Next Article in Journal
Artificial Intelligence in Ship Trajectory Prediction
Previous Article in Journal
Modes of Weather System-Induced Flows through an Arctic Lagoon
Previous Article in Special Issue
Optimizing Smart Energy Infrastructure in Smart Ports: A Systematic Scoping Review of Carbon Footprint Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimizing Infragravity Wave Attenuation to Improve Coral Reef Restoration Design for Coastal Defense

by
Benjamin K. Norris
1,*,
Curt D. Storlazzi
2,
Andrew W. M. Pomeroy
3 and
Borja G. Reguero
1
1
Institute of Marine Sciences, University of California, Santa Cruz, 115 McAllister Way, Santa Cruz, CA 95060, USA
2
US Geological Survey, Pacific Coastal and Marine Science Center, 2885 Mission Street, Santa Cruz, CA 95060, USA
3
School of BioSciences, Biosciences 4, The University of Melbourne, Royal Parade, Parkville, VIC 3052, Australia
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2024, 12(5), 768; https://doi.org/10.3390/jmse12050768
Submission received: 10 February 2024 / Revised: 26 April 2024 / Accepted: 30 April 2024 / Published: 1 May 2024
(This article belongs to the Special Issue Coastal Engineering: Sustainability and New Technologies, 2nd Edition)

Abstract

:
Coral reefs are effective natural flood barriers that protect adjacent coastal communities. As the need to adapt to rising sea levels, storms, and environmental changes increases, reef restoration may be one of the best tools available to mitigate coastal flooding along tropical coastlines, now and in the future. Reefs act as a barrier to incoming short-wave energy but can amplify low-frequency infragravity waves that, in turn, drive coastal flooding along low-lying tropical coastlines. Here, we investigate whether the spacing of reef restoration elements can be optimized to maximize infragravity wave energy dissipation while minimizing the number of elements—a key factor in the cost of a restoration project. With this goal, we model the hydrodynamics of infragravity wave dissipation over a coral restoration or artificial reef, represented by a canopy of idealized hemispherical roughness elements, using a three-dimensional Navier–Stokes equations solver (OpenFOAM). The results demonstrate that denser canopies of restoration elements produce greater wave dissipation under larger waves with longer periods. Wave dissipation is also frequency-dependent: dense canopies remove wave energy at the predominant wave frequency, whereas sparse canopies remove energy at higher frequencies, and hence are less efficient. We also identify an inflection point in the canopy density–energy dissipation curve that balances optimal energy losses with a minimum number of canopy elements. Through this work, we show that there are an ideal number of restoration elements per across-shore meter of coral reef flat that can be installed to dissipate infragravity wave energy for given incident heights and periods. These results have implications for designing coral reef restoration projects on reef flats that are effective both from a coastal defense and costing standpoint.

1. Introduction

Coastal hazards present a significant risk to their adjacent communities. Along tropical coastlines, coral reefs protect against flooding and erosion by efficiently dissipating incoming wave energy before it reaches shore [1,2,3,4,5,6,7,8,9]. Globally, thousands of reef-lined tropical coastlines are threatened by pressures such as climate change, sea-level rise, and coral degradation [10]. Most of these coastlines have critical infrastructure, housing, and agriculture within a few meters of mean sea level [11] and are vulnerable to increased flood risk posed by coral degradation offshore [12,13,14,15]. As reefs become increasingly impacted, there is growing concern over whether degraded coral reefs can maintain similar coastal protection benefits under present and future sea level scenarios [8,9,16].
Recent work suggests that maintaining reef structural complexity through restoration may be the best solution to mitigate present and future coastal hazards [17,18,19,20]. Structural reef reconstruction often involves the placement of individual rock piles or semi-porous concrete structures to provide the hard substrate onto which coral can be grown [21]. These individual elements could be arranged into submerged breakwaters to provide additional coastal defense benefits [18]. Although pilot reef restoration projects for coastal defense are still relatively scarce [22,23], limited evidence suggests this strategy is effective in reducing hazards [24]. Further, numerical modeling has identified where along the coastline, and where across the reef system, restoration projects could be most valuable. Along coastal Florida and Puerto Rico, Storlazzi et al. [20] determined that the greatest hazard risk reduction could be achieved through shallower restorations that were closer to shore. Roelvink et al. [19] analyzed over 30,000 global coral reef profiles to quantify how the across-shore placement of coral restorations modifies wave-driven runup. For the most common reef typologies, fringing and atoll reefs, the greatest flood-risk reduction could be achieved with coral restorations on the upper fore reef and middle reef flat, due to enhanced wave breaking and frictional dissipation, respectively. Indeed, many studies focusing on resolving modeled wave dynamics at large scales (i.e., across coral reefs) need to parameterize roughness (e.g., [19,25,26]). Similar studies that resolve the effects of waves on small-scale (i.e., coral colony) features are much less common as the computational effort to include such features scales with the resolved resolution of roughness elements.
The coastal defense benefit of coral reefs arises from two sources of wave height reduction: wave breaking and bottom friction. Most coral reefs have an offshore reef crest that is only a few meters deep, where the rapid change in bathymetry causes most (90–99%) incident sea-swell waves (“SS” > 0.04 Hz) to break, depending on the tidal stage [5,27]. However, not all SS wave energy is dissipated, and some are additionally transformed into infragravity (“IG” 0.004–0.04 Hz) and very-low-frequency (“VLF”, 0.001–0.004 Hz) waves through non-linear wave–wave interactions and breakpoint forcing on the fore-reef slope [28,29,30]. This process creates a mixture of short- and long-period waves that propagate shoreward across the reef flat [26,31], where gradually the remaining SS waves, and to a lesser extent IG waves, are attenuated by frictional dissipation [32]. Hence, unbroken IG waves tend to dominate the wave energy spectrum on the inner reef flat [26,29], and as a result, are often the primary contributor to wave runup and coastal flooding [3,9,33]. Therefore, one strategy to diminish coastal flooding in reef environments is to enhance the bottom roughness of the reef flat to dissipate IG wave energy before it reaches shore.
Although the coastal protection benefits of coral reefs have been the subject of several recent studies [8,9,19,20,34] one limitation has been how the bottom roughness created by reef-building organisms is described and quantified. For field studies, the geometry of large bottom roughness is hard to resolve, and hydrodynamic observations are typically sparse enough to necessitate the quantification of reef-aided coastal defense through wave energy flux gradients [29,31,35,36,37]. Similarly, laboratory and numerical modeling studies focused on reef-scale hydrodynamics may parameterize roughness with bulk drag coefficients [19,26,38] or neglect bottom roughness altogether [8,39,40]. As a result, these larger-scale studies often cannot resolve the dynamics of wave-driven turbulent flows within reefs, a process which drives many important physical and biological functions that are critical for maintaining healthy corals (Davis et al. [41] and references therein) and likely reef restorations. Recent advances in computational fluid dynamics (CFD) models have enabled the simulation of these complex flow patterns within reefs (e.g., [42,43,44,45]), though the focus of these studies was not for restoration.
Here, we focus on characterizing the small-scale (i.e., coral colony) hydrodynamics to optimize coral restoration designs for coastal protection. Specifically, we investigate whether there is an optimal balance between maximizing wave energy dissipation (a benefit of the restoration) with a minimal number of restoration elements (individual coral plantations). Using a three-dimensional Navier–Stokes equations solver to simulate cnoidal waves with IG periods over coral reef flats with idealized bottom roughness, we study different coral reef restoration strategies by characterizing how wave dissipation depends on wave conditions and the spacing of reef roughness elements (e.g., coral heads or artificial reef elements). Although restoration project cost is not the primary focus of this study, we additionally use the model results to determine a restoration design that maximizes wave energy dissipation while minimizing the number of restoration elements, an optimization that represents the most cost-effective design for a given set of wave conditions.

2. Materials and Methods

2.1. Relevant Hydrodynamic Parameters

Across relatively horizontal reef flats, the process of wave attenuation by bottom friction is predominantly controlled by the physical size and density of coral colonies. At the scale of individual corals, wave velocity attenuation varies inversely with the ratio of the wave orbital excursion, a length scale describing the horizontal displacement of water particles beneath successive waves, to the length scale of bed roughness, A b / k w [46,47,48,49]. Collectively, assemblages of corals over which wave and current energy is dissipated are referred to as “canopies”. The term k w is a canopy property and is commonly defined as either the circular diameter (d) of individual canopy elements (e.g., [45,50]) or as the center-to-center distance between adjacent canopy elements, Δ S . For circular objects, Δ S is related to the number of canopy elements per across-shore meter ( n ) by n d = d / Δ S 2 , and so Δ S = n 1 / 2 (e.g., [51]), and can be normalized by d to yield a non-dimensional roughness spacing D = Δ S / d .
Across a rough bottom, drag forces at the bed strip energy from wave flows, converting velocity into turbulence. To describe this process, Lowe et al. [46] defined a canopy attenuation parameter α w as the ratio of in-canopy to free stream wave velocities. Lowe et al. [46] demonstrated that A b is the single relevant wave parameter affecting flow in the canopy and identified three distinct flow regimes: (1) when canopy density is sufficiently sparse (i.e., when A b / k w ≤ 1), in-canopy velocity reduction is small ( α w is O(1)), flows are inertia-dominated and are not greatly affected by canopy elements; (2) a quasi-unidirectional limit when A b / k w > 100, in-canopy flows are severely reduced by drag ( α w is O(0.1)); and (3) a final regime where canopy drag, inertial, and shear forces all contribute to the momentum balance inside the canopy, leading to intermediate velocity attenuation. Wave energy dissipation, often parameterized with a friction factor fe, scales inversely with α w and hence decreases with proportional increases in A b relative to the physical roughness k w [52]. Here, we calculate A b and α w from the modeled free surface and velocity time series, respectively (see Section 2.4) and estimate k w as Δ S for each model.

2.2. Experimental Design

Experiments were performed with a series of numerical simulations using the open-source computational fluid dynamics modeling framework OpenFOAM, version-1912 [53]. Model scenarios were based on a range of values from observational and modeling studies of IG waves on reef flats reported in the literature (Table 1). Model cases were defined by combinations of incident IG wave heights Hs, incident IG wave periods Tw, the non-dimensional ratio of wave height to water depth ( γ ), and the ratio of wave height to wavelength (wave steepness; H s / λ ). In general, waves break on coral reef flats when γ = 0.4–0.6 [17] or when H s / λ = 0.142 [54], but the non-breaking IG waves considered here were well below both of these limits.
The model domains were designed as two-dimensional reef flat profiles ranging in across-shore lengths of 144, 235, and 564 m, with a height of 2 m, the still water line centered at zero, and a constant depth of 1 m. Model domain lengths varied with the IG wavelength (λ = 94, 157, and 376 m) to ensure an entire wave was captured in each domain. The domains began with a region equal to one-half wavelength with a smooth bottom to initialize conditions, followed by a region equal to one whole wavelength containing the roughness elements, which were repeated at regular intervals from x = x0 to x = Lm (Figure 1), where x is the across-model length in meters and Lm is the model domain length. Bed roughness was modeled with regular arrays of hemispherical objects with a constant diameter d = 0.4 m and height hc = 0.2 m. Hemispheres were chosen for simplicity as these approximate the shape of robust corals such as Porities, Diploastrea, Orbicella, Montastraea, and Siderastrea genera that are increasingly being outplanted [59] as well as artificial structural reef elements such as the ReefBallTM. The roughness density in the model domains varied with the across-shore distance between adjacent elements, D (Figure 1e). The final model scenarios combined one of six roughness densities: D = 2, 3, 4, 8, 16, and 32; with three incident wave heights: Hs = 0.05, 0.15, and 0.3 m; and three incident wave periods: Tw = 30, 60, and 120 s; for a total of 54 scenarios. These combinations of model scenarios simulated correspond to offshore wave heights H0 of 1–5 m, offshore wave periods T0 of 4–25 s, and reef flat widths Wf of 100–500 m (Table 2) based on the BEWARE database [8].
Despite the apparent similarity of the experimental design with studies of Bragg reflection (e.g., [60]), we note that the term 2ΔS/L is between 0.004 and 0.273 for the conditions considered in this study, and hence Bragg reflection (which is maximum at 2ΔS/L = 1) is not one of the dominant processes studied here.

2.3. Numerical Model Set-Up

All models described herein used the same set-up and configuration as the calibrated and validated models described in Norris et al. [61]. Below we provide a brief description of the model set-up, but please refer to Norris et al. [61] for further details.
All model domains were quasi-two-dimensional (i.e., varying dimensions along the x and z axes with the y-axis defined as a single cell layer). Initial tests for grid cell independence were determined by first running a series of empty tank wave models without turbulence and with increasing grid resolution for each set of wave conditions to determine the resolution at which (i) wave heights did not decay with distance along the x-axis of the numerical wave flume, and (ii) the space-averaged (i.e., between x0 and Lm; Figure 1) water velocity profiles converged. After choosing an appropriate resolution, the same exercise was repeated with turbulence activated to check the space-averaged turbulence profiles. The finest grid resolution was chosen for all wave conditions and was conservative for the largest waves. Hence, all domains started with a constant (x = z) mesh size of 0.2 m. Then, 3D STL models of the roughness elements (hemispheres) for each model domain were created in Blender, and were placed inside the model domains to be ‘snapped’ to the bottom of the domains using the snappyHexMesh utility in OpenFOAM. Mesh refinement was conducted in three levels down to 0.025 m (Figure 1e). At the bottom of the models, this mesh scaling corresponded to dimensionless wall distances (z+) of 50–150, where 30 < z+ < 300 defines the log-law layer where wall functions are applicable.
All models employed the following boundary conditions. The model inlet and outlet were set to waveVelocity for wave generation and absorption, respectively. The bottom wall and roughness elements used the zeroGradient condition, and the top of the model the inletOutlet condition. At the model inlet, outlet, and along the bottom, the fixedFluxPressure condition was applied to the hydrostatic field (pressure) to adjust the pressure gradient so that the boundary flux matched the velocity boundary condition. Active absorption was turned on at the inlet and outlet to reduce wave reflections inside the model domains. Turbulence was simulated using the k-ε model using a RANS approach. Since two-equation turbulence models are known to cause over-predicted turbulence levels beneath waves in numerical wave flumes, we stabilized the models using the method developed by Larsen and Fuhrman [62] using their default parameter values for the k-ε closure scheme.
Model turbulence parameters used respective wall functions to model the boundary layer profile near the bed. Initial values of the turbulence fields were specified using standard equations with Cµ = 0.09, turbulent intensity of 5%, depth-averaged velocity from the empty wave tank models, and turbulent length scale set by the still water depth above the roughness elements. Cnoidal waves with IG frequencies were generated in the models using the IHFOAM toolbox [63]. All models were sampled at 12 Hz.
Sampling was conducted during model runtime for the area encompassing the free surface to the bed within the region of the model domain containing the roughness elements. Data analysis was conducted on the last ten waves in each scenario. Longer time series were also investigated, but initial testing revealed little difference in the results between these and shorter time series, due to the simulated waves being monochromatic. Given that simulation times varied between 2.5 to 37.5 h, depending on the model domain length and wave period, preference was given for shorter runtimes that achieved the same result as longer runtimes.

2.4. Model Data Analysis

The significant IG wave height, Hs was computed from the free surface time series of each model simulation as:
H s ( x ) = 4 S η ( x ) d f
where S η ( x ) is the water surface elevation spectrum, which was estimated with Welch’s method and a Hamming window, and d f is the step-in frequency. Estimates of Hs were computed at the positions x0 and Lm (Figure 1d) to assess the change in wave height across each of the models.
To compare wave energy dissipation within the models, the change in the one-dimensional wave energy flux for shore-normal waves was computed as:
F j x = ε b , j ε f , j
where F j is the wave energy flux across the model domain, ε b , j is the wave dissipation due to wave breaking, ε f , j is wave dissipation due to bottom friction, and the subscript j indicates the jth frequency component of a spectral wave distribution [31]. In general, Fj at any across-shore location is given by F j = E j C g , j , where E j is the wave energy and C g , j is the group velocity for a given spectral frequency. Here, we modeled only IG wave frequencies, as SS waves are often a negligible component of the total wave energy on reef flats far from the reef crest [29]. As none of the conditions modeled here induced wave breaking, Equation (2) can be simplified to:
F j x = ε f , j = 1 4 ρ f e , j u b , r u b , j 2
where f e , j is the frequency-dependent energy dissipation factor, u b , r is a representative near-bottom wave orbital velocity, given by:
u b , r = j = 1 N u b , j 2
and u b , j is the velocity corresponding to the jth frequency component [31]. The mean wave orbital excursion A b was estimated from the mean value of u b , r and T w as A b = u b , r T w / 2 π for each model scenario. u b , j was estimated from the water surface elevation spectrum ( S η , j ) using linear wave theory:
u b , j = a j ω j sinh k j h
where k j is the wavenumber, ω j the wave radian frequency ( = 2 π f ), and the wave amplitude is a j   = 2 S η , j Δ f , with Δ f the spectral bandwidth. In Section 3.2, F j is presented at a single incident wave height for three incident wave periods and two across-shore element spacings. Equations (3)–(5) were used to estimate f e , j to assess frequency-dependent wave dissipation across the model domains from x0 to Lm.
To compare wave-frequency velocity attenuation within the canopy layer of the models, we define the wave attenuation parameter per unit frequency α w , j as
α w , j = S ^ U , j S U , j 1 / 2
where S ^ U , j is a depth-averaged velocity spectrum assessed within the canopy layer near the bed in the models, and S U , j is a depth-averaged velocity spectrum assessed just below the free surface [46]. To estimate α w , j , we used depth-averaged velocities from four vertical cells at each position: within the canopy layer (z = 0.8–0.90 m) and below the free surface (z = 0.1–0.2 m). Spectra were calculated with a Hamming window and 50% overlap to yield approximately 20 DOF.
The total energy dissipation in the models was assessed using the TKE dissipation rate (ε) calculated by OpenFOAM in each model grid cell. To simplify the model results, ε was binned along the x-z plane into 26 vertical (z) bins and between 1008 and 3984 horizontal (x) bins to keep the horizontal bin size consistent between model domains with different lengths. The binned estimates of ε were then time-averaged to produce a representation of the mean conditions for each model scenario. ε results are presented both in terms of depth-averaged with respect to the across-shore distance (x) in the models x/Lm, space-averaged with respect to depth (z), and depth-and-space averaged to compare with the number of elements per meter, n.
Finally, model performance in wave height reduction and energy dissipation is quantified as the percent change in these parameters across the models for each scenario, and was estimated as P e r c e n t   C h a n g e = β L m β x 0 β x 0 · 100 , where β is Hs or depth-averaged ε, β x 0 is the parameter value at the beginning of the bottom roughness, and β L m the value at the end. Greater negative values represent a larger percent change in each parameter for each model scenario.

3. Results

3.1. Infragravity Wave Attenuation

As IG waves propagate across the model domains, they are gradually attenuated as near-bed wave velocities are converted into turbulence. Raw output from the models demonstrates this conversion of wave energy above and within the hemisphere canopy (Figure 2). The greatest velocities occur under the wave crests, with larger accelerations along the top of the roughness elements (Figure 2a). Velocity shadows in the lee of the elements are indicative of eddying that drives turbulent energy dissipation. Accordingly, greater turbulence is concentrated beneath the wave crest where velocities are intensified, although high turbulence can also be observed downstream of the wave crest at the top and in between the roughness elements (Figure 2b).
The percent change in Hs across the models depends on the incident IG wave height H s 0 , the wave period Tw, and element spacing D, with greater attenuation for shorter element spacings, larger waves, and longer wave periods (Figure 3). For any Tw, the difference between the D = 2 and 32 roughness element spacings also grows with increasing H s 0 .
Changes in wave energy flux across the model domains are dependent on incident wave conditions. In general, the wave energy flux spectra (Fj) decrease from the beginning (x0) to the end (Lm) of the roughness, with increasing Tw and D at constant H s 0 (Figure 4). For the D = 2, Tw = 30 s case (Figure 4a), the decrease in wave energy occurs mainly between the incident frequency f = 0.035 and 0.23 Hz, and wave energy above f > 0.23 Hz shows little decrease from x0 to Lm. In contrast, wave energy in the D = 32, Tw = 30 s case (Figure 4b) is dissipated mainly at higher frequencies, f = 0.08 and 0.33 Hz, with more energy remaining at Lm at the incident frequency. A similar pattern repeats for the Tw = 60 and 120 s models (Figure 4c–f). Thus, dense bed roughness more efficiently removes energy from the waves as it acts on both the incident and higher frequencies, regardless of the incident wave period. Sparse bed roughness is less effective at removing wave energy as it acts mainly on higher frequencies.
Indeed, spectra of the energy dissipation factor fe,j (Equation (3)) for the same cases shown in Figure 4a,b demonstrates that the majority of wave dissipation in the D = 2 case occurs at approximately the predominant frequency fj = 0.035 Hz (T = 29 s), but occurs at higher frequencies fj = 0.13 Hz (T = 8 s) in the D = 32 case (Figure 5). Similarly, spectral decomposition of near-surface S U and in-canopy velocities S ^ U for the same cases indicate that in-canopy velocities are severely reduced at the predominant frequency in the D = 2 case (Figure 6a) yet are minimally reduced in the D = 32 case (Figure 6b). Consequently, α w , j is lowest close to the predominant frequency for the D = 2 case (Figure 6c) and remains α w , j ~1 across almost all frequencies for the D = 32 case (Figure 6d). These results indicate that the D = 2 model efficiently attenuates wave velocities at the predominant frequency, whereas the D = 32 model is much less efficient at dissipating energy at the predominant frequency.

3.2. Dependence of Turbulence on Bottom Roughness

Normalized across-shore profiles (x/Lm) of the depth-averaged TKE dissipation rate ( ε ) for every model scenario demonstrate that ε decreases with increasing Tw and D, and increases with increasing Hs0 (Figure 7). The decline of ε in the across-shore direction is generally monotonic for almost all models. However, models with dense element spacings (i.e., D = 2 to 4) have an inflection point around half the model domain length (x/Lm = 0.5) where the rate of change in ε decreases (for example, this pattern is especially apparent in the D = 2 case with Hs0 = 0.30, Tw = 120 s, Figure 7i). However, this inflection point becomes less obvious as the element spacing increases. One possible explanation for this behavior is that a wave will encounter more elements when they are spaced closer together, resulting in greater initial energy dissipation that weakens with cross-shore distance. Interestingly, although the Tw = 30 s models also exhibit a decline in energy dissipation across the models for shorter element spacings, there is a nearly constant dissipation with cross-shore distance for larger element spacings (e.g., compare D = 2 to D = 32 in Figure 7c).
Time-and-space averaged depth profiles of ε (Figure 8) reveal increasing variance in the vertical distribution of turbulence under increasing wave heights and periods. Under the smallest IG waves (Hs0 = 0.05 m), ε is enhanced only for the D = 2 cases, with peak ε occurring below the top of the roughness elements around z = 0.9 m depth. Similar to the across-shore profiles (Figure 7), this peak ε tends to decrease with increasing Tw at constant Hs0, from a maximum of 9 × 10−4 W/m2 under Tw = 30 s waves (Figure 8a) to a maximum of 1 × 10−4 W/m2 under Tw = 120 s waves (Figure 8g). As the incident wave height increases, enhanced ε also occurs in the D = 3 and D = 4 cases, with a greater distribution of ε higher in the water column above the roughness elements (Figure 8b,e,f). Finally, under the largest wave conditions, enhanced ε within and above the canopy layer is observed for the D = 2 to 8 cases, particularly for the Tw = 120 s models (Figure 8c,f,i).
Like the other turbulence metrics (Figure 7 and Figure 8), the depth-and-space averaged value of ε for each model scenario increases with Hs0, and decreases with Tw, n, and D (Figure 9). There is an inflection point on the curves of n and ε , where ε is nearly constant for further reductions in n (Figure 9). This inflection point occurs at much larger n for small incident waves (Hs0 = 0.05) and at smaller n as the incident wave height increases. Critically, these inflection points represent the maximum energy dissipation attained for a minimum number of restoration elements. This result also reveals an equivalency between different roughness densities under varying wave conditions: smaller incident waves passing over a denser canopy of restoration elements can produce similar levels of energy dissipation as larger waves passing over a sparser canopy of elements.
However, increasing incident wave conditions does not always result in greater energy dissipation, as identified by the percent change in ε (Figure 10). Increases in Hs0 from 0.05 to 0.15 m result in a greater change in energy dissipation, which also negatively scales with n (and D), and positively with Tw. Further increases in Hs0, from 0.15 to 0.30 m, result in a smaller change in energy dissipation, with minimal improvements for only the densest canopy spacings (see D = 2 through D = 4 in Figure 10b,c). This result indicates there are diminishing returns in energy dissipation with respect to wave conditions and canopy density.

4. Discussion

This study considered IG wave dissipation across a schematic reef flat with varying, uniform across-shore spacings of structural reef elements. We simulated coral reef roughness using idealized hemispherical forms, using a similar approach as other physics-based investigations of wave interactions with large roughness elements (e.g., [44,45]). These forms are a simplification of real reef bathymetry but approximate the shape of certain coral genera, as well as commercial artificial reef elements. The model results demonstrated that IG wave energy dissipation is enhanced both in the horizontal dimension (across-shore) and the vertical dimension (within the water column) when incident waves are large and wave periods are small (Figure 7 and Figure 8). For smaller waves (Hm0 = 0.05 m in our experiments), frictional dissipation rolls off towards nearly a constant value as the roughness spacing drops below D = 3, and similar patterns were only observed for larger waves when the roughness spacing was considerably greater (Figure 9), reminiscent of results obtained by Buccino et al. [64] for their work on ReefBallTM-assisted wave attenuation.
With respect to coral reef restoration design, there are several important considerations with the findings in the present work: (1) wave dissipation is not constant across all frequencies and varies with canopy density; (2) there are diminishing returns in dissipation with increasing wave height and canopy density; and (3) an efficient restoration design is one that generates maximal wave energy dissipation using a minimum number of canopy elements. Not only will this strategy minimize the cost of a restoration project but may also be beneficial when wave heights above the restoration area increase during storm conditions, as explained below.

4.1. Wave-Frequency Dissipation

The results demonstrated that wave energy dissipation is not constant across all frequencies and depends on both the incident wave conditions and the roughness density. This topic has been the focus of several pivotal studies in wave-canopy flow dynamics [31,46,65,66]. Theory from Lowe et al. [46] predicts the wave attenuation parameter α w , j will be large when canopy density is small relative to the length scale of bed roughness, kw. Using linear theory to estimate the near-bottom velocity for the cases presented in Figure 5 yields A b on the order of 11 m for the D = 2 case, and 18 m for the D = 32 case. If we take k w as Δ S for these cases, then A b / k w > 1 for the D = 2 case and A b / k w < 1 for the D = 32 case. From this exercise, it follows that α w , j should be small for the D = 2 case, when velocity reduction is greatest, and large for the D = 32 case, when velocity reduction is smallest (Figure 6c,d). A similar pattern follows for fe,j (Figure 5), except that estimates presented for the D = 32 case may be larger than in actuality. When A b / k w ≤ 1, inertial forces dominate canopy flow [46]. Since the inertial force is in quadrature with the water velocity and does no work, friction factors determined with Equation (3) likely overestimate the true value in this case [45]. Regardless, this reinforces the idea that a dense canopy is much more efficient in removing wave energy at the dominant frequency than a sparse canopy.

4.2. Diminishing Returns in Wave Dissipation

Another interesting finding is that increasing the roughness density of the seabed through restoration may not always lead to further increases in wave energy dissipation. Although the greatest energy dissipation was observed for the densest canopies under the largest waves (Figure 9c), the percent change in dissipation diminished as the incident wave height increased (Figure 10). Hence, utilizing a lower coral density design could be more beneficial because it could greatly enhance energy dissipation as incident wave heights increase (as demonstrated between the Hs0 = 0.05 and 0.15 m cases in Figure 10).
This effect of diminishing returns in wave dissipation can be explained by a simple model for wave dissipation based on a standard drag coefficient model for canopy friction [65]. The damping of wave velocities within bottom roughness is governed by the dimensionless damping parameter Λ 0 = C D a u b T w / 4 π , where u b is the near-bottom velocity, CD is the canopy drag coefficient, and the hemisphere frontal area is a = 1 / 8 π d 2 . Theory predicts when Λ 0 = 1.4, in-canopy velocities are substantially reduced by drag and wave dissipation is maximized. Dissipation rolls off either when Λ 0 << 1, because in-canopy velocities are minimally reduced by drag forces, or when Λ 0 >> 1, because in-canopy velocities decline nearly to zero. For the two cases presented in Figure 5, we estimate CD (as the total drag coefficient calculated by OpenFOAM for each scenario) as 0.26 for the D = 2 case and 0.25 for the D = 32 case. These CD values are generally consistent with those estimated by Yu et al. [45] in a similar modeling study of idealized, hemispherical bottom roughness. For the two cases presented in Figure 5, we estimate Λ 0 = 1.766 and 0.006 for the D = 2 and D = 32 cases, respectively, reiterating that energy dissipation is more efficient within a denser canopy for the incident wave conditions (recalling these are Hs0 = 0.15 m and Tw = 30 s). This result also demonstrates that the D = 2 case is already close to maximum dissipation at Λ 0 = 1.4. A doubling of wave height to Hs0 = 0.30 m would yield Λ 0 = 2.854 and 0.011 for the D = 2 and 32 cases, indicating that dissipation would decrease within the D = 2 canopy, suggesting no increased benefit could be achieved with such dense element spacing given these conditions. The use of a more moderate spacing, such as the ‘inflection point spacing’ identified in Figure 9, would allow for enhanced dissipation to occur if wave conditions became more intense during storm conditions.

4.3. Cost Implications in Optimizing Reef Restoration Design

For a given set of wave conditions, there is an optimal spacing of roughness elements that balances wave energy dissipation with element spacing (Figure 9). The relationship between the percent change in energy dissipation and the number of roughness elements per across-shore meter n is inversely related, such that similar levels of wave energy dissipation can be achieved with larger n under smaller waves, or smaller n under larger waves (Figure 10). It is reasonable to assume that the cost of a potential restoration project also varies with n, and the cost is also proportional to incident wave conditions (Figure 11). However, many other factors affect the engineering design and cost of a restoration project [23] and additional work is necessary to determine the true shape of the ‘cost curve’.

4.4. Biological Implications in Optimizing Reef Restoration Design

In this paper, we demonstrated that the intensification of turbulence within and above a canopy of roughness elements depends both on the incident wave conditions as well as the canopy density. This interplay between restoration element spacing and wave energy dissipation likely has important biophysical implications that would affect the ability of a coral restoration to be biologically efficient and thus self-sustaining and self-propagating, along with providing coastal protection. As sessile, benthic organisms, corals depend on mass transfer from and to the overlying water to facilitate many biologically important processes, including mitigating thermal stresses via advective transport, nutrient exchange, coral heterotrophy, and larval dispersal and settlement [67]. Indeed, many biological processes vital to coral health depend on turbulence near the coral surface, as it drives in-reef circulation (see the recent review by Davis et al. [41]). For example, nutrient uptake in coral reefs is highly dependent on turbulent energy dissipation, with maximal uptake rates occurring in shallow systems with large bottom roughness elements [68]. Similarly, stronger flow may help to mitigate the effects of bleaching; Nakamura and Van Woesik [69] showed that coral bleaching at high water temperatures can be partially suppressed if turbulent boundary layer flows are strong. Besides the transfer of nutrients, turbulent mixing aids in reef heterotrophy and larval transport. Uptake rates of picoplankton by coral communities have been shown to increase relative to greater turbulent transport within coral canopies [70]. Through hydrodynamic modeling, Reidenbach et al. [71] demonstrated that the intensity of fluid forces on coral larvae depends on the degree of bathymetric roughness, with greater shear stresses for wider-spaced roughness than tightly spaced roughness. If shear stresses at the seabed are too intense, the larva can become detached from the benthos and may not be able to resettle.
Hence, there is likely an optimum spacing of roughness elements (coral communities) that will promote turbulent processes to enhance the biological processes of reefs. If roughness elements are too closely spaced, intensified turbulence may inhibit larval settlement. If elements are too far apart, turbulent fluxes would be minimal and diminish mass transfer between adjacent elements (e.g., Figure 7). Another factor of the element spacing is the distribution of turbulence throughout the water column, where closer element spacings would likely result in greater mass exchanges to and from the canopy (e.g., Figure 8). Thus, balancing turbulence by optimizing element spacing would be a key step in maintaining the productivity of coral restorations and artificial reefs to ensure their health and hence their coastal defense capabilities.

5. Conclusions

Along coral fringing and atoll reef-lined coasts, low-frequency infragravity waves are often the primary driver of coastal flooding. Since recent research indicates that coral restorations on shallow reef flats may be most effective for reducing wave-driven runup and coastal flooding, this study of infragravity waves over-idealized hemispherical forms on shallow reef flats illustrates that the dynamics of wave-driven flow over high-relief topography are strongly dependent both on the incident wave conditions and the canopy density. As waves propagate over rough bathymetry, near-bed turbulence strips energy from the flows, which ultimately reduces wave height. Greater wave energy dissipation occurs for greater canopy density, larger infragravity wave heights, and shorter wave periods. Wave dissipation is also frequency-dependent: dense canopies are efficient in removing energy from waves at the predominant frequency, whereas sparse canopies are inefficient. This difference occurs due to a greater reduction in in-canopy wave velocities in dense canopies compared to sparse canopies. Larger wave heights and longer wave periods also generally increase the vertical distribution of turbulence within the water column above the canopy layer, as well as peak turbulence inside the canopy layer.
Collapsing dissipation estimates into a single, representative mean value for each model scenario demonstrated an ‘inflection’ or ‘balance’ point where enhanced dissipation is balanced by a minimum number of elements per across shore meter. Such balance represents the optimum canopy density to dissipate a given set of incident wave conditions. Although turbulent energy dissipation can be enhanced by either increasing canopy density for given wave conditions, or the incident wave conditions for a given density, there is a limit where further increases in either parameter leads to lower dissipation. These results, therefore, have implications for optimizing reef restoration design as projects could be designed with the best balance between costs and performance in terms of infragravity wave attenuation.
Restoration managers can use the results presented in this paper to design restoration studies that maximize IG wave energy dissipation for a minimum number of coral outplantings. Table 2 can be leveraged to determine the reef flat width and offshore wave conditions from the BEWARE database, and then the expected change in IG wave energy dissipation for a set number of restorations per across-shore meter is presented in Figure 10. We advise using a moderate number of restorations per across-shore meter as this will have a positive influence on IG wave dissipation across reef flats, while still reducing the total cost of the restoration project.

Author Contributions

Conceptualization, B.K.N., C.D.S. and A.W.M.P.; methodology, B.K.N. and C.D.S.; software, B.K.N.; validation, B.K.N.; formal analysis, B.K.N., C.D.S. and A.W.M.P.; investigation, B.K.N.; resources, C.D.S.; data curation, B.K.N.; writing—original draft preparation, B.K.N.; writing—review and editing, B.K.N., C.D.S., A.W.M.P. and B.G.R.; visualization, B.K.N.; supervision, C.D.S.; project administration, B.K.N. and C.D.S.; funding acquisition, C.D.S. All authors have read and agreed to the published version of the manuscript.

Funding

This project was funded by the U.S. Geological Survey Coastal and Marine Hazards and Resources Program, the U.S. Geological Survey Mendenhall Research Fellowship Program, and the Australian-America Fulbright Commission. B.G.R. was partly supported by an Early-Career Research Fellowship from the Gulf Research Program of the National Academies of Sciences, Engineering, and Medicine, although the content is solely responsibility of the author and does not necessarily represent the official views of the Gulf Research Program of the National Academy of Sciences, Engineering and Medicine.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The OpenFOAM model data sets analyzed in this manuscript are available from USGS ScienceBase.

Acknowledgments

The authors would like to thank Mark Buckley (USGS) and Stephen Henderson (Washington State University) for their helpful advice and guidance, as well as Babak Teranirad (USGS) for their peer review of an earlier version of this manuscript. Any use of trade, firm, or product names is for descriptive purposes only and does not imply endorsement by the U.S. Government.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

Abwave orbital excursion
awave amplitude
Cgwave group velocity
CDcanopy drag coefficient
Droughness element spacing
droughness element diameter
Etotal wave energy
Fwave energy flux
fewave energy dissipation factor
ffrequency
ggravitational constant (=9.81 m/s2)
Hs0incident significant wave height
Hssignificant wave height
H0offshore significant wave height
hwater depth
hcroughness element height
IGinfragravity wave band
jthe j-th frequency component of a spectrum
kwavenumber
kwbottom roughness
Lmmodel wave flume length
nnumber of roughness elements per across shore meter
ra representative value
rmsroot-mean-square
SSsea-swell wave band
Sηwater surface elevation spectrum
S ^ U velocity spectrum within the roughness layer
SUvelocity spectrum above the roughness layer
Twwave period
T0offshore wave period
ubnear-bottom orbital velocity
Uvelocity magnitude
Wfreef flat width
xacross-shore distance in models
x0start of roughness elements in models
zdepth in models
z+wall distance function
αwcanopy attenuation parameter (Lowe et al., 2005b)
Δfspectral bandwidth
ΔScenter-to-center distance between adjacent roughness elements
Δ x model cell size in the direction of the velocity
εturbulent kinetic energy dissipation rate
εbwave dissipation due to breaking
εfwave dissipation due to bottom friction
γcritical wave breaking parameter
κ c interface curvature (of the free surface)
Λ0wave damping parameter (Henderson et al., 2017)
λwavelength
ωwave radian frequency
ρwater density (=1025 kg/m3)

References

  1. Baldock, T.E.; Golshani, A.; Callaghan, D.P.; Saunders, M.I.; Mumby, P.J. Impact of Sea-Level Rise and Coral Mortality on the Wave Dynamics and Wave Forces on Barrier Reefs. Mar. Pollut. Bull. 2014, 83, 155–164. [Google Scholar] [CrossRef] [PubMed]
  2. Buckley, M.L.; Lowe, R.J.; Hansen, J.E.; Van Dongeren, A.R.; Storlazzi, C.D. Mechanisms of Wave-Driven Water Level Variability on Reef-Fringed Coastlines. J. Geophys. Res. Ocean. 2018, 123, 3811–3831. [Google Scholar] [CrossRef]
  3. Cheriton, O.M.; Storlazzi, C.D.; Rosenberger, K.J. Observations of Wave Transformation over a Fringing Coral Reef and the Importance of Low-Frequency Waves and Offshore Water Levels to Runup, Overwash, and Coastal Flooding. J. Geophys. Res. Ocean. 2016, 121, 3121–3140. [Google Scholar] [CrossRef]
  4. Escudero, M.; Reguero, B.G.; Mendoza, E.; Secaira, F.; Silva, R. Coral Reef Geometry and Hydrodynamics in Beach Erosion Control in North Quintana Roo, Mexico. Front. Mar. Sci. 2021, 8, 684732. [Google Scholar] [CrossRef]
  5. Ferrario, F.; Beck, M.W.; Storlazzi, C.D.; Micheli, F.; Shepard, C.C.; Airoldi, L. The Effectiveness of Coral Reefs for Coastal Hazard Risk Reduction and Adaptation. Nat. Commun. 2014, 5, 3794. [Google Scholar] [CrossRef] [PubMed]
  6. Franklin, G.L.; Torres-Freyermuth, A. On the Runup Parameterization for Reef-Lined Coasts. Ocean Model. 2022, 169, 101929. [Google Scholar] [CrossRef]
  7. Gourlay, M.R. Wave Transformation on a Coral Reef. Coast. Eng. 1994, 23, 17–42. [Google Scholar] [CrossRef]
  8. Pearson, S.G.; Storlazzi, C.D.; van Dongeren, A.R.; Tissier, M.F.S.; Reniers, A.J.H.M. A Bayesian-Based System to Assess Wave-Driven Flooding Hazards on Coral Reef-Lined Coasts. J. Geophys. Res. Ocean. 2017, 122, 10099–10117. [Google Scholar] [CrossRef]
  9. Quataert, E.; Storlazzi, C.; Rooijen, A.; Cheriton, O.; Dongeren, A. The Influence of Coral Reefs and Climate Change on Wave-Driven Flooding of Tropical Coastlines. Geophys. Res. Lett. 2015, 42, 6407–6415. [Google Scholar] [CrossRef]
  10. Souter, D.; Planes, S.; Wicquart, J.; Logan, M.; Obura, D.; Staub, F. Status of Coral Reefs of the World: 2020; Global Coral Reef Monitoring network (GCRMN) and International Coral Reef Initiative (ICRI): Townsville, Australia, 2021; pp. 1–12. [Google Scholar]
  11. UN-Habitat. Urbanization and Climate Change in Small Island Developing States; UN-Habitat: Nairobi, Kenya, 2015. [Google Scholar]
  12. Beck, M.W.; Pfliegner, K.; Quast, O. Ecosystem-Based Adaptation and Insurance: Success, Challenges and Opportunities; InsuResilience Secretariat: Bonn, Germany, 2019. [Google Scholar]
  13. Perry, C.T.; Alvarez-Filip, L. Changing Geo-Ecological Functions of Coral Reefs in the Anthropocene. Funct. Ecol. 2019, 33, 976–988. [Google Scholar] [CrossRef]
  14. Reguero, B.G.; Storlazzi, C.D.; Gibbs, A.E.; Shope, J.B.; Cole, A.D.; Cumming, K.A.; Beck, M.W. The Value of US Coral Reefs for Flood Risk Reduction. Nat. Sustain. 2021, 4, 688–698. [Google Scholar] [CrossRef]
  15. Sheppard, C.; Dixon, D.J.; Gourlay, M.; Sheppard, A.; Payet, R. Coral Mortality Increases Wave Energy Reaching Shores Protected by Reef Flats: Examples from the Seychelles. Estuar. Coast. Shelf Sci. 2005, 64, 223–234. [Google Scholar] [CrossRef]
  16. Reguero, B.G.; Secaira, F.; Toimil, A.; Escudero, M.; Díaz-Simal, P.; Beck, M.W.; Silva, R.; Storlazzi, C.; Losada, I.J. The Risk Reduction Benefits of the Mesoamerican Reef in Mexico. Front. Earth Sci. 2019, 7, 455350. [Google Scholar] [CrossRef]
  17. Harris, D.L.; Vila-concejo, A.; Power, H.E.; Kinsela, M.A.; Webster, J.M. Variability of Depth-Limited Waves in Coral Reef Surf Zones. Estuar. Coast. Shelf Sci. 2018, 211, 36–44. [Google Scholar] [CrossRef]
  18. Reguero, B.G.; Beck, M.W.; Agostini, V.N.; Kramer, P.; Hancock, B. Coral Reefs for Coastal Protection: A New Methodological Approach and Engineering Case Study in Grenada. J. Environ. Manag. 2018, 210, 146–161. [Google Scholar] [CrossRef] [PubMed]
  19. Roelvink, F.E.; Storlazzi, C.D.; van Dongeren, A.R.; Pearson, S.G. Coral Reef Restorations Can Be Optimized to Reduce Coastal Flooding Hazards. Front. Mar. Sci. 2021, 8, 653945. [Google Scholar] [CrossRef]
  20. Storlazzi, C.D.; Reguero, B.G.; Cumming, K.A.; Cole, A.D.; Shope, J.B.; Gaido L, C.; Viehman, T.S.; Nickel, B.A.; Beck, M.W. Rigorously Valuing the Coastal Hazard Risks Reduction Provided by Potential Coral Reef Restoration in Florida and Puerto Rico. In Open-File Report; U.S. Geological Survey: Reston, VA, USA, 2021. [Google Scholar]
  21. Fabian, R.; Beck, M.; Potts, D. Reef Restoration for Coastal Defense; University of California: Santa Cruz, CA, USA, 2013; pp. 1–59. [Google Scholar]
  22. Morris, R.L.; Konlechner, T.M.; Ghisalberti, M.; Swearer, S.E. From Grey to Green: Efficacy of Eco-Engineering Solutions for Nature-Based Coastal Defence. Glob. Change Biol. 2018, 24, 1827–1842. [Google Scholar] [CrossRef]
  23. Reguero, B.G.; Beck, M.W.; Bresch, D.N.; Calil, J.; Meliane, I. Comparing the Cost Effectiveness of Nature-Based and Coastal Adaptation: A Case Study from the Gulf Coast of the United States. PLoS ONE 2018, 13, e0192132. [Google Scholar] [CrossRef] [PubMed]
  24. Silva, R.; Mendoza, E.; Mariño-Tapia, I.; Martínez, M.L.; Escalante, E. An Artificial Reef Improves Coastal Protection and Provides a Base for Coral Recovery. J. Coast. Res. 2016, 1, 467–471. [Google Scholar] [CrossRef]
  25. Buckley, M.L.; Lowe, R.J.; Hansen, J.E.; Van Dongeren, A.R. Dynamics of Wave Setup over a Steeply Sloping Fringing Reef. J. Phys. Oceanogr. 2015, 45, 3005–3023. [Google Scholar] [CrossRef]
  26. Van Dongeren, A.; Lowe, R.; Pomeroy, A.; Trang, D.M.; Roelvink, D.; Symonds, G.; Ranasinghe, R. Numerical Modeling of Low-Frequency Wave Dynamics over a Fringing Coral Reef. Coast. Eng. 2013, 73, 178–190. [Google Scholar] [CrossRef]
  27. Pequignet, A.C.; Becker, J.M.; Merrifield, M.A.; Boc, S.J. The Dissipation of Wind Wave Energy across a Fringing Reef at Ipan, Guam. Coral Reefs 2011, 30, 71–82. [Google Scholar] [CrossRef]
  28. Nwogu, O.; Demirbilek, Z. Infragravity Wave Motions and Runup over Shallow Fringing Reefs. J. Waterw. Port Coast. Ocean Eng. 2010, 136, 295–305. [Google Scholar] [CrossRef]
  29. Pomeroy, A.; Lowe, R.; Symonds, G.; Van Dongeren, A.; Moore, C. The Dynamics of Infragravity Wave Transformation over a Fringing Reef. J. Geophys. Res. Ocean. 2012, 117, C11022. [Google Scholar] [CrossRef]
  30. Symonds, G.; Huntley, D.A.; Bowen, A.J. Two-Dimensional Surf Beat: Long Wave Generation by a Time-Varying Breakpoint. J. Geophys. Res. 1982, 87, 492–498. [Google Scholar] [CrossRef]
  31. Lowe, R.J.; Falter, J.L.; Bandet, M.D.; Pawlak, G.; Atkinson, M.J.; Monismith, S.G.; Koseff, J.R. Spectral Wave Dissipation over a Barrier Reef. J. Geophys. Res. C Ocean. 2005, 110, C04001. [Google Scholar] [CrossRef]
  32. Pomeroy, A.W.M.; Lowe, R.J.; Van Dongeren, A.R.; Ghisalberti, M.; Bodde, W.; Roelvink, D. Spectral Wave-Driven Sediment Transport across a Fringing Reef. Coast. Eng. 2015, 98, 78–94. [Google Scholar] [CrossRef]
  33. Cheriton, O.M.; Storlazzi, C.D.; Rosenberger, K.J. In Situ Observations of Wave Transformation and Infragravity Bore Development Across Reef Flats of Varying Geomorphology. Front. Mar. Sci. 2020, 7, 351. [Google Scholar] [CrossRef]
  34. Storlazzi, C.D.; Rey, A.E.; Dongeren, A.R.V. A Numerical Study of Geomorphic and Oceanographic Controls on Wave-Driven Runup on Fringing Reefs with Shore-Normal Channels. J. Mar. Sci. Eng. 2022, 10, 828. [Google Scholar] [CrossRef]
  35. Huang, Z.C.; Lenain, L.; Melville, W.K.; Middleton, J.H.; Reineman, B.; Statom, N.; McCabe, R.M. Dissipation of Wave Energy and Turbulence in a Shallow Coral Reef Lagoon. J. Geophys. Res. Ocean. 2012, 117, C03015. [Google Scholar] [CrossRef]
  36. Lentz, S.J.; Churchill, J.H.; Davis, K.A.; Farrar, J.T. Surface Gravity Wave Transformation across a Platform Coral Reef in the Red Sea. J. Geophys. Res. Ocean. 2016, 121, 693–705. [Google Scholar] [CrossRef]
  37. Monismith, S.G.; Rogers, J.S.; Koweek, D.; Dunbar, R.B. Frictional Wave Dissipation on a Remarkably Rough Reef. Geophys. Res. Lett. 2015, 42, 4063–4071. [Google Scholar] [CrossRef]
  38. Storlazzi, C.D.; Reguero, B.G.; Cole, A.D.; Lowe, E.; Shope, J.B.; Gibbs, A.E.; Nickel, B.A.; McCall, R.T.; van Dongeren, A.R.; Beck, M.W. Rigorously Valuing the Role of U. S. Coral Reefs in Coastal Hazard Risk Reduction. In Open-File Report 2019–1027; U.S. Geological Survey: Reston, VA, USA, 2019; Volume 42. [Google Scholar] [CrossRef]
  39. Buckley, M.; Lowe, R.; Hansen, J. Evaluation of Nearshore Wave Models in Steep Reef Environments Topical Collection on the 7th International Conference on Coastal Dynamics in Arcachon, France 24–28 June 2013. Ocean Dyn. 2014, 64, 847–862. [Google Scholar] [CrossRef]
  40. Lashley, C.H.; Roelvink, D.; van Dongeren, A.; Buckley, M.L.; Lowe, R.J. Nonhydrostatic and Surfbeat Model Predictions of Extreme Wave Run-up in Fringing Reef Environments. Coast. Eng. 2018, 137, 11–27. [Google Scholar] [CrossRef]
  41. Davis, K.A.; Pawlak, G.; Monismith, S.G. Turbulence and Coral Reefs. Annu. Rev. Mar. Sci. 2021, 13, 343–373. [Google Scholar] [CrossRef] [PubMed]
  42. Osorio-Cano, J.D.; Alcérreca-Huerta, J.C.; Mariño-Tapia, I.; Osorio, A.F.; Acevedo-Ramírez, C.; Enriquez, C.; Costa, M.; Pereira, P.; Mendoza, E.; Escudero, M.; et al. Effects of Roughness Loss on Reef Hydrodynamics and Coastal Protection: Approaches in Latin America. Estuaries Coasts 2019, 42, 1742–1760. [Google Scholar] [CrossRef]
  43. Osorio-Cano, J.D.; Alcérreca-Huerta, J.C.; Osorio, A.F.; Oumeraci, H. CFD Modelling of Wave Damping over a Fringing Reef in the Colombian Caribbean. Coral Reefs 2018, 37, 1093–1108. [Google Scholar] [CrossRef]
  44. Yu, X.; Rosman, J.H.; Hench, J.L. Boundary Layer Dynamics and Bottom Friction in Combined Wave-Current Flows over Large Roughness Elements. J. Fluid Mech. 2022, 931, A11. [Google Scholar] [CrossRef]
  45. Yu, X.; Rosman, J.H.; Hench, J.L. Interaction of Waves with Idealized High-Relief Bottom Roughness. J. Geophys. Res. Ocean. 2018, 123, 3038–3059. [Google Scholar] [CrossRef]
  46. Lowe, R.J.; Koseff, J.R.; Monismith, S.G. Oscillatory Flow through Submerged Canopies: 1. Velocity Structure. J. Geophys. Res. C Ocean. 2005, 110, C10016. [Google Scholar] [CrossRef]
  47. Madsen, O.S. Spectral Wave-Current Bottom Boundary Layer Flows. Coast. Eng. Proc. 1994, 1, 384–398. [Google Scholar] [CrossRef]
  48. Nielsen, P. Coastal Bottom Boundary Layers and Sediment Transport; World Scientific Publishing Company: Singapore, 1992; Volume 4, ISBN 978-981-02-0472-3. [Google Scholar]
  49. Swart, D.H. Offshore Sediment Transport and Equilibrium Beach Profiles; TU Delft: Delft, The Netherlands, 1974. [Google Scholar]
  50. Dalrymple, R.A.; Kirby, J.T.; Hwang, P.A. Wave Diffraction Due to Areas of Energy Dissipation. J. Waterw. Port Coast. Ocean Eng. 1984, 110, 67–79. [Google Scholar] [CrossRef]
  51. Mullarney, J.C.; Henderson, S.M. Flows Within Marine Vegetation Canopies. In Advances in Coastal Hydraulics; Panchang, V., Kaihatu, J., Eds.; World Scientific Publishing Ltd.: Singapore, 2018; pp. 1–46. [Google Scholar]
  52. Lowe, R.J.; Falter, J.L.; Koseff, J.R.; Monismith, S.G.; Atkinson, M.J. Spectral Wave Flow Attenuation within Submerged Canopies: Implications for Wave Energy Dissipation. J. Geophys. Res. Ocean. 2007, 112, C05018. [Google Scholar] [CrossRef]
  53. OpenFOAM. OpenFOAM User’s Guide; OpenCFD Ltd.: Bracknell, UK, 2020. [Google Scholar]
  54. Le Méhauté, B. An Introduction to Hydrodynamics and Water Waves; Springer: Berlin/Heidelberg, Germany, 1967; ISBN 978-3-642-85569-6. [Google Scholar]
  55. Sous, D.; Dodet, G.; Bouchette, F.; Tissier, M. Momentum Balance Across a Barrier Reef. J. Geophys. Res. Ocean. 2020, 125, e2019JC015503. [Google Scholar] [CrossRef]
  56. Masselink, G.; Tuck, M.; McCall, R.T.; van Dongeren, A.R.; Ford, M.; Kench, P.S. Physical and Numerical Modeling of Infragravity Wave Generation and Transformation on Coral Reef Platforms. J. Geophys. Res. Ocean. 2019, 124, 1410–1433. [Google Scholar] [CrossRef]
  57. Pomeroy, A.W.M.; Lowe, R.J.; Ghisalberti, M.; Winter, G.; Storlazzi, C.; Cuttler, M. Spatial Variability of Sediment Transport Processes Over Intratidal and Subtidal Timescales Within a Fringing Coral Reef System. J. Geophys. Res. Earth Surf. 2018, 123, 1013–1034. [Google Scholar] [CrossRef]
  58. Beetham, E.; Kench, P.S.; O’Callaghan, J.; Popinet, S. Wave Transformation and Shoreline Water Level on Funafuti Atoll, Tuvalu. J. Geophys. Res. Ocean. 2016, 121, 3372–3380. [Google Scholar] [CrossRef]
  59. Boström-Einarsson, L.; Babcock, R.C.; Bayraktarov, E.; Ceccarelli, D.; Cook, N.; Ferse, S.C.A.; Hancock, B.; Harrison, P.; Hein, M.; Shaver, E.; et al. Coral Restoration—A Systematic Review of Current Methods, Successes, Failures and Future Directions. PLoS ONE 2020, 15, e0226631. [Google Scholar] [CrossRef]
  60. Liu, W.; Liu, Y.; Zhao, X. Numerical Study of Bragg Reflection of Regular Water Waves over Fringing Reefs Based on a Boussinesq Model. Ocean Eng. 2019, 190, 106415. [Google Scholar] [CrossRef]
  61. Norris, B.K.; Storlazzi, C.D.; Pomeroy, A.W.M.; Rosenberger, K.J.; Logan, J.B.; Cheriton, O.M. Combining Field Observations and High-Resolution Numerical Modeling to Demonstrate the Effect of Coral Reef Roughness on Turbulence and Its Implications for Reef Restoration Design. Coast. Eng. 2023, 184, 104331. [Google Scholar] [CrossRef]
  62. Larsen, B.E.; Fuhrman, D.R. On the Over-Production of Turbulence beneath Surface Waves in Reynolds-Averaged Navier–Stokes Models. J. Fluid Mech. 2018, 853, 419–460. [Google Scholar] [CrossRef]
  63. Higuera, P.; Lara, J.L.; Losada, I.J. Realistic Wave Generation and Active Wave Absorption for Navier-Stokes Models. Application to OpenFOAM®. Coast. Eng. 2013, 71, 102–118. [Google Scholar] [CrossRef]
  64. Buccino, M.; Del Vita, I.; Calabrese, M. Predicting Wave Transmission Past Reef Ball TM Submerged Breakwaters. J. Coast. Res. 2013, 65, 171–176. [Google Scholar] [CrossRef]
  65. Henderson, S.M.; Norris, B.K.; Mullarney, J.C.; Bryan, K.R. Wave-Frequency Flows within a near-Bed Vegetation Canopy. Cont. Shelf Res. 2017, 147, 91–101. [Google Scholar] [CrossRef]
  66. Zeller, R.B.; Zarama, F.J.; Weitzman, J.S.; Koseff, J.R. A Simple and Practical Model for Combined Wave-Current Canopy Flows. J. Fluid Mech. 2015, 767, 842–880. [Google Scholar] [CrossRef]
  67. Stocking, J.B.; Rippe, J.P.; Reidenbach, M.A. Structure and Dynamics of Turbulent Boundary Layer Flow over Healthy and Algae-Covered Corals. Coral Reefs 2016, 35, 1047–1059. [Google Scholar] [CrossRef]
  68. Hearn, C.; Atkinson, M.; Falter, J. A Physical Derivation of Nutrient-Uptake Rates in Coral Reefs: Effects of Roughness and Waves. Coral Reefs 2001, 20, 347–356. [Google Scholar] [CrossRef]
  69. Nakamura, T.; Van Woesik, R. Water-Flow Rates and Passive Diffusion Partially Explain Differential Survival of Corals during the 1998 Bleaching Event. Mar. Ecol. Prog. Ser. 2001, 212, 301–304. [Google Scholar] [CrossRef]
  70. Ribes, M.; Atkinson, M.J. Effects of Water Velocity on Picoplankton Uptake by Coral Reef Communities. Coral Reefs 2007, 26, 413–421. [Google Scholar] [CrossRef]
  71. Reidenbach, M.A.; Stocking, J.B.; Szczyrba, L.; Wendelken, C. Hydrodynamic Interactions with Coral Topography and Its Impact on Larval Settlement. Coral Reefs 2021, 40, 505–519. [Google Scholar] [CrossRef]
Figure 1. Coral reef hemispherical element examples and model setup. (a) Pacific Porites solida. (b) Caribbean Siderastrea siderea. (c) A ReefBallTM encrusted with juvenile corals. (d) Full-scale model layout with vertical exaggeration to depict roughness. All models started with a smooth region equal to one-half wavelength, followed by a region between x0 and Lm containing roughness equal to one whole wavelength. The approximate location of the enlarged schematic layout in (e) is depicted with a shaded rectangle in (d). Degrees of roughness were created by varying the across-shore spacing of elements D = Δ S / d . In this figure, the still water line (SWL) is denoted as a dashed red line and the water surface elevation η as a solid blue curve. The wavelength has been shortened for illustrative purposes. In (e), the mesh refinement for the models is shown in the background. Photo credits: (a) Turak and Lyndon DeVantier, (b) Mary Stafford-Smith, (c) Reefball.org.
Figure 1. Coral reef hemispherical element examples and model setup. (a) Pacific Porites solida. (b) Caribbean Siderastrea siderea. (c) A ReefBallTM encrusted with juvenile corals. (d) Full-scale model layout with vertical exaggeration to depict roughness. All models started with a smooth region equal to one-half wavelength, followed by a region between x0 and Lm containing roughness equal to one whole wavelength. The approximate location of the enlarged schematic layout in (e) is depicted with a shaded rectangle in (d). Degrees of roughness were created by varying the across-shore spacing of elements D = Δ S / d . In this figure, the still water line (SWL) is denoted as a dashed red line and the water surface elevation η as a solid blue curve. The wavelength has been shortened for illustrative purposes. In (e), the mesh refinement for the models is shown in the background. Photo credits: (a) Turak and Lyndon DeVantier, (b) Mary Stafford-Smith, (c) Reefball.org.
Jmse 12 00768 g001
Figure 2. Example model output (a single timestep) depicting an infragravity wave crest passing over the bottom roughness from a model with water depth, h = 1 m, wave height, Hs0 = 0.05 m, wave period, Tp = 30 s, and element spacing, D = 3. In the figure, the wave propagates from left (offshore) to right (onshore). (a) Velocity magnitude. (b) Turbulent kinetic energy (TKE) dissipation rate, ε. Low parameter values are in blue and high values in red. The velocity magnitude demonstrates accelerating water at the top of the roughness elements and decelerating water in the wake of individual elements. These accelerations and decelerations generate turbulence above and behind the elements.
Figure 2. Example model output (a single timestep) depicting an infragravity wave crest passing over the bottom roughness from a model with water depth, h = 1 m, wave height, Hs0 = 0.05 m, wave period, Tp = 30 s, and element spacing, D = 3. In the figure, the wave propagates from left (offshore) to right (onshore). (a) Velocity magnitude. (b) Turbulent kinetic energy (TKE) dissipation rate, ε. Low parameter values are in blue and high values in red. The velocity magnitude demonstrates accelerating water at the top of the roughness elements and decelerating water in the wake of individual elements. These accelerations and decelerations generate turbulence above and behind the elements.
Jmse 12 00768 g002
Figure 3. Percent change in modelled wave height Hs across the roughness versus the number of elements per across-shore meter n, for three wave periods, Tw = 30 s, 60 s, and 120 s (colors); three incident wave heights, Hs0 = 0.05 (a), 0.15 (b), and 0.30 m (c); and six across-shore element spacings, D = 2–32 (shapes). Wave height decreases with increasing across-shore element spacing for all model scenarios, with greater wave attenuation for closer element spacings and larger waves.
Figure 3. Percent change in modelled wave height Hs across the roughness versus the number of elements per across-shore meter n, for three wave periods, Tw = 30 s, 60 s, and 120 s (colors); three incident wave heights, Hs0 = 0.05 (a), 0.15 (b), and 0.30 m (c); and six across-shore element spacings, D = 2–32 (shapes). Wave height decreases with increasing across-shore element spacing for all model scenarios, with greater wave attenuation for closer element spacings and larger waves.
Jmse 12 00768 g003
Figure 4. Spectra of wave energy flux (Fj) across different wave frequencies (fj) for a single wave height (Hs0 = 0.15 m), and three incident wave periods Tw = 30 s (a,b), 60 s (c,d), and 120 s (e,f), for the smallest (D = 2) and largest (D = 32) across-shore element spacings. Power spectra are shown for the beginning (x0) and end (Lm) of the simulated roughness. Loss of wave energy occurs at the predominant frequency for restorations with short element spacings (D = 2), and at higher frequencies for larger spacings (D = 32). Hence, denser restorations dissipate wave energy more efficiently than sparser restorations.
Figure 4. Spectra of wave energy flux (Fj) across different wave frequencies (fj) for a single wave height (Hs0 = 0.15 m), and three incident wave periods Tw = 30 s (a,b), 60 s (c,d), and 120 s (e,f), for the smallest (D = 2) and largest (D = 32) across-shore element spacings. Power spectra are shown for the beginning (x0) and end (Lm) of the simulated roughness. Loss of wave energy occurs at the predominant frequency for restorations with short element spacings (D = 2), and at higher frequencies for larger spacings (D = 32). Hence, denser restorations dissipate wave energy more efficiently than sparser restorations.
Jmse 12 00768 g004
Figure 5. Wave energy dissipation factor fe,j as a function of frequency, calculated from the wave energy flux and near-bottom wave-orbital velocity using linear wave theory (Equation (3)) for two cases, Hs0 = 0.15 m, Tw = 30 s and two across-shore element spacings D = 2 (a) and D = 32 (b). These values correspond to the energy flux spectra shown in Figure 4a,b. For the D = 2 case, fe,j is greatest around the predominant frequency fj = 0.035 Hz. Conversely, for the D = 32 case, fe,j is greatest at higher frequencies, fj = 0.13 Hz.
Figure 5. Wave energy dissipation factor fe,j as a function of frequency, calculated from the wave energy flux and near-bottom wave-orbital velocity using linear wave theory (Equation (3)) for two cases, Hs0 = 0.15 m, Tw = 30 s and two across-shore element spacings D = 2 (a) and D = 32 (b). These values correspond to the energy flux spectra shown in Figure 4a,b. For the D = 2 case, fe,j is greatest around the predominant frequency fj = 0.035 Hz. Conversely, for the D = 32 case, fe,j is greatest at higher frequencies, fj = 0.13 Hz.
Jmse 12 00768 g005
Figure 6. Spectra of wave energy attenuation parameters for two cases, Hs0 = 0.15 m, Tw = 30 s and two across-shore element spacings D = 2 and 32. (a,b) Decomposition of the velocity spectrum above the canopy S u and velocity spectrum within the canopy layer S ^ u for two cases, Hs0 = 0.15 m, Tw = 30 s and two across-shore element spacings D = 2 and 32. (c,d) Spectral wave attenuation parameter α w , j as a function of frequency (Equation (6)). When canopy elements are densely spaced (a,c), in-canopy velocities are highly attenuated by drag. When elements are sparse (b,d), in-canopy velocities are hardly attenuated and α w , j ~0.8 across most frequencies.
Figure 6. Spectra of wave energy attenuation parameters for two cases, Hs0 = 0.15 m, Tw = 30 s and two across-shore element spacings D = 2 and 32. (a,b) Decomposition of the velocity spectrum above the canopy S u and velocity spectrum within the canopy layer S ^ u for two cases, Hs0 = 0.15 m, Tw = 30 s and two across-shore element spacings D = 2 and 32. (c,d) Spectral wave attenuation parameter α w , j as a function of frequency (Equation (6)). When canopy elements are densely spaced (a,c), in-canopy velocities are highly attenuated by drag. When elements are sparse (b,d), in-canopy velocities are hardly attenuated and α w , j ~0.8 across most frequencies.
Jmse 12 00768 g006
Figure 7. Points of depth-averaged turbulent kinetic energy (TKE) dissipation ( ε ) at even positions within the roughness elements for three wave periods Tw = 30 s (ac), 60 s (df), and 120 s (gi); three incident wave heights Hs0 = 0.05 (top row), 0.15 (middle row), and 0.30 m (bottom row); and six across shore element spacings D = 2–32 (shapes). ε increases with incident wave height and decreases with increasing wave period and element spacing.
Figure 7. Points of depth-averaged turbulent kinetic energy (TKE) dissipation ( ε ) at even positions within the roughness elements for three wave periods Tw = 30 s (ac), 60 s (df), and 120 s (gi); three incident wave heights Hs0 = 0.05 (top row), 0.15 (middle row), and 0.30 m (bottom row); and six across shore element spacings D = 2–32 (shapes). ε increases with incident wave height and decreases with increasing wave period and element spacing.
Jmse 12 00768 g007
Figure 8. Depth profiles of the turbulent kinetic energy (TKE) dissipation rate ( ε ) averaged across the roughness elements for three wave periods Tw = 30 s (ac), 60 s (df), and 120 s (gi); three incident wave heights Hs0 = 0.05 (top row), 0.15 (middle row), and 0.30 m (bottom row); and six across-shore element spacings D = 2–32 (shapes). The dashed black line denotes the top of the reef elements in each subplot. As in Figure 7, turbulence increases with decreasing spacing and increasing incident wave height. Greater turbulence occurs higher in the water column under larger waves with smaller element spacings and is principally concentrated near the bed with larger element spacings.
Figure 8. Depth profiles of the turbulent kinetic energy (TKE) dissipation rate ( ε ) averaged across the roughness elements for three wave periods Tw = 30 s (ac), 60 s (df), and 120 s (gi); three incident wave heights Hs0 = 0.05 (top row), 0.15 (middle row), and 0.30 m (bottom row); and six across-shore element spacings D = 2–32 (shapes). The dashed black line denotes the top of the reef elements in each subplot. As in Figure 7, turbulence increases with decreasing spacing and increasing incident wave height. Greater turbulence occurs higher in the water column under larger waves with smaller element spacings and is principally concentrated near the bed with larger element spacings.
Jmse 12 00768 g008
Figure 9. The depth and space averaged turbulent kinetic energy (TKE) dissipation rate ( ε ) against the number of elements per across shore meter (n) for three wave periods Tw = 30–120 s (colors); three incident wave heights Hs0 = 0.05 (a), 0.15 (b), and 0.30 m (c); and six across-shore element spacings D = 2–32 (shapes). In general, ε decreases with n (and D), starting from a maximum value at D = 2 for all cases. The inflection point of these curves is a key balance point between TKE dissipation and element spacing, representing the restoration design that maximizes dissipation while minimizing the number of restoration elements, and hence the potential cost of the restoration, for each set of incident wave conditions.
Figure 9. The depth and space averaged turbulent kinetic energy (TKE) dissipation rate ( ε ) against the number of elements per across shore meter (n) for three wave periods Tw = 30–120 s (colors); three incident wave heights Hs0 = 0.05 (a), 0.15 (b), and 0.30 m (c); and six across-shore element spacings D = 2–32 (shapes). In general, ε decreases with n (and D), starting from a maximum value at D = 2 for all cases. The inflection point of these curves is a key balance point between TKE dissipation and element spacing, representing the restoration design that maximizes dissipation while minimizing the number of restoration elements, and hence the potential cost of the restoration, for each set of incident wave conditions.
Jmse 12 00768 g009
Figure 10. The percent change in depth and space averaged turbulent kinetic energy (TKE) dissipation rate ( ε ) against the number of elements per across shore meter (n) for three wave periods Tw = 30–120 s (colors); three incident wave heights Hs0 = 0.05 (a), 0.15 (b), and 0.30 m (c); and six across-shore element spacings D = 2–32 (shapes). The change in ε increases with n across all cases, however the greatest change was observed for Tw = 120 s waves. There is a noticeable change in ε between Hs0 = 0.05 and 0.15 m, and negligible change between Hs0 = 0.15 and 0.30 m. Hence, improvements in ε may not be achievable by increasing element density beyond certain limits.
Figure 10. The percent change in depth and space averaged turbulent kinetic energy (TKE) dissipation rate ( ε ) against the number of elements per across shore meter (n) for three wave periods Tw = 30–120 s (colors); three incident wave heights Hs0 = 0.05 (a), 0.15 (b), and 0.30 m (c); and six across-shore element spacings D = 2–32 (shapes). The change in ε increases with n across all cases, however the greatest change was observed for Tw = 120 s waves. There is a noticeable change in ε between Hs0 = 0.05 and 0.15 m, and negligible change between Hs0 = 0.15 and 0.30 m. Hence, improvements in ε may not be achievable by increasing element density beyond certain limits.
Jmse 12 00768 g010
Figure 11. A schematic of percent change in energy dissipation (red curves, see Figure 10) versus the number of structural reef restoration elements per across-shore meter versus the hypothetical cost of restoration (blue line). Here, we represent the restoration cost (from lower cost, $, to higher cost, $$$) as a linear function of restoration unit numbers for simplicity. Percent dissipation increases as incident conditions (Hs, Tw) increase (from the solid to dashed red lines), necessitating fewer reef elements and lower cost, such that the price of the restoration moves from orange to blue circle on the cost curve and thus decreases.
Figure 11. A schematic of percent change in energy dissipation (red curves, see Figure 10) versus the number of structural reef restoration elements per across-shore meter versus the hypothetical cost of restoration (blue line). Here, we represent the restoration cost (from lower cost, $, to higher cost, $$$) as a linear function of restoration unit numbers for simplicity. Percent dissipation increases as incident conditions (Hs, Tw) increase (from the solid to dashed red lines), necessitating fewer reef elements and lower cost, such that the price of the restoration moves from orange to blue circle on the cost curve and thus decreases.
Jmse 12 00768 g011
Table 1. Literature values from former field and numerical modeling studies for expected range of infragravity wave conditions on reef flats. Here, h is the water depth, Hs is the infragravity significant wave height, Tw is the infragravity wave period, γ = Hs/h, and Hs is the wave steepness.
Table 1. Literature values from former field and numerical modeling studies for expected range of infragravity wave conditions on reef flats. Here, h is the water depth, Hs is the infragravity significant wave height, Tw is the infragravity wave period, γ = Hs/h, and Hs is the wave steepness.
Authors (Year)h (m)Hs (m)Tw (s)γHs
Sous et al. [55]1.50.180–0.30025–2500.12–0.203.1 × 10−4–1.9 × 10−3
Masselink et al. [56]0.50.060-0.12-
Buckley et al. [2]0.10.006–0.00827–1000.06–0.088.5 × 10−5–2.4 × 10−4
Pomeroy et al. [57]2.10.080–0.20025–2500.03–0.101.8 × 10−4–7.1 × 10−4
Beetham et al. [58]1.2–2.00.100–0.400>250.08–0.201.2 × 10−3–3.6 × 10−3
Cheriton et al. [3]1.8–2.20.250–0.80025–2500.14–0.366.8 × 10−4–2.4 × 10−3
Van Dongeren et al. [26]1.00.180–0.20025–6000.18–0.201.1 × 10−4–2.3 × 10−3
Pomeroy et al. [29]1.5–1.80.040–0.0801000.03–0.041.0 × 10−4–1.9 × 10−4
Table 2. Extrinsic offshore oceanographic conditions and intrinsic reef parameters in the BEWARE database [8] that correspond to the cases simulated in this study. Here, h is the water depth, Hs0 is the infragravity wave height, and Tw is the infragravity wave period on the reef flat, H0 is the wave height, T0 is the wave period offshore, and Wf is the reef flat width for a given set of offshore conditions.
Table 2. Extrinsic offshore oceanographic conditions and intrinsic reef parameters in the BEWARE database [8] that correspond to the cases simulated in this study. Here, h is the water depth, Hs0 is the infragravity wave height, and Tw is the infragravity wave period on the reef flat, H0 is the wave height, T0 is the wave period offshore, and Wf is the reef flat width for a given set of offshore conditions.
h (m)Hs0 (m)Tw (s)H0 (m)T0 (s)Wf (m)
10.053015350
10.056047150
10.05120111400
10.1530210500
10.156016250
10.15120325500
10.3030516100
10.306055250
10.30120311150
10.60120423500
10.6036019150
10.60720111200
10.9012034200
10.9036025500
10.9072014200
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Norris, B.K.; Storlazzi, C.D.; Pomeroy, A.W.M.; Reguero, B.G. Optimizing Infragravity Wave Attenuation to Improve Coral Reef Restoration Design for Coastal Defense. J. Mar. Sci. Eng. 2024, 12, 768. https://doi.org/10.3390/jmse12050768

AMA Style

Norris BK, Storlazzi CD, Pomeroy AWM, Reguero BG. Optimizing Infragravity Wave Attenuation to Improve Coral Reef Restoration Design for Coastal Defense. Journal of Marine Science and Engineering. 2024; 12(5):768. https://doi.org/10.3390/jmse12050768

Chicago/Turabian Style

Norris, Benjamin K., Curt D. Storlazzi, Andrew W. M. Pomeroy, and Borja G. Reguero. 2024. "Optimizing Infragravity Wave Attenuation to Improve Coral Reef Restoration Design for Coastal Defense" Journal of Marine Science and Engineering 12, no. 5: 768. https://doi.org/10.3390/jmse12050768

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop