Chapter 1. Introduction

Similar documents
Prediction of Nearshore Waves and Currents: Model Sensitivity, Confidence and Assimilation

Pathways Interns: Annika O Dea, Ian Conery, Andrea Albright

Coastal Wave Energy Dissipation: Observations and Modeling

cbathy Bathymetry Estimation in the Mixed Wave-Current Domain of a Tidal Estuary

WAVE BREAKING AND DISSIPATION IN THE NEARSHORE

Determination Of Nearshore Wave Conditions And Bathymetry From X-Band Radar Systems

Directional Wave Spectra from Video Images Data and SWAN Model. Keywords: Directional wave spectra; SWAN; video images; pixels

CROSS-SHORE SEDIMENT PROCESSES

Rip Currents Onshore Submarine Canyons: NCEX Analysis

IMAGE-BASED FIELD OBSERVATION OF INFRAGRAVITY WAVES ALONG THE SWASH ZONE. Yoshimitsu Tajima 1

Southern California Beach Processes Study

Analysis of Wave Predictions from the Coastal Model Test Bed using cbathy

Appendix E Cat Island Borrow Area Analysis

Undertow - Zonation of Flow in Broken Wave Bores

Determination of Nearshore Wave Conditions and Bathymetry from X-Band Radar Systems

Waves. G. Cowles. General Physical Oceanography MAR 555. School for Marine Sciences and Technology Umass-Dartmouth

Creation of bathymetric maps using satellite imagery

Rip Currents Onshore Submarine Canyons: NCEX Analysis

THE SPATIAL VARIABILITY OF LARGE SCALE SAND BARS

Beach Wizard: Development of an Operational Nowcast, Short-Term Forecast System for Nearshore Hydrodynamics and Bathymetric Evolution

PUV Wave Directional Spectra How PUV Wave Analysis Works

IMPACTS OF COASTAL PROTECTION STRATEGIES ON THE COASTS OF CRETE: NUMERICAL EXPERIMENTS

Undertow - Zonation of Flow in Broken Wave Bores

Nearshore Wave-Topography Interactions

Wave Transformation and Setup from a Cross-shore Array of Acoustic Doppler Profilers (Poster)

IMAGE-BASED STUDY OF BREAKING AND BROKEN WAVE CHARACTERISTICS IN FRONT OF THE SEAWALL

Use of video imagery to test model predictions of surf heights

Unsteady Wave-Driven Circulation Cells Relevant to Rip Currents and Coastal Engineering

Wave Breaking, Infragravity Waves, And Sediment Transport In The Nearshore

14/10/2013' Bathymetric Survey. egm502 seafloor mapping

PROPAGATION OF LONG-PERIOD WAVES INTO AN ESTUARY THROUGH A NARROW INLET

TRANSPORT OF NEARSHORE DREDGE MATERIAL BERMS

MECHANISM AND COUNTERMEASURES OF WAVE OVERTOPPING FOR LONG-PERIOD SWELL IN COMPLEX BATHYMETRY. Hiroaki Kashima 1 and Katsuya Hirayama 1

High-Resolution Measurement-Based Phase-Resolved Prediction of Ocean Wavefields

Figure 4, Photo mosaic taken on February 14 about an hour before sunset near low tide.

Comparison of Predicted and Measured Shoaling at Morro Bay Harbor Entrance, California

SEASONDE DETECTION OF TSUNAMI WAVES

Small Footprint Topo-Bathymetric LiDAR

The History of Coastal Flood Hazard Assessments in the Great Lakes

G. Meadows, H. Purcell and L. Meadows University of Michigan

Salmon: Introduction to ocean waves

Wave-Current Interaction in Coastal Inlets and River Mouths

An experimental study of internal wave generation through evanescent regions

Challenges in determining water surface in airborne LiDAR topobathymetry. Amar Nayegandhi, Dewberry 15 th Annual JALBTCX Workshop, June 11 th 2014

RIP CURRENTS. Award # N

Wind Blow-out Hollow Generated in Fukiage Dune Field, Kagoshima Prefecture, Japan

Surf Survey Summary Report

Waves, Bubbles, Noise, and Underwater Communications

Town of Duck, North Carolina

OECS Regional Engineering Workshop September 29 October 3, 2014

MIAMI BEACH 32ND STREET HOT SPOT: NUMERICAL MODELING AND DESIGN OPTIMIZATION. Adam Shah - Coastal Engineer Harvey Sasso P.E.

Data Collection and Processing: Elwha Estuary Survey, February 2013

Emerging Subsea Networks

USE OF X-BAND RADAR FOR WAVE AND BEACH MORPHOLOGY ANALYSIS

A study of advection of short wind waves by long waves from surface slope images

DUXBURY WAVE MODELING STUDY

Inlet Management Study for Pass-A-Grille and Bunces Pass, Pinellas County, Florida

Utilizing Vessel Based Mobile LiDAR & Bathymetry Survey Techniques for Survey of Four Southern California Breakwaters

Wave behaviour in the inner surf zone

Currents measurements in the coast of Montevideo, Uruguay

Nortek Technical Note No.: TN-021. Chesapeake Bay AWAC Evaluation

An IOOS Operational Wave Observation Plan Supported by NOAA IOOS Program & USACE

Nearshore Placed Mound Physical Model Experiment

RAPID CHANGE IN COASTAL MORPHOLOGY DUE TO SAND-BYPASSING CAPTURED BY UAV-BASED MONITORING SYSTEM. Abstract

Appendix D: SWAN Wave Modelling

CMS Modeling of the North Coast of Puerto Rico

INTRODUCTION TO COASTAL ENGINEERING

Surface Wave Processes on the Continental Shelf and Beach

WOODFIBRE LNG VESSEL WAKE ASSESSMENT

STUDIES OF FINITE AMPLITUDE SHEAR WAVE INSTABILITIES. James T. Kirby. Center for Applied Coastal Research. University of Delaware.

ISOLATION OF NON-HYDROSTATIC REGIONS WITHIN A BASIN

Volume and Shoreline Changes along Pinellas County Beaches during Tropical Storm Debby

Evaluation of June 9, 2014 Federal Emergency Management Agency Flood Insurance Study for Town of Weymouth, Norfolk, Co, MA

THE POLARIMETRIC CHARACTERISTICS OF BOTTOM TOPOGRAPHY RELATED FEATURES ON SAR IMAGES

Wave Energy Atlas in Vietnam

Ongoing Research at the USACE Field Research Facility Jeff Waters, PhD Chief, Coastal Observations & Analysis Branch Coastal & Hydraulics Lab

RCEX: Rip Current Experiment

Comparisons of Physical and Numerical Model Wave Predictions with Prototype Data at Morro Bay Harbor Entrance, California

Acoustic Focusing in Shallow Water and Bubble Radiation Effects

Figure 1, Chart showing the location of the Breach at Old Inlet and sensors deployed in Great South Bay.

Regional Analysis of Extremal Wave Height Variability Oregon Coast, USA. Heidi P. Moritz and Hans R. Moritz

PARAMETRIZATION OF WAVE TRANSFORMATION ABOVE SUBMERGED BAR BASED ON PHYSICAL AND NUMERICAL TESTS

WAVE MECHANICS FOR OCEAN ENGINEERING

NEED FOR SUPPLEMENTAL BATHYMETRIC SURVEY DATA COLLECTION

Coastal Wave Studies FY13 Summary Report

Kelly Legault, Ph.D., P.E. USACE SAJ

Airborne Remote Sensing of Surface and Internal Wave Processes on the Inner Shelf

P2.17 OBSERVATIONS OF STRONG MOUNTAIN WAVES IN THE LEE OF THE MEDICINE BOW MOUNTAINS OF SOUTHEAST WYOMING

THE WAVE CLIMATE IN THE BELGIAN COASTAL ZONE

Wave research at Department of Oceanography, University of Hawai i

A PRACTICAL APPROACH TO MAPPING EXTREME WAVE INUNDATION: CONSEQUENCES OF SEA-LEVEL RISE AND COASTAL EROSION.

COMPARISON OF CONTEMPORANEOUS WAVE MEASUREMENTS WITH A SAAB WAVERADAR REX AND A DATAWELL DIRECTIONAL WAVERIDER BUOY

The Influence of Breaking at the Ocean Surface on Oceanic Radiance and Imaging

EVALUATION OF ENVISAT ASAR WAVE MODE RETRIEVAL ALGORITHMS FOR SEA-STATE FORECASTING AND WAVE CLIMATE ASSESSMENT

Chapter 8 Wave climate and energy dissipation near Santa Cruz Island, California

Bob Battalio, PE Chief Engineer, ESA September 8, 2016

Chapter 4 EM THE COASTAL ENGINEERING MANUAL (Part I) 1 August 2008 (Change 2) Table of Contents. Page. I-4-1. Background...

Gravity wave effects on the calibration uncertainty of hydrometric current meters

Comparison of data and model predictions of current, wave and radar cross-section modulation by seabed sand waves

Airy Wave Theory 1: Wave Length and Celerity

Transcription:

Chapter 1 Introduction 1.1 Motivation for study The nearshore is the coastal region close to land where surface waves rapidly evolve and break in the surf zone. It is also an area where the bathymetry is continuously evolving over time scales varying from hours to years. One of the most significant factors contributing to the vulnerability of coastal areas is the erosion of sediment from beaches. As waves break on the coast they are constantly moving sand, eroding beaches in some areas and depositing sediment in others. With over half of the US population living within 50 miles of the coastline, beach evolution can have a dramatic impact on coastal communities, economies, and industry (Helvarg, 2003). The National Oceanic and Atmospheric Administration (NOAA) estimates that $10 billion of infrastructure is currently at risk from coastal storm damage and this number is expected to greatly increase as people continue to migrate to coastal areas. Continuing research on the complex behavior of coastal erosion is needed to better understand how the nearshore responds to wave forcing in order to better predict the outcomes of interaction between society and this dynamic environment. 1.2 Background One of the dominating factors influencing surface wave transformation is the underlying topography and its spatial and temporal variability. Ocean waves are typically generated by winds blowing across the water surface creating orbital wave motion in the water column. In

deep water, these waves are unaffected by the ocean floor, but as water depths decrease toward shore, the impedance of the land disrupts this associated, circular motion (Figure 1.1). Figure 1.1: Schematic showing the process of wave transformation as waves propagate into the shallow water of the surf zone. This causes the wave to become steepened until reaching a critical point where it then spills over itself creating a breaker (Trujillo and Thurman, 2005). Topographies of varied slope greatly impact the nature of the propagating wave field, particularly close to shore where the slope increases and the water shallows. Therefore, knowing the topography is critical to understanding 2

the way waves and currents behave in the nearshore, and their impact on sediment transport and beach erosion. In the surf zone (often) powerful waves dissipate their energy on the shore, and thus obtaining beach profile measurements can be a difficult task. A range of approaches have been developed from conventional in situ surveys to technological advances allowing for remote optical estimation. Many of the traditional in situ methods using, for example, rod surveys (Figure 1.2) and amphibious vehicles with sonar and GPS technology (Figure 1.3) have logistical drawbacks. Only a relatively small area (of order 1 km² ) is typically surveyed because of the laborious, time-consuming process associated with these practices. Measurements are also only able to be taken when wave heights are relatively low limiting the ability to examine the dynamic evolution occurring during storms and high wave events. Figure 1.2: Example rod survey, a traditional methods for obtaining water depths. 3

Figure 1.3: Coastal Research Amphibious Buggy (CRAB) one of the more traditional methods for obtaining in situ bathymetric data. Recent advancements in remote sensing technology allow for the capability to greatly increase the scale of surveyed regions spanning regional topographic changes along large stretches of coastline. Light Detection and Ranging (LIDAR; Figure 1.4) is one such system being implemented since the early 1990 s (Irish et al., 2000). Figure 1.4: A modern Light Detection And Ranging (Lidar) system used to obtain water depths through the used of aerial laser pulses. 4

By using GPS and high frequency lasers pulsed from an aircraft, the system is able to measure water depths based on the amount of time taken for the beam to be emitted, reflected from the surface, and returned. Although this is a very precise and efficient method, there are still technical complications hindering the overall effectiveness of this technique (within 15 cm vertically; Irish et al., 2000). One major impediment is the presence of bubbles in the surf zone caused by wave breaking. Along with fog and rain, the laser is unable to penetrate turbid, bubbly water to the sea bed. With the high cost associated with LIDAR surveys, it is often inaccessible to many communities with beach monitoring needs. Due to logistical difficulty and high cost associated with established methods, the need to develop low cost, easily accessible techniques was recognized. One approach includes passive imaging systems combined with spatial and temporal properties of surface gravity wave behavior to remotely extract water depths (Dugan et al, 2001 b). The theoretical connection between bathymetry and the phase speed of surface waves was first modeled mathematically using linear wave theory (e.g. Stockdon and Holman, 2000). Based on this idea, water depths in intermediate and shallow water can be calculated by observing the spatial variation (wavenumber) in surface gravity wave energy distribution as a function of wave frequency. This relationship is given by the linear wave theory dispersion relation, σ 2 = g k tanh( k h) (1) 5

where frequency σ is 2π/T with T the wave period, g is the acceleration of gravity, wavenumber k is 2π/L with L the wavelength, and h is the water depth. Equation (1) can then be rewritten to solve for the h in terms of σ and k, tanh h = ( / g k ) 1 σ 2 k (2) Thus, if the surface wave energy distribution, E(σ, k) can be measured remotely, then the water depth can be inferred from (2) by determining the appropriate σ, k values from E(σ, k). This method was first used in remote sensing practices of nearshore research in the 1940 s during World War II with military applications (Williams, 1946; Seiwell, 1947). Still photographs were taken and the change in wavelength of swells was measured between two images separated by a short time. The use of this imaging technique later expanded to digital video systems in the 1990 s (Dugan et al, 2001 a). Video systems have also been used to record images from fixed cameras mounted on land (e.g. Lippmann and Holman, 1989, and many others). Video imaging methods have many scientific benefits because their imagery spans large spatial and temporal scales, they do not require emersion into an environment of damaging nature such as the surf zone, and they can obtain data at times when it would be impossible to make manual measurements, such as during storms (Stockton and Holman, 2000). For these reasons, video-based techniques have become internationally practiced for coastal studies in countries worldwide (Turner et al., 2006). Video cameras detect the luminance reflected off the sloping sea surface. E(σ, k) spectra needed for the dispersion relation can then be approximated with video through the use of 3D Fast Fourier Transforms (3D FFT). This technique was first explored in the work of Dugan et al, 6

(2001 a) with video imaging systems mounted on small airplanes (Figure 1.5). Their system, called the Airborne Remote Optical Spotlight System (AROSS), captured 8 minutes of video images of an area 500 x 500 m recorded at 2 Hz at an altitude of 2.8 km. Figure 1.5: Airborne Remote Optical Spotlight System (AROSS) used in the aerial imaging technique of Dugan et al, 2001. This plane, with a camera mounted on the lower nose in the lower right corner of the image, was used to record coastal video images to later be analyzed for estimated water depths. The plane circled the area and then subsequently moved landward with each successive circle being analyzed for individual water depth estimates. These images were then transformed onto the World Geodetic System 1984 (WGS84) grid and precisely known ground control points were used in the image to ensure accurate position and to correct distortion from the highly oblique camera angle. Spectral analysis using 3D FFT methods were then applied to selected 1 minute time series from the collected imageset allowing a water depth to be estimated. Results from this work were compared to in situ survey measurements. Water depths in 4-12 m from the bathymetric inversion algorithm had root-mean-square (RMS) errors typically between 0.5 1.0 m. Estimates in shallow water (less than 4 m) were not examined because the large spatial area 7

of the camera footprint was too large and spanned the rapidly varying bathymetry all the way to the shoreline. The goal of this research is to apply similar video imaging methods developed by Dugan et al, (2001b) to a land-based system and estimate the bathymetry in shallow water to as near the shoreline as possible. After modifying the airborne methods it will be shown that water depths can be estimated by the land-based system with RMS errors on the order of 0.5 m into depths as shallow as 1 m. In the following we describe our field methods for collecting video data and the inversion techniques for estimating the depth. We then present results from 12 surveyed comparisons with independently sampled bathymetry from established in situ methods. Results are then described in terms of wave conditions, wave nonlinearities, image resolution, and obstructions by a pier visible in the images. 8

Chapter 2 Methodology 2.1 Research Location Video data for this study are obtained from a camera positioned on a 20 m tower (Figure 2.1) mounted on the dune on the Outer Banks, North Carolina near the town of Duck (Figure 2.2). The camera faces northeast and overlooks the surf zone near the US Army Corp of Engineers (USACE) Field Research Facility (FRF) pier. Figure 2.1: Image of the tower with video camera at the top used to record coastal video images. The beach at Duck is an intermediate sloping beach (generally ranging from 0.05-0.015) and the wave field is primarily dominated by incident waves (Birkemeier et al., 1985, Lippmann et al., 1993). Bathymetric surveys over a 1.5 km section of coastline are regularly performed by the FRF staff using amphibious vehicles on an approximate monthly basis. These surveys will be 9

used as ground truth for our estimated bathymetry obtained from the video data. Ancillary data includes mean water level data measured by a NOAA tide gauge, and directional wave information from permanent bottom-mounted pressure sensors (Long et al., 1991). (a) (b) Figure 2.2: The research site location for this study in Duck, North Carolina, USA (Panel a) at the USACE Field Research Facility (Panel b). Video images are recorded three times a day typically for 30 minute durations at four hour intervals. These images are part of a large dataset for this location that began in October 1997. The video system was upgraded in November 2006 from analog cameras with 640-by-480 pixels to a digital camera with 1280-by-960 pixels. The 4-fold increase in pixels is expected to improve the results; therefore, the dataset used in this study only considers the images collected using the new system from November 2006 to October 2007. A total of 12 sample days were analyzed corresponding to days when the FRF surveyed the beach. Recorded images encompass an area of 1 km alongshore and 0.5 km cross-shore (Figure 2.3) that overlapped with 10

FRF surveys. Image-to-ground geometrical transformation parameters are determined for the camera orientations using methods described in Holland, et al (1997). Figure 2.3: Plan view of the area encompassed by the camera. The axes are in the FRF coordinate system. Each image is orthonormalized into an equally-spaced 1 m grid based on weighting the pixel intensities by neighboring values (Figure 2.4). Each pixel in the orthonormalized image corresponds to a ground location. For the analysis, a 100 m by 100 m region is sampled from an approximately 12 minute time series and used in the inversion algorithm to estimate a water depth at the center of the array (described next). The spatial and temporal variability of the wave field is quantified with the video, where the nature of the signal outside the surf zone is based on the changing reflectance characteristics of light on the front and back sides of the waves. In random waves, these spectral characteristics are used to represent the wavenumber-frequency variation in directional wave spectra. Within the surf zone where waves are breaking, the signal is based on the contrast between actively breaking waves and bores and the unbroken surrounding water. 11

(a) Figure 2.4: Panel (a) displays a sample snapshot image recorded from the video system. This image is time-averaged and placed in an ortho-normalized view shown in Panel (b). (b) 12

2.2 Inversion Methods Video systems record incoming waves and their breaking patterns through spatial and temporal variations in pixel intensities. Based on the linear wave theory dispersion relation (1), identified energetic waves at particular wavenumber and frequency can be used to solve for the water depth (2). A 3D FFT can be applied to the video data to estimate these wave variables. The 3D FFT transforms the square array of pixel time series into the frequency and wavenumber domain. In this case the 2 spatial dimensions of the square image array leads to wavenumber estimation in both the cross-shore direction, k x, and the alongshore direction, k y, at each frequency. The magnitude and direction, α, of the wavenumber vector (i.e., the direction of the propagation of that wave component) is given by 2 2 ( k x + k y 1 2 =k) (3) k 1 y α = tan (4) k x For this reason the frequency-wavenumber energy spectra, E(σ, k x, k y ), is sometimes referred to as the directional wave spectrum, E(σ, α), indication from which direction the wave energy is approaching. Integrating over all directions (or equivalently over all k x, k y ) gives the energy spectra of the wave field, E(σ) E σ ) = E( σ, k, k ) dk dk (5) ( In Dugan, et al. (2001b), E σ, k x, k ) is approximated with sea surface reflectance obtained ( y from video records, I σ, k x, k ). The assumption is that the spatial scales of energy distribution ( y as a function of frequency is retained, and that I σ, k x, k ) is proportional to E σ, k x, k ). x y x ( y y ( y 13

Similarly integrating over all wavenumbers provides an estimate of the image intensity distribution as a function of frequency, I (σ ). An example of observed I (σ ) from our data set is shown in Figure 2.5. Figure 2.5: Plot of frequencies associated with the wave field at a selected location on the video image. In the methods of Dugan, et al (2001b), the peak frequency, σ p, associated with the maximum in E(σ) is identified from the 3D FFT of a 2 minute time series of orthonormalized imagery. The k x and k y associated with the maximum energy in I( σ p,k x, k y ) determines the corresponding wavenumber to be used with σ p in (2) to determine h. The resolution in wavenumber spectra, dk, is determined by the size of the interrogation window; in Dugan, et al (2001b) the interrogation window was 500 m by 500 m, corresponding to dk = 0.002 1 m for both the cross-shore and alongshore wavenumber resolution. 14

In our methods, we follow a similar procedure with a few notable exceptions. The first is that the interrogation window is much smaller, 100 m by 100 m. This allows us to probe shallower water where the bathymetry is more spatially variable. Comparisons done later will determine to what depths the technique can be applied. The cost of a small area is decreased wavenumber resolution, dk = 0.01m 1. The wavenumber resolution is not fine enough to accurately determine k. However, assuming that the wave field is a Gaussian process, we can employ a Gaussians peak fitting algorithm to I( σ,k, k estimates. An example I( σ,k, k distribution. p x y p x y ) to give sub-wavenumber bandwidth ) is shown in Figure 2.6 with identified centroid of the energy Figure 2.6: Two wave number spectra selected from frequencies in Figure 2.5. Wavenumbers k, associated with the maximum energy determine k. x k y The second difference in our methods is that we can utilize longer time series of video images and further repeat the analysis over successive 4 minute intervals (ensembles). In this work we estimate wavenumbers for three consecutive 4 minute ensembles and average the results to give a single k for the given frequency. The third difference is that we will estimate an average k for each frequency in I(σ), not just the peak value. We can thus estimate wavenumbers as a function of frequency spanning the 15

range of sea and swell frequencies typical of open coast beaches (0.05 Hz < f < 0.33 Hz). We can then fit a curve based on the dispersion relation to the range of σ, k pairs estimated. A nonlinear least squares fit to the data determines the best estimate of the water depth. An example is shown in Figure 2.7. Figure 2.7: Plot of wave numbers used with frequency to estimate a water depth based on the best-fit curve of the linear dispersion relation to the data. After a water depth has been estimated, the window is subsequently moved 10 m landward and the 3D FFT is repeated to solve for another water depth at the new location. After repeatedly moving the region from well-offshore to the shoreline, an estimated cross-shore profile is produced (Figure 2.8) similar to methods in Dugan et al., (2001b). 16

Figure 2.8: This figure shows an example region of 100 x 100 m used to create a cross-shore profile at y = 410m by subsequently moving the region toward the shore. The analysis can be further repeated with additional transects alongshore offset by 20 m up the coast to yield a 2-D map of the bathymetry over a nearshore region spanning approximately 0.5 km across-shore and 1 km along the coast. When making these water depth estimates, there are several ways to solve the dispersion relation. One way is to assume shallow water waves, σ = k gh, which will require use of a straight line to fit through the frequencies. This assumes linear orbits of the waves, however, due to the dispersive nature of waves in intermediate depths we would not expect this method to perform well outside the surf zone. Regardless, analysis using this technique was explored to determine if favorable results would be produced (with mixed results). A better alternative is to simply solve the dispersion relation (equation 2) requiring a nonlinear least square approach to estimate h. The curved form of the solution is fit through the associated frequencies rather than 17

a straight line. This method better accounts for some of the frequency-dependent properties of waves in intermediate water depths. Another method explained in the analysis was to take the average of both shallow and general forms for evaluating the dispersions. All three approaches were implemented and evaluated in this work, results indicated that using the general form for the dispersion relation was generally more reliable and was thus used henceforth. Several assumptions in the analysis were necessary, including ignoring possible wavecurrent interactions and wave nonlinearity, and that the temporal variations of the incident wave field were accurately reflected in the data. The implications of these assumptions will be further discussed in chapter 4.. 18

Chapter 3 Results & Analysis In this chapter we compare results between the land-based video inversion technique with tradition in situ survey data collected by the FRF staff to evaluate the accuracy of the results. In total 12 sample days were used for this study. Representative results from 7 December 2006 are presented in detail, and then a summary of errors as a function of spatial position for all runs is presented. An oblique snapshot and an orthonormalized image of the study region are shown in Figure 3.1, depicting a clear day characterized by relatively low amplitude (Hs = 0.24 m), long wave period waves (T = 10.5 s) and a narrow surf zone very near the shoreline. Figure 3.1: This image is a snapshot from the sample date 7 Dec, 2006. 19

3.1 Survey Comparison Figure 3.2 shows the FRF bathymetry measured with traditional methods on 7 Dec 2006. The region of interest ranges from the shoreline to about 500 m from shore at a depth of about 6 m. There is a small sandbar at approximately 300 m cross-shore in the FRF coordinate system. Alongshore variability is approximately uniform except near the FRF pier located at 500 m alongshore. Figure 3.2: 2D map created by interpolating between transect lines measured using in situ FRF survey data. These data are used to compare with estimated results to determine accuracy of the inversion technique. 20

Waves interacting with the pier pilings cause the formation of a large scour hole around the structure. FRF transect lines were measured approximately every 50 m and, with interpolated results between measured lines, provide full coverage of the large scale bathymetry in the region. Estimated results from the video imaging technique are shown in Figure 3.3 for 7 Dec 2006. Similar to the survey transects in Figure 3.2, only cross-shore lines were calculated in the image and interpolated to the FRF measurement locations to allow direct comparisons. D C B A Figure 3.3: Estimated water depths obtained using the inversion technique. The black lines labeled A-D are associated with the individual profile lines shown in Figure 3.4. Four of these transects have been selected from the image and displayed in Figure 3.4 to demonstrate the performance of the technique at various distances throughout the image. 21

Underwater formations such as the sandbar location and the hole under the pier are also capable of being detected and can be seen in finer detail in the individual profile lines. A B C D C D Figure 3.4: These Panels shows individual selected transect lines chosen to display how the inversion technique performs at specific locations. Each panel corresponds to lines in Figure 3.3. The black line represents the in situ data surveyed by the FRF staff and the estimated bathymetries are represented by the red dots. Mean sea level is also shown by the straight, dashed line. 22

Differences (in meters) between the in situ and video estimated profiles are shown in Figure 3.5. Differences were created by subtracting the estimated water depths from the FRF in situ measured values. Figure 3.5: Difference map between the estimated results and the in situ survey data. Positive values correspond to underestimates and negative numbers represent overestimated water depths using the inversion technique. Positive values on this plot indicate that the inversion technique underestimated the actual water depth in meters while negative numbers are associated with overestimated depths. Bathymetry errors for this day are generally much less than 0.5 m over most of the region, including at large distances (about 1 km) from the camera. Larger errors are observed closer to the shoreline, and 23

in regions influenced by the pier structure itself (e.g., the pier pilings and pier deck and infrastructure), and are discussed in more detail later. 3.2 Statistical Analysis RMS errors were calculated to provide a quantitative assessment of results as a function of cross-shore and alongshore location. That is, along specific cross-shore and alongshore transects, RMS errors are calculated and plotted as a function of alongshore and cross-shore position, respectfully (Figure 3.6). RMS errors associated with the cross-shore position are generally less than 0.5 m within 300 m of the shoreline, beyond which errors increase to about 0.75-1.0 m as a result of diminishing image resolution, edge effects in the image, and influences from the pier. Figure 3.6: RMS errors as a function of cross-shore and alongshore positions for the sampled day 7 Dec 2006. 24

Efforts to reduce errors caused by the pier were attempted by excluding estimates known to be contaminated by the pier (within the blue lines greater than 300 m in Figure 3.3). However, shadows and reflections from the pier often extend beyond this area and impact the results. RMS errors as a function of alongshore coordinate are typically less than 0.5 m except in the region between 400-550 m which is the region strongly influenced by the pier. Between 650-700 m the deck of the pier blocks the wave field completely and is eliminated from the analysis. RMS errors from all 12 sample days spanning the 1 year period have been contoured as a function of cross-shore and alongshore positions (Figures 3.7 and 3.8). RMS errors are generally less than 0.5-0.75 m within 300 m of the shore and often extended to distances as much as 500 m from shore. Alongshore errors are also generally less than 0.5-0.75 m except where the pier influences the imagery. Periods of poor bathymetric inversion are a result of poor image quality (such as rain or fog). Figure 3.7: Contoured RMS errors as a function of the cross-shore position for each of the 12 samples days spanning the 1 year period from November 2006 October 2007. 25

Figure 3.8: Contoured RMS errors as a function of the alongshore position for each of the 12 samples days spanning the 1 year period from November 2006 October 2007. 26

Chapter 4 Discussion 4.1 Influence of FRF Pier The region from approximately 500 700 m in the orthonormalized image reflects the area where the FRF pier partially obscures the wave field causing the introduction of significant inaccuracies in the analysis. Without the presence of the pier it can be assumed that continuous water depths could be estimated with similar accuracies before and beyond the pier location, but it plays an important role in this study by demonstrating how the technique can be used when structures are present. To reduce pier biases in the RMS errors, in all calculations the region encapsulated by the thin blue lines greater than the 300 m cross-shore coordinate (such as in Figure 3.3) were not included in the analysis. The region of the image with the pier region shoreward of the 300 m coordinate often gave good results because the propagating wave field was detected through the pier pilings. However, with increasing distance from the shore, the camera viewing angle changes and the pilings become a barrier in the images blocking views of the wave field under and beyond the pier. Some of the sampled days appeared to be more impacted by the pier due to factors such as sky luminance and the position of the sun. Bright sunshine and calm waves increases the reflectance of the pier off the water and creates more visible shadows underneath and behind the pier. Figure 4.1 shows a worse-case scenario where errors arising from pier influences inhibit the ability to estimate for water depths. 27

Figure 4.1: Worse case example where the influence of the pier inhibits the ability to produce and estimated water depth. Figure 4.2 shows a more optimal situation where water depths are able to be accurately estimated through the pilings suggesting that results can be obtained under favorable conditions. Figure 4.2: Example favorable scenario where we are able to accurately estimate water depths through the pier. 4.2 Wave Conditions While video data was recorded daily, the bathymetric inversion technique was only applied to specific days when the in situ surveys were performed by the FRF staff. These were typically completed at monthly intervals, but since these surveys require submersion into the surf 28

zone, measurements were only made on days when wave conditions were generally mild ( H < 1.0 m). As a consequence, results presented however are biased to small wave days. It is of interest to examine the technique to determine limitations over the range of conditions anticipated in the field. This requires further field studies where ground-truth survey data was collected under high waves, most likely with jet-skis or other highly invulnerable watercraft operating in the surf zone. 4.3 Currents The presence of surface currents in the nearshore was not taken into consideration in this study. Mean currents can interact with the wave field causing the waves to refract and the wave celerity to decrease in speed (depending on the direction of the current) changing the estimated water depth. It is possible to include mean current interactions in the dispersion relation (1) as in Dugan, et al (2002). To extract the flow requires very high resolution of the higher frequency wave motions and with higher than 8-bit (255 shades of gray) pixel depth (for example, 12-bit cameras used by Dugan, et al, 2002). This effort is being pursued in ongoing studies. Nevertheless, results for the low wave days examined suggest that this effect is weak, and likely due to the absence of strong wave driven flows under these conditions. 4.4 Wave nonlinearities Linear wave theory is used to estimate depths in all regions despite the fact that waves associated in shallow water, closer to the shoreline evolve nonlinearly and have celerities that include amplitude dispersion (i.e., the wave phase speed depends also on the height of the wave), and in fact, we see increases in errors very near the shoreline. Theoretically, the linear dispersion relation should break down once waves begin to develop nonlinearity; however, in this study it is ignored and linear wave theory is applied all the way to the shoreline. Again, rms 29

under the relatively small wave days examined, it appears that the linear dispersion relation works well over much of the nearshore with RMS errors typically within 0.5m. 4.5 Spatial and Temporal Variations It is assumed that spatial and temporal variations of the incident wave field are being accurately reflected in the data. By using video systems, we are not concerned with the absolute energy associated with the waves. Instead, we are examining the spatial and temporal scales of the wave field and assume that video pixel intensities reflect the wavenumber-frequency distribution of the wave field accurately (as in Dugan, et al 2001a). Utilizing the k, σ pairs throughout the spectrum of sea and swell waves appears to minimize differences that may arise and results suggest that the video representation works well under most conditions. 30

Chapter 5 Conclusions The technique described herein relys on shore-based video imaging and linear wave theory and demonstrates a new method for estimating bathymetric changes in the nearshore on a daily basis from very near the shoreline to about 5m water depths. Such techniques have already been developed and tested for airborne systems (Dugan, et al., 2001a, 2001b, 2002), and the methods used in this study follow these but have been modified for fixed land-based deployments. Individual water depths are determined by wavenumber-frequency analysis obtained using a 3D FFT of orthonormalized pixel intensities associated with a selected 100 x 100 m square region of 12 minute duration with video images sampled at 3.75 Hz. This process is then repeated by subsequently moving the region shoreward at 10 m increments until a profile of water depths has been estimated. A total of 12 sample days between November 2006 October 2007 were used in this study. The accuracy of the technique was determined by comparing depth estimates with in situ surveys collected on corresponding days by the USACE FRF staff. RMS errors were calculated based on the differences between the inversion technique and the in situ surveys in both the alongshore and cross-shore directions. Errors are generally less than 0.5-0.75 m within 300 m of the shore and often out to as much as 500 m from shore. Alongshore errors are also generally less than 0.5-0.75 m except where the pier influences the imagery. 31

Despite the assumptions of a linear wave field and no nearshore currents the techniques appear to work well from 6 m water depths to regions very near the shoreline. While there may be discrepancies associated with the pier (e.g. shadows, reflections, blockages) that introduce errors into the analysis, the overall results demonstrate the successful capability of video imaging techniques when combined with linear wave theory to accurately estimate nearshore water depths. However, the results are obtained only when the in situ surveys were conducted under generally mild wave conditions ( H rms < 1.0m). Further investigations are required to determine under what range of environmental scenarios the technique can be applied. 32

References Birkemeier, W.A., Miller, H.C., Wilhelm, S.D., DeWall, A.E., and Gorbics, C.S., 1985, User s guide to CERC s Field Research Facility. Instructions Report-85-1, Coastal Engineering Research Center, US Army Corps of Engineers Waterwasy Experiment Station, Vicksburg, Mississippi. Dugan, J.P., C.C. Piotrowski, and J.Z. Williams Water Depth and Surface Current Retrievals from Airborne Optical Measurements of Surface Gravity Wave Dispersion, Journal of Geophysical Research, 106, 906-915, 2001 a. Dugan, J.P., G.J. Fetzer, J. Bowden, G.J. Farruggia, J.Z. Williams, C.C. Piotrowski, K Vierra, D. Campion, and D.N. Sitter, Airborne Optical System for Remote Sensing of Ocean Waves, Journal of Atmospheric and Oceanic Technology, 18, 1267-1275, 2001 b. Helvarg, David, Coasts at Risk: Coastal Sprawl at the Shore, The Multinational Monitor, 24, 2003. Holland, K. Todd, Holman, Robert A., Lippmann, Thomas C., Practical Use of Video Imagery in Nearshore Oceanographic Field Studies, IEEE Journal of Oceanic Engineering, 22, 81-92, 1997. Holman, R.A., A.H. Sallenger, T.C. Lippmann, J.H. Haines, The Application of Video Image Processing to the Study of Nearshore Processes, Oceanography, 6, 78-85, 1993. Irish, J.L., McClung, J.K., and Lillycrop, W.J., Airbourne LIDAR Bathymetry: The SHOALS System, 2000. Lippmann, T.C., Holman, R.A., and Hathaway, K.K., Episodic, Nonstationary Behavior of a Double Bar System at Duck, North Carolina, USA, 1986-1991, Journal of Coastal Research, 15, 1993. Long, C.E., and Oltman-Shay, J.M., 1991, Directional Characteristics of Waves in Shallow Water, Tech. Report CERC-91-1, US Army Corp of Engineers Waterways Experiment Station, Vicksburg, Mississippi. Seiwell, H.R., Military Oceanography in the World War II, Military Engineering, 39(259), 202-210, 1947. Stockdon, H., and R.A. Holman, Estimation of Wave Phase Speed and Nearshore Bathymetry from Video Imagery, Journal of Geophysical Research, 105, 22015-22033, 2000. Turner, Ian L., Aarnikhof, Stefan G.L., Holman, Rob A., Coastal Imaging Applications and Research in Australia, Journal of Coastal Research, 22, 37-48, 2006. 33

Williams, W.W., The Determination of Gradients on Enemy-held Beaches, Geog. Journal, XIC, 76-93, 1946. 34