Download as pdf or txt
Download as pdf or txt
You are on page 1of 44

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/342615490

Hydrodynamic pressures over modular porous reef breakwaters

Article  in  Journal of Hydro-environment Research · July 2020


DOI: 10.1016/j.jher.2020.06.001

CITATIONS READS
0 106

2 authors:

Srineash V K Kantharaj Murali


Indian Institute of Technology Madras Indian Institute of Technology Madras
11 PUBLICATIONS   12 CITATIONS    86 PUBLICATIONS   652 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

LARGE MULTIPURPOSE PLATFORMS FOR EXPLOITING RENEWABLE ENERGY IN OPEN SEAS (PLENOSE) View project

Final Year B-Tech Project View project

All content following this page was uploaded by Srineash V K on 02 July 2020.

The user has requested enhancement of the downloaded file.


1 HYDRODYNAMIC PRESSURES OVER MODULAR POROUS REEF

2 BREAKWATERS

3 V. K. Srineash (1,*), K. Murali (2)

(1)
4 Research Scholar, Dept. of Ocean Engineering, Indian Institute of Technology Madras, Chennai, India.

(2)
5 Professor, Dept. of Ocean Engineering, Indian Institute of Technology Madras, Chennai, India.

6 Corresponding author: srineash@gmail.com, 2murali@iitm.ac.in.


(1,*)

7 Abstract

8 Low crested structures built with homogeneous stones or precast blocks serving the purpose of

9 wave attenuation are often referred to as reef breakwaters. The present work involves laboratory

10 investigation of hydrodynamic pressures on reef breakwaters made of gabion boxes in

11 submerged conditions. The study provides an understanding of nature and magnitudes of

12 pressures in and around the reef structures depending on the configuration of choice. The

13 knowledge on the wave induced pressure on porous structures is important to better understand

14 the stability and flow around the coastal structures. Such reef based structures are also being

15 adopted as mitigation against sea-level rise and to reduce loads exerted on existing maritime

16 structures. The pressure reduction achieved on the leeside of the structure is also examined in the

17 present study. This is of immediate interest for preserving the existing coastal structures in a

18 serviceable condition, which are prone to more severe conditions than considered in the design.

19 The wave-induced dynamic pressures on the submerged porous structure are analyzed and

20 discussed in the present work for varying reef and wave parameters. This experimental study

21 involves analysis of wave-induced pressure on the seaside, leeside and at the mid-section of the

22 porous reef breakwaters. The investigation has been carried forward to bring out major

1
23 parameters of concern for the present problem and their ranges. The study demonstrates that a

24 pressure reduction of about 70% is achievable by the use of reef breakwaters and this clearly

25 indicates the application of such structures for coastal resilience.

26 Keywords: Artificial Reefs, Coastal Structures, Gabions, Low-Crested Structures,

27 Hydrodynamic Pressures, Reef Breakwaters.

28 1.0 Introduction

29 Low crested structures, built predominantly with homogeneous stones facilitating wave

30 attenuation, are often classified as reef breakwaters. These structures will be well suited for

31 locations where only a moderate degree of wave attenuation is required (Dattatri et al. 1978;

32 Ahrens 1987). Such reef breakwaters are shown to be effective in protecting shorelines and

33 coastal structures, that are vulnerable. Sometimes, they can also be adopted for to reducing loads

34 on the existing coastal structures (Pilarczyk 2003; Jeng et al. 2005; Koriam et al. 2014). These

35 structures, being low crested, preserve the aesthetics of the coast while also serving their function.

36 Further, the reef breakwaters tend to enhance marine life in the vicinity (Firth et al. 2014).

37 These sub-aerial reef breakwaters become economical when compared with the existing

38 conventional structures (such as seawalls, revetments and breakwaters) as they consume less

39 resources (Johnson et al. 1951; Dattatri et al. 1978; Ahrens 1987). Though there are many studies

40 relating to the wave transmission over reef breakwaters (Tanaka 1976; Ahrens 1987;

41 d’Angremond et al. 1996; Seabrook and Hall 1998; van der Meer et al. 2005), there is hardly any

42 literature available focusing on the hydrodynamic pressures on reef breakwaters. On the other

43 hand, the knowledge on wave-induced pressures on porous structures is important to derive an

44 understanding of the stability and the flow around the coastal structures (Oumeraci and

2
45 Partenscky 1990). The pressures also play a major role in foundation stability. Further, the

46 studies focused on pressure measurements are essential to derive insights about the magnitudes

47 of wave loads induced on the structures subjected to wave action (Allsop et al. 1997; Hull and

48 Muller, 2000). It is also emphasized in the works of Groot et al. (1995); Pinto and Neves (2006)

49 that the understanding of the hydrodynamic pressures over the rubble mound structures is vital

50 for the economic design of structures. The information derived from the pressure measurements

51 may be related to the sliding stability and wave transmission aspects (Groot et al. 1995, Muttray

52 and Oumeraci 2005; Pinto and Neves 2006; Sumer et al 2011; Sumer et al. 2013; Jensen et al.

53 2014). It can be perceived from the existing literature (Groot et al. 1995; Allsop et al. 1997;

54 Muttray and Oumeraci 2005; Pinto and Neves 2006; Cantelmo et al. 2010; Cuomo et al., 2010;

55 Jensen et al. 2014), that there is adequate studies focused on the hydrodynamic pressures over

56 rubble mound breakwaters. There are also studies where pressures over coastal structures such as

57 semicircular breakwaters (Sundar and Ragu 1998; Dhinakaran et al. 2002), pile-supported

58 breakwaters (Sundar and Subba Rao 2002) have been investigated. However, there is still a gap

59 in knowledge on understanding the hydrodynamic pressures acting over the reef breakwaters.

60 Due to the frequent occurrence of storm surges and also ensuing sea level rise, coastal

61 structures face the hazard of excessive loading than considered in the design (Sasikumar et al.

62 2018). This leads to structural damage on the existing structures owing to the excessive loading

63 caused due to the above-mentioned aspects due to climate change. In recent times, reef

64 breakwaters are used to reduce loading on the existing structures (Pilarczyk 2003; Reddy and

65 Neelamani 2005; Jeng et al., 2005; Koraim et al. (2014); Sasikumar et al. 2018; Srineash et al.

66 2020) in order to counteract the above-mentioned effects. These submerged structures when

67 placed in tandem with the existing structures modify and reduce the incident waves. The

3
68 hydrodynamic force reduction attained on the existing structures due to the presence of

69 submerged structures on the sea-side (also called as tandem arrangement) has been demonstrated

70 in the works of Madrigal and Prud’Homme (1990) and Reddy and Neelamani (2005). It is worth

71 mentioning the fact that the work of Madrigal and Prud’Homme (1990) gives an insight into the

72 feasibility of incorporating reef breakwaters for reducing the forces on existing structures. As

73 pointed out in the works of Madrigal and Prud’Homme (1990) ; Pinto and Neves (2006), there is

74 still a scope and need for an extensive study with varying reef and wave parameters, to better

75 understand the interaction of waves with reef breakwaters. Reddy and Neelamani (2005) have

76 studied the effects of the low crested breakwater in reducing the forces experienced by a vertical

77 sea wall. It is insightful to note that the findings of Reddy and Neelamani (2005) concludes that

78 the effect of pool length (distance between the reef breakwater to the seawall) is not significant

79 in altering the wave force experienced by the vertical seawall. This observation gives an

80 impression that the hydrodynamic pressure on the leeside of the reef breakwater is not expected

81 to have considerable spatial variation. Therefore, the above-mentioned literature reinforces the

82 fact that reef breakwaters can be considered to be a viable solution to protect a structure from the

83 increase in wave-induced loads and to enhance coastal resilience.

84 The Narrow neck reef in Gold coast, Australia is a classic example where the

85 construction of a submerged reef has addressed the prevalent coastal erosion problem and led to

86 beach formation (Black 2001; Black and Mead 2001). Such solutions have also been tried and

87 found to be successful in many parts of the world. Further, there are examples where reef

88 breakwaters are used to protect or to reduce loads on existing structures. One such example, is

89 the use of submerged breakwater to protect existing breakwater in Santa Monica, USA (Adams

90 and Sonu 1986). There is also a recent example at Kiberg, Norway (Sasikumar et al. 2018) where

4
91 a reef breakwater in tandem (on the seaside) to an existing breakwater is considered as a climate

92 change mitigation measure. It has also been brought out in the study of Cox and Clark (1992)

93 that the tandem arrangement of reef breakwater provides safety against extreme events and result

94 in an economical design. Further, the analysis of pressures exerted on such reef based structures

95 may be considered vital to understand the amount of pressure reduction attained due to the

96 presence of the structure (for given reef and wave parameters). This is taken up in the present

97 research. The literature study reveals that there is hardly any literature available where the

98 pressures exerted on reef breakwaters are examined. On the other hand, the theoretical analysis

99 of the wave-induced pressures over such submerged porous reef breakwaters is not viable

100 (Reddy and Neelamani 2005). To address this research gap, the present work is focused to

101 understand and quantify the hydrodynamic pressures acting on gabion based reef breakwaters.

102 This experimental study involves the analysis of wave-induced pressure on the seaside, leeside

103 and at the mid-section of the porous reef breakwaters. The investigation has been carried forward

104 to bring out major parameters of concern for the present problem and their ranges. A detailed

105 parametric investigation has been performed in the study involving various reef configurations

106 and wave conditions. The results from the present study is compared and referenced with other

107 similar works from the literature. Further, predictive equation capable of estimating the wave-

108 induced pressure at the crest of the structure is proposed based on this experimental study .

109 2.0 Methodology

110 The present study is based on laboratory investigation of model reef breakwaters made of

111 gabion boxes in submerged conditions. These porous stone gabion units are known for their

112 superior energy damping characteristics (Chinnarasri et al. 2008) while also providing shelter for

113 marine organisms. Use of pre-filled gabions for the construction of coastal structures (Thomas et

5
114 al. 1986) facilitates faster construction due to its modular nature. Further, the use of gabions is

115 considered to be economical in comparison with the traditional riprap (Thomas et al. 1986). The

116 test program involved measurements of pressures on the seaside (p1), at the mid-section (p2), at

117 the leeside (p3), and at the seaside crest (p4) of the structure (Fig. 1). Simultaneously,

118 measurements of wave elevations have also been carried out for the purpose of correlating the

119 pressures with wave characteristics. The detailed results on wave transmission and reflection

120 have been presented elsewhere (Srineash and Murali, 2019).

121 2.1 Dimensional analysis

122 It is essential to identify important variables and parameters that influence the physics of the

123 problem considered. In the present research, the following variables are considered to be

124 essential for describing wave interaction with porous reef breakwaters. These are the depth at the

125 toe of the structure (d), crest width (B), depth of submergence (d’), wavelength (L) at the toe of

126 the structure, incident wave height (H), specific weight of water (γ). The hydrodynamic pressure

127 acting on the reef breakwater can be described by following independent variables.

128 p = f (d, B, d’, L, H, γ) (1)

129 In order to obtain the required nondimensional parameters, Buckingham’s Pi theorem is used.

130 Therefore the equation (1) is expressed as,

131 f’ (p, d, B, d’, L, H, γ)=0 (2)

132 In accordance with Buckingham’s Pi theorem, n variables with m dimensions can be expressed

133 as (n-m) nondimensional terms or π terms. Therefore, for the present case containing n=7

134 variables and m=3 dimensions. Therefore, the physical quantities can be expressed with n-m=4

6
135 nondimensional numbers. The pressure is expressed in nondimensional form as 2p/γH (Sundar

136 and Ragu 1998; Dhinakaran et al. 2002; Sundar and Subba Rao 2002; Cuomo et al., 2010

137 Alkhalidi et al., 2015). In the present study, the pressure response factor Kp is introduced (Dean

138 and Dalrymple, 1984) to account for the variations in the vertical levels of the pressure

139 transducers (for each reef configuration). Hence, the dependence of nondimensional pressure

140 (p/KpγH) on the above-mentioned variables may be associated nondimensionally as given below:

141 (p/KpγH) = f (d/L, d’/d, d’/H, B/d) (3)

coshk d+z
142 Where, p is the hydrodynamic pressure acting over the structure; Kp = coshkd
is the

143 pressure response factor (k- wave number [k=2π/L]). This method of parametrization is

144 convenient for the present study involving reef breakwater configuration with varying d’. The

145 wavelength (L) is considered to be the governing variable for studies involving wave-induced

146 dynamic pressures on coastal structures. This can be noticed from the existing literature

147 (Dhinakaran et al., 2002; Reddy and Neelamani 2005; Dhinakaran et al., 2009; Alkhalidi et al.,

148 2015) where, the dynamic pressures on coastal structures are studied as a function of d/L.

149 Therefore, in the present study, the hydrodynamic pressures exerted over the reef breakwaters

150 are studied as a function of wavelength, L and this is represented in nondimensional form as d/L.

151 However, the effect of other nondimensional parameters (B/d, d’/d, d’/H) is also considered in

152 the present research. It is worth mentioning here that in the related investigation by Reddy and

153 Neelamani (2005), the crest width ratio, B/d ratio was maintained constant but in the present

154 study, the effects due to variations in B/d are also investigated. Further, The submergence ratio

155 (d’/d) is also expected to affect the dynamic pressure variations and hence, the investigation is

156 also focused to bring out the effects due to the changes in d’/d. The ranges of dimensional

157 variables and dimensionless parameters are provided in Table 1 and Table 2 respectively.

7
158 Table 1 Range of variables considered in the study

Variable Definition Range Unit


B Crest Width 0.30 – 0.90 m
d’ Depth of Submergence 0.05 – 0.15 m
H Wave Height 0.03 – 0.14 m
L Wave Length 1.50 – 5.00 m
T Wave Period 1.06 – 2.98 s
159

160 Table 2 Range of nondimensional parameters considered in the study

Non-dimensional parameter Definition Range


B/d Crest width ratio 1.00 – 3.00
B/L Relative crest width 0.06 – 0.60
d/L Relative water depth 0.06 – 0.20
d’/H Relative depth of submergence 0.34 – 5.30
d’/d Submergence ratio 0.17 – 0.50
H/d Relative wave height 0.10 – 0.48

161 3. Experimental Study

162 3.1 Model description

163 The model reef breakwater was created in the laboratory scale with model scale gabion boxes. A

164 slope of 1:2 and 1:1 was maintained on the seaside and leeside respectively during the

165 experimental investigations. The structure takes a stepped profile owing to the use of gabion box

166 models which are cuboidal in shape. The dimension of each gabion box unit measures 0.05 x

167 0.05 x 0.15 m and comprises stones of size ranging from 7.5 mm to 12.5 mm. The average

168 porosity maintained throughout the study is 0.37. Generally, in field conditions, the value of

169 porosity varies from 0.30 to 0.40 for gabions. More details about the gabion box models may be

170 found in Srineash and Murali (2015 a, b). Readers may refer to Srineash and Murali (2019) to

171 view the snapshots taken during the experiments. In the present research where gravity waves are

8
172 involved, Froude’s scaling law has been used to achieve similitude between model and prototype

173 with a scale ratio of 1:20.

174 3.2 Experimental Setup

175 The experiments involving the interaction of waves with reef breakwaters have been

176 conducted in the 72 m long, 2 m wide and 2.7 m deep wave flume at the Department of Ocean

177 Engineering, Indian Institute of Technology Madras, India. The schematic sketch of the

178 experimental setup is depicted in Fig. 1. A water depth (d) of 0.3m was maintained throughout

179 the study. During the experiments, the reef variables, d’ and B were varied along with the wave

180 variables L and H. The waves are made to shoal over submerged slope provided in the seaward

181 direction of the porous structure. This process makes the waves to attain required steepness

182 replicating the real field condition. The submerged slope used in the study extends to a length of

183 8.1 m horizontally and 0.37 m vertically, as illustrated in Fig. 1. This arrangement of the sloping

184 bottom gives rise to a gradient of about 1in22. The shoaling of waves over the submerged slope

185 in the test facility has been studied in the earlier work (Srineash and Murali 2018). The wave

186 flume is demonstrated to have good wave absorption capability (Murali and Mani 1996;

187 Krishnakumar et al. 2010; Srineash and Murali 2018). The gabion based reef breakwaters were

188 rested over the sand bed in order to bring in the real-time effects of a porous bed. The sand bed

189 was confined on all sides and compacted leading to no subsidence (Srineash and Murali 2019).

190 The reef breakwater model was placed at a distance of 34.15 m from the wave-maker. The

191 functional performance characteristics (such as wave transmission and wave reflection) of the

192 considered gabion based reef breakwaters has already been published (Srineash and Murali 2019)

193 and hence the present investigation focuses on the hydrodynamic pressures acting over the reef

194 breakwaters.

9
195
196 Fig. 1. Schematic view of the experimental setup.
197 3.3 Test Procedure

198 The gabion reef breakwater was installed in the wave flume as depicted in Fig.1. An

199 enlarged view of the porous structure is also brought out for better clarity. The investigation of

200 wave interaction with the reef breakwaters has been carried forward with 10 geometric

201 configurations (with varying B and d’). The instantaneous wave elevations ( ) were recorded

202 with resistance type wave probes. The locations of the wave gauges (WP1 and WP2) can be

203 observed from Fig. 1. These wave gauges were used to measure the wave elevations and thereby

204 to estimate the incident wave height. More details on the wave elevation measurements, wave

205 transmission and wave reflection aspects can be found from (Srineash and Murali 2019).

206 Measurements of wave-induced pressures was carried out using diaphragm type underwater

207 pressure transducers (Make: KISTLER RTC 28) with a range of 0–0.2 bar. In total, four

208 pressure transducers were used for the study. The location of the pressure transducers installed

209 on the structure is depicted in Fig. 1. It is to be noted that the pressure transducer p4 measure the

210 pressure in the crest of the structure (d-d’). The transducers p1, p2 and p3 measures the pressure at

211 the mid-height of the structure [i.e., at (d-d’)/2]. Further, the transducer p1 measures the pressure

212 in the seaside of the structure while p2 measures the pressure in the mid-section of the structure.

10
213 The transducer p3 measures the pressure in the leeward direction of the structure at the mid-

214 height. A sampling frequency of 40 Hz was adopted for acquiring data from wave gauges and

215 pressure transducers. This corresponds to a Nyquist frequency of 20 Hz. For the experimental

216 investigations, regular waves with combinations of wave periods and wave heights were

217 generated and the corresponding pressure response on the structure is studied.

218 4 Analysis of Data

219 The pressure measurements warrant extraction of characteristic pressures of wave crest and wave

220 trough along with their correlation at different locations. This helps us understand the peak crest

221 and trough pressures with various nondimensional parameters. Therefore, the hydrodynamic

222 pressure data (excluding the hydro-static pressure) from the present measurements were analyzed

223 in the time domain using zero up-crossing approach. For this analysis, the steady-state part of

224 wave-induced pressure time-history (care has been taken to exclude the transient and re-

225 reflections) was considered for analysis (Reddy and Neelamani, 2005). The mean of measured

226 pressure amplitudes above the reference level is taken as crest pressure (pc). Similarly, the mean

227 of measured pressure amplitudes below the reference level is taken as trough pressure (pt). The

228 wave height incident on the structure is ascertained using two probe approach proposed by Goda

229 and Suzuki (1976). The wave probes WP1 and WP2 are located such that they satisfy the

230 requirements proposed by Goda and Suzuki (1976) [0.05 ≤ Δl/L ≤ 0.45 where, Δl is the distance

231 between the two wave probes] to resolve the measured wave elevation into the incident and the

232 reflected components. The incident wave height is required to determine the theoretical pressure

233 at the given location. This theoretical pressure is used in the study for nondimensionalisation of

234 the measured hydrodynamic pressures.

11
235 5 Results and Discussion

236 Understanding the hydrodynamic pressures over the porous reef breakwater is the main objective

237 of the present study. Typical time histories of p1, p2, p3 and p4, along with their corresponding

238 energy spectra and the location of the pressure measurement are presented in Fig. 2.

(a) 1 (b) 10
0 (c)
10-1

S/SO, p1 (p1 )
-2
10
0.5
p1 /  H

10-3
-4
10
0 10-5
10-6
10
-7 B/d= 2 ; d'/d= 0.34
-0.5 10-8 d/L= 0.12; d'/H= 1.28
9 10 11 12 13 14 15 16 0 1 2 3 4 5 6
t/T f/fO
(d) 1 (e) 10
0 (f)
10-1

S/SO, p1 (p2 )
-2
10
0.5
p2 /  H

10-3
-4
10
0 10-5
10-6
10
-7 B/d= 2 ; d'/d= 0.34
-0.5 10-8 d/L= 0.12; d'/H= 1.28
9 10 11 12 13 14 15 16 0 1 2 3 4 5 6
t/T f/fO
(g) 1 (h) 100 (i)
10-1
S/SO, p1 (p3 )

-2
10
0.5
p3 /  H

10-3
-4
10
-5
0 10
-6
10
10-7 B/d= 2 ; d'/d= 0.34
-0.5 10-8 d/L= 0.12; d'/H= 1.28
9 10 11 12 13 14 15 16 0 1 2 3 4 5 6
t/T f/fO
(j) 1 (k) 100 (l)
-1
10
S/SO, p1 (p 4 )

10-2
0.5
p4 /  H

-3
10
10-4
0 10-5
-6
10
10-7 B/d= 2 ; d'/d= 0.34
-0.5 10
-8 d/L= 0.12; d'/H= 1.28
9 10 11 12 13 14 15 16 0 1 2 3 4 5 6
t/T f/fO
239
240 Fig. 2. Typical time series of pressure measurements and its corresponding spectral densities at
241 various locations of the reef breakwater

242 The figure pertains to a d/L of 0.12, d’/H of 1.28, d’/d of 0.34 and B/d of 2 as indicated

243 in the figure. While, the pressure-time histories of p1, p2, p3 and p4 are brought out in Figs. 2(a),

244 2(d), 2(g) and 2(j) respectively, the corresponding normalized power spectra (S/So,p1)

12
245 corresponding to each pressure time series is presented in Figs. 2(b), 2(e), 2(h) and 2(k)

246 respectively. In these plots, the spectral density (S) is normalized with the spectral density

247 corresponding to the fundamental frequency of p1 (So,p1). Likewise, the frequency of the spectra (f)

248 is normalized with the fundamental frequency (fo) of the same spectrum to obtain the normalized

249 frequency (f/fo).

250 In general, it is observed that the pressure decays within the reef (p2) and behind the reef

251 (p3) in comparison with the wave maker side pressure, p1. As expected, the crest pressure (p4) is

252 higher in comparison to p1, as p4 is close to the still water level. The reduction in the dynamic

253 pressure at the mid-section (p2) and in the leeside (p3) of the reef, with respect to the seaside

254 pressure (p1), is evident from Figs. 2(a), 2(d) and 2(g) respectively. The spectra corresponding to

255 the location of the pressure transducers p1, p2 and p4 [Figs. 2(b), 2(e) and 2(k)] reveal that the

256 energy in the second harmonic is two orders less than the energy in the fundamental frequency

257 (first harmonic). The same is not observed in case of p3, where the second harmonic itself is

258 merely one order less than the fundamental frequency. This denotes a considerable energy

259 redistribution to higher harmonics in case of the leeward pressure transducer, p3, due to a strong

260 interaction (Srineash and Murali 2019) of the reef breakwater with the wave train. This

261 occurrence has also been observed earlier (Hall and Seabook 1998, Dattatri et al. 1978 and

262 Johnson et al. 1951) with regard to transmitted waves. But the earlier efforts did not quantify the

263 same. To explore this further, the values of S/So,p1 (at p1, p2, p3 and p4) for the first four

264 harmonics are brought out in Table 3. The table shows peak values of normalized power spectra

265 at different locations (p1, p2, p3 and p4). It is evident from Table 3 that the presence of higher

266 harmonics are prominent in p3 in comparison with p1, p2 and p4. Further, the energy in the 2nd, 3rd

267 and 4th harmonics of p3 are noticed to be about one, two and four orders less than the energy in

13
268 its fundamental frequency. On analyzing Table 3, it is inferred that at the mid-section of the reef

269 (p2), there is a reduction in the energy density of about 41% in comparison with p1, in the

270 fundamental frequency. Similarly, about 71% of the energy is attenuated at pressure

271 measurement pertaining to p3 located in the leeside of the reef. These aspects reinforce that such

272 reef breakwaters can be suitable for wave energy reduction and thereby to reduce loads on the

273 leeside structures. Thus, making such structures ideal to enhance the resilience of the coast.

274 Similar behaviour is observed for other reef parameters (B/d and d’/d). This analysis of variation

275 in dynamic pressure with d/L and other relevant parameters is explained in the forthcoming

276 discussions.

277 Table 3 Variation in the harmonics of the pressure measurements for d/L =0.12 [d’/H=1.28,
278 d’/d=0.34 and B/d=2]
f/fo S/So,p1
p1 p2 p3 p4
1 1.00x100 5.92x10-1 2.87x10 -1
1.4 x100
2 1.40x10-2 1.22x10-2 2.76x10 4.4 x10-2
-2

3 2.10x10-4 9.12x10-4 1.49x10-3 1.4 x10-3


4 3.70x10-5 6.51x10-5 5.18x10-5 2.1 x10-4
279

280 5.1 Effects of relative water depth (d/L) on hydrodynamic pressure

281 The variation of magnitudes of hydrodynamic pressures is studied in this section as a

282 function of relative water depth (d/L) for all the ranges of B/d and d’/d. The measured dynamic

283 pressures are nondimensionalized with the theoretical dynamic pressure (KpγH/2) as explained in

284 section 2.1.

285 5.1.1 Nondimensional crest pressure

286 The variations in the nondimensional crest pressure (2pc/KpγH) are depicted as a function

287 of relative water depth (d/L) in Fig. 3. Figures 3(a), 3(b) and 3(c) portray the effects of relative

14
288 water depth on nondimensional crest pressure corresponding to the transducer p1 for

289 submergence ratios d’/d =0.17, 0.34 and 0.5 respectively. Similarly, the effects of relative water

290 depth on other locations of pressure transducers (p2, p3, p4) are illustrated in Figs. 3(d) - 3(f), Figs.

291 3(g) - 3(i) and Figs. 3(j) - 3(l) respectively. As evident, each figure brings out the effects due to

292 changes in the crest width ratio, B/d. For better clarity, the location of the pressure transducer is

293 embedded in the figures. In general, a higher magnitude of pressure is obtained for lower d/L

294 (<0.1). This is expected, as longer wavelengths lead to higher dynamic pressures. When the

295 wavelengths are long, the value of d/L becomes close to shallow water conditions and hence, the

296 effects due to dynamic pressures will be significant for longer wavelengths. A peak in crest

297 pressure is also observed at d/L=0.08. There is also an occurrence of the secondary peak at half

298 of the above wavelength (d/L=0.16). A similar trend is noticed in the earlier publications of

299 Dhinakaran et al. (2002), Sundar and Subba Rao (2002) where the occurrence of the peak in crest

300 pressure is noticed in multiples of d/L. The peak crest pressures occur at d/L=0.45 and 0.9 in the

301 work of Sundar and Subba Rao (2002) and d/L= 0.1 and 0.2 in the related investigation

302 conducted by Dhinakaran et al. (2002). These aspects suggest that the variation of crest pressure

303 is oscillatory in nature with its dependence on d/L. This oscillatory trend in the hydrodynamic

304 pressures may be due to the result of complex wave-field present in the vicinity of the reef

305 breakwater which includes incident and reflected wave components (Young and Testik 2011)

306 along with wave transmission effects. It is worth to recall the wave reflection characteristic of the

307 reef breakwaters from the earlier work of Srineash and Murali (2019), where, an oscillatory

308 behaviour is perceived when wave reflection is studied as a function of d/L. Further, this

309 observation was found to be in accordance with past studies (Lan and Lee, 2010; Mei and Black,

310 1969). Further, it can also be noticed that the maximum value of the wave reflection occurred for

15
311 d/L=0.08 (Srineash and Murali, 2019), suggesting that the oscillatory behaviour in the wave-

312 induced pressure is due to the complex wave field in the vicinity of the reef breakwater.

313 However, it is essential to conduct further studies for a broader range of d/L to ascertain if the

314 oscillatory trend of hydrodynamic pressure is periodic and if it has got a diminishing effect for

315 higher d/L. An overall examination of all plots for each location of pressures indicates a similar

316 range of dynamic pressure across all d’/d.

317 A closer examination of Fig. 3(a) reveals that the value of 2pc1/KpγH is around 1.25 for

318 the d/L range of 0.1 to 0.2. This is true for the range of B/d considered (B/d=1-3). When d/L falls

319 below 0.1, there is an increase in the crest pressure by about 40%, on comparison with higher d/L.

320 This is because, for longer wavelengths, the value of d/L becomes close to shallow water

321 conditions and hence the effects due to dynamic pressures will be significant at p1. Owing to the

322 dependency of the Kp on the wavelength, short-period waves (lower wavelengths) tend to have

323 small Kp values towards the bottom, while for the waves with longer wavelengths can take values

324 up to unity. Therefore, longer wavelengths lead to higher dynamic pressures at p1 which is a

325 function of Kp. The observations from Fig. 3 reveals that the effect of B/d is not appreciable for

326 any particular d’/d. Further, inter-comparison of plots corresponding to p1 (Figs. 3(a) - 3(c))

327 suggests that the crest pressure does not vary significantly as the submergence of reef changes.

328 This gives a perception that the crest pressure on the seaside of the structure is a major function

329 of d/L. Vertical distribution of particular symbol represents varying wave heights and the effects

330 due to the wave height variations were found to be less significant. In order to substantiate this

331 aspect, the pressure measurements are studied as a function of d’/H in the forthcoming section

332 (section 5.2).

16
(a) 4 (b) 4 (c) 4
B/d= 1 B/d= 1 B/d= 1
B/d= 2 B/d= 2 B/d= 2
B/d= 3 B/d= 3 B/d= 3
3 3 3

2 pc1 /KpH

2 pc1 /KpH

2 pc1 /KpH
2 2 2

1 1 1

d' /d=0.17 d' /d=0.34 d' /d=0.50


0 0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L d/L
(d) 4 (e) 4 (f) 4
B/d= 1 B/d= 1 B/d= 1
B/d= 2 B/d= 2 B/d= 2
B/d= 3 B/d= 3 B/d= 3
3 3 3
2 pc2 /Kp H

2 pc2 /Kp H

2 pc2 /Kp H
2 2 2

1 1 1

d' /d=0.17 d' /d=0.34 d' /d=0.50


0 0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L d/L
(g) 4 (h) 4 (i) 4
B/d= 1 B/d= 1 B/d= 1
B/d= 2 B/d= 2 B/d= 2
B/d= 3 B/d= 3 B/d= 3
3 3 3
2 pc3 /KpH

2 pc3 /KpH

2 2
2 pc3 /KpH2

1 1 1

d' /d=0.17 d' /d=0.34 d' /d=0.50


0 0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L d/L
(j) 4
(k) 4 (l) 4
B/d= 1 B/d= 1 B/d= 1
B/d= 2 B/d= 2 B/d= 2
B/d= 3 B/d= 3 B/d= 3
3 3 3
2 pc4 /KpH

2 pc4 /KpH

2 pc4 /KpH

2 2 2

1 1 1

d' /d=0.17 d' /d=0.34 d' /d=0.50


0 0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L d/L
333
334 Fig. 3. Effect of relative water depth (d/L) on normalized crest pressures

335 The effects of d/L on 2pc1/KpγH is also investigated for d’/d=0.34 and 0.5 respectively in

336 Figs. 3(b) and 3(c). It is evident that the occurrence of the secondary peak in nondimensional

337 crest pressure at d/L=0.16 becomes more prominent with a decrease in submergence ratio (d’/d).

338 This could be due to the fact that the interaction of the wave with the structure becomes

17
339 pronounced at lower d’/d. This behaviour leads to higher magnitudes in the secondary peak for

340 lower d’/d. The occurrence of maximum crest pressure is noticed for d/L =0.08 for d’/d = 0.17,

341 0.34 and 0.50 with the 2pc1/KpγH taking values of 2.20, 2.23 and 2.06 respectively.

342 A similar analysis as carried out for p1 has been carried forward to the pressure

343 transducer p2 [Figs. 3(d), 3(e) and 3(f)]. The phenomenon of increase in nondimensional crest

344 pressure with a decrease in d/L remains unaltered for p2 at d/L<0.1 (as discussed earlier). The

345 occurrence of the secondary peak in nondimensional crest pressure is observed to be marginal for

346 p2 as observed from Figs. 3(d), 3(e) and 3(f). It is interesting here to note that the effect of the

347 crest width ratio (B/d) is prominent for lower submergence ratios of 0.17 and 0.34. This indicates

348 that the effect of crest width has its role in reducing the wave-induced pressures at the mid-

349 section of the structure.

350 The nondimensional crest pressure measured at the location p3 is analyzed and the

351 recordings are observed to be generally below the theoretical wave pressure. The values of

352 2pc3/KpγH are below unity for most cases as perceived in the figure [Figs. 3(g)-(i)]. This

353 demonstrates the utility of such reef based structures in reducing the pressure on the leeside.

354 Further, the prominence of B/d at lower d’/d is observed from the measurements pertaining to p3.

355 From the results, it is witnessed that the pressure reduction by the reef breakwater is pronounced

356 for shorter waves (i.e., d/L>0.1). A nondimensional pressure ratio of about 0.32 is measured for

357 d/L=0.2 for B/d=3 at d’/d=0.17. This denotes about 70% reduction in the wave-induced pressures

358 for the above-mentioned reef and wave parameters. It is insightful to notice here that this

359 corresponds to the highest B/d ratio and the lowest d’/d ratio considered in the present campaign.

360 It is noted that for higher wavelengths (d/L<0.1), the reef breakwater considered in the present

361 study is not noticed to provide wave pressure reduction. Indeed, it is seen to amplify the leeside

18
362 pressure for higher submergence ratio and at lower relative water depth (d/L). This may be due to

363 the presence of dominant nonlinear effects for d/L<0.1 (Srineash and Murali 2018; Srineash and

364 Murali 2019), wave reflection combined with the phenomenon of wave shoaling over the reef

365 (Walmsley et al. 2002) leading to the amplification of p3 at higher submergence ratios. The

366 effects of d’/H on nondimensional crest pressure of p3 will also be discussed in the study for

367 better understanding in the consequent discussions.

368 Further, the effects of d/L on p4 is discussed in Figs. 3(j), 3(k) and 3(l)]. The pressure

369 record at p4 is noticed to have the highest values of dynamic pressures in comparison with the

370 other pressure transducers (p1, p2, p3) as p4 is closer to the still water level. On comparing the

371 results from, p1 and p4, similar trends of the dynamic pressures are noticed. The highest value of

372 nondimensional pressure recorded at p4 is 2.4 at d/L=0.08 and at d’/d=0.17. It is worth

373 mentioning that this highest pressure is recorded for the B/d=1, which corresponds to the lowest

374 crest width considered. Further, the nondimensional crest pressure of (2pc4/KpγH) 2.4 signifies

375 that there is about 2.4 times increase in pressure due to the introduction of the structure in the

376 wave field. It is insightful to notice that the maximum value of secondary peak occurs when

377 d/L=B/L and d’/H~1 [ Fig 3 may not bring out the variations due to d’/H and hence the effects

378 due to d’/H is brought out in section 5.2]. When d’/H~1, the depth of submergence equals to the

379 incident wave height (d’~H). This above-mentioned condition represents a scenario with the

380 water depth at the toe being equal to the crest width of the structure and wave height being equal

381 to the depth of submergence. Therefore, under such conditions (of wave and reef parameters),

382 the magnitude of the secondary peak is noticed to amplify due to prominence of the wave

383 structure interaction process, including the wave reflection effects, thereby leading to the

384 oscillatory behaviour (Lan and Lee, 2010; Young and Testik 2011; Srineash and Murali, 2019).

19
385 5.1.2 Crest–trough pressure ratio

386 The study on crest-trough pressure ratio signifies the net change in momentum on the reef

387 elements for the given wave parameters. This could further be correlated to the forces acting on

388 the reef breakwater. Such studies help in understanding the variation or symmetry in the crest

389 and trough pressure distribution. This is attempted in Fig. 4. In general, it is noticed that when

390 d/L<0.1, the crest to trough pressure ratio is higher. This is consistently observed for p1, p2 and p4.

391 Such a behaviour is not observed in p3 at lower d’/d ratios, as there is energy redistribution to the

392 higher harmonics in case of transmitted wave (Hall and Seabook 1998, Dattatri et al. 1978 and

393 Johnson et al. 1951) due to the strong interaction of structure with the wave. Further, the effects

394 of wave heights (H/d or d’/H) become relatively appreciable at lower d/L due to the presence of

395 nonlinear effects in these regions i.e, d/L<0.1 (Srineash and Murali 2018; Srineash and Murali

396 2019). At higher d/L (d/L>0.1), the crest-trough pressure ratio seems to remain close to unity

397 indicating equal magnitudes in the crest and the trough pressures.

20
(a) 4 (b) 4 (c) 4
B/d= 1 B/d= 1 B/d= 1
B/d= 2 B/d= 2 B/d= 2
B/d= 3 B/d= 3 B/d= 3
3 3 3

pc1 /pt1

pc1 /pt1

pc1 /pt1
2 2 2

1 1 1

d' /d=0.17 d' /d=0.34 d' /d=0.50


0 0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L d/L
(d) 4 (e) 4 (f) 4
B/d= 1 B/d= 1 B/d= 1
B/d= 2 B/d= 2 B/d= 2
B/d= 3 B/d= 3 B/d= 3
3 3 3
pc2 /pt2

pc2 /pt2

pc2 /pt2
2 2 2

1 1 1

d' /d=0.17 d' /d=0.34 d' /d=0.50


0 0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L d/L
(g) (h)12 (i) 12
12
B/d= 1 B/d= 1 B/d= 1
11 11 11
B/d= 2 B/d= 2 B/d= 2
10 10 10
B/d= 3 B/d= 3 B/d= 3
9 9 9
d' /d=0.34 d' /d=0.50
8 8 8
pc3 /pt3

pc3 /pt3

7 7 pc3 /pt37
6 6 6
5 5 5
4 4 4
3 3 3
2 2 2
1 1 1
d' /d=0.17
0 0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L d/L
(j) 4
(k) 4 (l) 4
B/d= 1 B/d= 1 B/d= 1
B/d= 2 B/d= 2 B/d= 2
B/d= 3 B/d= 3 B/d= 3
3 3 3
pc4 /pt4

pc4 /pt4

pc4 /pt4

2 2 2

1 1 1

d' /d=0.17 d' /d=0.34 d' /d=0.50


0 0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L d/L
398
399 Fig. 4. Effect of relative water depth (d/L) on crest-trough pressure ratio (pc/pt)
400

401

402

21
403 5.1.3 Crest pressure ratio

404 The ratio of crest pressure at p2, p3 and p4 is studied with reference to the crest pressure at

405 p1 in this section. This exercise has been performed for developing a correlation of wave-induced

406 pressure values at various locations (p2, p3 and p4) with reference to p1. From Fig. 5, it is noticed

407 that the values of pc2/ pc1 and pc3/pc1 remains less than unity demonstrating the wave attenuation.

408 The crest pressure ratio of pc2 in relation to pc1 is brought out in Figs. 5 (a), 5(b) and 5(c) for

409 d’/d= 0.17, 0.34, 0.50. Further, it is observed from Fig. 5(a), that the effect of B/d becomes

410 evident at higher d/L (d/L>0.1). This suggests that at lower submergence ratios and at higher d/L,

411 the effect of B/d becomes distinct. This indicates that when the crest of the structure is close to

412 still water, the effects of B/d becomes appreciable for longer wavelengths. However, the effect

413 of B/d on pc2/pc1 is noticed to be relatively less significant for larger submergence ratios as

414 noticed from Figs. 5(b) and 5(c). Analogous behaviour of crest pressure ratio is obtained for p3,

415 which is on the leeside of the structure. The minimum value of pc3/pc1 among all the cases

416 considered is noticed to be 0.30 which denotes a 70% reduction in pressure on the leeside of the

417 structure in relation to the seaside. This is observed for the reef parameters with d’/d=0.17 and

418 B/d=3. Hence, it becomes evident that the maximum pressure reduction is achieved for the case

419 with highest crest width and the lowest submergence ratio. This, in turn, signifies the role of reef

420 parameters in aiding reduction in wave-induced pressures.

421 The crest pressure ratio of p4 is brought out in Figs. 5(g), 5(h) and 5(i). It has to be noted

422 here that in order to take into account the difference in the vertical levels of p1 and p4, the crest

423 pressure ratio is divided by its corresponding pressure response factor (Kp). As expected, the

424 crest pressure ratio of p4 is predominantly greater than unity indicating that the pressure in the

425 crest level is greater than the wave-induced pressure at the mid-level (p1) of the structure.

22
426 However, this effect is less significant in the case of d’/d=0.50 as perceived in Fig. 5(i), as the

427 structure is submerged sufficiently below still water. Further, the mean crest pressure ratio of p4

428 is about 1.22, 1.14 and 1.02 for d’/d =0.17, 0.34 and 0.50 respectively. This indicates that the

429 amplification of crest pressure gets prominent with the reduction in d’/d.

(a) 4
(b) 4 (c) 4
B/d= 1 d' /d=0.17 B/d= 1 d' /d=0.34 B/d= 1 d' /d=0.50
B/d= 2 B/d= 2 B/d= 2
3 B/d= 3 3 B/d= 3 3 B/d= 3
pc2 /pc1

pc2 /pc1

pc2 /pc1
2 2 2

1 1 1

0 0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L d/L
(d) 4
(e) 4
(f) 4
B/d= 1 d' /d=0.17 B/d= 1 d' /d=0.34 B/d= 1 d' /d=0.50
B/d= 2 B/d= 2 B/d= 2
3 B/d= 3 3 B/d= 3 3 B/d= 3
pc3 /pc1

pc3 /pc1

2 2
pc3 /pc1 2

1 1 1

0 0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L d/L

(g) 4 (h) 4 (i) 4


B/d= 1 d' /d=0.17 B/d= 1 d' /d=0.17 B/d= 1 d' /d=0.50
pc4 /pc1 (Kp1/Kp4)

pc4 /pc1 (Kp1/Kp4)

pc4 /pc1 (Kp1/Kp4)

B/d= 2 B/d= 2 B/d= 2


3 B/d= 3 3 B/d= 3 3 B/d= 3

2 2 2

1 1 1

0 0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L d/L
430
431 Fig. 5. Effect of relative water depth (d/L) on crest pressure ratio

432 5.2 Effect of the relative depth of submergence (d’/H) on dynamic pressures

433 The preceding discussions above does not bring out the effects of wave height

434 sufficiently. Therefore, the wave-induced pressure measurements are studied as a function of

435 d’/H in this section for a given d/L. In each plot, the value of B/d may be related to B/L as the

23
436 plots correspond to a constant d/L. Therefore, the change in B/d essentially means a change in

437 B/L. Also, the influence of d’/H for a particular d’/d and B/d can be related to H/d as the present

438 measurements pertain to constant water depth. The values of the relative depth of submergence

439 (d’/H) tend to be greater for larger d’/d (=0.50) and smaller for lower d’/d (=0.17), as a similar

440 range of relative water depth H/d was chosen across all the configuration of reef breakwaters.

441 During experiments, reef induced wave breaking is observed for d’/H<1 (Srineash and Murali

442 2019). However, breaking pressure may not have been recorded due to limitations in the

443 instrumentation.

444 5.2.1 Nondimensional crest pressure of p1

445 The effect of nondimensional crest pressure for d/L=0.2-0.06 is brought out in Figures

446 6(a)-6(h). Generally, for each d/L, it is observed that the effect of change in crest width ratio (B/d)

447 or relative crest width (B/L) has a marginal bearing on the pressure. This is in concurrence to the

448 earlier observation as discussed in section 5.1.1.

449 The mean value of 2pc1/KpγH for d/L of 0.2 is about 1.1. A similar behavior is noticed for

450 d/L of 0.18, 0.16, 0.14 and 0.12 [Figs. 6(b)-6(e)]. It is prudent to recall here the earlier

451 observation that for higher d/L (>0.1) there would be no salient increase in nondimensional

452 dynamic pressure. A close observation of Fig. 6(c) pertaining to d/L =0.16 reveals a distinct peak

453 pressure with values approaching close to 1.8. This is noticed to be 30% higher than the mean

454 wave-induced pressure for the given relative water depth. Further, it is noted that this distinct

455 occurrence of crest pressure is noted when d/L=B/L and d’/H~1. This behaviour is consistent at

456 all locations and results when d/L = 0.16 which is due to the prominence in the wave structure

457 interaction process for the above-mentioned wave and reef parameters.

24
(a) (b)
(b)
3
(c)
d/L=0.20 d/L=0.18 P=Pc4 d/L=0.16
2.5

2 2 2
2
2pc1 /Kp H

p H

2pc1 /Kp H
1.5

Kp
/v H 
2pc1P/K
1.5 11.5 1.5

0.5

d/L=0.18
1 0 1 1
0 1 2 3 4 5 0 0 1 1 22 33 44 55 0 1 2 3 4 5
d'/H d'/H
d'/H d'/H
(d) (e) (f)
d/L=0.14 d/L=0.12 d/L=0.10

2 2 2
2pc1 /Kp H

2pc1 /Kp H

2pc1 /Kp H
1.5 1.5 1.5

1 1 1
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
d'/H d'/H d'/H

(g) (h) (i)

2 2
2pc1 /Kp H

2pc1 /Kp H

1.5 1.5

d/L=0.08 d/L=0.06
1 1
0 1 2 3 4 5 0 1 2 3 4 5
d'/H d'/H
458
459 Fig. 6. Effect of relative depth of submergence (d’/H) on normalized crest pressure of p1

460 5.2.2 Nondimensional crest pressure of p2

461 The results of the pressure transducer, (p2) against d/L is brought out in Fig. 7. Two

462 distinct trends are noted in Fig. 7: (i) for d’/d ≥0.34, the effect of d’/H and B/d is not noted to be

463 dominant (as observed in case of p1 in Fig. 7), (ii) for d’/d =0.17, the effects due to B/d or B/L is

25
464 noticed to be appreciable. It is recalled here that when pc2 was studied in relation to d/L as similar

465 behaviour is noticed for lower d’/d values. Overall, the effect of d’/H is found to be less

466 significant considering the nondimensional crest pressure. Moreover, the effect of B/d seems to

467 be pronounced at lower d’/d and the influence of d/L on nondimensional crest pressure becomes

468 noticeable for d/L<0.1.

(a) 2 (b)
(b)
3 2
(c) 2
P=Pc4
2.5

1.5 1.5 1.5


2pc2 /Kp  H

H

2pc2 /Kp  H
2
p
/ H Kp/K

1 1.5 1 1
v c2
P 2p

0.5 0.5 0.5


0.5

d/L=0.20 d/L=0.18
d/L=0.18 d/L=0.16
0 0 0 0
0 1 2 3 4 5 0 0 1 1 22 33 44 55 0 1 2 3 4 5
d'/H d'/H
d'/H d'/H
(d) 2 (e) 2 (f) 2

1.5 1.5
2pc2 /Kp  H

2pc2 /Kp  H

2pc2 /Kp  H
1.5

1 1
1

0.5 0.5

0.5
d/L=0.14 d/L=0.12 d/L=0.10
0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
d'/H d'/H d'/H

(g) 2 (h) 2 (i)

1.5
2pc2 /Kp  H

2pc2 /Kp  H

1.5

1
1

0.5

0.5
d/L=0.08 d/L=0.06
0
0 1 2 3 4 5 0 1 2 3 4 5
d'/H d'/H
469
470 Fig. 7. Effect of relative depth of submergence (d’/H) on normalized crest pressure of p2

26
471 5.2.3 Nondimensional crest pressure of p3

472 The nondimensional crest pressure pertaining to the pressure transducer in the leeside of

473 the reef breakwater (p3) is studied in this portion. The effect of B/d on lower d’/d (=0.17)

474 becomes perceptible in Fig. 8 for p3.

(a) (b)
(b)
3
(c)
P=Pc4
2.5

1 1 1
2
2pc3 /Kp H

/v HK H

2pc3 /Kp H
1.5
p
2pc3P/K p

0.5 10.5 0.5

0.5

d/L=0.20 d/L=0.18 d/L=0.18 d/L=0.16


0 0 0 0
0 1 2 3 4 5 0 0 1 1 22 33 44 55 0 1 2 3 4 5

d'/H d'/H
d'/H d'/H
(d) (e) (f)

1 1 1
2pc3 /Kp H

2pc3 /Kp H

2pc3 /Kp H

0.5 0.5 0.5

d/L=0.14 d/L=0.12 d/L=0.10


0 0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
d'/H d'/H d'/H

(g) (h) (i)

1 1
2pc3 /Kp H

2pc3 /Kp H

0.5 0.5

d/L=0.08 d/L=0.06
0 0
0 1 2 3 4 5 0 1 2 3 4 5
d'/H d'/H
475
476 Fig. 8. Effect of relative depth of submergence (d’/H) on normalized crest pressure of p3
477

27
478 The occurrence of the distinct peak in nondimensional crest pressure is detected for d/L

479 =0.16 [Fig. 8(c)], when d/L=B/L and d’/H~1. This phenomenon is noticed to be preserved on the

480 seaside, leeside and mid-section pressure records. Also, there are regions where the wave

481 pressure reduction on the leeside of the reef breakwaters is not significant. Especially at lower

482 relative water depths (d/L<0.1) and at higher depth of submergence (d’/H>1.5), where the

483 dynamic pressures observations are close to unity. This interprets that the reef may not be

484 considered effective in the pressure reduction in the above-mentioned range of reef and wave

485 parameters.

486 5.2.4 Nondimensional crest pressure of p4

487 The effect due to the variations in d’/H on the nondimensional crest pressure of p4 is

488 portrayed in Fig. 9. There is a trend of increase in nondimensional crest pressure (2pc4/KpγH)

489 with a reduction in d’/H at p4. This is because the changes in the wave-induced pressure due to

490 the wave height variations become appreciable close to the still water level and hence this

491 observation is perceived at p4. A difference of about 22% is noted across the mean of maximum

492 and minimum d’/H for d/L=0.2 [Fig. 9(a)]. However, the effects of B/d (or B/L) and d’/H is

493 noticed to be less influential. A similar trend is noticed for d/L =0.18 - 0.1 as seen in Figs. 9(b)-

494 9(f). It is interesting to note from Fig. 9(c) that the occurrence of a distinct peak in p4 (for

495 d/L=0.16 at d’/H~1 and B/L=d/L) is similar to the observations in other locations of pressure

496 measurements. This substantiates that this combination of B/L, d/L and d’/H is noted to amplify

497 the wave pressure on the structure. Further, the examination in Figs. 9(g) and 9(h) reveal that the

498 effect of B/d and d’/d is significant at d/L<0.1. This is in concurrence to the observations at p1.

499 This could be due to the strong interaction of the reef with the wave at lower d/L making the

28
500 effects of crest width (B/d or B/L) and the submergence ratio (d’/d) to influence the dynamic

501 pressures.

(a) (b)
(b)
3
(c)
2.5 d/L=0.20 2.5 d/L=0.18 P=Pc4 2.5 d/L=0.16
2.5
2pc4 /Kp  H

H

2pc4 /Kp  H
2
2 2 2

p
/ H Kp/K
1.5

v c4
P 2p
1.5 11.5 1.5

0.5

1 1 d/L=0.18 1
0
0 1 2 3 4 5 0 0 1 1 22 33 44 55 0 1 2 3 4 5
d'/H d'/H
d'/H d'/H
(d) (e) (f)
2.5 d/L=0.14 2.5 d/L=0.12 2.5 d/L=0.10
2pc4 /Kp  H

2pc4 /Kp  H

2pc4 /Kp  H
2 2 2

1.5 1.5 1.5

1 1 1

0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
d'/H d'/H d'/H

(g) (h) (i)


2.5 2.5
2pc4 /Kp  H

2pc4 /Kp  H

2 2

1.5 1.5

1 d/L=0.08 1 d/L=0.06
0 1 2 3 4 5 0 1 2 3 4 5
d'/H d'/H
502
503 Fig. 9. Effect of relative depth of submergence (d’/H) on normalized crest pressure of p4

504 Overall, there is a marginal rise in nondimensional crest pressure of p4 with a reduction in

505 d’/H, which is about 20% when d’/H reduces from 5.2 to 0.34. However, such effects are not

506 reflected in p1, which is sufficiently submerged below the free surface. Nevertheless, for p4, the

29
507 effects due to d’/H has got its significance as the effects of wave height is felt at the crest of the

508 structure (which is close to the free surface/ still water level).

509 5.3 Comparison of present measurements with published work

510 The present pressure measurements are compared with the pressure measurements from

511 the earlier works. Such an exercise increases the confidence in adopting the inferences derived

512 based on the study. The work of Alkhalidi et al. (2005) has been chosen here, and this research is

513 focused on wave-induced pressures over vertical slotted wave barrier. Structures such as wave

514 barriers have similar functionality as that of breakwaters, which is wave attenuation. These

515 structures offer resistance to wave motion and reduce wave transmission. Also, as noted earlier,

516 the hydrodynamic pressure measurements for reef breakwaters with similar wave parameters are

517 unavailable. Therefore, the results from the present study involving the hydrodynamic pressures

518 are compared with the results of Alkhalidi et al. (2005). Further, the range of relative water depth

519 used in the study of Alkhalidi et al. (2005) was similar to the present study. The hydrodynamic

520 pressures measurements from Alkhalidi et al. (2005) is portrayed in Fig. 10 along with the

521 measurements from the present experimental study. Further, the experimental data (from

522 Alkhalidi et al., 2005) chosen for the comparison correspond to the measurements closer to the

523 free surface. Hence, the measurements of p4 is chosen for the comparative study. This

524 comparative study has been carried out for measurements with three crest widths B/d=1-3 and for

525 three submergence ratios d’/d=0.17, 0.34 and 0.50 as depicted in Figs. 10 (a), (b) and (c)

526 respectively. The study conducted by Alkhalidi et al. (2005) consists of measurements pertaining

527 to 20% and 30% porosity and this is brought out in Fig. 10. On analyzing Figs. 10 (a), (b) and (c)

528 it is evident that the results presented in Figs. 10(c) has got a higher correlation. This may be due

529 to the fact that when d’/d=0.5, the effect of the structure on the wave becomes less prominent

30
530 (relatively). This could be a reason to perceive a higher correlation at d’/d=0.5. Overall

531 comparisons reveal that the dynamic pressure ranges observed in the present experimental

532 campaign are comparable with the past studies (Alkhalidi et al. 2005). However, this exercise is

533 performed to compare the wave-induced pressure exerted on structures meant for similar

534 functionality and hence a higher/accurate correlation of the present results with the earlier study

535 is not essential.

(a) 4 (b) 4
B/d= 1 B/d= 1
B/d= 2 B/d= 2
B/d= 3 B/d= 3
3 3
+ 20% porosity + 20% porosity
2 pc4 /Kp H

2 pc4 /Kp H
Alkhalidi, 2015 Alkhalidi, 2015
30% porosity 30% porosity

2 + 2 +
+ +
+ +
1 + 1 +

d' /d=0.17 d' /d=0.34


0 0
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2
d/L d/L
(c) 4
B/d= 1
B/d= 2
B/d= 3
3
+ 20% porosity
2 pc4 /Kp H

Alkhalidi, 2015
30% porosity

2 +
+
+
1 +

d' /d=0.50
0
0.05 0.1 0.15 0.2
d/L
536

537 Fig. 10 Comparison of present pressure measurements with earlier researchers

538 5.4 Empirical equation for prediction of wave-induced pressure exerted on reef

539 breakwaters

31
540 In the present research, the empirical equation for prediction of wave-induced pressure

541 exerted on the crest of the reef breakwaters is derived based on the experimental data. This is

542 considered crucial for design purposes and hence the derivation of predictive equation is taken

543 up at the crest of the structure (p4). Therefore, the derived equation based on the experimental

544 campaign is presented below in equation (4). The derived equation is a function of

545 nondimensional parameters considered in the study. The significance of each of the

546 nondimensional parameters in governing the hydrodynamic pressures have been elaborated in

 B d' 
547 the earlier section (2.1). In addition, the term   is introduced to relate the mass of water
L H

548 stored above the reef and the notional mass of water incident on the structure (Srineash and

549 Murali 2019). Further, to derive the predictive equation, a nonlinear regression analysis has been

550 performed based on nonlinear least-squares method. The measured and predicted wave-induced

551 pressures are compared in Fig. 11. The proposed equation was subjected to statistical analysis

552 and the Root Mean-Square Error (RMSE) and R2 values of equation (4) was found to be 0.15

553 and 0.73 respectively. The equation is applicable within the range of parameters considered in

554 the present study as discussed in Table 2. The derived equation does not hold good when

555 d/L=0.16; d’/H=1 and B/d=1. This is due to the occurrence of the secondary peak in this region.

556 Therefore, the equation is not valid at the above-mentioned combination of reef and wave

557 parameters.

0.53 2.46 0.27 0.028 2.89


 pc 4   B d'   d'   d'
   0.36 d  B
 0.64   0.4   0 . 8   1 . 44  (4)
 K H  L L L H H d
 p 

32
2.8
2
2.6
R = 0.73
RSME = 0.15

2Pc4 /K pH- Predicted


2.4
2.2
2
1.8
1.6
1.4
1.2
1
0.8 d'/d= 0.5 - 0
0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8
2P c4 /K pH- Measured
558

559 Fig. 11 Measured versus predicted wave-induced pressures

560 6 Conclusions

561 A systematic experimental study involving the interaction of waves with model

562 submerged reef breakwaters made out of modular gabion units are presented and discussed. The

563 pressure measurements from the leeside of the structure provide evidence for the reduction in

564 hydrodynamic pressure. The study brings out the pressures as a function of relative water depth

565 (d/L) and relative depth of submergence (d’/H) for varying crest width ratio (B/d) and

566 submergence ratio (d’/d). The study suggests that d/L is the major parameter governing dynamic

567 pressures for reef based structures followed by B/d and d’/d.

568 A general trend of higher magnitudes in wave-induced pressures is realized for lower d/L

569 (<0.1), due to dominant shallow water effects in these regions. The wave-induced pressures were

570 found to have an oscillatory trend when studied as a function of d/L which may be due to the

571 result of complex wave-field present in the vicinity of the reef breakwater which includes

572 incident and reflected wave components along with wave transmission effects. The pressure

33
573 values are noticed to have peak corresponding to d/L=0.08 and there is also an occurrence of the

574 secondary peak at d/L=0.16 which becomes prominent at lower d’/d.

575  The results pertaining to the non-dimensional crest pressure of p1 (seaside), reveal that the

576 crest pressures are about 40% higher for d/L<0.1 when compared to higher d/L. This was

577 noticed as the relative water depth in these regions are close to the shallow water

578 conditions.

579  The nondimensional crest pressure of p2 (mid-section) at the mid-height of the structure

580 (d-d’)/2 indicates the dependence of B/d at lower d’/d. A clear pressure reduction is

581 witnessed for higher B/d at p2.

582  A distinct pressure reduction is noticed across the reef breakwater, as observed from the

583 results pertaining to the leeside pressure measurements (p3). A maximum of about 70%

584 reduction in incident pressure is realized during the study for d/L=0.2 for B/d=3 at

585 d’/d=0.17 when compared to theoretical pressure. The effect of B/d becomes pronounced

586 for lower d’/d values. The study demonstrates that a pressure reduction of about 70% is

587 achievable by the use of reef breakwaters and this promotes the application of such reef

588 based structures to enhance coastal resilience.

589  The crest pressure at p4 indicates higher magnitudes of pressure than observed in p1.

590 However, the increase in crest pressure for d/L<0.1 is consistent as noted for p1. The

591 effects of B/d and d’/d become prominent when d/L<0.1, thus indicating the pronounced

592 effect of the wave structure interaction process in these regions.

593 Finally, an empirical equation capable of predicting the wave-induced pressure at the crest

594 of the structure is derived based on the experimental data. This may be considered to be

595 crucial for the design of coastal structures subjected to wave action.

34
596 Acknowledgment

597 The authors are thankful to the Department of Ocean Engineering for the provision of the

598 flume and experimental facilities. The first author would like to thank fellow researchers and

599 masters students for their support rendered during the experiments. The authors acknowledge the

600 constructive comments and feedback from the anonymous reviewers, which has significantly

601 improved the quality of the manuscript.

602 Notations

603 The following symbols are used in this paper:

604 B Crest width of the reef (m);

605 d Water depth at the toe of the structure (0.30 m);

606 d1 Water depth before the submerged ramp (0.67 m);

607 d’ depth of submergence (m);

608 H Incident wave height (m);

609 k wave number;

610 L Wavelength at the toe of the reef (m);

611 T Wave period (s);

612 WP Wave probe;

613 p Pressure (kgf/m2);

614 η Wave Elevation (m);

615 Δl Distance between the wave probes (m);

616 Kp Pressure response factor (m/m);

617 γ Specific weight of water (N/m3);

35
618 z Vertical coordinate (defined positive in the upward direction);

619 The following subscripts are used in this paper:

620 c Crest;

621 t Trough;

622 APPENDIX A

623 EXPERIMENTAL ERROR ANALYSIS

624 The measurements during the experiments are prone to uncertainties due to various factors.

625 Though great care is taken to make sure the measured quantities are accurate, there are certain

626 uncertainties which cannot be ruled out during experiments. During the present experimental

627 investigations, a personal computer was used to acquire the data and hence this eliminates the

628 possible observational and recording error. Further, acquiring the data in digital form enables one

629 to have measurements accurately with highest possible decimals. However, the wave probes and

630 pressure transducers require calibration which involves observation and recording of quantities.

631 There may be uncertainties associated with the above-mentioned process and during the process

632 of other measurements. An attempt has been made here to evaluate and quantify uncertainties

633 associated with various measurements performed during the study.

634 A.1 UNCERTAINTY ANALYSIS

635 In an experimental measurement, the uncertainty associated with a quantity ‘R’ as a function of

636 various measured variables such as X1, X2,… Xn each subjected to uncertainty of W1, W2,… Wn

637 can be estimated as follows,

36
(A.1)
+ +֠֠֠

638 The above expression is in accordance to Holman and Gajda (1989). The uncertainties

639 associated with the measured variables such as pressure, wave elevation, crest width, wave

640 height and other relevant variables is presented in Table 4.

641 Table 4 Uncertainty analysis

Sl. No. Variables Uncertainty


1 Pressure, p ±1.95%
2 Wave height, H ±1.00%
3 Wavelength, L ±0.17%
4 Wave period, T ±0.01%
5 Water depth, d ±0.04%
6 Crest width, B ±0.67%
7 Depth of submergence, d’ ±0.67%
642

643 7 References

644 Adams, C. B., and Sonu, C. J. (1987). “Wave transmission across submerged near-surface

645 breakwaters.” Coastal engineering, B. L. Edge, ed., ASCE, New York, 1729–1738.

646 Ahrens, J. P ., 1987. “Characteristics of Reef Breakwaters,” Technical Report CERC-87-17, U.S.

647 Army Corps of Engineers, Waterways Experiment Station, Vicksburg.

648 Alkhalidi, M., Neelamani, S., & Assad, A. I. A. H. (2015). “Wave pressures and forces on slotted

649 vertical wave barriers”. Ocean Engineering, 108, 578-583.

650 Allsop, N. W. H., McKenna, J. E., Vicinanza, D., & Whittaker, T. T. J. (1997). “New design

651 methods for wave impact loadings on vertical breakwaters and seawalls”. In Coastal

652 Engineering 1996 (pp. 2508-2521).

37
653 Black, K. (2001). Artificial surfing reefs for erosion control and amenity: theory and

654 application. Journal of Coastal Research, 1-14.

655 Black, K., & Mead, S. (2001). Design of the Gold Coast reef for surfing, public amenity and

656 coastal protection: surfing aspects. Journal of coastal research, 115-130.

657 Cantelmo CL, Allsop WI, Dunn SC. “Wave pressures in and under rubble mound breakwaters”.

658 (2010), Fifth International Conference on Scour and Erosion 2010. San Francisco,

659 California.

660 Chinnarasri, C., Donjadee, S., & Israngkura, U., 2008. "Hydraulic Characteristics of Gabion-

661 Stepped Weirs", J. Hydraul. Eng., (134), 1147–1152.

662 Cox, J. C., & Clark, G. R. (1992). “Design development of a tandem breakwater system for

663 Hammond Indiana”. Coastal structures and breakwaters: ICE London, 6&8 November

664 1991 (pp. 111-121). Thomas Telford Publishing.

665 Cuomo, G., Allsop, W., Bruce, T., & Pearson, J. (2010). “Breaking wave loads at vertical

666 seawalls and breakwaters”. Coastal Engineering, 57(4), 424-439.

667 Dattatri, J., Sankar, N.J., Raman, H., 1978. “Performance characteristics of submerged

668 breakwaters.” Proceedings of the 16th Coastal Engineering Conference, ASCE, (130),

669 2153–2171.

670 De Groot MB, Yamazaki H, Van Gent MR, Kheyruri Z. (1995). “Pore pressures in rubble mound

671 breakwaters”. In Coastal Engineering (pp. 1727-1738).

672 Dhinakaran, G., Sundar, V., Sundaravadivelu, R., & Graw, K. U. (2002). “Dynamic pressures

673 and forces exerted on impermeable and seaside perforated semicircular breakwaters due

674 to regular waves”. Ocean Engineering, 29(15), 1981-2004.

38
675 Dhinakaran, G., Sundar, V., Sundaravadivelu, R., & Graw, K. U. (2009). “Effect of perforations

676 and rubble mound height on wave transformation characteristics of surface piercing

677 semicircular breakwaters”. Ocean Engineering, 36(15-16), 1182-1198.

678 Firth, L. B., Thompson, R. C., Bohn, K., Abbiati, M., Airoldi, L., Bouma, T. J., … Hawkins, S. J.

679 2014. “Between a rock and a hard place: Environmental and engineering considerations

680 when designing coastal defence structures”. Coastal Eng., 87, 122–135.

681 Goda, Y., and Suzuki, Y., 1976. “Estimation of Incident and Reflected Waves in Random wave

682 Experiments”. Proceedings of 15th Coastal Engineering Conference, 1, 828-845.

683 Hall KR, Seabrook SR. (1998) “Design Equation for Transmission at Submerged Rubblemound

684 Breakwaters.” Journal of Coastal Research. Apr 1:102-6.

685 J.P. Holman, W.J. Gajda Jr. (1989). “Experimental methods for engineers” McGraw-Hill

686 International Editions, New York.

687 Hughes, S. A. (1993). ”Physical models and laboratory techniques in coastal engineering” (Vol.

688 7). World Scientific.

689 Hull, P., & Müller, G. (2002). “An investigation of breaker heights, shapes and

690 pressures”. Ocean Engineering, 29(1), 59-79.

691 Jeng, D. S., Schacht, C., & Lemckert, C. (2005). “Experimental study on ocean waves

692 propagating over a submerged breakwater in front of a vertical seawall”. Ocean

693 Engineering, 32(17-18), 2231-2240.

694 Jensen B, Christensen ED, Sumer BM. (2014) “Pressure-induced forces and shear stresses on

695 rubble mound breakwater armour layers in regular waves”. Coastal Engineering. Sep

696 1;91:60-75.

39
697 Johnson, J.W., Fuchs, R.A. and Morison, J.R., 1951. “The damping action of submerged

698 breakwaters”. Trans. Am. Geophys. Union, 32(5): 704-718.

699 Koraim, A. S., Heikal, E. M., & Zaid, A. A. (2014). “Hydrodynamic characteristics of porous

700 seawall protected by submerged breakwater”. Applied Ocean Research, 46, 1-14.

701 Krishnakumar, C., Sundar, V., & Sannasiraj, S. A. 2009. “Hydrodynamic performance of single-

702 and double-wave screens”. J. Waterw. Port Coastal Ocean Eng. 136(1), 59-65.

703 Lan, Y. J., & Lee, J. F., (2010). “On waves propagating over a submerged poro-elastic

704 structure”. Ocean Engineering, 37(8-9), 705-717.

705 Madrigal, B. G., & Prud'Homme, J. O. (1991). “Reduction of wave forces and overtopping by

706 submerged structures in front of a vertical breakwater”. In Coastal Engineering 1990 (pp.

707 1348-1361).

708 Mei, C. C., & Black, J. L. (1969). “Scattering of surface waves by rectangular obstacles in waters

709 of finite depth”. Journal of Fluid Mechanics, 38(3), 499-511.

710 Murali. K., Mani, J.S., 1996. “Performance Analysis of a Cage Floating Breakwater”. PhD

711 Thesis, Dept. of Ocean Engg., IIT Madras, India.

712 Muttray, M., Oumeraci, H., (2005). “Theoretical and experimental study on wave damping

713 inside a rubble mound breakwater”. Coast. Eng. 52, 709–725

714 Oumeraci H, Partenscky HW. (1990) “Wave-induced pore pressure in rubble mound

715 breakwaters”. In Coastal Engineering (pp. 1334-1347).

716 Pilarczyk, K. W. 2003. “Design of low-crested submerged structures—An overview.” 6th Int.

717 Conf. on Coastal and Port Engineering in Developing Countries, Pianc-Copedec,

718 Colombo, Sri-Lanka, 1–16.

40
719 Taveira-Pinto, F., & Neves, A. C. (2006). “Dynamic Pressure Evaluation over Submerged

720 Breakwaters Slopes”. Proceedings of AFM, 203-213.

721 Reddy, M. M., & Neelamani, S. (2005). “Hydrodynamic studies on vertical seawall defenced by

722 low-crested breakwater”. Ocean Engineering, 32(5-6), 747-764.

723 Sasikumar, A., Kamath, A., Musch, O., Lothe, A. E., & Bihs, H. (2018). “Numerical Study on

724 the Effect of a Submerged Breakwater Seaward of an Existing Breakwater for Climate

725 Change Adaptation”. In ASME 2018 37th International Conference on Ocean, Offshore

726 and Arctic Engineering (pp. V07AT06A027-V07AT06A027). American Society of

727 Mechanical Engineers.

728 Seabrook, S., & Hall, K. 1998. “Wave transmission at submerged rubble-mound breakwaters,”

729 Proceedings 26th Coastal Engineering Conference, ASCE, 2000 - 2013.

730 Srineash, V. K., & Murali, K., 2015a. “Hydrodynamic Performance of Gabion Box Artificial

731 Reefs”, Proceedings of the 25th International Ocean and Polar Engineering Conference

732 (ISOPE), Hawaii, USA, 1438-1442.

733 Srineash, V. K., & Murali, K., 2015b. “Pressures on Gabion Boxes as Artificial Reef Units”,

734 Procedia Engineering, 116(1), 552–559.

735 Srineash, V. K., & Murali, K., 2018. “Wave Shoaling over a Submerged Ramp: An Experimental

736 and Numerical Study” J. Waterw. Port, Coast. Ocean Eng., 144 (2), 04017048-1-12.

737 Srineash V. K., Murali K 2019. “Functional performance of modular porous reef breakwaters”

738 Journal of Hydro-environment research, 27, 20-31.

739 Srineash V. K., Kamath A., Murali K., Bihs H., 2020. “Numerical Simulation of Wave

740 Interaction with Submerged Porous Structure and application for coastal resilience”

741 (Accepted for publication) Journal of Coastal Research.

41
742 Sumer, B.M., Sen, M.B., Karagali, I., Ceren, B., Fredsøe, J., Sottile, M., Zilioli, L., Fuhrman,

743 D.R., (2011). “Flow and sediment transport induced by a plunging solitary wave”. J.

744 Geophys. Res. 116, 1–15.

745 Sumer, B.M., Guner, H.A.A., Hansen, N.M., Fuhrman, D.R., Fredsøe, J., (2013). “Laboratory

746 ob- servations of flow and sediment transport induced by plunging regular waves”. J.

747 Geophys. Res. Oceans 118, 6161–6182.

748 Sundar, V., & Ragu, V. (1998). “Dynamic pressures and run-up on semicircular breakwaters due

749 to random waves”. Ocean engineering, 25(2-3), 221-241.

750 Sundar, V. & Subba Rao, B. V. V., (2002). “Hydrodynamic pressures and forces on quadrant

751 front face pile supported breakwater”. Ocean engineering, 29(2), 193-214.

752 Tanaka, N. 1976. “Effects of submerged rubble-mound breakwater on wave attenuation and

753 shoreline stabilization.” Proc., Japanese Coastal Engineering Conf., JSCE, Tokyo, 152–

754 157 (in Japanese).

755 Thomas, J. L., 1986. “Use of Gabions in Coastal Environment”, Technical Report CETN-III-31,

756 U.S. Army Corps of Engineers, Waterways Experiment Station, Vicksburg.

757 Van der Meer, J.W., Briganti, R., Zanuttigh, B., Wang, B., (2005). “Wave transmission and

758 reflection at low crested structures: design formulae, oblique wave attack and spectral

759 change”. Coastal Eng., 52.

760 Wamsley, T. V., Kraus, N. C., & Hanson, H. (2002). “Wave transmission at detached

761 breakwaters for shoreline response modeling”. Technical Report, U.S. Army Corps of

762 Engineers, Waterways Experiment Station, Vicksburg.

763 Young, D. M., & Testik, F. Y. (2011). “Wave reflection by submerged vertical and semicircular

764 breakwaters”. Ocean Engineering, 38(10), 1269-1276.

42
765 FIGURE CAPTION LIST:

766 Fig. 1. Schematic view of the experimental setup (Srineash and Murali, 2019).

767 Fig. 2. Typical time series of pressure measurements and its corresponding spectral densities at

768 various locations of the reef breakwater

769 Fig. 3. Effect of relative water depth (d/L) on nondimensional crest pressures

770 Fig. 4. Effect of relative water depth (d/L) on crest - trough pressure ratio (pc/pt)

771 Fig. 5. Effect of relative water depth (d/L) on crest pressure ratio

772 Fig. 6. Effect of relative depth of submergence (d’/H) on nondimensional crest pressure of p1

773 Fig. 7. Effect of relative depth of submergence (d’/H) on nondimensional crest pressure of p2

774 Fig. 8. Effect of relative depth of submergence (d’/H) on nondimensional crest pressure of p3

775 Fig. 9. Effect of relative depth of submergence (d’/H) on nondimensional crest pressure of p4

776 Fig. 10 Comparison of present pressure measurements with earlier researchers

777 Fig. 11 Measured versus predicted wave-induced pressures

778 TABLE CAPTION LIST:

779 Table 1 Range of variables considered in the study

780 Table 2 Range of nondimensional parameters considered in the study

781 Table 3 Variation in the harmonics of the pressure measurements for d/L =0.12 [d’/H=1.28,

782 d’/d=0.34 and B/d=2]

783 Table 4 Uncertainty analysis

43

View publication stats

You might also like