Application of the Goda Pressure Formulae for Horizontal Wave Loads on Elevated Structures

Similar documents
WAVE LOAD ACTING ON HORIZONTAL PLATE DUE TO BORE

LABORATORY EXPERIMENTS ON WAVE OVERTOPPING OVER SMOOTH AND STEPPED GENTLE SLOPE SEAWALLS

LABORATORY EXPERIMENTS FOR WAVE RUN-UP ON THE TETRAPOD ARMOURED RUBBLE MOUND STRUCTURE WITH A STEEP FRONT SLOPE

EXPERIMENTAL RESEARCH ON COEFFICIENT OF WAVE TRANSMISSION THROUGH IMMERSED VERTICAL BARRIER OF OPEN-TYPE BREAKWATER

Yasuyuki Hirose 1. Abstract

ISOLATION OF NON-HYDROSTATIC REGIONS WITHIN A BASIN

WAVE PRESSURE DISTRIBUTION ON PERMEABLE VERTICAL WALLS

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

A New Generator for Tsunami Wave Generation

Aalborg Universitet. Published in: Proceedings of Offshore Wind 2007 Conference & Exhibition. Publication date: 2007

INTRODUCTION TO COASTAL ENGINEERING AND MANAGEMENT

WAVE INTERACTION WITH DECK OF JETTY ON A SLOPE

Wave forces on bridge decks and damping techniques to reduce damages

INTERNATIONAL JOURNAL OF CIVIL AND STRUCTURAL ENGINEERING Volume 1, No 4, 2010

MODELING OF CLIMATE CHANGE IMPACTS ON COASTAL STRUCTURES - CONTRIBUTION TO THEIR RE-DESIGN

EXPERIMENTAL STUDY ON THE HYDRODYNAMIC BEHAVIORS OF TWO CONCENTRIC CYLINDERS

Survival Analysis of Elevated Homes on the Bolivar Peninsula after Hurricane Ike Tori Tomiczek 1 ; Andrew Kennedy 1, Spencer Rogers 2

2. Water levels and wave conditions. 2.1 Introduction

Technical Brief - Wave Uprush Analysis Island Harbour Club, Gananoque, Ontario

Experimental study of tsunami wave load acting on storage tank in coastal area

PHYSICAL MODELING FOR MEASURING THE EFFECTIVENESS OF SINGLE CURTAIN PILE FOUNDATION BREAKWATER IN INTERMEDIATE WATER DEPTH

Experimental Investigation of Clear-Water Local Scour at Pile Groups

CRITERIA OF BOW-DIVING PHENOMENA FOR PLANING CRAFT

Broken Wave Loads on a Vertical Wall: Large Scale Experimental Investigations

COMPARISON OF ROCK SEAWALL AND DUNE FOR STORM DAMAGE REDUCTION

DETRMINATION OF A PLUNGER TYPE WAVE MAKER CHARACTERISTICE IN A TOWING TANK

Wave Breaking and Wave Setup of Artificial Reef with Inclined Crown Keisuke Murakami 1 and Daisuke Maki 2

ITTC Recommended Procedures and Guidelines

Ocean waves and shock waves

Using sea bed roughness as a wave energy dissipater

CHAPTER 135. Influence of the core configuration on the stability of berm breakwaters. Nikolay Lissev 1 AlfT0rum 2

PHYSICAL AND NUMERICAL MODELLING OF WAVE FIELD IN FRONT OF THE CONTAINER TERMINAL PEAR - PORT OF RIJEKA (ADRIATIC SEA)

WAVE RUNUP ON COMPOSITE-SLOPE AND CONCAVE BEACHES ABSTRACT

Analysis of Shear Lag in Steel Angle Connectors

Effect of channel slope on flow characteristics of undular hydraulic jumps

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

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

BRRAKING WAVE FORCES ON WALLS

SUSTAINABLE DEVELOPMENT: CIVIL ENGINEERING AND THE BUILT ENVIRONMENT. INTERNATIONAL CONFERENCE Malta 5 th May 2011

Aerodynamic Analysis of a Symmetric Aerofoil

PHYSICAL EXPERIMENTS ON THE HYDRODYNAMIC RESPONSE OF SUBMERGED FLOATING TUNNEL AGAINST THE WAVE ACTION

Wave Overtopping of Seawalls. Design and Assessment Manual. HR Wallingford Ltd. February R&D Technical Report W178. R&D Technical Report W178

BEACH EROSION COUNTERMEASURE USING NEW ARTIFICIAL REEF BLOCKS

TRANSPORT OF NEARSHORE DREDGE MATERIAL BERMS

COASTAL HAZARDS. What are Coastal Hazards?

Wave Load Pattern Definition

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

International Journal of Innovative Research in Advanced Engineering (IJIRAE) ISSN:

STUDY ON TSUNAMI PROPAGATION INTO RIVERS

Australian Journal of Basic and Applied Sciences

WAVE MECHANICS FOR OCEAN ENGINEERING

SPECTRAL CHARACTERISTICS OF FLUCTUATING WIND LOADS ON A SEPARATE TWIN-BOX DECK WITH CENTRAL SLOT

SUBMERGED VENTURI FLUME. Tom Gill 1 Robert Einhellig 2 ABSTRACT

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

WAVE OVERTOPPING OF RUBBLE MOUND BREAKWATERS

REVETMENTS. Purposes and Operational Constraints. Purposes Erosion control o o. Revetment Design 4/5/2016. CE A676 Coastal Engineering

OECS Regional Engineering Workshop September 29 October 3, 2014

Evaluation of Tsunami Fluid Force Acting on a Bridge Deck Subjected to Breaker Bores

ESTIMATION OF TSUNAMI FORCE ACTING ON THE BLOCK ARMORED BREAKWATER DUE TO SOLITON FISSION

WAVE REFLECTION AND WAVE RUN-UP AT RUBBLE MOUND BREAKWATERS

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

An experimental study of internal wave generation through evanescent regions

[Barve, 4(7): July, 2015] ISSN: (I2OR), Publication Impact Factor: 3.785

The Failure of the Kamaishi Tsunami Protection Breakwater

SPH applied to coastal engineering problems

DISTRIBUTION OF WAVE LOADS FOR DESIGN OF CROWN WALLS IN DEEP AND SHALLOW WATER. Jørgen Quvang Harck Nørgaard 1, Thomas Lykke Andersen 1

EFFECT OF MEASUREMENT SYSTEMS ON IMPACT LOADS ON RIGID STRUCTURES

Numerical modeling of refraction and diffraction

EVALUATION OF TSUNAMI FLUID FORCE ACTING ON THE BRIDGE DECK

Wave Force on Coastal Dike due to Tsunami

Stability of Cubipod Armoured Roundheads in Short Crested Waves Burcharth, Hans Falk; Andersen, Thomas Lykke; Medina, Josep R.

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

Influence of rounding corners on unsteady flow and heat transfer around a square cylinder

ANSWERS TO QUESTIONS IN THE NOTES AUTUMN 2018

LONG WAVE EFFECTS ON BREAKING WAVES OVER FRINGING REEFS

Numerical Simulation of Wave Loads on Static Offshore Structures

Stress and deformation of offshore piles under structural and wave loading

CFD Simulation and Experimental Validation of a Diaphragm Pressure Wave Generator

DESIGN OPTIMIZATION FOR A PASSIVE MESH SCREEN WAVE ABSORBER FOR THE CCOB

EXAMPLES (OPEN-CHANNEL FLOW) AUTUMN 2018

Dynamic Response of Jacket Structures to Breaking and Nonbreaking Waves: Yesterday, Today, Tomorrow

Proceedings of the ASME st International Conference on Ocean, Offshore and Arctic Engineering OMAE2012 July 1-6, 2012, Rio de Janeiro, Brazil

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

LOCALLY CONCENTRATED SEVERE BEACH EROSION ON SEISHO COAST CAUSED BY TYPHOON T0709

CHAPTER 113 Impact Loading and Dynamic Response of Caisson Breakwaters

Technical Brief - Wave Uprush Analysis 129 South Street, Gananoque

Offshore platforms survivability to underwater explosions: part I

SUPERGEN Wind Wind Energy Technology Rogue Waves and their effects on Offshore Wind Foundations

Undertow - Zonation of Flow in Broken Wave Bores

The Adequacy of Pushover Analysis to Evaluate Vulnerability of Masonry Infilled Steel Frames Subjected to Bi-Directional Earthquake Loading

Tutorial. BOSfluids. Relief valve

Wave Assessment on Loch Ericht

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

Wave Breaking. Wave Breaking

CHAPTER ONE HUNDRED NINETY FIVE

Wave Dragon A slack moored wave energy converter

Three-Dimensional Physical Modeling on Perforated Skirt Breakwater

WAVE IMPACTS DUE TO STEEP FRONTED WAVES

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

CHAPTER 132. Roundhead Stability of Berm Breakwaters

Transcription:

Application of the Goda Pressure Formulae for Horizontal Wave Loads on Elevated Structures Wiebe, D. M., Park, H., & Cox, D. T. (2014). Application of the Goda pressure formulae for horizontal wave loads on elevated structures. KSCE Journal of Civil Engineering, 18(6), 1573-1579. doi:10.1007/s12205-014-0175-1 10.1007/s12205-014-0175-1 Springer Accepted Manuscript http://cdss.library.oregonstate.edu/sa-termsofuse

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 Application of the Goda Pressure Formulae for Horizontal Wave Loads on Elevated Structures Dane M. Wiebe *, Hyoungsu Park**, and Daniel T. Cox*** * Graduate Research Assistant, School of Civil and Construction Engineering, Oregon State University, Corvallis, Oregon, 97331, USA (E-mail: Wiebe@oregonstate.edu) ** Graduate Research Assistant, School of Civil and Construction Engineering, Oregon State University, Corvallis, Oregon, 97331, USA (Corresponding Author, E-mail: Hyoungsu.park@gmail.com) *** Professor, School of Civil and Construction Engineering, Oregon State University, Corvallis, Oregon, 97331, USA (E-mail: dan.cox@oregonstate.edu) Abstract Small-scale physical experiments were conducted to investigate the application of the Goda wave pressure formulae modified to predict the horizontal wave loads on elevated structures considering non-breaking, broken, and impulsive breaking waves. The air gap defined as the vertical distance from the still water level to the base of the structure played a key role in the reduction of wave impact forces. Physical model results using random waves confirmed that the modified application of the Goda wave pressure formulae provided a good estimate of the horizontal forces on elevated structures for both broken and impulsive breaking waves. As the air gap was increased, the resulting forces decreased, and the estimated values became increasingly conservative. When the ratio of the air gap to water depth, a h, increased from -1.0 to 1.5, the reduction in force was approximately 75% when the wave height to breaking water depth ratio, H/h b, was equal to unity. Keywords: wave force; wave pressure; elevated structure; air gap; random waves; Goda; jetty 1. Introduction Hurricanes can devastate coastal communities along the U.S. Gulf Coast. In 1900, a category 4 hurricane destroyed Galveston, Texas, killing 6000 people, and remains the single most deadly natural disaster in U.S. history. In 2005, Hurricane Katrina, a category 3 storms struck New Orleans, 1

27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 Louisiana, killing 1800 people, and caused an estimated $81 billion in damages, the most expensive natural disaster in U.S. history. On average, there are 6.2 hurricanes per year in the Atlantic Ocean, 1.7 of which make landfall along the U.S. coast (NOAA, 2012). These events often cause widespread damage as many structures near the coast are subjected to unexpected hydrodynamic loads from storm surge and waves which accompany hurricanes, and has been documented as the cause of wide spread failure of coastal highway bridges (Cuomo, et al., 2009; Robertson, et al., 2007). Even structures sited well inland, such as residential dwellings, are susceptible if not elevated well above the maximum storm surge elevation. Building elevation has been found to be the critical parameter in determining a structures survival. For large wave climates, structures elevated above the storm surge are generally capable of survival and suffer relatively little damage, whereas structures located below it are generally completely destroyed. The relation between damage and elevation is so sensitive, that in some areas the difference between survival and destruction is only 0.5 m in elevation (Kennedy, et al., 2011). Early research on elevated structures was performed by Bea et al. (1999), who studied the performance of platforms in the Gulf of Mexico which were subjected to several hurricane events. According to the design guidelines provided by the American Petroleum Institute, the decks of many of these platforms were lower than the storm wave crest heights, and should have been destroyed; however, some of these platforms survived while others failed. Based on the performance of these platforms, modifications to the design guidelines were suggested. The total force was a summation of buoyancy, horizontal slamming, horizontal hydrodynamic drag, vertical hydrodynamic uplift, and acceleration dependent inertia forces. A more complicated mathematical model based on momentum was developed by Kaplan (1992), and Kaplan et al. (1995) to predict the time history of impact loadings on offshore platforms, and the wave impact force from large incident waves. The technique was similar to that used for modeling the ship slamming phenomena, based on Morrison s equation, and accounted for hydrodynamic inertial forces, buoyancy forces, and drag forces. The theoretical horizontal force was combined total of the inertial momentum and drag. 2

54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 More recently, Cuomo et al. (2007) conducted 1:25 scale model test of wave forces on exposed jetties. These tests focused on the physical loading processes, for both quasi-static and impulsive loading condition. Impulsive forces were found to reach values three times that of the corresponding quasi-static forces. Physical model results were compared to existing predictive formula (the momentum model) and found that the previous method had gaps and was inconsistent with the physics; therefore, new dimensionless predictive equations which are consistent with the physics were developed. Inspired by the failure of coastal highway bridges during extreme storm events, Cuomo et al. (2009) performed large scale experiments, 1:10, on coastal highway bridges and determined the dynamics of wave loadings, the effects of openings in bridge decks, and derived predictive methods for both quasi-static and impulsive wave loads. The new predictive equations are intended for design, and account for the effects of impact duration. Cuomo et al. (2010) also developed a predictive method for quasi-static and impact wave forces on vertical walls, which was derived from recent laboratory data collected for the Violent Overtopping by Waves at Seawalls (VOWS) project. The results were compared to previous studies and were found to provide a relatively good prediction. The new equations are similar in form to previous work by Cuomo et al. (2007) which are applicable to exposed jetties. Wave forces on a 1:5 scale reinforced concrete causeway-type coastal bridge superstructure were investigated by Bradner et al. (2011) for a range of random and regular wave conditions, and water levels. The effect of wave height, wave periods, and water levels for both horizontal and vertical forces were investigated, and vertical force were found to be approximately four times greater than the horizontal force. Formal design guidance for nearshore structures are published by both the American Society of Civil Engineers (ASCE) and the Federal Emergency Management Agency (FEMA). ASCE (2005) published minimum design standards for buildings and other structures, which include wave loads (Sections 5.4.2 to 5.4.5). In the ASCE standards, non-breaking and broken waves are treated as hydrostatic and hydrodynamic loads, with the hydrodynamic loads converted to an equivalent 3

81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 hydrostatic load based on a drag coefficient and the wave velocity. Breaking waves are treated as depth limited waves, and a combination of hydrostatic and dynamic pressures. FEMA (2011) published a design manual for residential coastal dwellings, which includes guidance on both wave forces, similar to those presented in the ASCE standards, and minimum structure elevations, which are set by local regulatory agencies and are often the minimum elevation required by the National Flood Insurance Program. In summary, there is little consistent guidance for determining design wave loads on elevated structures in the nearshore, such as jetties, coastal highway bridges, or raised dwellings subject to wave loads. Most theories have been developed for either nearshore vertical walls or offshore elevated structures, and application of either of these methods to elevated structure in the nearshore may not be appropriate. Therefore, this paper investigates how well the existing and well-accepted formulae by Goda originally developed for caisson structures can be modified and applied for elevated structures. This paper is outlined as follows. Section 2 presents the modified application of the Goda wave pressure formulae. Section 3 compares the observed physical model results to theory. Section 4 concludes the paper with a concise summary of results. 96 97 98 99 100 101 2. The Goda Wave Pressure Formulae Goda (1974; 2010) developed one of the most widely accepted methods for calculating wave forces on caissons, which assumes a trapezoidal pressure distribution (Fig. 1A). The formulae predict a maximum pressure at the still water level, p 1, which is directly proportional to the wave height, H, and is given by the following relation: 102 103 104 105 106 1 2 1 (1 )( 1 1 2 2 ) p cos cos gh (1) 2 where β is the angle of wave incidence, ρ is the density of the water, g is the acceleration due to gravity, λ 1 and λ 2 are modification factors for structure geometry, and α 1 and α 2 are wave pressure coefficients. The pressure decreases linearly from p 1 at the SWL to p 2 at the depth in front of the breakwater, h. The pressure at the base of the armor layer, p 3, is determined by linearly interpolating 4

107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 between p 1 and p 2. The pressure at the structure crest, p 4, is determined by linearly interpolating between p 1 and the theoretical elevation above the still water where the pressure goes to zero, η. The total horizontal force, F h, acting on the vertical face of the caisson is calculated by integrating the pressure distribution over the corresponding area and is given by the following (for a structure where the freeboard exceeds η ): 1 1 Fh p1 * p1 p3 h' (2) 2 2 where h is the water depth at the base of the armor layer. The original formulae (Goda, 1974) did not address impulsive wave breaking. Therefore, Takahashi et al. (1994) modified the wave pressure coefficients, α 2, to account for impulsive conditions, which modified the pressure at the still water level, p 1 (Fig. 1A). While calculating pressure coefficients, α 2, we assumed that the berm width is zero, since our experiment performed on the plane slope without rubble mound. In modifying Goda s formulae for elevated structures, two cases are considered. The first case shown in Fig. 1B is when the base of the structure is elevated above the bed but remains submerged below the SWL. The second case shown is Fig. 1C is when the base of the structure is elevated above the bed and is emergent above the SWL. Similar to the original case (Fig. 1A), the wave pressures p 1, p 2, p 3, and p 4 are assumed to remain constant. The pressure at the base of the elevated structure, p 5, is linearly interpolated using the following relation: 125 p 5 a 1 p * 1 a 1 p p p h' 1 3 3 for a 0 a 0 (3) 126 127 128 129 130 where a is the air gap defined as the distance from the SWL to the base of the structure and the SWL such that a < 0 when submerged and a > 0 when emerged. These equations are intended for the case of zero overtopping, h c > η. Fig. 1D shows a detail of the elevated structure for a > 0 with the parameters labeled. To calculate the total horizontal force, F h, acting on the structure, the pressure distribution is integrated over the corresponding area. Note that the Goda wave pressure formulae are 5

131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 intended for sizing caisson breakwaters, and the input wave is supposed to be the highest wave height in the design sea state, H max ; however, for this work the individual wave forces from each wave are of interest, so H max has been replaced with H. As well, note that the use of Eq. (3) with a h = -1 gives the original formula with the structure starting at the mudline (ie. no rubble mound base). Fig. 2 shows the theoretical curves for both non-impulsive and impulsive non-dimensional 2 horizontal wave force, F h ρgh b, as a function of non-dimensional wave height, H h b, for six non- dimensional air gaps, a h, ranging from -1 to 1.5 using Eq. (3) and the assumption of no overtopping. Both the force and wave height are non-dimensionalized by the depth at breaking, and the air gap non-dimensionalized by the depth of water at the base of the structure. In both equations, there was a change in curvature near H h b = 0.8. This was attributed to the α 2 parameter which was taken as the minimum of two equations, and was the intersection of the two curves. As observed in Fig. 2, as the structure was elevated the trends of these lines remained well behaved, and the change in elevation simply provided a reduction in force. When the wave height to breaking depth ratio, H h b, was equal to unity, the theoretical reduction in force for each air gap case, a h = -0.5, 0.0, 0.5, 1.0, and 1.5 with respect to the resting on the bed cases, a h = -1.0, are 17, 35, 51, 66 and 77%, respectively. For each air gap case, the percent reduction in force decreased as wave heights increased (H h b ratio increased). 148 149 150 151 152 153 154 155 156 3. Experimental Description In order to validate the modified Goda wave pressure formulae, a small-scale experiment was designed to measure the horizontal loads on an elevated structure. The experiment was conducted in a narrow flume which measured 487.5 cm long, 13.7 cm wide, and 32.2 cm deep, and had a piston type wave maker (Fig. 3). For these experiments, the bathymetry comprised of a horizontal section approximately 236.5 cm in length, followed by a 1:10 slope 190.0 cm in length. The load cell was suspended over the sloped portion of the bathymetry, approximately 332.5 cm from the maximum stroke position of the wave maker. Acoustic wave gages were placed over the horizontal portion of 6

157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 the flume, between the wave maker and sloped bathymetry, and over the sloped bathymetry, approximately 30.0 cm from the load cell (5H S ). Random waves using the JONSWAP spectrum with γ = 3.3, H S = 5.0 cm, and T P = 1.5 s, were tested for 200 waves with a run time of 300 s. For this experiment ten different spectra with the same JONSWAP parameters were randomly generated, and the five spectra with the most similar significant wave heights (variation less than 0.7%) were chosen for further analysis, for a total of 1000 waves generated for each air gap. The start and end of each trail were truncated to eliminate transient effects. Therefore, approximately 193 of the 200 waves per trial for a total of 964 waves were used in the analysis at each of the six air gaps tested. The input wave conditions had a significant wave height, H S = 5.79 cm, a peak period, T P = 1.59 s, and a maximum wave height, H max = 8.90 cm (Table 1). For the experiment, the flume was held at a constant water depth, h = 17.0 cm. A load cell was placed in the middle of the breaking region, defined as the region were the majority of waves break, as to record non-breaking, broken, and breaking waves. The air-gap of the load cell was varied in 1.0 cm increments, from a = -2.0 cm (resting on the bed) to a = 3.0 cm above the still water level. Two acoustic wave gages were used to measure the free surface elevation at the flat section of the flume and approximately 5 wave heights in front of the load cell. Video records of each trial were used to classify the largest waves. The load cell was comprised of a flat aluminum plate which spanned the width of the flume and backed by four strain gages in each corner. In this way, the load cell gave a direct measure of the integrated pressure distribution although we did not account for the dynamic effects of the plate mass which was assumed negligible. The sampling frequency of load cell was 0.01 and mean strain gage value was calibrated at the start and end of the experiment. Fig. 4 shows a scatter plot of the individual 964 waves plotted as h gt 2 and H gt 2 as measured by wave gage 1 for a/h = 1.5. The results are similar for all a/h ratios tested, and the value a/h = 1.5 chosen as it was the highest air gap tested and had the least reflection. From this figure it is evident that the measured waves are at a transitional depth, and approaching the depth limited condition. The waves are scattered across multiple theories including Solitary, Cnoidal, and 7

184 185 186 187 188 189 190 191 192 193 194 Stokes II, III, and V order. The majority of the waves are categorized as Stokes II order, representative of realistic conditions. From this diagram it can be concluded that the waves were breaking due to the depth limited conditions rather than steepness limited conditions as would be expected in shallower water. Fig. 5 shows a time series of non-dimensional free surface displacement, η/h b, measured at wave gage 2 and non-dimensional forces, F h ρgh b 2, for a/h = 1.0 for the first realization of waves. This figure shows the stochastic behavior of the horizontal wave forces in relation to the wave heights. The greatest forces are not necessarily associated with the largest waves, large waves may produce insignificant forces, and small waves are capable of generating significant forces. This variability is due to the location of breaking, and the compression of the entrained air pocket against the wall. 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 4. Results and Discussions The following section presents observations, and a comparison between values from the physical experiment to the modified theory. Using video analysis, the highest one-tenth of waves (ranked by the corresponding force recorded on the load cell) were classified as non-breaking, broken, or breaking waves. Non-breaking waves seldom ranked within the highest one-tenth, and generated relatively insignificant forces. When these waves contacted the base of the elevated structure, the crest of the waves was sheared off while the remainder of the wave continued to propagate by. As the air gap ratio, a/h, increased from -1.0 to 1.5, the overall number of non-breaking waves within the highest one-tenth increased from 2 to 14, and dramatically decreases the maximum horizontal force, 2 F h ρgh b, 81% from 1.96 to 0.37. Broken waves generated the second largest forces, after breaking waves. Depending on the distance between the break point and the structure, the wave either broke in front of the structure and impacted as a violent splash, or broke some distance before the structure and arrived as a bore of water. As the air gap ratio, a/h, increased from -1.0 to 1.5, the overall number of 8

210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 breaking waves within the highest one-tenth increased from 51 to 74, and substantially decreased the maximum horizontal force, F h ρgh b 2 by 85% from 3.94 to 0.58. As expected, breaking waves generated the largest forces of all wave types. When the wave broke just before the structure, the wave crest trapped a pocket of air against the wall even for these small-scale tests. The water compressed the air pocket as the wave collapsed until the air could burst upwards, which generated significantly more force than non-breaking and broken waves which did not trap air. As the air gap ratio, a/h, increased from -1.0 to 1.5, the overall number of breaking waves within the highest one-tenth decreased from 47 to 12, and dramatically decreased the maximum horizontal force, F h ρgh b 2, 75% from 6.12 to 1.50. Interestingly, the maximum number of breaking waves occurred for an air gap ratio of a/h = -0.5, with 74 out of 100 waves. It is speculated that the slightly raised structure minimized wave reflection, provided a cleaner surf zone for incoming waves, and prevented premature breaking. Due to the compression of air against the structure, breaking waves are capable of increasing the force by a factor of three when compared to broken waves of 2 2 similar height (H h b = 1.4, F hbrg ρgh b = 6, F hbkn ρgh b = 2). Fig. 6 shows the non-dimensional force, F h ρgh b 2, as a function of non-dimensional wave height, H h b, for the six air gap cases. Fig. 6F corresponds to the Goda formulae with no air gap, a/h = -1.0, where the structure starts at the mudline, and Fig. 6E to A represent increasing air gap ratios from a/h = -0.5 to 1.5. The experimental results show general agreement with the modified Goda formulae when compared at a/h = -1.0. The modified non-impulsive formulae provide a reasonable estimate of non-breaking and broken wave forces, and the modified impulsive formulae provide a reasonable estimate of breaking wave forces (Fig. 6F). As the air gap is increased from resting on the bed to elevated well above the SWL, both of the modified theories continue to provide accurate estimates of the horizontal force (Fig. 6E to A). The modified non-impulsive theory became increasingly conservative and overestimated the horizontal force for non-breaking and broken waves as the air gap ratio increased. The divergence was readily apparent at the largest air gaps (Fig. 6A and B). Due to the scarcity and scatter of breaking 9

236 237 238 239 240 waves at the highest air gaps ratios, it is uncertain whether the modified impulsive theory also overestimates the breaking wave forces. For the higher air gaps ratios, the fundamental physical impact processes are changed. Therefore, instead stopping the full momentum of non-breaking and broken waves, it passes underneath the structure. For breaking waves, instead of air being trapped and compressed against the structure, the air is forced out underneath the structure. 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 5. Summary and Conclusions The present paper investigated the application of the Goda wave pressure formulae to elevated structures, by carrying out small scale physical model tests. The waves were classified into three types, non-broken, broken, or breaking using video analysis and forces measured. Based on these observations from the small-scale flume with relative air gaps ranging from -1.0 < a/h < 1.5, the air gap was found to play a key role in the reduction of horizontal wave forces. As expected, when the air gap increased, the resulting forces decreased, and the estimated value became increasingly conservative. As the a/h ratio increased from -1.0 to 1.5, the reduction in force was approximately 75%, for H h b equal to unity. The physical model results confirmed that the modified application of the Goda wave pressure formulae provide a good estimate of the horizontal forces on elevated structures for impulsive breaking waves for all air gaps. For non-breaking and broken waves, the formulae were found to provide a good estimate for air gaps ranging from -1.0 < a/h < 0, and provide a conservative estimate for a/h > 0. Overall, the results are encouraging but large scale model tests would be needed to confirm the applicability of Goda s formulae modified for elevated structures. 257 258 259 260 261 6. Acknowledgement This material is based upon work partially supported by the National Science Foundation under Grant No. 13XXXXX. Any opinion, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the National 10

262 263 264 Science Foundation. The authors thank two anonymous reviewers for their constructive comments. (NSF Grant No. is not currently unavailable due to the Shut Down. It will be updated before the final revision.) 11

265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 7. References ASCE, 2005. Minimum design loads for buildings and other structures, ASCE/SEI 7-05, Reston, VA: ASCE. Bea, R. G., Xu, T., Stear, J. & Ramos, R., 1999. Wave forces on decks of offshore platforms. Jounral of Waterway, Port, Coastal, and Ocean Engineering, Vol. 125, No. 3, pp. 136-144. Bradner, C., Schumacher, T. C. D. & Higgins, C., 2011. Experimental setup for a large-scale bridge superstructure model subjected to waves. Journal of Waterway, Port, Coastal, and Ocean Engineering,, Vol. 137, No. 1, pp. 3-11. Cuomo, G., Allsop, W., Bruce, T. & Pearson, J., 2010. Breaking wave loads at vertical seawalls and breakwaters. Coastal Engineering, Vol. 57, No. 4, pp. 424-439. Cuomo, G., Shimosako, K. & Takahashi, S., 2009. Wave-in-deck loads on coastal bridges and the role of air. Coastal Engineering, Vol. 56, No. 8, pp. 793-809. Cuomo, G., Tirindelli, M. & Allsop, W., 2007. Wave-in-deck loads on exposed jetties. Coastal Engineering, Vol. 54, No. 9, pp. 657-679. FEMA, 2011. Coastal Construction Manual: Principles and Practices of Planning, Siting, Designing, Constructing, and Maintaining Residential Buildings, s.l.: FEMA. Goda, Y., 1974. New wave pressure formulae for composite breakwater. Copenhagen, ASCE, pp. 1702-1720. Goda, Y., 2010. Random Seas and Design of Maritime Structures, 3rd Edition. s.l.:world Scientific. Kaplan, P., 1992. Wave impact forces on offshore structures: Re-examination and new interpretations. Houston, Texas, Offshore Technology Conference, pp. 79-86. Kaplan, P., Murray, J. J. & Yu, W. C., 1995. Theorectical analysis of wave impact forces on platform deck structures. Proceedings of the 14th International Conference on Offshore Mechanics and Arctic Engineering, pp. 189-198. Kennedy, A. et al., 2011. Building destruction from waves and surge on the Bolivar peninsula during Hurricane Ike. Journal of waterway, port, coastal, and ocean engineering, Vol. 13, No. 7, pp. 291 132-141. 12

292 293 294 295 296 297 298 299 NOAA, 2012. Hurricane Research Division: Atlantic Oceanographic and Meteorological Laboratory. Available at: http://www.aoml.noaa.gov/hrd/tcfaq/e11.html Robertson, I. N., Riggs, H. R., Yim, S. C. S. & Young, Y. L., 2007. Lessons from Hurricane Katrina storm surge on bridge and buildings. Journal of waterway, port, coastal, and ocean engineering, Vol. 133, No. 6, pp. 463-483. Takahashi, S., Tanimoto, K. & Shimosako, K., 1994. A proposal of impulsive pressure coefficient for the design of composite breakwaters. Proclamation of the International Conference on Hydro-Technical Engineering for Port and Harbor Construction, pp. 489-504. 13

300 Notation List 301 302 303 304 a F H g H air gap horizontal force gravity wave height 305 H max maximum wave height in the design sea state 306 307 H S h significant wave height depth of water in front of the breakwater 308 h depth of water above foundation 309 310 311 312 313 314 315 316 317 318 319 320 321 h b h c L N N R p T P α β γ λ η ρ depth of water five wave heights seaward of the breakwater freeboard wavelength number of waves number of realizations pressure peak period wave pressure coefficient angle of wave incidence JONSWAP parameter modification factors for structure geometry distance above SWL where pressure goes to zero density 14

322 323 324 325 326 327 328 329 330 331 332 333 334 List of Figures Fig. 1 Schematics of the Goda wave pressure distribution applied to: A) original application to a caisson type breakwater; B) modified for a structure submerged below sea level (a h 0); and C) modified for a structure elevated above sea level (a h > 0). D) Detail of C for a structure elevated above sea level with labeled parameters. Fig. 2 Theoretical curves for non-impulsive (dash-dot) and impulsive (solid) horizontal wave forces for the range of air gaps, -1 < a/h < 1.5. The Goda formulae correspond to a/h = -1.0. Fig. 3 Schematic of experimental configuration. Fig. 4 Wave conditions for a/h = 1.5 (964 waves) Fig. 5 Time series of free surface displacement (top) and force (bottom) for a/h = 1.0 Fig. 6 Non-breaking (triangle), broken (square), breaking (circle), and unclassified (dot ) wave forces as a function of wave height for the six air gap cases compared to the modified theoretical impulsive (solid line) and non-impulsive (dash dot line) forces. 15

335 336 337 338 339 Fig. 1 Schematics of the Goda wave pressure distribution applied to: A) original application to a caisson type breakwater; B) modified for a structure submerged below sea level (a h 0); and C) modified for a structure elevated above sea level (a h > 0). D) Detail of C for a structure elevated above sea level with labeled parameters. 16

340 341 342 Fig. 2 Theoretical curves for non-impulsive (dash-dot) and impulsive (solid) horizontal wave forces for the range of air gaps, -1 < a/h < 1.5. The Goda formulae correspond to a/h = -1.0. 17

343 344 Fig. 3 Schematic of experimental configuration. 18

345 346 Fig. 4 Wave conditions for a/h = 1.5 (964 waves) measured at wave gage 1. 19

347 348 349 Fig. 5 Time series of free surface displacement (top) and force (bottom) for a/h = 1.0, measured at wave gage 2. 20

350 351 352 353 Fig. 6 Non-breaking (triangle), broken (square), breaking (circle), and unclassified (dot ) wave forces as a function of wave height for the six air gap cases compared to the modified theoretical impulsive (solid line) and non-impulsive (dash dot line) forces. Wave heights measured at wave gage 2. 21

354 355 356 List of Tables Table 1. Summary of wave properties for the six test cases 22

357 358 Table 1. Summary of wave properties for the six test cases Case a/h N R N H s T p H L ξ (cm) (s) 1-1.0 5 964 5.99 1.57 0.0505 0.4449 2-0.5 5 964 5.97 1.60 0.0495 0.4494 3 0.0 5 964 5.83 1.59 0.0493 0.4505 4 0.5 5 964 5.68 1.59 0.0485 0.4540 5 1.0 5 964 5.71 1.59 0.0486 0.4535 6 1.5 5 964 5.58 1.59 0.0482 0.4557 Avg. - - 964 5.79 1.59 0.0491 0.4513 23