are prohibitively expensive, and there are very few satellites designed to pass over the arctic because geostationary satellites must be located along the equator in order to remain in orbit. Another issue is that satellites have difficulties measuring ice thickness. Some satellites have mounted altimeters, but the accuracy of this data is fairly low, and generally only good for long time scales of a month or longer. There are some satellites, such as the NOAA Pathfinder satellite [24] (see Figure 2-3), that can measure ice thickness via an Advanced Very High Resolution Radiometer (AVHRR) using knowledge of microwave emissivity. This, however only 26 Figure 2-4: The USNS Able, or T-AGOS-20 is a SURTASS ship in service to the US Navy's Military Sealift Command [25]. Platforms such as this can deploy many different types of sensors to study the ocean environment, but are expensive to operate. works when the satellite is overhead and no cloud cover [12] [40]. Finally, aerial monitoring gives no hints as to the resources that lie in and below the Arctic seabed. An alternative solution is deploy ships and submarines to monitor the area. Figure 2-4 shows a picture of the USNS Able, an oceanographic survey ship in the service of the United States Navy's Military Sealift Command [25]. Able is equipped with a number of different sensors, including a Surveillance Towed Array Sensor System (SURTASS), a large towed array deployed from the back which can provide much information about the environment, as well as other information in terms of seabed mapping, resource discovery, and general climate change research. The presence of ships and submarines in the Arctic however, even smaller ones, could be very costly. Operating these platforms is not inexpensive, and neither are the personnel that man them. One last solution is research using autonomous underwater vehicles, or AUVs. AUVs, like ships and submarines, can be equipped with a number sensors, including acoustic arrays, that are all very useful in gathering information for Arctic research. And precisely because they are unmanned, they could prove to be a persistent, rel27 atively inexpensive method for doing so. The main problem with AUVs however, is that they represent a relatively new and risky technology, and much research about them before they can be reliably deployed to measure the environment in a useful way. The feasibility of using AUVs and associated technology to evaluate the environment is one of the main focuses for this research. 2.3 The Arctic Acoustic Environment The primary focus in this research is studying the Arctic through the use of acoustics as measured by AUVs, and in particular, the three-dimensional (3D) soundscape, which will be discussed later. There are two important differences from most other ocean environments that present themselves in the Arctic. The first is the unique sound speed profile that is the result of salinity changes near the surface, and the second is the ice cracking noise at the ice edge. There have been some studies that give us hints as to the noise content we will see. The biggest area of interest is along the arctic ice edge, because there is constant moving and cracking which creates a significant amount of noise. This is noted in Diachok's paper [11], where he measured the ambient noise levels at distances away from the marginal ice zone (MIZ) to see the effects of the acoustic environment at different frequencies. His results can be seen in Figure 2-5. We can see from this figure that the spectrum levels for the different frequencies increase significantly at and near the MIZ, for the frequencies of 100 Hz, 315 Hz, and 1000 Hz. At these frequencies, Diachok observed ambient noise levels up to 20 dB (re (1ptbar) 2 /Hz ) higher than noise levels far under the ice field, and 12 dB higher than noise levels in the open water. For a diffuse ice edge, these numbers are somewhat less. These effects can be observed for long distances, especially at lower frequencies such as 100 Hz, where we see the effects extend more than 50 km. Diachok concluded that the noise mechanisms were probably related to wave and swell interactions with individual ice floes, in a continuous distribution along the ice edge. Further studies however, have indicated that this is not necessarily the case. In 28 A A-0I 1% 400 - %im1q0 IM*d b -30 /Ao - A YAA 00VO dip -40 1- -I- -s0- -o' i it0 I s0 I I I I I IOMCE I I I 40 0 40 I MET) Figure 2-5: A measure of the ambient noise levels in the Arctic for different frequencies at different distances from the ice edge, as measured in Diachok's paper [11]. 29 30' DiSTANT ICE EDGE DOECTON 10' C 40' v0 *8 11. Figure 2-6: Yang's experiments [39] showed that the ice edge noise, rather than being a continuous line source, is actually the result of separate ice cracking "hot-spots", which can be easily resolved at distances of 15 km away. 30 1983, a ship with a towed array of 64 hydrophone elements and an aperture of 290m, listened to the noise field over a 24 hour period north of Jan Mayen Island. The ship tracks, as well as some of the results of the experiment, are illustrated in Figure 26 [39]. While it was expected that the noise from an ice field came from a continuous line source across the ice front, this experiment showed that the noise was actually arising from a few localized individual ice cracking "hot-spots" along the ice edge, which persisted for nearly 1 day, with a nominal separation of 50 km. The arrowed lines in this figure represent the track of the ship, while the ice edge is represented by the solid black lines. The points labeled A, B and C represent the separate areas where ice cracking noise could be observed. 2.4 ICEX 2016 In the Spring of 2016, the US Navy is planning an ICe EXercise (ICEX), to which it has invited MIT to participate for research with AUVs. This invitation represents a renewal in funding for research in the Arctic Environment. During this exercise, a team from the Laboratory for Autonomous Marine Sensing Systems (LAMSS) will set up an ice camp somewhere on the polar ice cap above the Alaskan peninsula. From there, they will deploy an AUV equipped with an acoustic towed array to learn what sort of information can be learned about the environment through the use of acoustics, or in particular the three-dimensional (3D) soundscape. The lessons that could be learned from this experiment could prove significant. Natural noise from the ice cracking could be used as a natural sound source. This sound could not only be potentially utilized to track the ice edge, but for all significant areas of research. The noise bouncing off the bottom could give hints as to the depth of the arctic seabed, and also possibly the resources below. One method for imaging the acoustic environment, will be the use of 3D acoustic pictures, or the 3D soundscape. While currently no analysis on the soundscape exists for the Arctic, we have learned information from the acoustic picture in other environments. The next section will give just such an example. 31 Marginal Ice Zone (MIZ) Open Water Figure 2-7: During the ICEX 2016 experiment with the US Navy, MIT will launch an AUV equipped with a passive towed array in order to study the Arctic acoustic environment. The goal is to use anthropomorphic ice cracking noise as a natural source to obtain information to include bathymetry data, sub-seabed composition, and ice edge location. 2.5 A Sample 3D Image Using the array's information during ICEX 2016, the team from LAMSS hopes to create a number of acoustic "pictures" which can give information about the surrounding environment. Figure 2-8 is an example of one such image. In 1993 the SACLANT Undersea Research Center (SACLANTCEN) ran a number of towed array experiments listening to the ambient noise in shallow water areas using their ship (R/V Alliance) as a towing platform. R.A. Wagstaff, in a paper [38], looked at several of these experiments, and used an algorithm to develop a series of 3D acoustic images using the methods that will be described later in this thesis (see Section 3.3.1). The array in this case had an angle relative to the horizontal plane of approximately 1-2', and the image was developed using a processing frequency of 100 Hz. The image shows the full sound picture, with vertical angle on the y-axis and azimuthal angle on the x-axis. A lot of information can be gained from this image by comparing to the surroundings to the image. For example, the noise arriving at near 0' vertical angle and approximately 1500 azimuthal comes from a major northern 32 Distant Shipping Nearby Island Island Blockage Noise - I I Noise go 60 1 -30 -490 0 4 lei I so Upslope propagation paths, allowed by bathymetry 1351|180 225 I S AZ~r9WM4iE (dbg) I7 I 270 315 I- -- o am0 Major Northern Access for coastal steamers Figure 2-8: An 3D acoustic picture analysis of a shallow water area from Wagstaff and Newcomb's 1997 paper [38]. 33 access for coastal steamers to a nearby port. From azimuthal angles 160' to 185', the sound is mostly blocked by a series of small islands and relatively shallow bathymetry. From azimuth's 245' to 330', the noise is arriving from a chain of nearby islands and shipping which is bouncing off the sea bottom and the surface many times, which gives rise to the vertical noise seen coming from these directions. Lastly, we can see more noise from coastal steamers from the azimuths of approximately 25' to 60', where the verticalness of the noise seen from the bathymetry from that direction is indicative of the significant upslope propagation paths in that direction. 2.6 Types of AUVs Here we will discuss the different types of AUVs available for general research today, their capabilities, and a discussion on the appropriate vehicle choice for the ICEX experiment. While there are many different models of AUVs available both commercially and militarily, the AUVs can generally be divided into three categories: (1) gliders, (2) torpedo shaped AUVs, (3) and non-torpedo shaped AUVs. A glider is an AUV that uses small changes in buoyancy in conjunction with a set of wings to convert its vertical motion into horizontal motion. Examples of these include: e The Webb Research Corporation's Slocum glider (named after the first person to sail around the world solo)[35], e The University of Washington's Applied Research Laboratory's "Seaglider" [1] 9 The Scripps Institution of Oceanography "Spray" [27] (Joshua Slocum's boat when he sailed around the world, see Figure 2-9), which is sold by the Bluefin Corporation Gliders are advantageous in many ways to other AUVs in that they require very little power to propel themselves through the water. They use very small changes in buoyancy in conjunction with their wings to convert vertical motion to horizontal. 34 Figure 2-9: A Picture of a Bluefin "Spray" Glider[9] The result is that they can conduct very long endurance missions, although they are less maneuverable. Some, such as the Slocum Glider, can travel at 0.4 m/s and operate in the water for 4-12 months. However, because of their slow speed and generally small design size (they are typically designed to be handled by 1-3 people) they cannot handle a particularly large payload, and therefore are limited on the types of sensors they can carry. They operate at a variety of designed max depths (200 m to 6000 m depending on the type). Another advantage is that they are generally less expensive than other types of AUVs. For long term surveillance in the Arctic these types of AUVs could prove to be the most useful. The second type of AUVs are actively propelled and torpedo shaped. Examples of these AUVs are: " Bluefin 9", 12", and 21" from Bluefin Robotics [6]. " Remote Environmental Monitoring UnitS (REMUS) 100, 600, 3000, and 6000 from the Wood Hole Oceanographic Institute (WHOI) [18]. " Gavia - in types of (1) Offshore Surveyor, (2) Defense, (3) and Scientific classes [13]. These types of AUVs right now are considered the work horses of the community, and are probably the most common type of AUV. One disadvantage is that they tend to have shorter endurance ranges (on the order of hours or days) than the gliders, however this can be extended if power saving methods are used, such as limiting propulsion use and keeping the computing power to a minimum. They use active 35 propulsion generally by means of a propeller and a small fin, and the option to carry larger payloads and more sensors than a glider, such as a towed array is an advantage. These sorts of AUVs are particularly good for experiments that occur on the order of hour or days, and can be deployed from a mother ship when doing so. The last type of AUV is non-torpedo shaped AUVs. For a variety of reasons, AUVs may be designed to a different shape. This may be for the purpose of carrying unique payloads, such solar panels, or so that they can use more control surfaces or propellers for increased maneuverability. These types of AUVs are typically only used when their payloads require. Examples of non-torpedo shaped AUVs include: 9 The Autonomous Benthic Explorer (ABE) at WHOI [17] e Solar AUV (SAUV II) [14] 9 Hovering Autonomous Underwater Vehicle (HAUV) [28] 2.7 The Bluefin-21 This research will focus on the Bluefin-21 torpedo shaped AUV because of its ability to haul a towed array, it is actively propelled, and it is best suited for the ICEX experiment. Named for its 21" diameter hull, the Bluefin-21 has been used for many types of missions, including offshore surveying, search and salvage, archaeology and exploration, oceanography, mine countermeasures, and handling unexploded ordnance [5]. The Bluefin-21 that will be used for the ICEX is maintained by LAMSS and is named "Macrura". 2.7.1 General Specifications Macrura is approximately 5 m long, weighs 785 lbs (dry), and contains nine lithium polymer batteries which power its propulsion system. The top speed of the vehicle is 4.5 knots, with an endurance time of 25 hours at 3 knots. Launch and recovery of the vehicle is fairly straightforward: the launching platform must be equipped 36 Figure 2-10: A Picture of a Bluefin-21 on deck during a transit [6]. This is the vehicle that will be used for the ICEX experiment, and an example of an actively propelled, torpedo-shaped AUV. with an A-frame, which launches the vehicle from a lift point mid-frame on the body. Included in the on board navigational sensors are an Inertial Navigation System (INS), a Doppler Velocity Log (DVL), Sound Velocity Sensors (SVS), and for when surfaced, a Global Positioning Satellite (GPS) system. These sensors together allow the vehicle to maintain real-time circular area probable (CEP) 50 accuracy of <0.1% of distance traveled (DT) [5]. This means that 50% of the time the vehicle's navigation systems will be accurate within a circular range of 0.1% of DT. The primary form of communication while the vehicle is underwater is via underwater acoustics. While the AUV is underwater, the ship or controlling platform for the AUV receives data and updates every couple minutes for near real-time information updates. Acoustic communication, or accoms, is handled by a system developed by MIT under the open source software project Gobysoft.org [33]. In this communications paradigm, the transmissions of each node of an acoustic communications network (such as VLA buoys, controlling ship, or the AUV), are given time windows in which they can send their data. The data is encoded on the sending side and decoded on the receiving side in order to reduce bandwidth, because the data transfer rate is limited to about 1800 bps [30]. The reason for this is that acoustic communications must use lower frequency ranges so that the signal does not attenuate after just a 37 few meters, even though the bandwidth at those frequencies is limited. While on the surface, the Bluefin-21 is able to communicate through the RF or iridium antennas on board for much higher data transfer rates [5]. Equipped Acoustic Array - DURIP 2.7.2 The array to be used in the ICEX experiment, as well as the one modeled in this research, is the Defense University Research Instrumentation Program (DURIP) array. The DURIP array has a total of 32 hydrophones, with a total acoustic aperture of 30 m. 21 of the hydrophones are spaced at 1.5 m apart, while the remaining 11 are nested in the center, as shown in Figure 2-11. The hydrophones have a sensitivity of -176 dB/V/pPa, with a sensitivity tolerance of +/- 1 dB [26]. The optimal frequencies are measured when the acoustic wavelength is equal to half the spacing d of the array, such that: d= 2 A = c/f 2 (2.1) where A is the wavelength of the acoustic wave, c is the speed of sound in the water, and f is the optimal frequency for the array spacing. Using these equations, we can calculate the optimal frequency for an array spacing of d as: ft C = (2.2) In our acoustic array, we essentially have two arrays with a spacing of 0.75 m and 1.5 m. Using a sound speed c = 1500 m/s, this array can be operated for frequencies up to 1000 Hz and 500 Hz, respectively for the two apertures. Lower frequencies can also be measured by choosing elements that are spaced further apart, however, the resolution on your aperture decreases. This is why larger arrays are needed to resolve lower frequencies. We can say the array has a bandwidth of approximate 100 Hz to 1000 Hz. Above 1000 Hz, we can expect to see aliasing, as the wavelength will become significantly smaller than the array spacing, and the noise directionality becomes less reliable. 38 DURIP Array 20 m 36 m 20 m * 0 30 m Aft Compass Aft Depth Sensor Forward Compass 32 Pressure Sensor Elements * Drogue 1.5m spacing 1.5m spacing 0.75m spacing Tow Line Data Acquisition System (DAS) macrura Figure 2-11: A general schematic of each of the DURIP array. Beginning at the attach point of the AUV and working from right to left, the full array consists of a 20 m tow line, a data acquisition system (DAS), a forward compass, 32 pressure elements, an aft compass, an aft pressure sensor, and finally a 20 m drogue. The part of the array that contains the actual elements is what's called a "nested" array, and has two spacing sizes, 0.75 m and 1.5 m. These spacings allow it to cover a frequency range anywhere from 100 to 1200 Hz. Figure 2-12: The complete DURIP array at LAMSS [7], which will be used in the ICEX experiment in the spring of 2016, and modeled for simulation in this research. 39 Each of the hydrophones are sampled at a rate 12 kHz, and data is collected on each hydrophone in 2 second intervals. The array also has some other features - on both the forward and aft ends a compass is mounted to help determine array positions, and another pressure sensor is located just aft of the acoustic portion of the array. The pressure and compass readings are sampled at a lower data rate than the hydrophones [26]. Figure 2-12 shows a photo of the array. 40 Chapter 3 Overview of Methods This chapter discusses the methods used to create the virtual environment, and in particular the tools used to process the resulting data from that environment. This includes beamforming techniques in two and three dimensions, how a noise field is modeled, and how we can extract a pseudo-stationary noise field from the beam response patterns. This section also discusses how I quantify the "verticalness" in the array. 3.1 3.1.1 Virtual Environment MOOS-IvP and LAMSS There are many different autonomy middlewares that are used for robotic applications. Of these, some of the more common middlewares include Mission Oriented Operating Suite (MOOS) [2] [3], Robot Operating System (ROS) [15], and Lightweight Communications and Marshalling (LCM) [23]. The code architecture used in this research is based on the open source project MOOS-IvP, from the MOOS autonomy middleware. Launched originally at MIT in 2005, MOOS-IvP includes more than 150,000 lines of code and 30+ core applications dedicated to controlling marine vehicles, mission planning, debugging, and post-mission analysis. The software has been run on more 41 MOGS Appkedon Figure 3-1: A MOOS community is a collection of applications, each publishing and subscribing to variables published to a central MOOS database. A MOOS community typically operates on a single vehicle or computer. Image courtesy of Mike Benjamin [4]. than a dozen different platforms and is being used at the Office of Naval Research (ONR), the Defense Advanced Research Projects Agency (DARPA), and the National Science Foundation programs at MIT. MOOS-IvP can be used for a simulated environment, or for fielding the vehicles in a real environment. MOOS contains a core set of modules that provide a middleware capability based on a publish-subscribe mailing architecture. Processes in the MOOS database are defined by what messages they write (publishing), and what messages they consume (subscribing). The key idea for MOOS is that it allows for applications that are mostly independent, and that any application can be easily replaced or upgraded with an improved version. The only requirement for the improved application is that only its interface matches [4]. Figure 3-1 shows a MOOS community, which typically runs on a single machine, and the structure of processes. MOOS communities set-up on on multiple platforms are also capable of communicating with one another. On top of the MOOS-IvP architecture, there is another architecture known as the LAMSS environment - where LAMSS stands for the Laboratory for Autonomous Marine Sensing Systems. The LAMSS repository is an additional build of MOOS modules that have been structured to allow easy transfer between simulation of an autonomy system on a vehicle and moving that autonomy system to an actual experiment. In simulation, LAMSS already has a set of tools that allow a use to simulate 42 1vP Function 1vP Function Informiation| Decision Figure 3-2: The IvP helm is a single MOOS application. It uses a behavior-based architecture, in which a mode structure is used to determine which behaviors are active. Each of the active behaviors are then reconciled using a multi-objective optimization solver, or the IvP solver. The resulting decision is then published to the MOOS database. Image courtesy of Mike Benjamin [4]. an acoustic environment. Many of these simulations have been leveraged for the ideas presented in this research. 3.1.2 Vehicle Behaviors IvP (Interval Programming) is a single application that runs inside the MOOSDB. IvP uses a behavior-based architecture for implementing autonomy. These behaviors are distinct modules, each of which are dedicated to a specific aspect of autonomy, such as following a set of waypoints or for collision avoidance. If multiple behaviors are active, the IvP uses a solver to reconcile the desires of each behavior using an objective function [4]. Figure 3-2 shows how the IvP system is structured. There are two relatively simple behaviors to control the vehicle in this research. The first behavior, BHViLoiter, directs the vehicle to move along the points of a polygon. These points can be specified individually, or as a collection of points with a center and a radius. To create a circular pattern using this behavior, one simply increases the number of points in the loiter polygon so that it approximates a circle. The second behavior, BHVSmartYoyo, is used to control depth. This is a behavior 43 available inside the LAMSS repository, and by specifying a yoyo length period in meters, the AUV is directed to move up and down in the water column between two specified depths. By controlling the period of the yoyo pattern, we can control the pitch of the path taken by the AUV, and of the towed array behind it. In a real world experiment, many other behaviors would need to be utilized to control the vehicle safely, specifically including such behaviors as an emergency surface behavior, a minimum altitude behavior so that the vehicle doesn't ground, a travel to waypoint behavior, and a set of behaviors for vehicle recovery. Fortunately, many of these behaviors exist in the standard MOOS-IvP and LAMSS repositories, and they have been tried and tested in both simulation and real world experiments. 3.1.3 Building a Virtual Acoustic Environment As discussed in the previous section, the code base used in this research utilizes existing code built in the MOOS-IvP and LAMSS repositories. Figure 3-3, shows the key tools and programs used to generate the virtual acoustic environment in this thesis. The crux of the data processing happens in a piece of code called uSimPassiveSonar. uSimPassiveSonar takes in elements from several other applications that define the environment, and ultimately outputs the pressure sensor data for the elements of the simulated array. The other utilities and processes provide interfaces, inputs, and do calculations that allow uSimPassiveSonar to operate. The application uses a ray-tracing model, which simulates ambient noise, noise from sources, and also allows for specifications of the vertical noise directionality. The model for the system is discussed in Appendix A. uSimPassiveSonar relies on two other processes - uSimNoiseCovariance, which develops the noise covariance matrix required for producing the pressure patterns, and iBellhop, which provides a MOOS interface to an acoustic modeling program called Bellhop. This program is a Gaussian/finite element beam code written in Fortran, and is part of a larger acoustic toolbox available online from the Ocean Acoustics Library [22]. 44 pGen~ealce uSimNoise ce uSiMptarge UsimCTm uSimBathy - -- - uSIMIassive or IBeffhop uSimTowedArray Figure 3-3: The MOOS-IvP Autonomy Code Architecture for the experiments run in this research. The purple blocks represent the building blocks for the acoustic picture during the simulation in real time, and the blue block is the resulting data, including the time, array position, and pressure readings, recorded in real time. Lastly, the orange block is the 3D acoustic noise code, generated after the experiment is completed. 45 Code pGenSealce uSimTargets uSimCTD uSimPassiveSonar uSimNoiseCovariance iBellhop Description Defines the characteristics of the noise source(s) being used in the experiment and provides visual cues in the simulation environment, including source strength, location, depth, and frequency range Takes definitions from pGenSealce and produces the sources in a format that can be consumed by uSimPassiveSonar Gives inputs for temperature, salinity, and pressure that ultimately define the sound speed profile Provides bathymetry (sea bottom) data for the environment Provides array element positions using a dynamic cable model for the array Coordinates inputs from other various modules to produce the pressure and array location data used for post-simulation processing Builds the covariance matrix based on the environmental inputs and vehicle location to produce the acoustic data A MOOS interface to the Bellhop program which performs the ray tracing calculations needed to provide accurate noise data Table 3.1: A brief description of each of the key code modules used to build the virtual acoustic environment. 46 Figure 3-4: A snapshot from the display of a simulated acoustic environment. The ice edge and hot-spot visuals are created by pGenSealce, while the IvP helm controls the AUV Macrura with a series of behaviors. There are other pieces of code that feed inputs into uSimPassiveSonar. pGenSealce defines the noise source sources, including source frequencies, location, and depth. pGenSealce also then publishes these definitions as messages to the MOOSDB, which is subscribed to by uSimTargets. uSimTargets then generates the information and data necessary to simulate the noise source, and feeds this data into uSimPassiveSonar for processing. uSimCTD and uSimBathy each describe different parts of the environment. uSimCTD simulated the acoustic environment in terms of the salinity, temperature, and pressure in order to develop a sound speed profile which is used in the passive sonar processing. uSimBathy provides data on the bathymetry of the acoustic environment. Finally, uSimTowedArray gives the position of each of the elements on the array based on the vehicles position using non-linear cable mechanics. Table 3.1 gives a brief description of each of the processes involved, and Figure 3-4 shows a snapshot of the simulation in progress. 3.1.4 Modeling Array Shape The shaping of an array is a crucial part of acoustic processing, and therefore so is the program uSimTowedArray. uSimTowedArray provides those positions using a 47 nonlinear dynamic cable model to describe the motion through the water, and the background for these mechanics are discussed in Appendix C. The exact positions of the array are known and calculated, and used both for receiving the signal in the simulated acoustic environment, and for processing afterwards. In real world experiments, however, the actual array position that receives the signal will not necessarily match the assumed element positions used for signal processing. Convention tells us that if the position of the elements is known to within 10% of the wavelength, the signal will still be reasonably accurate. For the frequency used in this research, that means the position of the elements must be known to be within approximately 17 cm: PositionError<= (0.10)- ~ 17cm (3.1) This metric can be achieved in a number of different ways. Arrays that deal with higher frequencies ( 10+ kHz) can be made at much shorter lengths to achieve the same aperture as an array designed for lower frequencies. These smaller arrays can be held in a rigid fixed line, such as off the nose of an AUV. However, this proves impractical when using larger arrays that necessitate the use of towed line. Another solution is modeling with constraints. Based on the position of the vehicle, the model used to simulate the array position in this research could be used to estimate the actual position. Also, recall that the DURIP array has several other sensor inputs apart from the arrays - two compasses and a pressure sensor on the aft portion of the array. Using the model subject these constraints could achieve the desired accuracy, however, more testing on this subject would need to be done. There are methods in real world experiments such as the use of high frequency transponders or Ultra Short Baselines (USBLs) [34], which are used to actively track accurate distances to different locations along the array. These USBLs can be fairly accurate - one such USBL available commercially is the iXBlue fourth generation GAP, which advertises a 0.06% accuracy at ranges of up to 4,000 m [19]. Since the farther element on the array would be approximately 53 m, by these standards we 48 could see positional accuracy to within ~3 cm. This USBL is not equipped to be installed on a Bluefin-21, however, the technology for measuring the elements within the required accuracy is available. 3.2 3.2.1 3D Acoustic Pictures 2D Beamforming The purpose of beamforming is to find the angle of incidence from an array to a noise source (or a noise field created by a number of sources) by measuring the phase differences of the arriving acoustic waves between a series of pressure sensors along that array. To begin understanding three-dimensional (3D) beamforming, it helps to begin with standard 2D beamforming. Let z be the direction of a straight array of i = 1 : N elements on a horizontal plane, at the locations d, = [0, s, 2s, 3s, . . , Ns], where s is the spacing of the array. If we are observing at frequency f, we can calculate the wave number to be: k = 27rf /c, and describe the wave number k in terms of the azimuthal direction 0 as: kZ =eik cos O (3.2) where 0 is the angle away from perpendicular to the array in the domain of 100 1800], as shown in Figure 3-5. This wave number can be used to calculate the beam for that direction for this array orientation: N b(0) = N eikdi (3.3) i=1 To get the phase information from the array, we take the Fourier Transform of a snapshot of pressure data pi at frequency f to get the signal delay vector in complex space for each array element i where i = 1 : N: Vi = TT(pi, f) 49 (3.4) 0 kZ Pressure wave fronts Figure 3-5: The "look" direction 0 in 2 dimensions (typically on the horizontal plane) away from perpendicular to the array. Finally, the beam pattern response is generated by multiplying the delay vector of the signal by the beam for each direction 0: N B(6) = N vi b() (3.5) i=1 This beam response gives the picture of the noise directionality seen by the array. It is known as Bartlett beamforming [20]. The ambiguities of this sort of beamforming are that positive and negative angles (6's) cannot be distinguished from one another. For example, if the direction of the wave number k. was 0 = 30', the noise directionality calculated by Equation 3.5 would not be able to distinguish that direction from 0 = -30'. 3.2.2 Extension to 3D Beamforming 3D beamforming is similar to 2D beamforming, except that we realize the fact that noise is not just ambiguous for 0 vs. -6, but for a whole cone of directionality with an axis along the array. In the case of a line array as described in Section 3.2.1, the direction 9 actually describes a conical angle /3 in three dimensions, as shown in Figure 3-6. In conventional 2D beamforming, this concept is lost. Noise sources coming from above or below the horizontal plane will be spread over many azimuthal directions, giving a false representation of the actual noise field. It therefore follows that a 3D method will in fact give a more accurate picture. Figure 3-7 shows an example of a simple broadside beam response pattern rep50 # Figure 3-6: The noise coming onto a straight array from the beam formed at angle is coming from all the directions of the cone projected onto a sphere. resented in two and three dimensions. 0 represents the azimuthal bearing, where 00 is directly North and 1800 is directly South, and # represents the vertical direction, where 90' is directly up from the horizontal plane and -90' is directly below. In the 3D noise picture at the bottom of Figure 3-7, 0 is on the horizontal axis, 0 is on the vertical, and the beam response is represented by the color on a scale where red is the maximum beam response and blue is the minimum. The figure represents a three dimensional pattern of a 10-element horizontal array with its ends aligned from 0 = -900 to 0 = 900 with a source at 0 = 0' corresponding to 3 = 90'. In the 3D representation, note that the maximum of the pattern is on the area represented by the cone defined by j = 90'. This illustrates that not only does the beam pattern max out at 0 = 0 and 0 = 180', but also from directly above and below, as we would expect. In the following sections, we will explore a few other beam patterns, as well as the mathematical process for producing them. 3.2.3 Creating the 3D Beam Response Let the x, y, and z positions of the array in three dimensional space be represented by the vectors xn, yn, and zn for the nth array element of N array elements. Using these numbers, we define a distance vector dn(0, #), which represents the distance from the nth element to a sphere in the far field. The location of the center of the coordinate system is unimportant, because it is only the relative distances from each direction (0,0) that are needed to measure the phase difference of the arrival pressure waves. d,,(0, #) can be calculated by the following equation based on the geometry presented in Figure 3-8: 51 I 0.8 -\ 0.6 0.4 0.2 0 50 100 150 200 250 300 350 M 5 CU iznt IUU r MU _Iuu horizontal (azimuthal) bearing, degrees Figure 3-7: A comparison of the beam pattern response for a horizontal array broadside to a source in 2D and 3D. d,(0, 0) = -(x, sin 0 + y, cos 0) cos #- z, sin 0 (3.6) Once the distance vector is calculated for each element, and the wave number k = 2wrf/c is found, the beam pattern for the array orientation can be calculated, similar to the way it is calculated in 2D beamforming: b,(0, 0)= e ikdn (3.7) Finally, the 3D beam response r(O, #) is calculated by summing the product of the delay vector (Equation 3.4) and the calculated beam (Equation 3.7) over each element, and dividing by the total number of elements: N r(o, =V'Cikd,,(0,<0) 1) (3.8) n=1 This picture is generic for any array orientation, and any shape, including curved 52 z d.(OO) 00A Figure 3-8: The distance vector da(O, #) represents the distance from the x, y, z position of the array element to a sphere located in the far field that represents the 0 and # directions. arrays, as long as the positions of the individual elements are known. Note, however, that the conical ambiguities discussed in Section 3.2.2, will only apply if the array is configured straight. 3.3 Beam Response Patterns To get an understanding of different types of beam response patterns, it is important to look at a few examples of the beam pattern responses for different source locations and array orientations, as shown in Figure 3-9. For these examples, all the data was normalized, and the array held in a straight line at different orientations to a point source. The top left picture illustrates the resulting beam pattern for a source broadside to the array. This clearly shows that the ambiguities exist at elevation +/- 90 , and also in the azimuthal direction 1800 degrees away from the source. If we think about it as a conical pattern around the array, the pattern is the cone defined by 0 = 90', or a circle in a plane perpendicular to and centered on the array. In the example in the top right corner of Figure 3-9, we see the beam response with the array tilt still set to 0 , but now the source is closer to endfire. This would be the projection of the cone specified by 3 = 150 for a horizontal array. Note that when the source is closer to the endfire of the array as opposed to broadside, the beamwidth of the response is much larger, resulting in an increase in the uncertainty of our directionality measurements. In the bottom two pictures of Figure 3-9, we introduce tilt into the array. The 53 Source ?eer Endfire 80 GO. tilt I . - -- 1 0.9 60 4C 0.7 41 20 0.6 2' 03 0.4 -2C -4( -2 0.3 0.2 -6( 0.1 -8C 50 100 uu 250 200 150 Azimuthal Bearing > 100 150 200 Azimuthal Bearing 'CUU Azimuthal Bearing au uv as I 028 0,7 0.6 0.5 04 013 0.2 0.1 Off-horlhantal Source Tilted Arry A.. -... A-m .. nea 111U 0.9 I I I 0.9 0.8 0.7 0.6 0.5 0.4 03 02 0.1 >: Azimuthal Bearing I 1 0.9 0.8 0.7 0.6 05 0A 0.3 02 0.1 Figure 3-9: Beam response patterns for multiple array orientations and source locations. Top Left: A horizontal array broadside beam response. Top Right: A horizontal array with the source near endfire. Bottom Left: Source broadside to a tilted array. Bottom Right: Tilted array with a source not at broadside. 54 picture on the bottom left represents an array with 300 tilt and a source placed broadside, and again with the conical "circle" defined by # the circle is tilted. This removes the ambiguity away from = 90'. This time, however, +/- 90' elevation. In the final picture, we see a more complicated beam pattern, with 100 tilt in the array and the source off broadside. In general, if the location of the source is known and the array is assumed to be straight, we can calculate the angle / by the following equation [38]: cos-- cos # cos 0 cos a + sin[cos -- (cos # cos 0)] x sin a sin tan -1 sin (cosqs cos 0 sin0 # # f where a is the tilt of the array. The full geometric derivation for calculating this angle is available in Appendix D [36]. 3.3.1 Deconvolving the Noise Field The method we chose to deconvolve the noise field was that originally proposed by Wagstaff to use 3D directionality principles to increase the quality of the 2D noise rosette. The first inclination when trying to understand a noise field, is to run the vehicle in as many directions as possible, and then take average of all the beams produced from the different directions. This method however presents problems, particularly due to the presence of sidelobes, which will be included in the picture, but don't actually make a part of it. Instead we use a method for extracting a noise field originally proposed by Wagstaff [37]. Wagstaff uses a model of the 3D noise field that includes the beam of the array, and then uses an iterative method to extract the pseudo-stationary noise field from it. Pseudo-stationary means that the measurements assume the noise field does not change over all the positions of the array from which the measurements are taken, and also does not change over the span of time 55 Figure 3-10: The technique used for deconvolving the noise field. On the left side the data and the shape of the array are used to produce a measured beam response, or "directional measure", and on the right hand side, the beam of the array and a noise field estimate is used to create an estimated beam response, or "directional estimate". The estimate and response are then compared, and a new noise field estimate is generated so that the process can be repeated. during which the measurements are taken. Here I will present the method used to extract the noise field. This method is slightly different than the one Wagstaff presented because Wagstaff calculated all his beams in advance, converted them to 2D using Equation 3.9, and then compared those to the measured 2D beam patterns of the array to produce the noise field. This conversion was necessary because at the time, calculating the 3D beam patterns was computationally expensive. However, now we can now simply calculate the beam response for each set of data in real time for each iteration, and then compare it to the 3D estimated noise field for each iteration. This has the added benefit that curved arrays can now be measured, as the 2D conversion can only be applied to straight arrays. The full iterative technique for this process is illustrated in Figure 3-11. 3.3.2 Modeling the Noise Field In Wagstaff's 1997 paper, he presented an iterative method for developing the noise field based on a model [38]. To begin, we assume that the noise field is developed with Equation 3.10: rij = - - dtj T 0 27T 2 0 ,r nitotai(, 4,t)bj (6,1) -do 2 cos # do (3.10) _,/2 In the above equation, rij is the 3D beam response for the ith beam (a (0, #) combination), and the jth array orientation (or data sample). t is time over the total 56 # time T of the measurement, 0 is the azimuthal angle in the horizontal plane, and is the vertical angle in elevation. bj(0, #) is the beam response pattern for the jth array orientation (or data sampling), as calculated for that particular array position in Equation 3.7, and nrtoti (0, q, t) is the complete, time dependent noise field. This equation assumes that the range of positions of the array are all small compared to the overall size of the noise field in such a way that the noise field doesn't change based on array position. Now, let us take a look at the overall noise field as a sum of components: ntotal(O, q, t) = n(O, 9) + ((0, 0, t) + E(0, #, t) In Equation 3.11, nrtotal (0, n(0, #, t) #) is the pseudo-stationary (3.11) is the overall time-dependent measured noisefield. noise in the environment, ((0, 9, t) is the time-dependent component of the noise field due to fluctuations in the acoustic propagation, noise source movement, changes in noise source, etc., and c(0, #5, t) is the error induced by tow ship noise, flow noise, array position error, etc. In simulation as well as in real time testing, we can consider ((0, E(O, 9, t) 9, t) to be negligible with temporal averaging, and to be small enough to be negligible by using noise reduction techniques and array grooming [38]. With these assumptions, we can re-model the noise field as: ntotai(O, #, t) r n(0, 9) (3.12) Using this approximation we can reduce Equation 3.10 to the following: ri = 4x 0 in(O, #)bj (0,9 ) cos #d# dO (3.13) -_r/2 This is the model we use to solve for the noise field. 3.3.3 Iteratively Solving for the Noise Field Equation 3.13 has three main components. The first, the 3D beam response rij (9 5), can be calculated from the array positions as in Section 3.2.3. The second component 57 Assume a Constant Noise Field Calculate the Delay Vector Calculate Difference between current and last Pooled Std. Devs. Calculate the Estimated Beam Response s Find the Pooled Standard Deviation Find the Difference between the Estimate and Measured Output NoiseField Divide by two, add the lp average difference to Estimated Noise Field Figure 3-11: A Process Diagram for the iterative process used to deconvolve the noise field from the beam response patterns. is the beam for each array orientation b-(9, 0), which can also be calculated using Equation 3.7. The remaining component, n(O, 0), is the last unknown, and what we are trying to solve for. This is done with the iterative process prescribed by Wagstaff [38], where we begin with an estimate for the noise field, use it to calculate the estimated beam pattern response, compare it to the actual response, and then based on the difference make a new guess for n(O, #). This process is repeated until the noise field settles on a certain position. This is illustrated in a process diagram in Figure 3-11. The following paragraphs describe this process in greater detail. The process begins with an initial estimation for the noise field in dB Nk=o, which we set to any constant C, which could just be an approximation of the ambient noise field. The choice doesn't really matter as the end result is relatively insensitive to the initial estimate. This guess is taken and converted to the pressure intensity domain: Nk=0(0, 0) = C hnk(0, ) 10Nk/10 - (314) where rik is the estimated noise field in the pressure intensity domain for kth iteration in the process. In general, upper case letters will be used to describe fields in the dB 58 2 ) domain, while lower case letters will be used to describe fields in the intensity (p| domain. Based on the noise field, we now want to find the estimated beam response pattern. This is done by calculating the signal noise delay vector for each element: 7/2 Vi,j,k,n j /27r7r/2 (O, jO)e ikdi,j' cos k OdfdO (3.15) _ where Vi,j,k,n is the delay vector for the ith direction of I total directions, the jth array orientation or data sample of J total data samples, the kth iteration, and the nth array element of N total elements. With the delay vector, the estimated beam response is calculated and then converted back to dB space: ik i,j,k,feikdi,jn ij,k 10 log1 0 ij,k) (3.16) n=1 The final step in the iterative process is to find the difference between the estimated field response and the measured one, divide by 2, and then add the result to the estimated noise field from the previous iteration: Ai,j,k = Rij - Ni,j,k 1 (0, 0) + -k+ I Ai~~ 2 J (3.17) j=1 This process is then repeated until the standard deviation of the solution converges, as described in Section 3.3.4. 3.3.4 Statistical Convergence To determine when the process has been repeated enough times and has stabilized on a noise field, we measure the standard deviation of the estimated beam responses relative to the actual beam responses. A quantitative measure can be derived by calculating the pooled standard deviation o-, of the differences between the measured beam response and calculated guessed response [37]. 1 J (= j) E j=1 59 1/2 (3.18) where the standard deviation o- for the jth array orientation or data sample is: [ j) 1/2 ~~ I =1_ (Rij . i_1 - I-1 - (3.19) RJ)-1R1/2 and Rj is the mean difference for the jth array orientation: - Rj = 1' (Ni,j - Ri,) (3.20) i-1 After each iteration, Equation 3.18 is used to calculate the pooled standard deviation op. This value in general becomes smaller and smaller as the estimation for the noise field becomes more accurate. For each iteration, we can track the change 6 p in ap. 6 p,k = Up,k - Up,k-1 (3.21) Finally the iteration process is discontinued when 6p,k reaches some pre-determined arbitrarily small threshold. At this point, the resulting h(O, q) is accepted as the final estimate for n(O, q). 3.4 Measuring Vertical Directionality One of the key components for understanding the quality of the 3D picture, as we will see in Chapter 4, is in understanding how adding "verticalness" into a towed array as it is given tilt in the water column will affect the 3D acoustic picture. Quantifying how much "verticalness" one can have in a towed array for a period of different array orientations and tilts is crucial. There are many ways to get verticalness into an array. The yoyo pattern used here is one of those ways. However, the vehicle could also be maneuvering through the depth column for a specific purpose, such as to establish acoustic communications, or to rise to the surface. Because of this, we need a way to quantify the amount of verticalness we find in an array over time that works in all situations. This section 60 L h Figure 3-12: The measurements for finding the verticalness of an array by its elements. h represents the height of the array, from the lowest z-direction element to the highest element, where L represents the maximum distance between any two elements on the array projected on to the x-y plane, which is typically the horizontal distance between the first and last element. shows how we can quantify verticalness so that it can be used as a benchmark to test how to achieve the best 3D acoustic picture. 3.4.1 Instantaneous Verticalness At any instant in time, we are able to measure the amount of verticalness we have in our array. This is done by measuring the height of the array as compared to the length, and using the following equation: w(t) = 2 - wr sin 1 h (--) L (3.22) where w(t) is the verticalness of the array, h is the height from the lowest element to the highest element, and L is the total length of the array, as show in Figure 3-12. In this method, if an array is perfectly horizontal, its verticalness is w(t) = 0, and similarly if it's perfectly vertical, its verticalness is w(t) = 1. In this way, we can track the verticalness of an array at a given instant time. 61 3.4.2 Tracking Verticalness over Time Creating a single 3-dimensional picture involves measuring the values over a length of time, something on the order of minutes or longer. Therefore, it's not the instantaneous verticalness in the array that's important, but the amount of verticalness that has been seen over the most recent amount of time. This section shows a method to do this, using the instantaneous verticalness measurements and applying it to an impulse response model. If we keep track of the verticalness of the array over time with Equation 3-12, we can keep track of a "vertical score", which I'll call v(x), with the following equation: e-(tx-)/w(x)dx v(t) = (3.23) Jt In the above equation, T, is a decay constant, t is the current time and to is the time the measurements began. Verticalness more recently in the array becomes more important than verticalness a longer time ago, using an exponential decay. Figure 313 shows an example of one of the experiments conducted by showing vehicle depth, the measured array verticalness, and finally the vertical score. For this example and all the experiments, the value r, = 100 was used. 3.4.3 uVertScoreKeeper In order to keep track of the vertical score, a matlab program call uVertScoreKeeper was created to run in along with the vehicle to keep track of the verticalness of the array. Using iMatlab as an interface, it provides real time tracking and visual representation of the ship position, array positioning, the instantaneous verticalness, and the vertical score. In addition, it saves all the information so that it can be used for processing and analysis after the virtual experiment is completed. shows an example output running in simulation. 62 Figure 3-13 XY Position 1000 r 800- Array Visualization 0- Hot Spot 1 SL = 120 TL = 65.036 SE = 14.964 -100 -100 -60 -80 600- -40 0 -20 Verticalness of Array UU 0.2- Verticalness = 0.21006 . 0.1 400- 0 200- 500 2000 1500 1000 Time(t) Vertical Score - cu 0.3 0 Vertical Score 0.2 acrura 0 = 0.18652 0.1 -600 -400 -200 0 200 400 0 600 500 1000 1500 2000 Time(t) Figure 3-13: An example of the real time tracking provided by uVertScoreKeeper. The left window shows the path taken by the AUV in the x-y plane of the AUV, the top right shows the depth of the AUV along with a visualization of the array projected behind the vehicle, the middle right window is a graph of the instantaneous verticalness, and the bottom right window is a measure of the vertical score. The vertical score measured at the end is the overall vertical score for the run. 63 64 Chapter 4 Noise Field with a Point Source 4.1 Quantifying Resolution One of the major points of this thesis is the exploration or the increase in the resolution of a noise picture based on the the verticalness of the array. In this section, we will run experiments that will demonstrate that increasing verticalness in the array can significantly increase the resolution of the 3-dimensional acoustic picture. In addition, we will see the time for convergence of a noise picture, and finally demonstrate the feasibility of using 3D pictures to find the range to a target. 4.1.1 Experimental Setup To explore the changes in quality of the noise picture, a simple experiment was devised. In this experiment, we set a simple source directly North of the vehicle, at a distance of 5km and at 5 m depth, with a frequency of 900 Hz. The source level is set to 120 dB. The frequency was chosen because that is similar to some of the noise frequencies that have been seen previously in the Arctic. It was also chosen because it matches well with the frequency range of the DURIP array. The noise source level was chosen to be high enough such that the signal could be seen relatively well above the ambient noise present. A lower source signal could be used, however, the goal is to distinguish how the vertical directionality of the array affects the resolution, not 65 Experiment 1 2 3 4 5 6 7 8 Yoyo-Period 300 500 600 700 1000 1500 2000 N/A Table 4.1: Different periods were used for the yoyo patterns for each experiment to introduce different levels of vertical directionality into the array. For the 8th run, the vehicle was held at a constant depth, and therefore has no period. see the limits of can and can not be seen. The AUV was run in a circular loiter pattern with a radius of 100m. The vehicle was started at a depth of 50 m and allowed to run one full circular loiter pattern. Figure 4-1 gives the general setup. For the constant depth run, the vehicle remains at 50m, but for the other runs, the vehicle is commanded to perform different yoyo pattern with different periods. The period specified is the distance the vehicle would travel along its horizontal path before it completed a full up-down-up pattern. The upper limit of the yoyo pattern is set to 20 m depth and the lower 100 m depth. Shorter periods correspond to higher degrees of vertical directionality, while larger periods correspond to lower. Table 4.1 gives a list of the different yoyo periods used in the experiment. The pixel resolution of the 3D acoustic images produced is 30 in the vertical direction and 3' in the azimuthal. Higher pixel resolutions were not measured because small changes in the pixel resolution lead to great changes in the time to calculate. Also, because the algorithm uses large multi-dimensional matrices as opposed to loops to make the calculations, the program runs faster but is susceptible to memory clog-ups at higher resolutions. Through trial and error experiments, the 3 0 -by-3' resolution is good for data post-processing. For real-time measurements, a 5 0 -by-5' resolution is more appropriate. The resolution of the azimuthal and vertical directions do not necessarily need to be the same. When trying to get more information in one direction over another, the resolution in that direction can be increased while the 66 Ice Edge 5km 1O0m Loiter Pattem Figure 4-1: The setup for the experiment involving a single simple source. The AUV will conduct a loiter pattern with a 100 m radius, 5 km South of a noise "hot spot" along the ice edge, simulated as a point source with a source level of 120 dB at 900 Hz. other is decreased to maintain similar calculation speeds. This method is used in Chapter 5. 4.1.2 Noise Field Results Figure 4-2 shows the result of the experiment with a period of 300 m. The top left graph in the figure shows the noise field estimate in decibels. The bottom three graphs, from left to right, show the path of the vehicle on the horizontal (x-y) plane, the depth of the vehicle as a function of time, and the vertical score of the array tracked over time. The figure to the right of the acoustic picture shows the vertical noise profile. The black solid line corresponds to the average vertical noise profile (averaged in the intensity domain, not dB), and the dashed blue line corresponds to the vertical noise profile at the source location. We can observe from the results that for this particular experiment, that the vehicle did not have enough time to travel all the way between the depths of 20 m and 67 3D Noise Field - Period 300 80 80 80 75 60 a 40 7 70 20 .r 0, 65 0 60 -20 -40 60 40 20 0n 0 0 -20> 55 -40 50 -60 -80 45 Azimuthal Bearing, degrees 150 100 - -' -100-0 x (M) 100 = 0.20003 / -60 -100 0.2Vert Score -20 - - - - - - - - - - - - -- - - 40 -0.2 .50 0 -100 -150 0.3 3 0 ----------------------- 0 100 50 Ambient Noise Level (dB) 0 100 -----------------200 300 Time (s) 400 0 0 100 200 300 Time (s) 400 Figure 4-2: The noise picture results from a yoyo period of 300 m. The top left graph is the acoustic picture, with azimuthal angle 0 on the horizontal axis, and vertical angle # on the vertical axis. The bottom left graph depicts the path of the vehicle in the x-y plane, the bottom center is the depth of the vehicle as a function of time, and the bottom right shows the vehicle's vertical score over time, as well as the ending score. The graphic on the right shows the average vertical noise field (black line) and the vertical noise field at the source location (blue dotted line), which is at 0 = 0'. 68 100 m. This is due to the physical limitations on the Bluefin-21 AUV's maneuvering, set in the vehicle simulator uMVSBluefin. Therefore, a period of 300 m with the depth limits set will correspond to approximately the maximum verticalness that can be achieved in the array for the Bluefin-21. The vertical score achieved is v = 0.21. We can also see that the vertical score tracks well with the depth profile. The vertical score increases while the vehicle is on its up or down-slope paths, and tends to decrease as soon as the vehicle makes a turn to go back in the other direction. This is a result of the array leveling out at the peaks and troughs of the yoyo pattern. In the measured noise field, we can see very clearly the horizontal and vertical location of the source - at 0' azimuth and near 0' elevation. This corresponds well with the placement of the source. To measure the resolution of the noise source (the quality of the picture, not the number of pixels per degree), we measure the peak of the response and the 3-dB down bandwidth. The peak is simply the maximum of the beam response at the source location. The 3-dB down bandwidth of the response is measured in both the horizontal and the vertical directions by taking the peak response, subtracting 3, and finding where the beam pattern intersects on either side. For this particular experiment, we can see a maximum dB level of 79.4 dB, and a 3-dB down bandwidth of approximately 10.5' in the vertical direction, and 4.5' in the horizontal direction. If we remove some of the vertical directionality from the array by increasing the yoyo pattern period, we can see some degradation in the resolution of the picture. Figure 4-3 shows the results for a yoyo period of 1000 m, with a resulting vertical score of v = 0.123. This is a little more than half of the verticalness achieved by the 300 m period yoyo experiment. The maximum response of the source is 3.1 dB lower at 76.3 dB, and the vertical 3-dB down bandwidth at the source is 5.5' wider at 16'. However, if we look at the horizontal 3-dB bandwidth at the horizontal axis we see no degradation, which is still at approximately 4.5'. This means that increasing the verticalness from 0.123 to 0.21 does not degrade the horizontal arrival structure. Eventually at some level of verticalness we would assume that adding verticalness will degrade this direction, but it does not happen within the maneuvering limits of the 69 3D Noise Field - Period 1000 80 N 80 80 75 60 60 40 70 40 0 1 20 20 65 0 60 Ca2C -- 20 0 -20 > 55 -60 -60 50 -80 -80 45 -15u -IUU -U U iuu u Azimuthal Bearing, degrees - -- -20 50 - - --- - - 0.3 - - 0 Vert Score= 0.12258 - 150 100 0.2 -40C 0 -50 80 40 60 Ambient Noise Level (dB) Lau -60 -80 -100 -1501 -100 0 x (m) 100 0 01 - -100 --------------- 100 300 200 Time (s) 400 100 0 300 200 Time (s) 400 Figure 4-3: Results of the 3D acoustic picture for a horizontal circular pattern combined with a yoyo maneuver with a specified period of 1000 m. 3D Noise Field - Const depth 80 80 75 a a OE a 60 40 70 65 OE 20 / a 0 EU 60 .2 -20 ( a 55 a -40 - 50 -60 -80 -100 -150 150 50 -10 -50 -60- -100 -150 -80 -100- --- -' -100 0 x (m) 100 0 a -100 300 400 Time (s) Vert Score 80 40 60 Ambient Noise Level (dB) =0.018588 0.1 _ 0L ----------------200 45 150 0.3 0 -----------------------20 - - - - - - - - - - - - - - - - V -4 0.2 -40- 100 .. 100 50 0 -50 Azimuthal Bearing, degrees 0 100 300 Time (s) 200 400 Figure 4-4: Results of the 3D acoustic picture for a horizontal circular pattern with no change in depth. 70 . - 0 - - - Yoyo Period 300 Yoyo Period 1000 Constant Depth -20-1 ~-40,-. -1 -60 00 y-irection -5-5 -100 -100 x-direction Figure 4-5: A comparison of 3D path plots taken by the AUV for three different experiments. The black dashed line is the vehicle path taken at constant depth, the blue line is the vehicle path taken for the same loiter pattern, but with a yoyo maneuver with a period of 300 m, and the red line is a yoyo pattern with a shallower angle, corresponding to a period of 1000 m. AUV. Finally, Figure 4-4 shows the results where there is minimal vertical directionality, or where the vehicle was still run in the circular pattern, but without the yoyo maneuver. Even when traveling at a constant depth, some vertical directionality still exists because the array in the model is slightly buoyant, and the array rises slightly in the water column behind the vehicle. This experiment achieved a vertical score of v = 0.019. The measured noise field in this experiment is even less desirable than what was observed from the two previously discussed experiments. The peak dB level of the response is significantly lower, at 71.3 dB, more than 8 dB lower than the response from the experiment with the period of 300 m. In addition, the 3-dB down vertical bandwidth at the source is larger - at approximately 21.5'. This is more than double the bandwidth from the period 300 m yoyo pattern. And similar 71 80- Veia - 0A1gl 70 --- 65 -- Vert -0.123 Vr-0.5 460 0 75 0 30 0 6000 40er 00 0-420 t45 40 -60 -20 0 20 40 60 80 Vertical Angle Figure 4-6: Vertical noise profiles at the source location for multiple levels of array verticalness. As the vertical score decreases, there is a clear pattern where the bandwidth increases and noise peak decreases. to the result we observed with the period 1000 m yoyo experiment, the 3 dB down horizontal bandwidth still remains relatively constant at the same 4.5'. For 3D plot path comparison of the constant depth path vs a period of 300 m vs. a period 1000 m, refer to Figure 4-5. Let us now take a closer look at the comparison between all the experiments run from Table 4.1. A full listing for the results from each of the pictures is listed on Table 4.2. Figure 4-6 shows the vertical profiles for each of the acoustic pictures at the source location (0 = 0) side-by-side, while Figure 4-7 shows the horizontal profile at the source location. 3D acoustic pictures for all the results can be found in Appendix E. We can see from observing the vertical ambient noise profiles that as the vertical score increases, the resolution of the vertical arrival structure of the noise from the source increases as well. This is both in the form of the bandwidth and the peak response. Table 4.2 categorizes the results from the measurements. While the limited pixel resolution of these pictures (3' in both the 0 and q5 directions), does remove some of the preciseness from the results, the general trends in the data appear to be clear. With an increased verticalness, within the limitations of the AUV's maneuvering capabilities, we see a general increase in the peak dB, a decrease 72 Experiment Vertical Score Peak dB 1 2 3 4 5 6 7 8 0.019 0.057 0.080 0.123 0.136 0.155 0.193 0.200 0 2.0 3.1 5.0 4.4 6.2 6.9 8.1 Vertical 3- Horizontal on dBrDown 33-dB Down dB Down 21.80 6.80 6.20 27.4" 26.00 5.30 16.30 5.6 5.2" 15.60 13.80 4.50 4.20 12.50 10.30 5.90 Table 4.2: Vertical noise profile resolutions for experiments looking at a simple source within the maneuvering limits of the Bluefin-21. 80 -_Vert = 75 - 70 0.200 0.193 Vert-0155 Vert 0.136 -_Vert- Vert =0.123 -Veft-0.080 __Vert- 0D57 -Vert -0.019 65 60 0 10.55 CU E 50 S45 40 35 30 -80 -60 -40 -20 0 20 Azimuthal Angle 40 60 80 Figure 4-7: Horizontal noise profiles for multiple levels of array verticalness looking at a simple source. We can see from the graph that the 3-dB down bandwidth of the noise field does not change considerably with increased vertical scores. 73 in the vertical 3-dB down bandwidth, and little to no change in the horizontal. In the next section, we will test the verticalness of the array beyond the maneuvering limits of the Bluefin-21 AUV, and define general relationships between the verticalness and these three metrics which define the resolution. 4.2 Experimental Testing on a Simple Source 4.2.1 Experimental Setup In examining the verticalness of the array within the maneuvering limits of the Bluefin-21, we have learned that increasing the verticalness will increase the resolution of the picture as well as the vertical bandwidth of the response. But intuition tells us that this will not necessarily be the case if we continue to make the array more vertical. At some point, these measurements can no longer continue to get better, and the horizontal bandwidth may suffer. We know that, for example, a perfectly vertical array would have no horizontal information. And so the question arises, how much verticality in a towed array is too much? We can answer this question with a series of well-designed experiments. There are many ways this could be set up though the use of constraints, and many ways this could be tested. The goal of this research, however, is to test the improvement that verticality would have on the picture. The constraints on the experimental process were developed as follows: 1. The towing AUV will always conduct a full circle in the horizontal plane, and a yoyo pattern in the vertical direction 2. The duration of time over which the measurements are made will remain constant 3. The speed of the vehicle will remain constant 4. The max depth and period of the AUV yoyo maneuver can be changed to manipulate the verticality of the array, but the AUV must always remain in the 74 1theoreical d 4 A/2 Figure 4-8: The geometry used to calculate the total path distance S taken by the AUV as a function of the horizontal distance traveled X'. surface duct so that the noise picture can remain pseudo-stationary Using these constraints, we can design a series of experiments. We control the verticalness of the experiment by changing the yoyo period and the minimum and maximum depths of the behavior. To ensure that the vehicle always travels the same distance, we calculate the radius of the circle in the horizontal plane that the AUV will need to take using a few assumptions. These are: (1) the vehicle will always travel in a straight line, (2) can instantly change pitch after each up-down leg of the yoyo maneuver, and (3) maneuvers in a perfect circle in the horizontal direction. With these assumptions, the radius of the circle can then be calculated using geometry. Let us assume that X is the distance the vehicle has traveled in the horizontal plane along the circumference of a circle. From the constraint that the vehicle must travel in a full circle, we can calculate the full distance xO the AUV will travel in terms of the radius of the circle: = 27rr -O (4.1) Now let us look at the maneuvering in the depth column. If the depth through which the AUV travels before it turns to its next leg of the yoyo is d and the period is 75 A, then using the geometry presented in Figure 4-8, we can calculate the total distance ' the vehicle has traveled as a function of the distance traveled in the horizontal plane () = A 2 d2 + () 2_ (4.2) Now from the constraints that the vehicle travels over a fixed time at a fixed velocity, to and vo, and that the total distance traveled is S' = tovo, we can rearrange Equations 4.1 and 4.2 to solve for the radius: r =v Atov0 (4.3) 47r d 2 +(A) This is the radius used in the experiment given the other parameters. Obviously, even in the simulation, the vehicle cannot change pitch instantly. Also the vehicle does not maneuver in a perfect circle all the time, particularly at higher pitches when the azimuthal angle is difficult to control, and as such the vehicle may go a bit over or under the circle in the fixed time allotted for each experiment. However, as we will see in Section 4.3, the acoustic picture will tend to converge before that, and so a little less or more will not affect the picture significantly. Through a series of experiments we were able to test vertical scores of up to almost 0.54. Higher verticalities could not be tested for two reasons. The first is that even inside of simulation, the vehicle was difficult to control at higher pitches. Similar to an actual in-water experiment, for an AUV at high pitches, small changes in the rudder of an AUV result in large changes in the azimuthal direction of the vehicle, which causes the simple PID controller designed to handle the azimuthal direction to become unstable. With a redesign of the PID controller, this could be fixed, but they are well outside the actual limits of the AUV, and are therefore not included in this work. The second is the shape of the towed array with a yoyo maneuver. Each time the AUV turns at the top or bottom of the yoyo pattern, the array becomes more or less horizontal. So because of these direction changes, even with very high pitches in 76 Set Parameters Time [s [ l Calculated Parameters peed Period [m] Min Depth Max Depth [im] Calculated # Direction Theoretical [m] Radius Changes Pitch [m/s] 1 450 1.5 N/A 50 50 107.4 0.0 0.0 2 450 1.5 2000 20 100 107.1 1.0 4.6 3 450 1.5 1500 20 100 106.8 1.0 6.1 4 450 1.5 1000 20 100 106.1 2.0 9.1 5 450 1.5 800 20 100 105.3 2.0 11.3 6 450 1.5 600 20 100 103.8 3.0 14.9 7 450 1.5 500 20 100 102.3 3.0 17.7 8 450 1.5 400 20 100 99.7 4.0 21.8 9 450 1.5 500 20 140 96.9 3.0 25.6 10 450 1.5 500 20 170 92.1 3.0 31.0 11 450 1.5 500 20 200 87.2 3.0 35.8 12 450 1.5 450 20 200 83.9 3.0 38.7 13 450 1.5 350 20 200 74.9 3.0 45.8 Table 4.3: 13 Experiments conducted to test the verticality of the array. The parameters speed and time were fixed, and the period and min/max depths were varied in order the control the vertical score. The radius of the horizontal loiter pattern was calculated so that the vehicle always completes one full circle in the time allotted. the middle of the yoyo legs, the overall vertical score will be lower. The full design of experiments can be found in Table 4.3. The yoyo period and the maximum depth were the parameters used to control the verticality. A general sense of the verticality can be understood by the theoretical pitch and the number of direction changes from the yoyo maneuver, also listed in the table. The theoretical vertical angles are calculated using Figure 4-8 with the following equation: -tan- (jd2 (4.4) ) Otheoretical and the number of direction changes is simply nchanges = voto/(A/2) rounded up. Higher theoretical tilt angles will correspond to higher vertical scores, and more direction changes will result in lower vertical scores. 77 Relative Max dB 3-dB Down Bandwidth 25 8-i a 0 I 7 Ve rtical1 Horizontal 20 6- ~15 S0 al 0 0.1 0.2 tclScr 0.3 0.4 0.5 0 0.6 cr etia Ca 0.1 0.2 0.3 0.4 0.5 0.6 Vertical Score Vertical Score Figure 4-9: These graphs show the improvement trends in the resolution of the measured noise field for a range of vertical scores. The graph on the left shows the peak DB of the noise field, while the graph on the right shows the trends in the bandwidth. On the right graph, the blue dots are the vertical 3-dB down bandwidth at the azimuthal source location and the green dots are the horizontal 3-dB down bandwidth. The vertical dashed lines represent the approximate maneuvering limits of the Bluefin-21 AUV. 4.2.2 Results and Discussion The results from the experiments can be seen in Table 4.4, and the are graphed in Figure 4-9. From these results we can devise a general set of conclusions about how the vertical score of the array can affect the resolution of the picture. It should be noted that the vertical and horizontal bandwidths are approximate. The actual pixel resolution on the picture is only three degrees, and so a linear interpolation method was used to calculate them. Our observations of the range of vertical scores lead us to the following conclusions. On the left graph, we can observe that increasing the vertical score will increase the peak dB of the noise field, up to a point. The peak dB tends to increase until a vertical score of approximately 0.25 or 0.3, and then tends to level off at approximately 6.5 dB higher than what was measured in the constant depth experiment. From the right graph, we can see the vertical (blue dots) and horizontal (green dots) 3-dB down bandwidths measured. The 3-dB down vertical bandwidths tend to improve up until a vertical score of about 0.3, and then level off to about 3'-5'. The set of points 78 # 1 2 3 4 5 6 7 8 9 10 11 12 13 Vert Score Horizontal Bandwidth Vertical Bandwidth 0.021 0.075 0.092 0.117 0.138 0.169 0.202 0.239 0.305 0.413 0.491 0.539 0.524 6.70 6.4* 6.30 5.50 4.60 5.00 4.1* 4.40 5.30 5.50 7.30 6.6* 7.00 20.80 22.40 17.80 15.40 17.80 14.10 11.20 10.00 4.80 3.50 3.70 3.00 2.70 Table 4.4: The results from the 13 Experiments tested over a full range of array verticalness. appears to generally follow a quadratic relationship. Lastly, over the entire range of the horizontal bandwidth, there doesn't seem to be much loss over of the entire range of vertical scores. To draw general relationships in the domain of vertical scores tested, one can take a look at the trend-lines developed. With the peak dB, I've separated the response generally into two domains, the first is a linearly increasing relationship, and the second is a constant, generally level response. There is some variation in these results, particularly for the higher vertical scores where the peak dB deviates from the average at that point. The equations with their constants and domains are as follows: peakdBrei v 6 [0 , 0.24) cV (4.5) v e [0.24 , 0.54] C2 c 2 = 6.64 c = 30.0 Note that we've adjusted the dB scale to a relative scale, where 0 is the peak dB response for a horizontally towed array (no yoyo maneuver), such that peakdBrei = 79 peak-dB - peak-dBconstant-depth. a 1 and a 2 are constants, and v is the vertical score for the run. We also have relationships for the bandwidths. For the horizontal bandwidth, we'll assume that the bandwidth is constant over the full domain, and so we set our trend line to the average: Horizontal_Bandwidth= C3 C3 V E [0 , 0.54] (4.6) = 5.75 Finally, for the vertical bandwidth, we find that a quadratic relationship makes the best fit. The following is the best fit line for the vertical bandwidth: Vertical _Bandwidt h = C4 4.3 =75.2 c4 v2 + c 5 v + c6 c5 =-82.1 v E [0 , 0.54] (4.7) c 6 = 25.1 Convergence of a Noise Field When developing a 3D noise picture, the natural question arises: how much time, data, and maneuvering is required for an AUV with a towed array to obtain a reliable noise picture? It was clear from the previous experiments that a full rotation is in fact enough data to converge on a noise field, but what about less? To study this, the AUV was run in a full circle, while the estimated noise field was calculated different points throughout along that run using the data that had been collected up to that point. This field was then compared to the noise field measured after from the full loiter pattern to test if it matches. These two pictures were compared via a correlation factor, calculated by the following equation: 80 base(Oi, q5) - field(i, #5) 1 Correlation= where j E n I Z?- csQ5~ ~ mE> is the jth noise field part in the a e2 cos(#5) E' cos(5)(48 # direction of n total vertical directions, (4.8) i is the ith potion of the noise field in the 0 direction of m total azimuthal directions, and a is the standard deviation. In this way, each pixel of the acoustic noise field is measured against and compared to the measured noise field, and given a value based on a normal distribution curve. For this calculation, Correlation = 1 means the picture matches the measured noise field, and Correlation - 0 would be several standard deviations away. A standard deviation of o- = 3dB was chosen because in practice this is what we tend to see when calculating up, which the pooled standard deviation between the estimated and measured beam responses as described in Section 3.3.4. The reason for the cos(#) term in Equation 4.8 is to take into account the fact that not all the pixels in the noise picture are equal. The noise field as we display it is actually a Mercator projection, or a flat cylindrical projection of a spherical field. The cos(#) makes sure that pixels at high vertical angles are less significant than pixels at or near the horizontal plane, because those pixels represent a smaller area from which the noise arrives. Lastly, when calculating the signal, a minimum dB floor was set. The reason for this is that we want to measure the correlation in the response (i.e. the areas with higher dBs) and not the random differences of the noise. Therefore, for the purposes of measuring the correlation, if a pixel of the noise picture fell below the floor, it was set to that floor. The dB floor used here was set to 20 dB, which is still well below the ambient noise level. Figure 4-10 shows the noise field generated at certain points throughout the loiter pattern circle. The last picture (bottom right) of the figure represents the noise field generated after the full circle is complete, which took 402 seconds. The first (top left) picture represents the noise field generated through the algorithm based after just a single data sample was taken. 81 t=2s t=13s t=36s t=80s t=201 s t=321 s t=48s t=402s Figure 4-10: A series of snapshots of the noise fields generated after different points of time as more and more data is gathered. Clearly, we can see that one data sample is not enough for the algorithm to generate a complete noise picture. The iteration technique converges on a field that is not realistic and the beam response scattered in many directions. As more data is collected, however, the picture does get better. In the t = 36 s picture, we can see a smoother more realistic picture is starting to develop, although this is not necessarily reliable yet. Later, at t = 48 s a mix of the data once again appears pixelated, with high noise levels coming from directions that aren't the direction of the source. Note the high noise content present near the bottom of the picture. As time moves on, however, it becomes much more likely that the noise algorithm converges on the true noise pattern. At t = 80 s, it is almost there, and all the pictures after that have reasonable correlation with the full picture achieved after t = 402 s. We can better understand the trends if we look at the correlation scores of the noise fields as a function of time around the circle - this is shown in Figure 4-11. We can see it three domains over which we see different behaviors of the algorithm. In the first domain, from t = 0 s to about t = 40 s, we see no convergence of the picture, the generated noise field is not going to work. In the second domain, from t = 40 s to about t = 120 s, we see intermittent convergence. Here, occasionally the noise field iteration algorithm will converge on a picture which is relatively close to the true noise field, however, it's not reliable. Finally, in the third domain, from about 82 Acoustic Picture Convergence 0.90.80.70.60 0 U 0.40.30.20.10 0 50 100 150 200 Time (s) 250 300 350 400 Figure 4-11: The correlations calculated based on Equation 4.8 for generated noise fields with the available data up to that time. t = 120 s and onward, we have fairly reliable convergence on the true noise field. There are many other factors that play a roll in establishing the convergence of a picture. First, it is important that the array turns and faces different azimuthal and vertical directions relative to the source in order to reduce the overall noise ambiguity. If the array always stays in the same orientation relative to the source, the resulting noise picture will be poor, no matter how much time and data is gathered. Another thing to consider is that each beam formed in the simulation is based off of 2 seconds of data. For example, t = 60 s is also equivalent to evaluating 30 different beam patterns being used in the iterative process, and this may also be more of a factor in determining the time. In an effort to capture all this information, the following is a list of the factors that should be taken into consideration when attempting to get an accurate picture: " The percent completion in which the array has turned " The number of data samples taken 83 40 km Water Edge Source 2,000 m Acoustic Ray AUV Figure 4-12: The experimental set up for the range finding experiments. Here, the source was 40 km away, and the vehicle conducted its yoyo patterns at a depth of approximately 2000 m. This depth was chosen so that the only noise directions arrivals would be the result of the reliable acoustic path (RAP) ray. o The time over which the data samples are taken From our experiments, we observed that the picture converges after approximately 120 seconds, or 60 data samples, or about 30% of a circle. 4.4 4.4.1 Range Finding on a Source Experimental Setup In this experiment, we examine the feasibility of finding the range to a source using the vertical arrival structure. Here, we want the AUV to be further away from the source, and below the surface duct. This is so the only sound arriving from the source to the array is from the reliable acoustic path (RAP), whose direction is well below the horizontal plane. For this experiment, the vehicle is set deeper at 2000 m, at a distance of 20 km South of the source. Based on these new depths, a series of experiments was run, with same periods specified in Table 4.1. The source level was raised to 140 dB so that the sound still comes through the water even with the increased transmission loss associated with the additional distance. Figure 4-12, show the general setup for this experiment. Before we examine the results, we will give a short discussion on the calculations used to estimate the arrival of the rays. 84 4.4.2 Range Finding Calculation A simple calculation for the range to a target can be made if some simplifying assumptions are made in the calculation. According to Snell's law, a ray of sound propagates such that: CosO(Z)= Const. (4.9) c(z) Now, if we assume that sound speed is linear with depth, such that: (4.10) c(z) = c(0) + gz where c(0) is the sound speed at the water surface (z = 0) and g is the sound speed gradient, we can say that rays will travel in circular paths, where the center of the circle will be where the sound speed goes to zero, or at the location: z (4.11) C(0) 9 By the geometry shown in Figure 4-13, the radius of the circle given the arrival angle of the ray 0 and the depth of the receiver d, is: (4.12) c(0)/g + d, cos 0 , Again, by the geometry of the problem, we can again calculate the variable a1 and similarly a 2 by the following equations, and ultimately find our range between the source and receiver: a, = r2 - (C() + d,) a2 - range = a, + a 2 r2 _ c() + (4.13) (4.14) For the purposes of this particular calculation, we approximate the constant sound speed gradient to be g = 0.022m/ m and c(0) = 1440 m/s. In the measured sound 85 a, a2 d,, Figure 4-13: This figure shows the geometry used to calculate the range of the source given the angle of arrival 0. For these calculations, the simplifying assumption that the sound speed profile is linear was used. speed profile, the gradient above z = 1000 m is g = 0.0196m/ and for below 1000m is g = 0.024 m/s. The actual sound speed at the surface is c(0) = 1440m/s. For more m precise range measurements, one could build a piece-wise array based on ray-tracing models currently in use [20]. However, for the purposes of this thesis only a simple calculation to demonstrate the feasibility of ranging using the 3D acoustic picture. 4.4.3 Results and Discussion To get a basic understanding of these properties, let us first compare two of the experiments for this discussion: the first experiment is with the yoyo pattern with a period of 300 m, and the second the experiment with no yoyo pattern. The results of the noise field calculations can found in Figures 4-14 and 4-15. We'll first discuss the difference between the two pictures. A comparison of the vertical arrival structure at the source location can be found in Figure 4-16. We can see clearly that the picture with the yoyo pattern better resolution in the vertical direction (smaller 3-dB down bandwidths), as well as higher peak responses ( 2.5 dB). The 3-dB down bandwidth can be a tool for measuring the uncertainty in the angle of arrival measurement. Larger bandwidths would correspond to higher uncertainties, and the uncertainty in the range can be calculated by finding the difference between 86 3D Noise Field - Period 300 75 80 80 60 70 60 40 40 65 20 20 60 0 C0 -20 -20 55 -40 -60 50 -60 -80 -80 45 0.3 Vert Score= 0.1953 - 150 100 50 E -2020 E0 -4r -2040 00.2 -2060. -50 -150, -100 0 Si0.1 -2080 -------2100' 0 100 -- -100 100 80 60 40 Ambient Noise Level (dB) -50 0 50 Azimuthal Bearing, degrees -1980 -2000 - - --- - - - - - - - - - - x (m) - - ---- 200 300 Time (s) 0 400 0 100 200 300 400 Time (s) Figure 4-14: The resulting 3D acoustic pictu re from the range finding experiment conducted with a yoyo period of 300 m. 3D Noise Field - Const depth 75 80 80 60 SI 4) Ci 40 CU 20 40 65 20 60 0 'U 60 70 -20 -20> 55 Si cy 00 -40 -40 -60 50 -60 -80 -80 45 Azimuthal Bearing, degrees -1980 150 50 E 0i -50 z -2040 a. -2060 40 Ambient Noise Level (dB) Vert Score = 0.021741 0 0.2 -2020 - -100 -150, 60 0.3 -2000 -- 100 80 -100 t CU 0.1 -- -2080------------ -0 x (m) 100 -2100- 0 100 200 300 Time (s) 400 0- 0 100 200 300 Time (s) 400 Figure 4-15: The resulting 3D acoustic picture from the range finding experiment conducted at constant depth. 87 Measured angle of arrival Suncertainty Vert 0.195 Vert=0.022 70CO65 C 600 C. 55 - c E 5045 40 -80 -60 -40 -20 0 20 40 60 80 Vertical Angle Figure 4-16: A comparison of the vertical noise arrival structures at the azimuthal location of the sound source for two experiments. The blue line is the arrival structure from the experiment with a yoyo pattern, and the red line is the same but without the yoyo pattern. The peaks of the vertical noise profiles represent the measured direction the reliable acoustic path (RAP) sound rays, which can be used to predict the distance to a source. the two distances that correspond the angles at the edges of the 3-dB down band. We can also observe that even though we expect the rays to arrive only from a negative elevation angle, there is very clearly a mirror image that can be seen above the horizontal plane. That differential, however, is less apparent in the yoyo pattern experiment. Now let's compare the measured vertical angles, the calculated ranges, and the uncertainties for all 8 experiments run. The results are graphed in Figure 4-17. We can see that for all the calculations, the range is approximately correct with some variation. The ranges for these experiments measure at 46.1 km, which is approximately a 15% error from the actual distance of 40 km. Further observation of the results shows that the distances fall into only 3 discrete values. Recall that the pixel resolution of these pictures is only 3'. While linear approximation is used for the 3-dB down bandwidth, the peaks will still always occur at a particular pixel, resulting in discrete values for the range calculations. Finer resolutions (i.e. more 88 80 1 70 09 0.8 060.7 C 50 40 - - - - - - vi 30 E 0.5 0.4 - QJ F 0.6 0.320 0 U 0.20.1 10 - 0 0.05 0.1 0.15 0 0.2 Vert Score 0.05 0.1 0.15 0.2 Vert Score Figure 4-17: The graph on the left shows the results of the range measurements. Left: The red squares represent the calculated range of the source based on the location of the angle of arrival, while the vertical red lines represents the error range specified by the 3-dB down uncertainty measurements. The dashed black line shows the actual distance of the noise source. Right: The uncertainties, as compared to the vertical score of the experiment. These values are simply the magnitudes of the red vertical lines in the graph to the right. computing power) would be needed for more varied results than the ones presented here. It should also be noted that the results are also sensitive to the parameters used, particularly the sound speed gradient, which, as mentioned in Section 4.4.2, is only an approximation of the actual, more complicated sound speed profile. Despite these difficulties, there are still some conclusions we can reach. The first is that it is feasible to measure distance with a towed array using the vertical angle of arrival structure, and that the angle of arrival is clearer, or has a higher signal excess, when the yoyo pattern is used. The second is that the uncertainty in the measurement is fairly large, with values ranging around 30% to 60% of the distance being measured. Part of this is due to the relatively low resolution in the measurements, and the other part is due to the fact that small changes in the angle of arrival mean large changes in the distance estimation. By finding methods to increase the accuracy and precision of this measurement, it may be possible to achieve better results. Finally, it does appear that perhaps higher vertical scores correspond to better measurements and smaller uncertainties, but with the limited data presented here, we cannot positively 89 state that this is the case. The work presented here shows that range finding is in fact feasible, however, further study should be done to better understand the effects of verticalness on the distance measurements. 90 Chapter 5 Horizontally Isotropic Vertical Noise Fields 5.1 Motivation In addition to just looking at simple sources, we will also take a look at the ability of a towed array to evaluate horizontally isotropic vertical fields. One major motivation for this research is related target to tracking. A unique characteristic of the Arctic environment is the presence of a noise 'notch' in the vertical ambient noise profile, where the ambient noise from that particular direction is much lower than the noise from other vertical directions, as shown in Figure 5-1. This notch is the result of the unique Arctic sound speed profile. The reason for the noise notch is a unique result given the arctic sound speed profile, where a bend in the gradient occurs. The bend causes some sound rays to be trapped above and turn back, while others will bend down and move lower before turning up - leaving a direction from which there is less ambient noise. The bend in the sound speed profile and the presence of the notch are shown in Figure 5-2. This noise notch is important in target tracking and acoustic communications, because it could be exploited in order to increase the signal-to-noise ratio (SNR) of a signal, which is directly related to the ability of a receiver to distinguish a signal, as per the sonar equation [20]: 91 900 50 30 00 Ambient Noise Notch -90 Figure 5-1: A simple illustration (not from real data) of what the vertical ambient noise profile may look like. SNR = SL - TL - N (5.1) where, SL is the source level, TL is the transmission loss, and N is the ambient noise level. If a vehicle is trying to listen to a target or an acoustic communication signal, the vehicle could maneuver in such a way that the vertical angle at which the signal is coming from is the same as the direction of the vertical noise "notch". This will result in a reduction in the noise level N, and thus a higher SNR. Thus, if the vertical notch does exist, it could be useful if a towed array was able to identify it. In this chapter, we discuss this possibility. 5.2 Experimental Setup In this experiment, no noise sources were present. Using the existing LAMSS software, a noise "notch" was created at an angle of -30' independent of depth, which is a reduction in the ambient noise form that direction. 92 The vehicle was then run at a BELLHOP- Arctic Environment, incoherent 0 -req900Hz '*0 50 Bend in the SSP 1000 60 1500- 65 2000 - 70 S2500- 75 3000- 80 3500 85 4000 90 4500- 95 50CC 1440 - ; 1460 140 500 1520 Sound Speed (mis) 1540 1500 0.5 1 1.5 2 2.5 Range (m)10 3 3.5 4 4.5 5 100 Figure 5-2: The presence of a "notch" is expected in the ambient noise profile given the affects of the sound speed profile on the array paths taken. number of different periods corresponding to different levels of vertical directionality, so that they may be compared against one another. The different periods specified are the same as those presented in Table 4.1, with the exception that a constant depth experiment was not run. Because the primary direction we are concerned with is vertical, the pixel resolution of the 3D pictures produced from these results was changed to 1' degree increments in the vertical # direction, and 100 increments in the azimuthal 0 direction. In order to get a true picture, a virtual experiment wherein an AUV with a vertical array of the same size as the DURIP array was used. The AUV with the vertical array was run in the same pattern as the other experiments, with no change in the depth column, and the same 3D algorithm was run. The base picture, against which we compare the results from the other virtual experiments, is presented in Figure 53. The notch is exaggerated to what we would likely see in the environment for experimental purposes, dropping -25 dB below the ambient noise level. 93 so 80 50 60 60 45 440 40 20 (0 0 20o 0 -20 -20 30 -40 -40 -60 -60 23 -80 -80 -150 -100 50 0 -50 Azimuthal Bearing, degrees 100 150 20 40 60 Noise Level (dB/deg) Figure 5-3: The noise notch as measured by a vertical array. Experiment 1 2 3 4 5 6 7 8 Period 2000 1500 1000 700 600 500 300 N/A Vertical Score 0.059 0.082 0.123 0.139 0.155 0.190 0.206 1 Min Beam 42.4 dB 42.4 dB 42.6 dB 43.2 dB 41.9 dB 42.6 dB 41.6 dB 21.4 dB Min Location -4.0 -1.0 -1.0 -10.00 -25.00 -27.00 -22.50 -28.50 Table 5.1: A summary of the results for the towed array experiments in an attempt to identify the vertical noise notch. While the minimum location does in fact change, the minimum peak does not approach anywhere near to the minimum of the field measured by a vertical array. 5.3 Initial Results The results from 3 of the 8 experiments run can be seen in the measured field noise pictures of Figure 5-4. Pictures for the results of all the experiments can be found in Appendix F. The vertical average noise profiles plotted against each other can be found found on Figure 5-5, and a summary comparing results of the experiments is listed in Table 5.1. The first and most obvious observation is that the measurement of the vertical notch by the 3D algorithm was not particularly successful. We can see that the total drop in decibels is very small compared to the drop that we see in the measured vertical array. However, if we look more closely, there are some hints of the measured 94 Vueticalness 0.05 50 so so5 60 v 40 10 cij 20 C 0 4S 40 -20> 0 0 35 U -40 .50 E -60 30 -80 2 -150 -100 5Sos25 150 - 0 ee 4 45 Noise Level (dB) 4^^-, 0-..4... Verticalness 0 50 0 -50 0.12 80 60 so 50 404 20 .r 0 40 3545 00 Q 3 -20 -40 .So UJ 30 -60 -80 -150 50 0 -50 -100 100 ISO 25 A-- 0-0- Verticalness * 0.21 45 so s5 Noise Level (dB) 55 80 50 40 C C 45 C S20 40 0 9 -20 0 o 3 -.40-5 -80 -150 -100 A -50 -,.. 50 0 l f 4% 100 .. , 150 25 So 48 46 Noise Level (dB) Figure 5-4: The vertical notch noise field as measured by 3 towed array experiments. The top experiment is the picture resulting from a yoyo period of 1500 m, the middle with a period of 1000 m, and the bottom with a period of 300 m. While the pictures really don't capture the actual notch, the towed array experiments above a vertical score of approximately 0.1 seem to show some evidence of a notch. The actual location of the notch is denoted by the dashed line on each of the pictures. 95 55 vs. Vertical Array -Towed - 40 z 45- --- vert - - 0.026 0.10 ___Vert ert 0 12 30- - 25 80 60 -40 -20 0 20 40 60 80 Vertical Angle Figure 5-5: The vertical notch as measure by 3 towed array experiments. The yellow line is the result for a yoyo maneuver with a period of 300 m, the red for 1000 m, and the blue 1500 m. The towed array appears to give poor measurements of the presence of a notch. noise field present in the results. Figure 5-6 shows a zoomed version of the results for the experiments running with vertical scores of 0.2, 0.12, and 0.08. In the experiment with the highest vertical score, we can see some evidence that a notch is detected. There is a gradual drop in dB across the entire vertical domain and we can see that the minimum peak bottoms out at -27', which is very close to the actual bottom at -30'. As the vertical score drops, that measured response minimum gradually moves its way into the center. The experiment with a vertical score of 0.12 shows a minimum at -22', and when we move and lower, to a vertical score of 0.08, the minimum moves all the way to the center. Any vertical scores below -0.10 appear to have this same behavior, where the notch cannot be detected at all. Measuring the location of the minimum does not, however, mean that these noise pictures make a good representation of the field. The 3D acoustic picture generated by these experiments do not present themselves as a notch picture from the actual noise field, but rather as a gradual descent to the minimum. Also, the drops to the minimums are small, on the order of about -5 dB, about 20% of the actual drop. Also that drop is spread over the entire vertical domain, rather than concentrated at the actual location of the notch. 96 Towed vs. Vertical Array 52 Actual Location (-30*) 51 I 50 InI nl I Vertert 47 46-II 45III -80 -60 -40 -20 0 Vertical Ange 20 40 60 80 Figure 5-6: A zoomed view of the average measured vertical noise profile for 3 towed array experiments and a vertical array. This zoomed view of Figure 5-5 shows that there is some evidence of the location of the noise notch in the resulting noise field. However, at the break between the periods of 1000 m and 1500 m, or at about a vertical score of ~0.1, any evidence of the location of the notch disappears as the minimum of the measured noise response moves to the horizontal plane. 5.4 Exploring a One-Sided Vertical Noise Profile In an effort to better understand why the picture was not being generated, a stronger one sided ambient noise profile was generated and the same experiments were run again. This profile was measured by a vertical array in a method similar to that discussed in Section 5.2. Figure 5-7 shows the noise picture as measured by a vertical array. In addition, several new experiments were run. While the limitations of the vehicle only allow a vertical score of -0.20 for the towed array, it is possible within the virtual environment to define completely fixed arrays (like volume arrays) relative to the coordinate system of the vehicle. Two of these volumetric arrays were created and tested against the one-sided vertical noise profile. These had the same dimensions as the normal towed array, but were held at fixed angles of 45' and 75' relative to the horizontal plane, extending in an upward and backward direction from the AUV. The verticalness of these volumetric arrays was measured just like the towed arrays with vertical scores of 0.48 and 0.81 respectively. Experiments with these arrays, as well 97 One SIded - VgdtIcal Ary 80 50 80 60 45 60 40 40 20 20 -20 35 ) -40 1 >40 30 - 0 0 -20 -40O -60 -s0 -80 -150 -100 100 50 Azimuthal Bearina. dearees -50 0 150 20 40 60 Noise Level (dB/deg) Figure 5-7: A full one sided ambient noise profile was developed to re-test the ability of the towed array to match the ambient noise profile. as with a towed array run with a yoyo period of 300 m (for maximum verticalness), were conducted and the results can be seen in Figure 5-8. The comparison of the average vertical profiles for each vertical score can be seen in Figure 5-9. The results show that the towed array data (top picture in Figure 5-8), even with the full exaggerated one-sided noise profile, does not seem to be able to capture the noise field. There is some evidence of a dip, and the average vertical field of the profile is slightly lower on the side that is in fact lower. However, this data does little to capture the magnitude of what is actually occurring in the noise field (see Figure 59). The fixed arrays on the other hand, do begin to capture the magnitude of what's happening. For the 450 tilted array, there is significant evidence of the existence of the one-sided profile. The only problem is that the drop is somewhat smeared across the transition. In the base picture measured by the vertical array there is a distinct drop in the vertical profile between above the horizontal and below the horizontal, while the 45' array tends to start high and then gradually work its way down, even to a point where it drops below the actual low side noise level. In the experiment where we see a tilt of 75', the results are significantly better, except that it tends to over-exaggerated. On the high side, the results are higher than those obtained by the vertical array, and on the low, the results are lower. While we can now see the general shape of the ambient profile, we can still see some discrepancies between the actual picture and the measured ones. We will discuss 98 4 so s0 60 40 45 j2 20 . 40 0 35 .4 20 30 as Azimuthal Barino. dearees 60 50 Noise Level 55 Tilt = 45* Vert = 0.5 40 (Wi) so s0 -80 40 4C 45 2C -20 40 3S **~ZZD -40 -41 =750 25 Tilt 60 0 _6C Azimuthal Beana. dearues 100 50 Noise Level 0 (Em) 55 8I Tilt = 75* \ Vert = 0.8 i so 6 4C 40 45 20C 21 40 -20 CS-21 3S -4( 30 -60 4( 25 Azimuthal Beadna. dearees 100 s0 0 Noise Level (Em) Figure 5-8: The 3D pictures resulting from three experiments run in one-sided ambient noise profile. (Top) A towed array at the limitations of its yoyo maneuver, (Middle) a fixed array held at 45', and (Bottom) a fixed array held at 75 . se, 15 Maneuvering vert limit: -0.2 99 60 vs. Vertical Array -Towed 50 vert - 0. 20 vert = I - ---- - $40 20 10 0 -80 -60 -40 -20 0 Vertical Angle 20 40 60 80 Figure 5-9: The different average vertical noise profiles as measured by the arrays for an exaggerated one sided ambient noise profile for testing purposes. The four experiments shown are (blue) a towed array with a period of 300 m, (red) a fixed array held at an angle of 45', (yellow) a fixed array held at 750, and (purple) a vertical array. some of the possible reasons for these results in Section 5.6. But before that, let us examine how the fixed arrays perform in measuring just the noise notch. 5.5 Revisiting the Notch Profile with Fixed Arrays The section examines the experimental results when using the fixed tilted arrays described in the previous section to again measure the noise notch. The results are shown in Figure 5-10. We can see with the fixed arrays with the higher levels of verticalness that the measured noise fields begin to be measured with vertical scores of greater than 0.80, however, we can still see some significant discrepancies between the measured fields. It appears from these results that very high levels of verticalness, essentially vertical arrays, are required to resolve these sorts of noise fields. 5.6 Beam Responses for a Vertical Notch In order to understand why the vertical responses do not match the actual noise profiles, we can look back at and examine the different beam responses used to create these images. Figure 5-11 shows those beam responses to different levels of vertical 100 55 ---- 4o- ert - 0.21 -Yr-04 80 -60 -40 0 -0 20 40 80 Vertical Angle Figure 5-10: A comparison of the average vertical noise fields generated by a towed array (blue) with a yoyo period of 300 nm, and three fixed diagonal arrays held at vertical angles of 450 (red), 60~ (yellow), and 750 (purple), compared against the vertical array measurement of the noise field (purple). array tilt. To better illustrate the points, I've exchanged the noise "notch" for a noise "bump" and normalized the responses (i.e. the delay vector used to create these pictures has an amplitude of 1). The top left picture shows the normalized responses to a vertical noise field to an array tilt of 20 , the top right 450, the bottom left 750, and the bottom right a fully vertical array. It is clear from observing these pictures that there is no clear indication in the beam response of the bump for lower tilt angles. Even with an array response of 750, the beam response is still only ~-'40% of the what the full beam response should appear to be. The flatness of these beams are the reasons for the poor noise patterns. The iterative method used to create the 3D pictures will not be able to find the notch if the beam responses do not present evidence of the notch or bump themselves. To further amplify the point, we can look at the maximum beam response for a number of array tilts ranging from perfectly horizontal to perfectly vertical. These responses are shown in Figure 5-12. In this figure the blue dots represent the maximum normalized beam responses observed for different levels of array tilt when attempting to resolve the notch, while the red-line represents the best fit exponential curve. The maximum beam responses are small and don't tend to increase until we see array tilts of 80 +, which is why the ability of a towed array to positively identify the location 101 Array Tilt = Aro Til a 45 20* 0.8 so 0.8 50 0.6 0.6 0 M0 0.4 0.4 0.2 -50 0 100 200 300 0.2 >-50 0 0 200 100 300 Azimuthal Bearing Azimuthal Bearing Array Ti1t = 7S* 1 Ar 0.8 so0 I a0 0.8 2 50 0.6 0.6 0 0 0.4 -0.4 02 -50 0 100 200 0 0.2 -50 0 300 100 200 300 Azimuthal Bearing Azimuthal Bearing Figure 5-11: The beam responses for a horizontally isotropic vertical noise field with a bump a -30', as measured by arrays with angles of tilt. of the notch is poor. These results are likely because of a mix between two facts. First is that the circular pattern on the horizontal plane no longer help to break any of the vertical ambiguity the way it would for a point source. For a point source the cones developed from arrays pointing in different azimuthal direction can help resolve the vertical directionality, which is not the case for horizontally isotropic fields by their very nature. The second reason is that the conical beams that would be able to differentiate the noise field, or the ones that don't intersect the noise field and can break the ambiguity, are almost always near endfire, where the beam resolution is poor. This point is illustrated in Figure 5-13. Furthermore, we should discuss the "wavy" pattern we see in the measure of the maximum beam pattern response in Figure 5-12. To begin, in the bottom left picture of Figure 5-11, the "bump" does not present itself as a single bump, but as two individual bumps around the actual location of the bump in the noise field. Remember that the beam pattern is created by convolving the actual noise field with 102 - 1 0.9- 0.8 - 0 a 0.70.6 - , tu CL 0.5- E CU 0.4E :: 0.3- M 0.2 0.1 0 0 10 20 30 40 50 60 70 80 90 Array Tilt Figure 5-12: A graph of the maximum normalized beam responses received from a horizontally isotropic vertical noise field with a "bump" at -30' for different array tilts. The blue dots are data points while the red line is the best fit exponential curve. The "waviness" in the data points is a consequence of the side lobes of the array passing over noise "bump" location at different degrees of array tilt. 103 Conical Beams Array Tilt = 30* Conical Axis Noise "Bump" Figure 5-13: A horizontally isotropic vertical noise field represented on a sphere, as measured by a tilted array. The red dashed line represents the conical array axis, while the white circles represent the lines of conical ambiguity. The color of the sphere is the representation of the noise field with the "bump" at 30' below the horizontal plane. In order for the array to be able to resolve the noise field in the beam response, the separate cones must be able to differentiate themselves from one another by intersecting the noise field by different amounts. In this case, many of the cones do intersect the noise bump approximately the same amount, and the only ones that don't are close to endfire. This is why we see poor resolution for these types of fields when we measure them with a towed array. 104 the array's beam pattern, which has it's own individual peaks and valleys in the form of side lobes. If, for a particular array tilt, the actual noise notch exists between two side lobes of the beam pattern for the relative array tilt, a response similar to that bottom left picture will result, with the maximum response lower than if the side lobes matched up perfectly with the noise bump. As such, small changes in the tilt of the array will result in slightly higher or lower beam responses, which we see evidence of in the form of the wavy graph shown in Figure 5-12. Revisiting the question left off in Section 5.4, this is the explanation for the mismatch in the pictures measured by the vertical array and the diagonal arrays, particularly the array held at an angle of 75'. In a towed array with changing verticality, this is not a problem. However, when we hold the arrays fixed at certain angles, as we did for some of these experiments, we will begin to see some of the effects of the quirks of the beam pattern on the measured noise picture. This phenomenon will affect not only the diagonal arrays, but the vertical ones as well, and we should bear that in mind when evaluating the results. 105 106 Chapter 6 Future work and Conclusions Real World Testing 6.1 All of the experiments conducted for the purposes of this thesis have been conducted in a virtual environment. While the various models used have been tested in the past as reliable, real world testing is the next clear step to moving this research forward. In this section, I will propose a series of 4 real world experiments that will be done to verify and test 6 different research objectives that have been demonstrated in this research through simulation. The ICEX experiment planned for the Spring of 2016 would be an ideal setting to perform these experiments. As we have shown in Chapter 5, a towed array is a poor tool to measure horizontally isotropic noise fields, so some of these objectives may only be accomplished through the use of a vertical line array (VLA). 6.1.1 Objectives The objectives for these real world experiments would be as follows: 1. Show that the increase in towed array verticalness through the use of a yoyo maneuver significantly improves the resolution and signal excess of the 3D acoustic picture. 107 Objective # Ojcie InTime Water Speed Depth Range Range fIe Array yp Description Demonstrate that 1 Verticalness Resolutionvs. 1.5 hrs 1.5 m/s 20 - 100 m 5 km Towed increased levels of verticalness increase the resolution of the noise field Verify the predicted 5, 10, 15, VLA vertical beam pattern at Ice Edge 5 min per drop 0 ~500 m 3 Convergence 30 min 1.5 m/s 20 - 100 m 5 km Towed 4 Ranging 2 hrs 1.5 m/s - 500 m 10 km Towed 5 Bottom Response 5 hrs 1.5+ m/s 20 - 100 m 2-5 km Towed source and observe the 2 Vertical Pattern to 6 Identify Notch I different distances away from the ice edge Demonstrate time to convergence of the noise field. Verify that the distance to Ice cracking noise can be measured by the vertical arrival structure Find the response from a reflected patterns identify and measure the 0 10 min I 20, 25 km - 700 m _I N/A I VLA location of the predicted noise notch Table 6.1: The various objectives that can be accomplished with the ICEX experiment and the approximate time requirements for each. 2. Show that a vertical line array can detect the presence of ice cracking noise in the Arctic. 3. Positively identify and quantify the presence of the noise notch. 4. Use the 3D acoustic picture to identify the range to an ice cracking event. 5. Observe the returns for bottom bounce at different frequencies in and around a source to learn how bathymetry and bottom type affects the noise propagation. 6. Identify and measure the location of the predicted horizontally isotropic vertical noise notch. Demonstrate Adding Verticalness Increases the Resolution The goal of this objective is to show the different pictures created in an actual Arctic environment using the AUV with a towed array. This can be done simply with one 108 experiment by placing the AUV in the water in the Arctic environment, preferably near an ice-cracking event if one can be identified. Then run the vehicle in the water in a number of circular patterns (a radius of 100 m will do), with each circular pattern having a different level of verticality. Finally, analyze the data in the same manner done in Section 4.1.2. Vertical Pattern to Ice Edge The next goal is to measure the vertical arrival structure of the noise coming from the ice edge. These noise patterns have been predicted in a new version of the OASES code [31]. Figure 6-1, gives the predicted noise patterns at 75 Hz for a number of different distances away from a tensile crack source. The different vertical arrivals correspond to noise arriving in the surface duct, noise incoming from the Reliable Acoustic Path (RAP), and reflected noise from the surface and the seabed. Note that in the simulations in this thesis, the reflections from the seabed are removed by placing the bottom below the path of the rays. Identifying Convergence In the same way we identified the time required for convergence in Section 4.3, we should also seek to accomplish this in a real world experiment. The process for this is simple - place the AUV in the water, preferably near some ice cracking noise, and run the vehicle with a yoyo pattern. Then, at multiple points throughout the maneuvering, measure the predicted noise field against the noise field obtained at the end of the run. Ranging This objective is to find the range to a source using the measured vertical arrival structure, using the method described in Section 4.4. Preferably, the location of the source is already known and the goal here will be to verify it with the data. For this test, the AUV should be run significantly deeper, preferably below the typical surface duct, so that the only arrivals happen from the bottom bounced and RAP rays (see 109 OASE S BeamPattern 150 m array 1 m target 75 Hz 20 dB surface noise 80 100 60 90 40% 180 20 1 0 70 < -20 ........ 60 -40 50 -60 40 -80 5 10 15 30 25 20 Target Range (km) 35 40 Figure 6-1: The predicted vertical noise arrival structure for a noise source at 75 Hz [29]. The different vertical arrival structures correspond to reflected noise of the sea bed, surface duct rays, and reliable acoustic path (RAP) rays, which can be exploited in the 3D pictures for ranging. 110 BELLHOP- Arctic Environment. Sound Speed Profie 500- 1000 1500- - 3500 4000- 0 05 1 1.5 2 2.5 3 Range (in) 3.5 4 4.5 1 101 Figure 6-2: The position of the AUV below the surface duct in order to predict ranging from a source. Ranging can also be tested in the surface duct as well, the analysis can be done from other experiments. Figure 6-2. The circular yoyo pattern should be run to observe the noise field and the resulting data evaluated. Bottom Response The objective here is to observe the bottom bounce for a number of different frequencies in order to observe the effects that the bathymetry, and possibly the sub-seabed composition (for example natural gas or oil), will have on the acoustic picture. It helps to have an already known bottom so that it can be compared against the picture, but it could also be run without knowing the bottom. After the locations of the sound sources along the edge are known, the vehicle should be run with circular yoyo patterns at different azimuthal angles away from the source. At each location, the 3D picture from the source can be evaluated to help understand the bottom bounce. Comparing this against the actual known bathymetry in the area will give significant data. With a vehicle speed of approximately 1.5 m/s and a total travel distance of approximately 15 km, plus additional time built in for the vehicle to make the noise run, this experiment could take upwards of 5 hrs. 111 Ice Eg Hot Spot 5km Figure 6-3: To measure the bottom response, the goal will be to run the vehicle in a pattern in which it can develop a 3D noise picture at different azimuths from the noise source. The response from the negative elevation angles at multiple frequencies will be evaluated to see what information about the seabed can be extracted from the picture. 6.1.2 Identifying the Notch This objective is based on the discussion in Section 5.1, and is to find the notch in the vertical arrival structure. The notch will be in the vicinity of the depth ranges from approximately 500 m to 1000 m based on predictions from bellhop. As we learned in Chapter 5, a towed array will not be sufficient for measuring a horizontally isotropic vertical notch, so this will need to be done with a vertical line array, deployed at the depth of the predicted notch. 6.1.3 Proposed Experiments The following are the series of proposed experiments that could be conducted in order to cover the objectives listed in Table 6.1. Experiment 1 In the first experiment, the AUV with the towed array should be placed in the water near (5 km) from the location of where hot-spots are expected. From this location, one could run the AUV in one location in a circular pattern at relatively shallow 112 depths, for example, between 20 and 100 m. This circular pattern will be run many times, but after each consecutive completion of the pattern (or every other circle, if time permits), the period of the yoyo maneuver will be changed to add more and more verticalness into the array for each run. From this data we will be able to test for the convergence (objective 3) of the overall noise pattern for any period we wish, as well as check the verticalness vs. the resolution (objective 1) of the noise picture for any identified noise sources in the area. In addition, post processing of this data will allow us to identify the location of any hotspots. If a hot spot is identified, then we could also attempt to see if there is evidence that shows ranging (objective 4), although this may be difficult because the vehicle is in the surface duct. If, during the experiment the vehicle travels at 1.5 m/s, each circular pattern will take approximately 7-8 minutes, depending on the additional yoyo maneuvers involved. If we test each of the periods presented in Table 4.1 with a single circular loiter pattern, that's 8 circles for approximately one hour of in water testing time, or if each experiment is done with two circles, approximately two hours. When planning also allow an additional hour or two to account for the time needed for transit, launch and recovery, getting to depth, and possible troubleshooting. I total, one should allow a total of 3-4 hours for the experiment. Experiment 2 In the second experiment, it is helpful to have already identified the location of an ice cracking "hotspot" from experiment 1. In this experiment, the AUV should be run in a pattern similar to the path described in Figure 6-3. This path consists of circular patterns with a yoyo maneuver (with as much yoyo as the vehicle limits allow) at different locations but at the same range from an identified hotspot or noise source. Each loiter pattern should be run at a relatively shallow depth, for example, 20 m to 100 m, with the exception of two patterns, in which the vehicle should descend below the surface duct at approximately 500 m to conduct a full circular pattern with a yoyo maneuver there. 113 The purpose of gathering this data is two fold. First, we will be observing the bottom bounce of the acoustic waves (objective 5), and match it against the bathymetry and known seabed type to see what can be learned from the bottom by observing the pictures at different frequencies. This can be done at deep or shallow depths. In addition, we can measure the vertical arrival structure from any noise sources to see if we can use that information to infer the range, either from the bottom bounce and the known bathymetry, or from the RAP (objective 4). This experiment will take significantly longer than the previous one. A vehicle traveling at 1.5 m/s in a semi-circular arc with a radius of 5 km while ascending and descending in the water column to get information at lower depths, will take approximately 3.5 hours. If we add an additional 7-8 minutes for each circular pattern run at each location, and if 8 locations are chosen, that would be an additional hour. Taking these into account, and the the time required for traversing, launch and recovery, and troubleshooting, the expected experimental time would likely take upwards of 6 hours. Experiment 3 This experiment is to test the remaining objectives, 2 and 6. Objective 2 is to verify the vertical noise pattern (i.e. identify the noise notch), and 6 is to observe the predicted noise notch. This experiment will require the stationary VLA. The placement of the array is fairly simple, a simple drop at various distances away from the ice edge. The array does not need to be deployed for long, a couple minutes of data should be sufficient to build a reliable vertical noise profile. For identifying the noise notch in objective 6, the array should be placed at a depth around 700 m in any location, but if possible away from other noise sources. 6.2 Better Ranging Algorithms As we have shown, it is possible using simple geometric methods to find the range to a sound source by measuring the angle for the maximum reliable acoustic path. 114 BELLHOP- ArcUc Environment, Incoherent 200 400 600 1000 12001400 1600 0 0.5 1 1.5 Range (m) 2 2.S 3 -10" Figure 6-4: When ranging to a vehicle, the reliable acoustic path may not be the only direction given the sensor and source locations, but in fact the sound can arrive from multiple directions in the form of eigenrays. In this case, three eigenrays have been found. The red line shows the reliable acoustic path, while the two black rays show the arrival from two rays that have reflected from the surface. However, this measurement makes a number of assumptions. It assumes a constant gradient in the sound speed profile, and also that the ray path arrives only from the reliable acoustic path (RAP). In reality, the sound speed profile is not linear and the arrival structure of the sound may be the result of multiple rays called eigenrays, which can be calculated [20]. Also, depending on the environment, the RAP may not be a path at all, particularly if the bottom prevents it. Figure 6-4 shows an example where these eigenrays have been calculated using an Arctic sound speed profile. It is conceivable that a more robust algorithm could take all these factors into account. If bathymetry data, the sound speed profile, and the sensor and source locations were input into the system, it may be possible to develop more robust algorithms which would allow for more opportunities for ranging (for example, when the RAP isn't present) and with greater accuracy. 115 Strike-Slip Tensile Dip-Slip , - 100.0 Hz SO- 1.0mCONDR.IP P 100.0 Hz SO- 1.0 nCONDRFIP 10 F- 100.0 HS SO- 1.0 CONDR.FIP -55 -5 .0 .0 (0 .5 . 5. 0 - 0 0 Range (kin) .5 10 I .0 . Range (kin) .0 1010 0 Range (km) Figure 6-5: This figure shows a model of three different ice cracking types and their radiation patterns, as modeled by the work of Kim [21]. By measuring the 3D acoustic picture at different locations relative to an ice cracking event, these models could be verified and studied further outside of a modeled environment. 6.3 Ice Cracking Research As mentioned in Section 2.3, the noise sources are the result of multiple ice cracking "hotspots" along the ice edge. These hotspots last approximately for a day in duration. There has been research conducted at MIT modeling the acoustic properties of different types ice cracking noise [21]. Figure 6-5 shows 3 different types of ice cracking mechanisms, and their predicted acoustic radiation patterns. By maneuvering the vehicle in such a way that it develops multiple 3D acoustic pictures at various angles and locations relative to the noise source, we could achieve better understanding the mechanisms that occur in ice cracking and their acoustic properties. 6.4 Autonomy Solutions The ultimate goal for this and other research is to develop a capability for AUVs to be deployed for long amounts of time to survey the ice edge, map the bottom, or track targets. Because acoustic communication underwater is very limited, this ultimate capability will require a sophisticated and reliable autonomy solution. As we learn more about the environmental characteristics of the Arctic, we should endeavor to 116 40 20 -20 -40 -60 -40 -20 0 X Pos (km) 20 40 60 Figure 6-6: A potential utility solution for vehicle maneuvering given the location of multiple ice cracking "hotspots". After the hotspots are identified, the vehicle could attempt to maintain distances to multiple sources for optimum surveillance capability. Using the IvP behavior structure already existing in the MOOS-IvP database, this could be balanced with other objectives the vehicle is attempting to accomplish. use these characteristics to develop better autonomy systems. In the case of tracking an ice edge, an AUV could be programmed to seek and maintain a certain distance and depth from the ice cracking events. This of course would require the AUV to be capable of analyzing its own acoustic data to find these events. Figure 6-6 shows an example of a potential utility function that could be developed for an AUV, given the location and amplitude of ice cracking events. If the type of radiation pattern could also be identified identified (Section 6.3) this could also be included into a more sophisticated algorithm. Other consideration could be to include cooperation with other AUVs through acoustic networks, and minimum energy solutions to facilitate long deployment times. 6.5 Conclusions There are three main conclusions that can be taken from this research, relating to the 3 contributions originally stated in Chapter 1. The first is that adding verticalness to the array does in fact improve the acoustic picture for sound sources when tested 117 in the virtual environment. For a single sound source, the relative 3D picture peak response for experiments with yoyo patterns (vertical score 0.3 or higher) can be on average 6.5 dB higher than than experiments without yoyo patterns (vertical score of 0). In addition, the 3-dB down bandwidth of the vertical arrival structure from the noise source decreased to less than half the result obtained from a horizontal array, from more than 20' to less than 100. In addition, over the vertical scores tested, the horizontal 3-dB down bandwidth did not appear to suffer. The second conclusion is that range estimation is possible given the vertical arrival structure of the sound, and that the uncertainty in that measurement (also measured by the 3-dB down bandwidth) somewhat decreases with range. Using a number of assumptions and some simple geometric calculations, I was able to calculate the location of a source 40 km away to within about 15% accuracy. These measurements however are sensitive to the parameters chosen in the assumptions, such as the assumed constant sound speed gradient. Also the uncertainties in these measurements is fairly high, measured at 30% to 60% of the actual distance. Additional research and experiments could be developed to make these calculations more accurate if actual ray tracing models are used with a full sound speed profile, and multiple arrival paths are considered using eigenray calculations. Finally, I've shown that horizontally isotropic vertical noise fields cannot be measured by a towed array for any realistic levels of verticalness. Evidence of a vertical "notch" in the 3D beam response generally doesn't exist until array sees tilt angles of more than 75'. These types of fields are best measured and characterized by vertical line arrays. 118 Appendix A Modeling the Pressure Field A.1 Ray Tracing In Section 4.4.2, we developed a simple geometry based on ray tracing, however, the acoustic model for developing the acoustic signals are significantly more sophisticated. In this section, we will discuss the derivation of how rays are traced in the ocean based on the derivations provided in Computational Ocean Acoustics [20]. Rays are vectors normal to a wavefront. The premise for ray curvature begins with notion of Snell's Law, which states that a relationship between the grazing angle of a wave 0 and the sound speed of a medium will always be constant. cl C2 cos0 1 cos02 Constant (A.1) So, as the speed of sound in which the acoustic waves are traveling speeds up, so does cosine of the grazing angle go down. To make some more intuitive sense of this, use the definition for the wave number k: k 2rf C (A.2) With this information, we can rearrange Equation A. 1 to the following: k, cos 01S -k2 COS02(A3 119 (A.3) Figure A-1: The incident angle of a ray will change based on the sound speed of the media in which it travels. and since w = 27rf and the frequency of the wave does not change, k, cos 01 = k 2 cos (A.4) 02 As illustrated in Figure A-1 for a stratified media the incident angle of a ray will change as the the sound speed of the media changes. For a linear sound speed profile with a constant gradient in an ocean acoustic environment we would expect the rays to be traced in a circle, similar to the way they are traced in Section 4.4.2, with a center equal to the location where the sound speed is equal to zero. However, for any general sound speed variation we can perform ray tracing using a general set of coupled ordinary differential equations, derived from the Helmholtz equation: A 2 P +W P = -6(x - xo) c2(x) where x is the set of cartesian coordinates x = (A.5) [x, y, z]. We begin by defining the pressure from a point source in what is called the ray series: p(x) = eiWT(X) (A.6) j j=0 (iW)j If we take the second derivative of the pressure with respect to x, we can write: +A j=0 (iw)j 120 + AA + A 2A (+2)w j=0 j=0 (iw)j (A7) Plugging this result back into the Helmholtz Equation A.5, we can obtain a series of equations based on the order of each of the w terms. The two equations that are found on the order of O(w) and O(w'-j) are known as the transport equations, and the remaining equation, found on the order of O(w 2 ) is known as the eikonal equation. For the purposes of ray tracing, it is this eikonal equation that we are concerned with, and is written as such: AT 12 = 1 c2 (A.8) (x) From here, we need to define some further terms. In ray coordinates, we can define the ray trajectory x(s) as by the differential equation: dx = CAr ds (A.9) where s is the arc length along the ray. From the eikonal equation, the right hand side of Equation A.9 can be shown to be unity. If we consider only one direction x, and differentiate Equation A.9 with respect to s, we find: d ds - - cds) d /dx OT) - ds o)x 92 TT&X 02T(Y - 2 -9x - &s + (A. 10) Oxoyas Substituting this back in Equation A.9, this can be re-written again as: d ( ds dx c ds _ &2 TOT =+c OX2 OX 2 & T&o + 19x1y - y) cT C -- [T OT - 2 2 Ox -Ox) +(0 r 21 (A. 11) OY and since the term in the square brackets is equal to the left hand side of the eikonal Equation A.8, we find that: d (1dx) ds cds c 0 1 2Ox c2 1 ic c 2 Ox If we apply the same method we used for the x-direction for all direction x = [x, y, z], the vector equation for the for ray trajectories can be obtained: 121 d (1 dx ds c ds 1 (A.13) c2 This equation is the basis for the ray equations. If we transform the cartesian coordinates to cylindrical coordinates about the ray source, we arrive at the general ray-tracing equations: dr -= c ((S), ds d ldc ds c2 dr dz d( 1 dc ds ds c2 dz (A.14) (A.15) In these equations, r(s), z(s) is the trajectory of the ray in the range-depth plane ( in cylindrical coordinates, where s is the arc length along the ray. The variables and have been introduced so that the equations are expressed in first-order form only. These equations, along with a set of initial conditions for the source position (r, zO) and take-off angle A.2 00, are the basis upon which ray paths can be calculated. Finding the Phase Difference The phase difference is found by solving the eikonal equation in the cylindrical coordinate system of the rays [20]. From the eikonal equation, Equation A.8: AT - AT = 12 AT - 2 c ldx c ds 1 =d - I C2 (A.16) or dT 1 ds c (A.17) If we then take the integral: T(S) = T(0) + 1 ds' 122 To c(s') (A.18) we can find the travel time along the ray - and thus the phase difference of the ray is just delayed by the time it takes for the pressure wave to travel along the ray. A.3 Calculating Ray Pressures Recall that from equation A.7 we obtained three equations, one the eikonal equation, and the other two were the transport equations. Again, the method presented here is from Computation Ocean Acoustics [20]. To find the pressure, we are concerned with the equation resulting from the terms with w on the order of O(w): 2AT - AAO + ( 2 T)Ao = 0 (A.19) or A - (A0 AT) = 0 (A.20) Now, if we consider Gauss's theorem (or the divergence theorem) for an arbitrary field F in any volume V, we know that the integral of the divergence of F (A - F) is equal to the flux of the field running the volume V. Using this, we can rewrite Equation A.20 in the following form: A0 AT - ndS (A.21) where V is any volume with a surface S with the normal vector n. This is useful when we look at rays, because we want to define what we call a ray tube as the volume with which we're concerned, which contains a family of rays. On the ends of the ray tube, dx we define the normal vector n = -, ds' and since the rays follow the inside of the ray tube by definition, the termAr -n = 0 on all the sides of the ray tube. These are all illustrated in Figure A-2. Lastly, if we take the eikonal equation written in terms of the ray coordinate s, we see that AT - n = 1/c and we can obtain the following using the energy conservation law: 123 Vron=0 ayV family of rays 8V ------- n= dx /ds Figure A-2: An illustration of a ray tube. The flux of the rays through the sides is zero, and the normal vectors to the ends are defined as n = dx/ds. j A0dS = VOC av = const A2dS a V, A2C (A.22) where &V and &V1 represent each of the end caps of the ray tube. If we let the ray tube become infinitesimally small and set an arbitrary value for the amplitude of the pressure at s = 0, we can conclude that: Ao(s) = Ao(0) (A.23) c(S) J(0) c(0) J(s) where J(s) is a quantity proportional to the cross sectional area of the ray tube. J(s) can be calculated by the hypotenuse of dr and dz for a given section of the array: - )2- 2 (00 00 1/2(A.24) _ where r is the distance in cylindrical coordinates, assuming symmetry about the zaxis (depth in the water column). We can also see that the value J can also be found by taking the Jacobian with respect to (s, o, #0), where 0 and #o are the declination and azimuthal ray take-off angles, respectively [20]. With these relationships to the cross sectional areas to the tubes, we can now set some initial conditions to find the resulting pressures. We set the initial pressure conditions at the beginning of the ray to be as follows: po(s) = Ao(s) eiwTo(s) where 124 (A.25) Ao(s) = 1 and , To(s) = 2 (A.26) and co is the sound speed at s = 0. This poses a problem because for Equation A.23 the amplitude appears to go to infinity when s = 0. We can solve this problem, however, because the limit of A(s) J(s) 1/2 as s -s oc does in fact converge to a finite value: lim A(s) J(s)|1 / 2 s-+oo 1 47r cos 00 1 2 / (A.27) and we can therefore rewrite Equation A.23: Ao (s) =II c(s) cos 0o 47r (A.28) 1/2 c(0) Putting this amplitude equation together with the phase difference (Equation A. 18) and plugging it back into the original pressure Equation A.25, we can now solve for the pressure at the end of the ray: p(s) = A.4 c(s COS0 1/2 ei fo 1/c(s')ds' (A.29) Finding the Total Pressure Now that we have solved for the pressures of a generalized ray tube, we now want to find the pressure resulting from a source and receiver. We do this by finding the eigenrays between the two points using the ray tracing calculations and the definition of the bathymetry in the environment (because rays will bounce off the surface and the bottom). By this method, we identify all the rays that originate from the source and end at the receiver at their corresponding incident angles and pressure losses. Each ray makes a contribution to the complex pressure field and therefore we sum the contributions from each of the eigenrays, such that: 125 N(r,z) Ptot (r, z) = E Pj (r, z) (A.30) j=1 where N(r, z) represents the total number of eigenrays for that particular range and depth, and p1 represents the the pressure from the jth eigenray. This gives us the pressures we use in the acoustic model. 126 Appendix B Bellhop Inputs The following is the information that was used as inputs into the Bellhop Program: frequency = 800 Fitting type: SVF (Analytic or C-linear Interpolation) bottom = 3750 m Run Type = I (R = Ray, C = Coherent, I = Incoherent, S - Semi-coherent) Number Beams = 0 (Allows matlab to choose optimal number of beams) Number of Sources = 1 Source Depth = 5 m Number of Receiver Depth Steps = 201 (every 2 m) Receiver Depth Range = 0 400.0 (0 to 400 m) Number of Distance Steps = 501 (every 10 m) Receiver Distance Range = 0.0 5.0 (0 to 5 km) Step Size = 0 (automatic), ZBOX = 500, RBOX = 5 127 128 Appendix C Modeling of a Towed Array C.1 Modeling of an Array One of the major areas that needs to be understood is the modeling of an array. This is done in the virtual environment by the program uSimTowedArray, which models the x,y, and z coordinates. uSimTowedArray is based off a Matlab program pArraySim, which has been shown to match fairly consistently with real world experiments [7]. To begin describing the model, we first make a series of assumptions: Small Perturbations are Negligible Local perturbations of the cable from its overall shape (due do small local eddies and other environmental factors) are negligible. Torsion and Rotational Intertia are Negligible The rotation and torsion of the array about the axis normal to a cross section of the array cable are both negligible. Bending Stiffness is Negligible Bending the array cable does not produce a bending moment. Instead, the cable acts as a "loose string" being towed behind the AUV. No Effect From Waves Waves from the surface do not affect the motion of the array. This is a particularly good assumption given that the AUV will be 129 operating at depths of more than 20 m below the surface, where most surface waves have little effect. Based on these assumptions, a set of governing equations is developed [16]. While the actual equations work in three dimensions for the array, only two dimensions will be discussed here in order to keep the equations succinct. (On fU=m at F,=m at Da N v at T 1 -wcosa--pdCtulu| 2 as u +m, at ) at =T as -2sina- 2 pdCHvv (C.1) (C.2) In the above equations, Fu represents the forces in the tangential direction toward the cable and F, represents the force in the normal direction to the cable. These terms are expanded on the left hand side of the equations and represent the acceleration of motion of an element of the array. The right hand side is a sum of the forces on the OT array. In Equation C.1, the u-direction is the change in tension in the tangential Os 1 direction, the -pdCulul term is the drag on the array element due to the element 2 moving through the water, and w cos a is the effect of gravity on the array element transformed from the x-y coordinate system to the local u-v. In Equation C.2 for the v-direction, the terms are similar with the exception that there is an added mass Ov term, ma OV, that occurs due the need to account for the moving fluid around the at outside of the cable. The balanced forces on the element are shown in Figure C-1. The compatibility relations in terms of the velocities are expressed as follows: O- Oa =O Os as Ot Ov Os a Oa as at (C.3) (C.4) In order to separate the static and dynamic parts of the problem we expand the the tension T ~ T + T and the angle a ~ d + d. Using these decomposed forces we can account for the forces of the current in the water in Morison form: 130 U 7 Aa u-drag a S v t Yu v-drag Figure C-1: The individual forces that act on an element of the towed array. On the left side is a picture of an element on the array, and on the right are the forces (not including current) of the forces that act on the element. 1 Fe-rrentu-2pdCU sin ajUsina| 1+ 2 (C.5) TE 2EA (C.6) 2 1+ ) ( 1 ,IYT IT Fcurrent,v =-pdCtUcosaU cosa| T) 2 EA where U is the is the current in the x-direction. Next through a series of steps, we remove the static terms, and use the small angle assumption on d. We also remove nonlinear terms higher than the first order, allowing us to substitute OP for u and at dq for v. With these terms and the expanded forms of T and oz, we can derive the at resulting dynamic equations: a2p at m2 +wocoscoa Jatap - Ucosd 1+2EA ( pdCaP + 2Pt at - Ucos- 1 + 1pdCtU cos dIcos a 2 (1 2EA) + as 131 (C.-7) Parameter Nsub Length (in) Sector 1 20 Sector 2 30 Sector 3 20 20.0 36.0 20.0 0 0.0095 0.2 0 0.035 0.0024 0.5 0 0.0095 0.001 0.2 1.0e8 1.0e8 1.0e8 Weight (N/m) Diameter (in) 0.001 td nd E (N/m ) 2 Table C.1: A table of the variables used in modeling the DURIP array. - T ad ~ ad = - T 02 (m 1 ma) Ot 2 + IPdC O + -wsin a a g O +U ~) 1 pdCdUTsinTsin. 2 Pd snasn I+ kj ~ (i\ k '1 __ 2EA) (C.8) T 2EA} T The 1 + 2EA above comes from the definition of strain that c = T/EA and the compatibility relation: 1+ - =1+ 2 2EA (C.9) These are the governing equations which are used to solve the shape of the towed array. For modeling the DURIP array, the array is divided into three sections, the 20 m tow cable (sector 1), the 36 m acoustic body with the center 30 in as the towed array section (sector 2) , and the 20 m drogue (sector 3). The parameters used for each of these sectors in the cable model are shown in Table C.1. 132 Appendix D Spherical to Conical Transformation For an assumed straight array, a conical coordinate system can be used to compare the 2D beam pattern to the 3D one. In addition, one can find the conical angle 3 through this transformation [36]. The geometry for the transformation can be derived from Figure D-1. The array axis is represented by the line OP and # is the angle of the conical beam. To keep the calculations simple the sphere is presented as a unit sphere. To begin the derivation we geometrically define some relationships: Oa = cos$ sin $ -c= (D.1) (D.2) Ob = Oa cos 0 = cos $ cos 9 (D.3) ab = Oa sin 0 = cos $ sin 9 (D.4) 133 P Y conica\ ray I Pherical Noise Equator Figure D-1: The measurements needed for the geometric transformation between the spherical and tilted conical coordinates. The sphere represents the noise sphere, OP represents the axis of the array with tilt a, and 3 is the angle of the conical beam. 134 a-J cos-'(Ob) = cos 1 (cos # cosO) S=tan-1 = tan- ( (D.5) 0_ sin (D.6) Cos # sin 0 In the above equations the bar represents the distance of the line segment between the two points. Now, using the law of cosines for the spherical triangle, defined by - the points c, P, and X < PXc on the unit sphere with an interior angle of (90 at the angle < PXc, we can come up with the following equation: - cos Pc = cos Xc cos XP + sin Xc sin XP cos(90 (D.7) where the rocker (-) denotes the are length on the surface of the unit sphere. Now we define some additional relationships among the arcs: XP = a, cos(90 - ) PC =,3, = sin = sin and tan-1 Xc cos- 1 (cos # cos 0) = sin tan- 1 6 ( sin q5 cos # sin 0 (D.8) (D.9) Next we take these relationships and definitions, and substitute them back into Equation D.7. Taking the inverse cosine of that result yields the following: = cos- 1 cos # cos O cos a + sin[cos--1 (cos# cos 0)] x sin a sin Itani sin# cos# sin9)_ } (D.10) This completes the transformation between the tilted conical coordinates and the array coordinates. 135 136 Appendix E 3D Acoustic Pictures for a Simple Source 137 3D Noise Field - Period 300 80 80 75 60 40 70 L 20 O' 0 0 65 60 -20 > 55 -40 -60 50 -80 45 0 -50 Azimuthal Bearing, degrees L- ------ -Vert Score 150 0-- 100 -20-------------- 50 0 -50 -100 -150 0 100 50 Ambient Noise Level (dB) 0.3- 0.20003 = - - - ------ -60- 600.1 -80-100--- -- -- -- - --- - -0 -100 0 x (m) 100 0 100 200 7- I 300 0 400 100 200 300 400 Time (s) Time (s) Figure E-1: Simple Source Results for Period 300 3D Noise Field - Period 500 80 80 1 80 75 60 ; 60 a) 0) 40 4 40 70 20 20 65 C 0 0 60 -20> -20 S-40 -60 55 -40 50 -60 -80 -80 45 100 0--20 - 50 A-4 .50 0.3 --------------- -- - - - - - - - - - - - - -- > -80 -100-----100 0 x (m) 100 J -U -50 -100 -150 0 ---- 100 300 200 Time (s) 400 Vert Score= 0.19349 0.2 0.1 0 0 100 200 300 Time (s) 400 Figure E-2: Simple Source Results for Period 500 138 0 Ambient Noise Level (dB) Azimuthal Bearing, degrees 150 100 50 3D Noise Field - Period 600 80 80 75 40 60 40 70 20 20 65 0 0 60 -- -20> 20 -40 55 -40 -60 so -60 -80 -80 45 -150 -100 0 ------ ---- - ----- 50 0 -5 0 60 40 Ambient Noise Level (dB) 80 re CL -0-6. -80 -100 0 x (m) 100 = 0.15525 0.2 -4 1 - - --100 - - - - - - 0 100 200 300 Time (s) -L 0 400 100 300 200 Time s) 400 ( 0 150 0.3 15 0 10 -10 0 -15 i 100 50 -50 0 Azimuthal Bearing, degrees Figure E-3: Simple Source Results for Period 600 3D Noise Field - Period 700 so 80 75 40 c 70 60 40 M 7 20 65 20 00 0 60 -20 -40 - 20 55 -40 50 -60 -80 45 Azimuthal Bearing, degrees 0.3 150 100 -20 - - - - 50 E 0 400 0 0.2 t0.1 -50 -so- -100 -150, 013631 - - - - - - - - - - - -100 0 x (m) 100 -100- - - --100 0 -- - -- 300 200 Time (s) --400 0 0 100 300 200 Time (s) Figure E-4: Simple Source Results for Period 700 139 400 40 80 60 Ambient Noise Level (dB) > 3D Noise Field - Period 1000 80 80 80 75 60 40 60 40 70 20 / 65 20 CP 0 0 60 -20 -20 t-40 55 -60 50 -> -40 -60 -80 -80 -100 15 0 10 0 E 50 -S0 -10 0 -15 01 -50 0 50 Azimuthal Bearing, degrees L -100 0 100 - - - - - - - - -6c -- -60-80 -100-0 45 150 40 60 Ambient Noise Level (dB) 80 0.3 - -20 - - - - -40- x (m) 100 Vert Scar e 0.12258 0.2 t 0.1 0 - 0 --- -- - --100 200 300 400 Time (s) 0 100 200 Time ( -150 300 400 s) Figure E-5: Simple Source Results for Period 1000 3D Noise Field - Period 1500 80 80 80 75 60 70 40 4 S20 65 0 40 20 C 0 o 60 -20> -20 -40 -60 55 -40 50 -60 -80 -80 45 Azimuthal Bearing, degrees 150 100 0 ------------------ - - - - - -0 -20 - - - - - - - - - 50 0.3 Vert Score = 0.080045 0.2 -40----- 40t -50 -100 -150 V) -80 -100----------------------100 0 x (m) 100 0 100 0.1 - 8j 60 300 200 Time (s) 400 0 0 100 300 200 Time (s) 400 Figure E-6: Simple Source Results for Period 1500 140 80 60 40 Ambient Noise Level (dB) 3D Noise Field - Period 2000 80 80 75 60 40 70 01 20 65 :2 00 P 60 -20 c 55 -40 50 -60 -80 45 .LUU I.:)U Azimuthal Bearing, degrees 15 10 5 -5 -10 0 00 3 0.-0 -15 -100 0.3 0 -------------- - - - - - - --- S20 Vert Score > 0 100 0 100 x (m) = 0.057055 0.2 -40 80- -- 80 60 40 Ambient Noise Level (dB) 200 300 Time (s) 400 0,1 0,0 100 300 200 Time (s) 400 Figure E-7: Simple Source Results for Period 2000 3D Noise Field - Const depth 80 80 80 75 60 e 40 60 40 70 20 51 20 65 0 0 60 -20 -20 N U -40 55 -40 -60 50 -60 -80 -80 45 10 0 Azimuthal Bearing, degrees 150 100 0 ------------------------20 - - - - - - - - - - - - - - - - w' -40 50 0 Vert Score = 0.018588 0.2 -60t -50 -80> -100 -150' 0.3 ---------------------- -100 -100 0 x (m) 100 0 100 200 300 400 Time (s) 0 100 200 300 400 Time (s) Figure E-8: Simple Source Results for Constant Depth 141 80 60 40 Ambient Noise Level (dB) 142 Appendix F 3D Acoustic Pictures for Noise Notch 143 3D Noise Field - Vert array 55 80 80 50 Ln y 60 40 40 01 45 20 02 40 0 0 -20 -20 6C 35 -0, -40 30 A8C -60 -80 25 -50 0 50 Azimuthal Bearing, degrees 150 100 Score = 0.98833 200 300 Time (s) 400 Vert - -60 -80 -100---- 0.2 - -50 -100 -150' 0.3 0-----------------------20 - - - - - - - - - - - - - - -40 so 60 40 20 Ambient Noise Level (dB) -100 0 x (m) 100 0 0.1 -------- -100 200 300 Time (s) -- --400 0 0 100 Figure F-1: Vertical ambient noise field results for a vertical array. 3D Noise Field - Period 300 55 80 80 60 0 50 60 40 40 45 GO (U 20 20 40 0 Go -20 -20 35 -40 -40 -60 30 -60 -80 -80 -:)u U 25 )U Azimuthal Bearing, degrees 50 E 0 -60 -80 -50 -100 -150, - --- - --- - - -- / - ' -20--0 -40.- - 0 ------------------ 0.3 150 100 Vert Score = 50 48 46 Ambient Noise Level (dB) 0.20641 0.2~^ 0.1/ 0. -- -- - -100------------------------ -100 0 x (m) 100 0 100 200 300 Time (s) 400 0 100 200 300 400 Time (s) Figure F-2: Vertical ambient noise field results for period 300. 144 3D Noise Field - Period 500 55 80 80 50 60 c 60 40 40 45 CU 20 20 40 0 C 0 -20 > -20 35 -40 w -40 30 -60 -60 -80 -80 vUU 25 L)u Azimuthal Bearing, degrees 10 5 -5 -10 -15 0 E 0 0t 00 0 x (m) 100 Vert Score= 0.1903 40.2 -60 -1 -80 > ------- --------- 100 -100 0.3 0 ----------------------. - - - - - - - - - - ---20 - - -- 0 100 300 200 Time (s) - 15 46 48 50 Ambient Noise Level (dB) 400 0 100 0 200 300 Time (s) 400 Figure F-3: Vertical ambient noise field results for period 500. 3D Noise Field - Period 600 55 80 50 60 40 40 45 20 0 20 40 0 0 0 -20 > -20 35 -40 -40 30 -60 -60 -80 -80 25 -50 0 50 Azimuthal Bearing, degrees 0-----------------------20 - - ------- - -- 50 0 -50 -100 -1501 -400 -40,0. -100----------------100 0 x (m) 100 0 100 46 48 50 Ambient Noise Level (dB) 0.3- - - - - - - 150 100 - CU 60 Vert Score = 0.15467 0.2 C0.1 200 300 Time (s) ------ 400 0 100 200 300 Time (s) 400 Figure F-4: Vertical ambient noise field results for period 600. 145 3D Noise Field - Period 700 55 80 so 60 Si Si 40 45 Si V 20 40 0 VS @1 0 -20 > 35 -40 0) 30 -60 -80 25 -50 0 50 Azimuthal Bearing, degrees 50 o -50 0 ---------------------4---)-----20 --40 60t .-80. -----100 I - -100 -150' 0 -100 100 0 100 x (m) 0.3 Vert Score= 0.1388 0.2 0.1 - 150 100 45 50 55 Ambient Noise Level (dB) 200 300 Time (s) 0- 0 400 100 300 200 Time (s) 400 Figure F-5: Vertical ambient noise field results for period 700. 3D Noise Field - Period 1000 55 80 50 60 40 40 45 20 20 40 0 0 U -20 -20 35 -40 40 30 -60 -60 -80 -80 25 Azimuthal Bearing, degrees 150 100 50 0 -50 -100 -150 - -- - - - - -- Vert Score - 0.12349 o" 0.2 -j 0.1 - -60 - -20 - - - - - 45 50 55 Ambient Noise Level (dB) 0.3- 0 ------------------------ (U li -80-> -100- ------------------100 0 x (m) 100 0 100 200 Time 300 (s) 400 0 100 300 200 Time (s) 400 Figure F-6: Vertical ambient noise field results for period 1000. 146 3D Noise Field - Period 1500 55 80 80 i 60 50 60 40 40 45 CM 20 20 40 0 -20 > -20 35 -40 C -4C 30 -6C -60 -80 25 Azimuthal Bearing, degrees 150 -20 - - -40 - - - - - , - - - - - - - 02 - -60 -80 0 x (m) wi > ----------- -100 ----- - -100 100 0 100 0.082 134 Vert Score V. 0.1- ----- 300 200 Time (s) - 50 0 -50 -100 -1501 0.3- 0 -------------------. 100 45 50 55 Ambient Noise Level (dB) 0 400 100 300 200 Time (s) 400 Figure F-7: Vertical ambient noise field results for period 1500. 3D Noise Field - Period 2000 55 80 60 50 40 CU 40 45 20 20 01 40 0 0 -20 -20 35 -40 -i40 30 -60 -60 -80 25 -50 0 50 Azimuthal Bearing, degrees 150 100 -100 -150 0.3 0 ---------------------20 - - - - - - - - - ---------40 50 0 -50 -80 -100- -100 0 x (m) 100 40 50 60 Ambient Noise Level (dB) 0 Vert Score = 0.058683 -0 0-60 .1 --------------------- 100 200 0 300 Time (s) 400 0 100 200 300 Time (s) 400 Figure F-8: Vertical ambient noise field results for period 2000. 147 D 3D Noise Field - Diag 45 55 80 580 60 50 60 40 40 ( 45 20 20 0 40 0 -20 20 35 -40 -40 0 -60 30 -80 I25 -80 50 -50 0 Azimuthal Bearing, degrees -100 -150 150 100 E 50 0 50 f c > Ul - -- -- -- -- -0----------20 - - - - - - - - - - - - - - - - ~-404 _ _ _ _ 40 50 45 Ambient Noise Level (dB) 150 100 0.3 Vert Score= 04848 e 0.2 U__ t -60- -V0 >. -100-80 200 300 Time (s) - -150 -100 0 x (m) 100 0 100 400 0 100 400 200 300 Time (s) Figure F-9: Vertical ambient noise field results for 450 tilted array. 3D Noise Field - Diag 60 55 80 80 50 60 60 40 40 45 20 20 40 0 0 0 -20 > -20 W 35 m-40 A40 -60 30 .80 -80 -100 -150 -50 0 50 Azimuthal Bearing, degrees 150 0 ------- 100 50 45 40 Ambient Noise Level (dB) 0.3 ----------------- Vert 50 Score= 0.65381 0E0.2 at-60 t -8U 0.1 0 -- -100 -100 0 x (m) 100 1 0 100 300 200 Time (s) 400 - 0 - - 0 -50 -150 25 150 100 100 300 200 Time (s) 400 Figure F-10: Vertical ambient noise field results for 60' tilted array. 148 3D Noise Field - Diag 75 - 80 55 80 50 60 60 40 40 45 C 20 20 20 0 -- 0 40 20 -20 -40 -40 -60 30 -80 7. -60 -80 - 25 -150 -100 -50 0 50 100 150 60 Azimuthal Bearing, degrees 150 100 50 E c. -50 -100 -150 0 ----20 -- ------ - --- ------------ --- --- -60 - t 0.1 w -100---------------------0 x (m) 100 Vert Score= 0.82141 o0.2 > -80 -100 0 100 300 200 Time (s) 400 0 0 100 200 300 Time (s) 400 Figure F-11: Vertical ambient noise field results for 75' tilted array. 149 40 Ambient Noise Level (dB) 0-3 _ _ -40 - 50 0 150 Appendix G Noise Code function %% get-3D-noise Load in the settings = '-/thesis-data/simple-src/'; directory period_300'; period_500'; period_600'; 'period-700'; 'period-1000'; 'period-1500'; 'period_2000'; experiment{1} = experiment{2} = experiment{3} = experiment{4} = experiment{5} = experiment{6} = experiment{7} = experiment{8} = % experiment{l} % experiment{2} const-depth'; = 'vert-array'; % experiment{3} = = 'diag_45'; 'diag_60'; % experiment{4} = 'diag_75'; % Get the start and end files you want to test to and from file-start = file-end = 22C 1; % Picture Resoltion theta-resolution = 3; phi-resolution = 3; for expt-num = greatly affects calculation times 1:length(experiment) % Location of files = input-filepath % Location output-filepath for [directory,experiment{expt-num}, deposit = of ACOUSFIELD.mat file [directory,experiment{expt-num}, 151 '/acous/']; '/']; %Processing Information fs = 12000; % Inpute the sampling frequency % Frequencies to process = 900; freqs-to-proc %% Calculate/Initiate % Array Resolution some important in Hz numbers Information thetas= 0:theta-resolution:359; phis= -90:phi-resolution:90; % Environmental Information c = 1495; % Speed of sound in water % Get the f-axis and closest frequencies along it to processing % frequencies, as well as those indicies on f-axis faxis=linspace(0,fs,fs/10); (size freq-index=zeros for (freqs-to-proc)); i=1:length(freqs-to-proc), freq-index(i)=find(faxis>=freqs-to-proc(i),l, 'first'); end fc=faxis(freq-index); % Corrected frequencies to calculate k=2*pi*fc/c; % Wave number for fc's % Initiate the data file data. angles=thetas; data.phis=phis; data. freq=freqs-to-proc; data.info='angles,field,freq,position,info,phis,time'; %% Pull the data from NAS and ACO files, lrmsg = logs 0; fprintf(['*********** for i = add it to ',experiment{expt-num},' ************\n']); file-start:file-end % Get the acoustic and non-acoustic file data names = sprintf('%09d',i); files strcat(input-filepath,'ACO',filenum-str,'.DAT'); files strcat(input-filepath,'NAS',filenum-str,'.DAT'); filenum-str % acoustic acofile = % non-acoustic nasfile = % Get the non-acoustic data fidnas = fopen(nasfile); % Get the timestamp = textscan(fidnas, line-nine 8); 'headerlines', '%s', 1, timestamp = str2double(line-nine{1}); % Get the array positions 152 'delimiter', '\n', next-line = textscan(fidnas, '%s', 'headerlines', 0); element-pos = 0; array-positions = zeros(1,3); == 0 while isempty(next-line{l}) element-pos = element-pos + 1; line-cell-array 1, 'delimiter', = strsplit(next-line{l}{l}, '\n', '\s* \s*', 'DelimiterType', 'RegularExpression'); = str2double(line-cell-array{l}); array-y = str2double(line-cell-array{2}); array-x array-z = str2double(line-cell-array{3}); (element-pos, :) array-positions next-line = textscan(fidnas, 'headerlines', 0); = [array-x,array-y,array-z]; '%s', 1, 'delimiter', '\n', end fclose(fidnas); if i == file-start num-proc-files = file-end - file-start + timestamps-log = zeros(num-proc-files,1); pos-log = zeros (num-proc-files,element-pos,3); length(freqs-to-proc), fx-log = zeros(element-pos, 1; % picket fence! num-proc-files); end % Get the Acoustic Data fidaco = fopen(acofile); temp = fread(fidaco, [2*fs,element-pos], 'float32'); tempdata = zeros (element-pos,length(faxis)); for id=l:(2*fs/length(faxis)) todo= (id-1) *length (faxis); tempdata=temp(todo+(1:length(faxis)),:).'; end tempdata=tempdata./id; fclose(fidaco); freqx=fft(tempdata, [],2); clear tempdata; fx = freqx(:,freqcindex); % Double check if there is NaN in data. if sum(temp)>0 fprintf('found nan in data, skipping %3.0f\n',i); % Reduce the size of the vectors by 1 now. timestamps-log = timestamps-log(l:end-1); pos-log = pos-log(:end-l,:,:); fx-log = fx-log(:,:,l:end-1); continue; end %% Add the data to the logs % Record the fx, positions and the 153 times in a log for processing. % Log will be held back as far as the time % 1st - frame specified Shift the old data back timestamps-log(2:end) = timestamps-log(1:end-1); pos-log(2:end,:,:) = pos-log(1:end-1,:,:); fx-log(:,:,2:end) = fx-log(:,:,1:end-1); % 2nd - Record the new data timestamps-log(1) = timestamp; pos-log(l,:,:) fx-log(:,:,1) fprintf msg = fprintf limsg = = array-positions; fx; (repmat ('\b' ,1, 1msg)); sprintf('loaded file number %3.0f/%3.Of',i,file-end); (msg); = numel(msg); end %% Do the field calculation tic calculation fprintf(['\nStarting experiment{expt-num}, '\n']) % Now Calculate the for ACOUSFIELD field for the current time, for save the fieldN = calculate-noisefield3d(fx-log(:,:,:),freqs-to-proc, pos-log(:,:,:),thetas,phis,c); for ACOUSFIELD calculation Finished fprintf([' experiment{expt-num}, '\n'1) t = toc; fprintf('Time elapsed = %f seconds\n\n',t) data.field = for ', fieldN; size (data.field) data.time = timestamp; data.positions = squeeze (array-positions); save( [output-filepath, 'ACOUS-FIELD'], 'data'); end fprintf('FINISHED CALCULATIONS\n') end function [fieldN]=calculate-noisefield3d(fx, freqs, positions, angles, vert-angles, c) fprintf('building 3D acoustic pictures... threshold=.005;%.005; k=permute(2*pi*freqs(:)/c, [2,3,4,1); lf=length(freqs); lels=size(positions,2); 154 ... data lt=length(angles); lp=length (vert-angles); nsamples=size(positions,1); dd=permute (fx, [3, 4, 5,2, 1]); thetas=permute(angles(:), [2,1]); phis=permute (vert-angles (:), [2,3,1]); arraypos=permute(positions, [1,4,5,3,2]); % Center arrays on first element for i = 1:nsamples arraypos(i,:,:,l,:) = arraypos(i,:,:,l,:) arraypos(i,:,:,2,:) = arraypos(i,:,:,2,:) arraypos(i,:,:,3,:) = arraypos(i,:,:,3,:) - arraypos(i,:,:,1,1); - arraypos(i,:,:,2,1); - arraypos(i,:,:,3,1); end % Calculate the added dists for each phi/theta to the elements dists= (- repmat (arraypos (:,1,1,1, :), [1, lt, lp, 1, 1]) ... .*sind(repmat(thetas, [nsamples,1,lp,1,lels]))... - repmat(arraypos(:,1,1,2,:),[l,lt,lp,1,1]) ... .*cosd(repmat(thetas, [nsamples,1,lp,l,lels])))... .*cosd(repmat(phis, [nsamples,lt,1,1,lels])) ... + repmat(arraypos(:,1,1,3,:),[1,lt,lp,1,1]) ... .*sind(repmat(phis, [nsamples,lt,1,1,lels])); % Calculate the complex beam pattern m=1/lels.*sum(repmat(dd, [1,lt,lp,1,1]) .*exp(li.*repmat(k, [nsamples,lt,lp,1,lels]) [1, 1,1,1f, 1) ) , 5); .*repmat (dists, ... % Find the actual baem pattern in power M=real(abs(m));clear m; %this is r-true; % Find the beam pattern in dB R-true=10*loglO (abs (M)); fprintf('done.\nextracting noise... ') %% Iterate to find the Noise Field Nguess=ones(1,lt,lp,lf); deltalog=[]; endvals=1; prevsig=O; ii=O; msg = [1; while max(endvals)>threshold & ii<60, ii=ii+l; ntemp=10.^(Nguess./10); % Get the delay vector for each element (for each element) ... x-hat=l./(sum(cosd(phis),3)*lt) .*sum(sum(repmat(ntemp, [nsamples,1,1,1,lels]) ... 155 .*cosd(repmat(phis,[nsamples,lt,1,1,lels])) .*exp(-li.*repmat(k, [nsamples,lt,lp,l,lelsl) .*repmat(dists,[1,1,1,lf,l])),2),3); ... % Calculate the complex beam array based on the guessed noise ... M-guess=1/lels.*sum(repmat(x-hat, [1,lt,lp,1,1]) [nsamples,lt,lp,1,lels]) .*exp(li.*repmat(k, .*repmat(dists,[1,1,1,lf,1])),5); [sample#, in dB % Find the R-hat ... phi] theta, R-hat=10*loglO (abs (M-guess)); % Get the average R-bar (average difference) % each sample [sample#, 1] R-bar=1/lt/lp*sum(sum(R-hat-R-true,3),2); sig-j % Get the % [sample#, 1] sig-j=sqrt(1/(lt*lp) = 1/(lt*lp) * sig-p = - R-hat - R-bar)^2 .*(sum(sum((R-true-R-hat -repmat (R-bar, [1, % Get the (R-true value for for samples lt, 1p, 1] )).^2, sig-p = 2),3))) (1/nsamples)*sum(sig-j^2) sig-p=sqrt(1/nsamples*sum(sig-j.*sig-j,1)); Delta= (R-true-R-hat) /2; newN=Nguess+sum(Delta,1)./nsamples; %sig=sum(squeeze (sig-p))/size (sig-p,2); sig=max (sig-p); endvals=abs(prevsig-sig); prevsig=sig; Nguess=newN; deltalog=[deltalog;endvals]; fprintf(repmat('\b',1,numel(msg))) sig-p msg = sprintf('iteration: %g\n = %5.3f\n ... delta = %5.3f',ii,sig-p,endvals); fprintf (msg); end fprintf('\ndone.') length(deltalog); if max(endvals)>threshold, %fieldN=zeros(lt,lp,lf); disp('could not resolve noisefield in time alotted') else end freqs] fieldN=squeeze(Nguess) ;%[angles,vert-angles, end 156 [1 1] Bibliography [1] University of Washington Applied Physics Laboratory. SeaGlider. 2015. URL: http: //www. api. washington. edu/proj ects/seaglider/summary . html. [2] Michael R. Benjamin and Henrik Schmidt. Unmanned Marine Vehicle Autonomy, Sensing, and Communications. University Lecture. 2012. [3] Michael R. Benjamin et al. "An Overview of MOOS-IvP and a Users Guide to the IvP Helm - Release 13.5". In: (2013). [4] Michael R. Benjamin et al. "Nested Autonomy for Unmanned Marine Vehicles with MOOS-IvP". In: Journal of Field Robotics 27.6 (2010), pp. 834-875. [5] A Batelle Company Bluefin Robotics. Bluefin-21. 2015. URL: http: / / www. bluefinrobotics.com/products/bluefin-21/. [6] A Battelle Company Bluefin Robotics. Bluefin Products. 2015. URL: http: // www.bluefinrobotics.com/products/. [7] Roberto M. Borsotti. "Development of a System for Stable Towing of Hydrohpone Arrays for High Frequency Seabed Imaging". Master's Thesis. Massachusetts Institute of Technology, 2009. [8] Philip Budzik. Arctic oil and naturalgas potential. US Energy Information Administration, Office of Integrated Analysis, Forecasting, Oil, and Gas Division, 2009. [9] Autonomous Undersea Vehicle Applications Center. Scripps Spray Glider Configuration. 2015. URL: http: //auvac. org/configurations/view/70. [10] Navy Task Force Climate Change. U.S. Navy Arctic Roadmap 2014 - 2030. US Navy, 2014. [11] 0. I. Diachok and R. S. Winokur. "Spatial variability of underwater ambient noise at the Arctic icewater boundary". In: The Journalof the Acoustical Society of America 55.4 (1974), pp. 750-753. [12] R. Drucker, S. Martin, and R. Moritz. "Observations of ice thickness and frazil ice in the St. Lawrence Island polynya from satellite imagery, upward looking sonar, and salinity/temperature moorings". In: Journalof Geophysical Research (Oceans) 108, 3149 (2003), p. 3149. [13] Teledyne Gavia ehf. AUVs. 2015. URL: http: //www. teledynegavia. com/ product-dashboard/auvs. 157 Inc. Falmouth Scientific. Solar A UV (SA UV 11). 2010. URL: http: / / www . [14] falmouth.com/products/solarauv.html. [15] Open Source Robotics Foundation. ROS. 2015. URL: www. ros . org. [16] Frank S Hover, Mark A Grosenbaugh, and Michael S Triantafyllou. "Calculation of dynamic motions and tensions in towed underwater cables". In: (1994). [17] Woods Hole Oceanographic Institution. Autonomous Underwater Vehicle ABE. 2014. URL: http://www.whoi.edu/main/ABE. [18] Woods Hole Oceanographic Institution. Ships and Technology: REMUS. 2014. URL: http://www.whoi.edu/main/remus. [19] iXBlue. 4th Generation GAPS USBL. 2015. URL: http: //www. ixblue. com/ sites/def ault/f iles/downloads/ixblue-ps-gaps- 10-2014-web. pdf. [20] F.B. Jensen et al. Computational Ocean Acoustics. Modern Acoustics and Signal Processing. Springer, 2011. [21] Jea Soo Kim. "Radiation from Directional Seismic Sources in Laterally Strat- ified Media with Application to Arctic Ice Cracking Noise". PhD thesis. Massachusetts Institute of Technology, 1989. [22] Ocean Acoustic Library. Rays. 2015. URL: http: //oalib. hlsresearch. com/ Rays/index.html. [23] David Moore, Albert Huang, and Edwin Olson. LCM Reference Manual. 2008. URL: http://lcm.googlecode.com/svn-history/r432/www/reference/ lcm/intro .html. [24] NASA. A VHRR-Pathfinder. 2015. URL: https : //podaac . jpl . nasa . gov/ AVHRR-Pathf inder. [25] US Navy. USNS Able (T-A GOS-20). 2015. URL: http: //www. navsource. org/ archives/09/66/6620.htm. [26] Maria Alejandra Parra-Orlandoni. "Target Tracking Onboard an Autonomous Underwater Vehicle: Determining Optimal Towed Array Heading in an Anisotropic Noise Field". Master's Thesis. Massachusetts Institute of Technology, 2007. [27] Bluefin Robotics. Bluefin Spray Glider. 2011. URL: http: //auvac .org/uploads/ platform-pdf/Bluefin-Spray-Glider-Product-Sheet.pdf. [28] Bluefin Robotics. HA UV. 2015. URL: http: / /www. bluef inrobotics . com/ product s/hauv/. [29] Henrik Schmidt. "The Information Content of Ocean Noise: Theory and Experiment". Arctic Muri Project Proposal. 2013. [30] Toby Schneider. Acoustic Communications Lecture. 2014. URL: http: //oceanai. mit . edu/2.680S14/docs/2.680-13-acoustic-communications .pdf. [31] Henrik Scmidt. OASES: Version 2.1. User Guide and Reference Manual. Technical Report. Massachusetts Institute of Technology, 2015. 158 [32] Henrik Scmidt and Nick Saflund. The Soundscape of the Changing Arctic Autonomous Underwater 'SoundScope'. Presentation. 2015. [33] Gobysoft.org Aquatic Software. Projects. 2015. URL: http: //gobysoft . org/ index.wt/projects/autonomy. [34] Barbara J. Sotirin and John A. Hildebrand. "Acoustic navigation of a large- aperture array". In: The Journal of the Acoustical Society of America 87.1 (1990), pp. 154-167. [35] Inc. Teledyne Webb Research Teledyne Technologies. The Slocum Glider. 2010. URL: http: //www.webbresearch. com/slocumglider.aspx. [36] R. A. Wagstaff. "Horizontal directionality estimation considering array tilt and noise field vertical arrival structure". In: The Journal of the Acoustical Society of America 67.4 (1980), pp. 1287-1294. URL: http: //scitation. aip. org/ content/asa/journal/jasa/67/4/10.1121/1.384181. [37] R. A. Wagstaff. "Iterative technique for ambient-noise horizontal-directionality estimation from towed line-array data." In: Journal of the Acoustical Society of America 63 (1978), pp. 863 -869. [38] R.A. Wagstaff and J. Newcomb. "Three-dimensional noise field directionality estimation from single-line towed array data." In: Journal of the Acoustical Society of America 102.2 (1997), pp. 1023 -1031. [39] TC Yang et al. "Acoustic properties of ice edge noise in the Greenland Sea". In: The Journal of the Acoustical Society of America 82 (1987), p. 1034. [40] Y. Yu and R. W. Lindsay. "Comparison of thin ice thickness distributions derived from RADARSAT Geophysical Processor System and advanced very high resolution radiometer data sets". In: Journal of Geophysical Research (Oceans) 108, 3387 (2003), p. 3387. 159