Abstract

The legacy of structural artist Félix Candela is defined by his integration of thin hyperbolic paraboloid (hypar) shells within architecture across the Americas. One such form is the inverted umbrella, arising from the merger of four straight-edged hypar quadrants. The strength and elegance embodied by this geometry facilitated the recent conceptualization of kinetic umbrellas as an adaptable alternative to conventional floodwalls against surge-induced coastal inundation. Although the conceptual feasibility of such structures under hydrostatic inundation has been ascertained, their performance under combined surge and wave loading remains unknown. This paper used a three-dimensional (3D) numerical scheme integrating smoothed particle hydrodynamics with finite-element modeling for the structural analysis of kinetic umbrellas under the hydrodynamic regime. The technique was validated via dam-break testing involving 3D-printed specimens, and through empirical wave pressure formations at full scale. The behavior of kinetic umbrellas subject to surge and wave impact imparted by Hurricane Sandy (2012) at Monmouth Beach, New Jersey, was evaluated across different hypar geometries and angles of wave attack. Results showed the introduction of hypar geometry significantly enhances structural performance such that a 100-mm-thick umbrella successfully can resist hydrodynamic wave forces accompanying 2.7 m of inundation from landfalling hurricanes. Ultimately, this paper illustrates a creative yet practical structural engineering solution for mitigating the effects of climate change in coastal communities.

Introduction

In an era of urban renewal and revitalization, the design of structures for coastal resilience must adapt to the shifting cultural climate of metropolitan waterfronts. The threat of climate change coupled with increasing levels of cosmopolitanism within postindustrial harbors have encouraged planners to seek socially conscious solutions beyond traditional forms of coastal armor (Nordenson et al. 2018). Classical hard countermeasures including floodwalls and storm surge barriers exhibit numerous adverse consequences and offer no benefit to society except during extreme weather events (Dugan et al. 2008; Horn 2015; Kimura 2016). In response, deployable four-sided hyperbolic paraboloid (hypar) shells inspired by the designs of Spanish-Mexican structural artist Félix Candela (Garlock and Billington 2008) were conceptualized as an adaptable alternative to conventional armoring systems (Wang et al. 2019, 2020a). Such kinetic umbrellas derive from the union of four straight-edged hypar quadrants constructed using straight-line generators, resulting in their distinctive inverted umbrella form [Fig. 1(a)]. This doubly curved geometry exhibiting negative Gaussian curvature imparts impressive structural performance which has been reported widely in the literature (Candela 1955; Draper et al. 2010; Michiels et al. 2016; Wang et al. 2020b). The incorporation of kinetic umbrellas adjacent to urbanized waterfronts thus provides shade and shelter without limiting visibility or access to the shore during normal weather conditions [Fig. 1(b)]. A rotational hinge near the vertex permits the umbrella shell to tilt from its initial position, producing a physical barrier against surge-induced coastal inundation when simultaneously deployed in a row [Figs. 1(c and d)]. This seamless transition from structural art to coastal armor illustrates the philosophy of adaptable aquatecture, promoting a paradigm shift in coastal hazard engineering toward urban-conscious solutions yielding societal and environmental benefits over their entire operational lifespan.
Fig. 1. (a) Four hypar quadrants created via straight-line generators combined to form an inverted umbrella in the upright and deployed configuration; (b) architectural concept of kinetic umbrellas during normal weather; and (c and d) architectural concept of kinetic umbrellas during storm surge conditions. [Images (b–d) courtesy of FotosForTheFuture/Shutterstock.com.]
The practical implementation of kinetic umbrellas at the community level can be realized only after understanding its interaction with hydrodynamic forces associated with storm surge scenarios. Parametric studies by Wang et al. (2021c) revealed that a 100-mm-thick concrete umbrella [adopting the planar dimensions in Fig. 1(a)] may be designed to resist successfully over 7 m of hydrostatic inundation. However, its performance when subjected to wave impact associated with landfalling hurricanes remains unknown. Whereas the design of conventional armor such as levees and floodwalls rarely is force-governed due to their sheer mass (Reeve et al. 2018), the accurate quantification of wave loading on the superstructure becomes paramount for slender geometries embodied by the kinetic umbrella. Throughout engineering practice, the empirical formulations proposed by Goda (2000) are utilized widely to estimate the distribution of quasi-static wave pressures impinging vertical walls. Hence, current design methodologies cannot address wave forces interacting with geometrically complex systems (such as hypars) without substantial simplifications. Zhu and Scott (2014) introduced the particle finite-element method (PFEM) into OpenSees for structural analysis incorporating fluid–structure interaction (FSI) effects. However, such an approach may prove computationally prohibitive for complex three-dimensional (3D) domains which are required for the analysis of kinetic umbrellas. In response, a decoupled numerical scheme constituting compute unified device architecture (CUDA version 10.1)-accelerated smoothed particle hydrodynamics (SPH) and the finite-element method (FEM) was proposed by Wang et al. (2020a) for the treatment of arbitrary fluid forces acting upon tilted hypar manifolds. Although the method proved capable in capturing the spatial distribution of hydrostatic loads, its performance under the hydrodynamic regime has yet to be explored. The SPH-FEM technique therefore must be validated against dynamic fluid impact before its implementation within engineering design.
The objectives of this study were twofold. The first objective was to validate experimentally the SPH-FEM numerical scheme against dry-bed dam-break flows on scale-model kinetic umbrella specimens. Dam-break experiments involve the sudden release of water from an impounded reservoir, resulting in the generation of a turbulent bore propagating downstream. The relative simplicity and repeatability of such experiments has resulted in its widespread adoption for validating numerical flows involving FSI (Aureli et al. 2008; Biscarini et al. 2010; Ozmen-Cagatay and Kocaman 2011). the dam-break experiment was not intended to simulate a realistic coastal hazard event, but merely to impart hydrodynamic forcing strictly for purposes of numerical validation. Dynamic impact forces associated with the umbrella column were captured and compared with their numerical counterparts obtained via the decoupled SPH-FEM scheme across different levels of hypar warping, i.e., changes in umbrella rise, r [Fig. 1(a)]. As supplementary to the experimental study, the modeling approach also was validated against the forces of wave impact at full scale through comparisons with Goda’s empirical formulations pertaining to a flat inclined plate. This numerical study effectively complemented the aforementioned dam-break experiments to confirm the veracity of the SPH-FEM technique across all simulation scales.
The second objective was to adopt the validated numerical scheme to reproduce the effects of combined storm surge/wave loading on kinetic umbrellas under landfalling hurricanes as exemplified via a case study. Inundation and spectral wave characteristics at Monmouth Beach, New Jersey during Hurricane Sandy (2012) were utilized to impart hydrodynamic forcing on umbrellas across r/S ratios ranging from 0 to 0.255 [Fig. 1(a)]. Storm waves approaching at angles of incidence from 0° (i.e., perpendicular) to 45° also were investigated. Critical stresses within the shell and supporting column were assessed as a function of hypar warping and angle of wave attack to elucidate the mechanisms governing their behavior, and to ascertain the overall structural feasibility of kinetic umbrellas under extreme loading scenarios.
This research demonstrated the structural feasibility of thin-shell deployable hypar structures as an adaptable form of coastal armor in defense against storm surges. The work presents an example of how historical (heritage) forms and structures can inspire creative solutions to modern structural engineering challenges imposed by climate change. Beyond this global objective, we present to the structural engineering community a methodology to determine loads associated with hurricane surge and waves on structures of any complexity. With the recent incorporation of tsunami loading into ASCE 7 (Robertson 2020), this research serves as an important step toward the advancement and formalization of strategies for structural analysis in response to extreme coastal events.

Experimental Configuration

Test Setup and Specimen Design

All dam-break experiments were conducted within a channel constructed using transparent acrylic panels 25.4 mm thick elevated 0.63 m above ground level via an aluminum frame [Figs. 2(a) and 3(a)]. The inner dimensions of the channel measured 3×0.56×0.305  m (length×width×height). An aluminum gate at a distance of L0=0.616  m from the upstream wall provided containment for water within the reservoir [Fig. 2(b)]. This gate was slotted into aluminum channels embedded into each side wall of the flume, and petroleum jelly (white petrolatum) was applied to ensure a watertight seal. A 4.76-mm-diameter steel wire rope was connected to the gate, which fed into a pulley system and attached at the opposite end to a 25-kg drop weight suspended approximately 2 m above ground level [Fig. 2(a)]. The initiation of each dam-break test involved the release of this weight using a trigger mechanism built into the aluminum pulley beam. Once triggered, the weight experienced free fall over a vertical distance of 1 m before placing the wire rope into tension, thus raising the gate. This effectively increased its initial velocity, reducing the time required to clear the reservoir to satisfy the sudden dam-break condition (Lauber and Hager 1998). Within this study, three impoundment depths of water in the reservoir (d0), 100, 150, and 200 mm, were implemented to vary the force of hydrodynamic impact on the umbrellas.
Fig. 2. (a) Dam-break testing facility and monitoring equipment; (b) close-up of reservoir gate; (c) rear view of kinetic umbrella specimens; and (d) front view of kinetic umbrella specimens.
Fig. 3. (a) Schematic of dam-break testing channel depicting specimen location and monitoring setup; and (b) elevation of central umbrella specimen and plan view of typical hypar panel. All dimensions in millimeters.
Scale-model specimens of kinetic umbrellas were created for dam-break testing based on the geometry of their full-scale counterparts presented by Wang et al. (2021c). The full-scale prototypes and their models are summarized in Table 1. At full scale, it was envisioned that each square hypar umbrella would cover a projected area of A=64  m2, corresponding to each side measuring 8 m and with a diagonal distance between two corners S=11.3  m. The extent of geometrical shell warping therefore can be controlled via modification of the rise (r), defined as the distance between the hypar vertex and the point bisecting S [Fig. 1(a)]. This study considered three values of r at full scale, resulting in rise to area (r/A) ratios of 0.015, 0.030, and 0.045  m1. Because scaling must be dimensionless, the scaling was proportional to the ratio r/S. Hence, the r/A values corresponded to r/S ratios of 0.085, 0.170, and 0.255, respectively (Table 1). This resulted in a scaling factor of approximately 157. All models were fabricated via stereolithography (SLA) 3D printing using Accura Xtreme resin (3D Systems, Rock Hill, South Carolina) with solid density (ρs) and Young’s modulus (Es) of 1,200  kgm3 and 1,885 MPa, respectively. A constant projected thickness of 4.5 mm was maintained across all three geometries. The material properties are such that the umbrellas would remain essentially rigid under testing. Although the properties of the resin do not reflect those of the real structure, this would not influence any conclusions pertaining to the verification of the SPH-FEM technique.
Table 1. Variable geometric dimensions for kinetic umbrella specimens
Full scaleaModelb
r (m)r/A (m1)r/SScalingr (mm)r/Sh (mm)d (mm)
0.960.0150.08515716.80.0856163
1.920.030.17033.60.1705552
2.880.0450.25550.40.2554941
a
A=(8  m)2, and S=11.3  m [Fig. 1(a)].
b
θ=70°, A=(140  mm)2, and S=198  mm [Fig. 3(b)].
Because individual kinetic umbrellas were designed to operate as part of a continuous impermeable barrier, two additional hypars with the same r/S were placed adjacent to the central hypar to enforce this boundary condition [Figs. 2(c and d)]. Only the central umbrella was monitored in terms of the hydrodynamic force history. A 3D printed column was attached to the central umbrella via a stainless-steel binding barrel and screw slotted through a hole near the top of the column. This screw also passed through two cleats attached to the umbrella via holes offset 8 mm from the vertex on the rear face of the panel so that the column was free to rotate [Fig. 3(b)]. The base of the central column consisted of a V-shaped wedge 15 mm in length with the tip entering a 9.53-mm-diameter hole drilled through the channel floor. This tip rested within a slot cut into the end of a 9.5-mm-diameter aluminum rod sitting snugly within the hole, enabling free motion along the vertical axis [Fig. 3(b)]. The opposite end of this rod contacted a load cell attached to the underside of the channel via a steel U-bracket [Fig. 2(c)]. Hence, a pin–pin boundary condition effectively was imposed upon the ends of the column, at which the transducer captured only the vertical reaction at the umbrella vertex (vibrational responses were not considered). The total length of the central column was controlled such that umbrellas across each r/S ratio maintained a constant angle of inclination to the horizontal (θ) of 70° [Fig. 3(b)]. Table 1 summarizes the vertical distance h between the column hinge and floor level for each r/S.
To integrate the umbrella specimens into the dam-break channel, a 3-mm-diameter hole travelling through the lower edge was incorporated into each 3D-printed hypar [Fig. 3(b)]. This enabled a single 2.38-mm aluminum rod to feed though the base of both the boundary and the central umbrellas while passing through four aluminum cleats [Fig. 2(c)]. The cleats were attached to an aluminum baseplate 4.76 mm thick [Fig. 3(a)] recessed into the tank with the top surface flush with the channel floor. To maintain θ=70°, the cleats were positioned such that the distance between the base hinge and the front edge of the baseplate d (Fig. 3) adhered to that in Table 1 for a given geometry defined by r/S. For each boundary umbrella, a separate column was printed containing a flat end that could be fixed to the baseplate via machine screws [Fig. 2(c)]. Similar to the central column, its overall length was based on that required to maintain θ=70° for each r/S. To control seepage between each boundary umbrella and the central umbrella, a 1-mm-thick aluminum strip was attached to the side edges of each panel with cyanoacrylate. The small gap remaining between each umbrella subsequently was filled with petroleum jelly [Fig. 2(d)].

Monitoring Setup

To capture the force history constituting the reaction of the central column, a Loadstar Sensors (Fremont, California) RSB2 resistive load cell recording at 50 Hz was adopted [Fig. 3(b)]. This sampling frequency is consistent with that in other dam-break studies (Wei et al. 2015), and effectively filters out extremely short duration impulsive pressures because they do not affect the structural performance significantly (Goda 2000; Reeve et al. 2018). As previously stated, the purpose of this transducer was to capture the vertical column reaction at the umbrella vertex (Rvexp). The recorded column reaction effectively acted to balance the torque generated about the pin at the umbrella base by dam-break impact to maintain static equilibrium. Evidently, the magnitude of this reaction at any given time is governed by a dynamic fluid pressure field acting on the hypar. Hence, comparisons with its numerical counterpart enabled an indirect assessment of the numerical technique’s ability to reproduce accurately the spatial and temporal hydrodynamic force distribution on a range of hypar geometries. In addition, two Basler acA1440-220um (Exton, Pennsylvania) monochrome high-speed cameras were synchronized for flow visualization. The first camera (C1) was placed at the gate location to capture the exact instant of dam-break initiation, whereas the second (C2) was directed toward the specimen. A third camera (C3) was directed toward the upstream end of the channel to capture potential overtopping and flow around the specimen.

Validation of Numerical Scheme

SPH-FEM Modeling

The decoupled SPH-FEM scheme was described in detail by Wang et al. (2020a) and is discussed only briefly herein. In this work, the open-source CUDA-accelerated SPH solver DualSPHysics version 4.2 (Crespo et al. 2015) was implemented to resolve the governing equations as detailed in the Appendix. The dimensions of the dam-break channel [Fig. 4(a)] defined the SPH computational domain. Based on previous sensitivity studies of particle resolution (Wang et al. 2020a), fluid particles with an initial interparticle spacing (dp) of 2 mm constituted the impounded reservoir, whereas boundary particles with the same dp comprised the umbrella specimens and walls of the channel. For the three impoundment depths of 100, 150, and 200 mm, this corresponded to a total of 4,413,948, 6,577,648, and 8,741,348 fluid particles, respectively. Because the sudden dam-break condition was satisfied, the gate was not included explicitly within the simulation (St-Germain et al. 2014). The front surface of the central umbrella in contact with the initial bore, referred to as the shield [Fig. 4(b)], was defined according to the geometric formulations presented by Wang et al. (2020a) based on the parameters in Table 1. This surface was discretized into 36 SPH cells arranged in a 6×6 grid for each quadrant, thus producing 144 cells in total as optimal for square hypars [Fig. 4(c)]. Each cell represented a collection of boundary particles in which the Cartesian components of the hydrodynamic force history were computed as a function of time [Fig. 4(c), fsw]. The two boundary umbrellas flanking the central hypar were modeled in SketchUp Pro 2020 and imported into the computational domain via discretization into a triangular lattice by the DualSPHysics preprocessor before being converted into boundary particles. Each dam-break simulation spanned 5 s of physical time and was performed utilizing an NVIDIA GeForce RTX 2080Ti GPU (Santa Clara, California) with 4,352 CUDA cores. The average SPH simulation time pertaining to d0=100, 150, and 200 mm was 21.7, 42.2, and 65.8 h, respectively.
Fig. 4. (a) SPH computational domain of dam-break experiment; (b) close-up of front and rear views of the umbrella specimens; (c) FEM model with 144  elements/quadrant (constituting four elements/SPH cell) with hydrodynamic force vectors assigned to the central node of each cell (where x^, y^, and z^ represent the standard basis vectors); and (d) schematic of the FEM model with vertex detail. All dimensions in millimeters.
An overview of the SPH-FEM modeling framework is presented in Fig. 5. After the spatial and temporal distribution of hydrodynamic force vectors, fsw(t), were computed using SPH as previously described, they were mapped to an identical finite-element representation of the central kinetic umbrella based on the same geometry as the SPH model [Fig. 4(c)] to compute its structural behavior assuming small deformations. This assumption was shown to hold under both hydrostatic conditions (Wang et al. 2020a) and for the dam-break experiments conveyed in the section “Validation against Dam-Break Results at Model Scale.” It was necessary to ascertain that an accurate description of the dynamic structural response can be obtained when following the procedure outlined in Fig. 5. Therefore, the open-source finite element solver OpenSees version 2.5.0. was implemented to validate the SPH-FEM technique against the aforementioned dam-break experiments. Following each dam-break simulation, forces on the umbrella surface as output from DualSPHysics were input into OpenSees (Fig. 5). In OpenSees, the umbrella was modeled using eight-node hexahedral stabilized single-point integration brick elements (SSPbrick) utilizing the ElasticIsotropic material (McGann et al. 2011) adopting material properties consistent with the experiment. Each quadrant composed of 36 SPH cells was discretized into 144 elements [Fig. 4(c)], and fsw(t) was assigned to the central node of each four-element cluster mapped to each cell. To model the column, nine uniaxialMaterial links of arbitrarily large stiffness were adopted to connect the nine nodes [Fig. 4(d)] adjacent to the vertex on the rear surface (referred as the soffit) to a node offset 8 mm from the vertex. This node effectively acted as the column hinge [Fig. 3(b)] and was fixed in translation along the y- and z-axes while being free to rotate. Likewise, all nodes along the base were fully pinned (fixed in translation but free to rotate about the y-axis) to simulate the base hinge depicted in Fig. 3(b). The vertical reaction at the column hinge obtained via the FEM model RvSPH(t) therefore was the numerical counterpart to the experimental force time history Rvexp(t) measured by the load cell. Hence, by evaluating the synchronicity between Rvexp and RvSPH as a function of time, the structural model could be validated against experimental results under the hydrodynamic regime. Although any FEM software can be used for validation purposes, OpenSees was selected in this instance due to its support for the assignment of temporal force data and the ease with which the reaction time history can be obtained.
Fig. 5. Flowchart describing the decoupled SPH-FEM numerical scheme.

Validation against Dam-Break Results at Model Scale

Fig. 6 summarizes the comparison between experimental (Rvexp) and numerical SPH-FEM (RvSPH) hydrodynamic force histories associated with the central column reaction across r/S ratios of 0.085, 0.170, and 0.255 at each reservoir impoundment depth of 100, 150, and 200 mm. The initial flat segment of loading reflects only the structural self-weight (attributed to the hypar, hinge, and column) prior to dam-break impact. The self-weight was compared with its numerical counterpart to confirm that friction associated with the sealing between the umbrellas did not significantly affect the measured force. For each experimental configuration, three tests were conducted to ensure data repeatability (Fig. 6, shaded lines). The solid black line in each subplot is the averaged force history across all three trials. A total of 27 individual dam-break experiments across the three r/S ratios and impoundment depths were conducted.
Fig. 6. Experimental (EXP) and SPH-FEM numerical (SPH-FEM) column reaction force histories for r/S ratios of 0.085, 0.170, and 0.255 across impoundment depths of (a–c) d0=100  mm; (d–f) d0=150  mm; and (g–i) d0=200  mm. Fig. 7 shows the flow profile demarcated in Figs. 6(b, e, and h).
At an impoundment depth of 100 mm, the decoupled SPH-FEM scheme proved capable of capturing the hydrodynamic behavior of the central column [Figs. 6(a–c)]. The dual-peaked nature of the experimental force history was reproduced successfully, and the peak reaction was well described by the numerical technique (Table 2). However, at d0=150  mm [Figs. 6(d–f)], numerical reactions across all r/S ratios underestimated the first spike associated with the initial bore impact, and overestimated the final peak immediately preceding the gradual transition into the hydrostatic regime. The corresponding peak column reaction also was underestimated by the SPH-FEM method (Table 2). At the highest impoundment depth, 200 mm [Figs. 6(g–i)], simulated column reactions for r/S ratios of 0.170 and 0.255 compared well with their experimental counterparts across the entire monitored interval. However, at the lowest level of hypar warping at r/S=0.085, the peak force was overestimated.
Table 2. Peak force associated with column reaction for r/S ratios of 0.085, 0.170, and 0.255 across impoundment depths of d0=100, 150, and 200 mm
d0 (mm)r/SMax Rvexp (N)Max RvSPH (N)Error (%)
1000.0852.702.824.44
0.1702.492.461.20
0.2552.262.183.54
1500.08511.28.5823.4
0.1707.877.642.92
0.2557.536.2317.3
2000.08515.019.026.7
0.17013.414.04.48
0.25511.612.14.31
To explain the observations in Fig. 6, experimental and numerical flow profiles at various time intervals were examined. This comparison was facilitated by horizontal and vertical gridlines (spaced at 50 mm) applied to the inner surface of each side wall of the channel. Fig. 7 presents snapshots of the flow profile for r/S=0.170 across all impoundment depths at the times marked in Figs. 6(b, e, and h). These times were selected to capture the approximate peak force for both Rvexp and RvSPH. Interactions between the dam-break bore and kinetic umbrella models resulted in the formation of a plunging-type breaker in the opposing direction. As the bore contacted the specimen, the geometry of each inclined hypar shell initially directed the flow upward following its curvature profile. After an upper limit was reached, the crest curled toward the upstream direction before plunging back into the flow, resulting in highly turbulent mixing with significant air entrainment. The collapse of this breaker thus corresponded to the first peak of the force history measured on the central column (Fig. 6). After an initial reduction, a second force peak was reached from pressure accumulation on the shield as the reservoir continued to drain.
Fig. 7. Experimental and SPH flow profiles pertaining r/S=0.170 for impoundment depths of (a and b) 100 mm at times demarcated in Fig. 6(b); (c and d) 150 mm at times demarcated in Fig. 6(e); and (e and f) 200 mm at times demarcated in Fig. 6(h).
At d0=100  mm, experimental flow characteristics were well captured by the SPH model [Figs. 7(a and b)]. However, the initial runup induced by hypar geometry was underestimated for d0=150  mm [Fig. 7(c)]. Although the plunging crest had yet to contact the upstream flow in the experiment, simulations appeared to depict the flow at a later stage of breaking. Experimental results at d0=150  mm also reflected higher levels of overtopping that were not captured numerically [Figs. 7(c and d)]. The lower peak force produced by the SPH-FEM scheme at this impoundment depth therefore can be linked to its underestimation of the maximum level of initial runup, hence underpredicting the maximum moment exerted about the base hinge of the central umbrella. The subsequent overestimation of the final peak [Figs. 6(d–f)] followed from a larger quantity of water retained on the upstream side of the specimen supplementing the hydrodynamic pressures induced by the remaining incoming flow. At the highest impoundment depth, d0=200  mm, continuous overtopping was observed [Figs. 7(e and f)], and SPH results also predicted an earlier formation of the plunging breaker (similar to d0=150  mm). However, the synchronicity between experimental and numerical column force histories for r/S=0.170 and 0.255 [Figs. 6(h and i), respectively] demonstrated the capability of the SPH-FEM technique to capture the main features of hydrodynamic FSI for flows in which overtopping can be represented reasonably. Evidently, the observed discrepancies may result from the exclusion of the gaseous phase so that the effects of air entrainment on the flow cannot be captured. Multiphase SPH simulations (Grenier et al. 2013; Mokos et al. 2015) therefore would be required to advance further the study of violent hydrodynamic interactions with kinetic umbrellas. However, the corresponding increase in computational cost may not prove feasible from a design perspective (De Padova et al. 2020). For purposes of predicting the structural response to inform the design of kinetic umbrellas in engineering practice, the single-phase implementation of SPH as described herein likely would prove to be sufficient based on the correlations observed in Fig. 6.

Validation against Goda Equations at Full Scale

The previous section validated the SPH-FEM scheme against experimental data using model-scale and hypar surfaces. This section validates the SPH-FEM scheme at full scale on flat surfaces. To achieve full-scale validation against the forces of wave impact, Goda’s formulations were used as applicable to conventional sloped barriers without hypar warping (Tanimoto and Kimura 1985). Because there are no existing formulations dealing with wave impact on hypar surfaces, the use of a barrier without warping was necessary. An 8×8-m flat panel (with r=0) inclined at 65° was selected to maintain consistency with the geometric parameters previously adopted by Wang et al. (2021c), and placed at the top of a sloping beach 10 m long and 2 m high [Fig. 8(a)]. Three water depths (d), 3.8, 4.7, and 5.6 m, were considered, which translated to levels of inundation (dw) matching 0.25h*, 0.37h*, and 0.5h*, where h*=7.25  m is the total deployed height of the barrier. For each water depth, regular waves with wave height Hw propagating at frequency f (with period T=1/f) and wavelength Lw were selected based on the maximum Hw that can be produced via Stokes second-order wave theory defined by the Le Mehaute abacus (Le Mehaute 1976) [Fig. 8(b)] (summarized in Table 3). The maximum Goda wave pressures p1, p2, and p3 [Fig. 8(a)] were determined as follows (Goda 2000):
p1=(α1+α2)ρwgHwp2=α4p1p3=α3p1
(1)
where ρw = water density; g = gravitational acceleration; and
α1=0.6+12[4πdsinh(4πdLw)Lw]2,α2=min{hbdw3hb(Hwdw)2,2dwHw},α3=1dwd[11cosh(2πdLw)],α4=1min{η*,hc}η*
(2)
where hb = water depth at a distance 5Hw from the barrier toe; hc=h*dw [Fig. 8(a)]; and η*=1.5Hw. The wavelength Lw=2π/k is determined from the wave period or frequency through the dispersion relation
ω=gktanh(kd)
(3)
where ω=2πf=2π/T; and k = wavenumber, which equals the number of wave cycles per unit length (simply defined as 2π/Lw). Eq. (1) combined with the corresponding linear hydrostatic pressure distribution therefore enables determination of the maximum base shear (Rb) and vertex reaction (Rv) [Fig. 8(a)] for the inclined panel via principles of static equilibrium. This was compared against their numerical counterparts from SPH-FEM modeling to validate the numerical scheme. An offset of 0.3 m between the vertex hinge and panel centroid was imposed to mirror the geometry in Wang et al. (2021c).
Fig. 8. (a) Schematic of analysis domain showing the distribution of Goda wave pressures on the inclined plane; (b) adopted wave characteristics based on Le Mehaute abacus; and (c) SPH simulation of maximum wave loading on the panel across three depths.
Table 3. Wave parameters adopted for each water depth, and Goda and SPH-FEM reaction forces at each level of inundation
d (m)dw/h*Hw (m)f (s1)Lw (m)Max RvGoda (kN)Max RbGoda (kN)Max RvSPH (kN)Max RbSPH (kN)Error Rv (%)Error Rb (%)
3.80.2510.16035.71913071922950.523.91
4.70.371.250.14543.75215975456064.611.51
5.60.501.50.13053.41,1099911,1601,0014.601.01
The domain illustrated in Fig. 8(a) subsequently was simulated via SPH across a total width of 16 m. This effectively accommodated one full panel modeled using 144 SPH cells (36  cells/quadrant) flanked by two half-panels acting as the boundary condition [Fig. 8(c)]. An interparticle spacing of dp=0.12  m was implemented to maintain the same resolution as that of the dam-break simulation described in the section “SPH-FEM Modeling.” An open boundary was adopted at the upstream end of the domain at which buffer particles were used to create an inlet layer where the horizontal components of the fluid velocity, surface elevation, and pressure were imposed externally (Verbrugghe et al. 2019). The total length of the flat upstream section of the numerical flume was adjusted for each water depth to ensure that it was longer than 1 wavelength Lw [Fig. 8(a)]. To generate the waves, the horizontal velocity time history Vb(t) of all fluid particles constituting the open boundary were prescribed. This was achieved through Stokes second-order wave theory as defined by Madsen (1971) to suppress the generation of spurious secondary waves (Altomare et al. 2017)
Vb(t)=rt(t){Hwω2ξcos(ωtπ2)+2ω[Hw232d][3cosh(kd)sinh3(kd)2ξ]cos(2ωtπ)}
(4)
which is applicable only when UN<8π2/326, where UN=HwLw2/d3 is the Ursell number. Furthermore, ξ denotes the Biesel transfer function assuming irrotational and incompressible fluid behavior with constant pressure at the free surface
ξ=2[cosh(2kd)1]sinh(2kd)+2kd
(5)
Finally, to ensure numerical stability, rt(t) represents an artificial ramping function (Ning et al. 2017)
rt(t)={12[1cos(πtT)],tT1,t>T
(6)
Wave forces acting on each SPH cell for a given loading scenario subsequently were extracted and applied to its FEM counterpart (following the same modeling approach defined in the section “SPH-FEM Modeling” and Fig. 4) to obtain the maximum base shear and vertex reaction. Across all three water depths, absolute errors pertaining to Rv and Rb between Goda’s relations and the SPH-FEM scheme were less than 5% (Table 3). As a supplement to the experimental results from the section “Validation against Dam-Break Results at Model Scale,” the modeling technique presented herein may therefore be adopted with confidence for the analysis of full-scale kinetic umbrellas subject to wave attack. Because direct comparisons were not made of the hydrodynamic response between model and full-scale simulations, scaling effects (such as Froude similarities) do not apply.

Monmouth Beach Case Study

Site Description

The validated SPH-FEM scheme was implemented to assess the performance of kinetic umbrellas subjected to a realistic storm event to ascertain their structural feasibility in an engineering context. Hurricane Sandy, the costliest cyclone on record to impact the New York metropolitan area, was selected for this case study. Making landfall near Brigantine, New Jersey at 2330 UTC on October 29, 2012, Sandy induced record levels of storm surge across the Mid-Atlantic states, with the highest inundation, 2.7 m, observed at Monmouth, New Jersey (Blake et al. 2013). Therefore, the Atlantic shoreline of Monmouth Beach located at 40°20′04″ N 73°58′28″ W was adopted as the study site [Fig. 9(a)], and its corresponding beach profile during fall 2012 [Fig. 9(b)] was extracted from the New Jersey Beach Profile Network (NJBPN) database compiled by Coastal Research Center (2021). An existing seawall approximately 3.5 m high prevents visibility and access to the shoreline [Fig. 9(c)]. Figs. 9(d and e) illustrate the proposed use of kinetic umbrellas as an adaptable alternative to conventional armoring at this site, the structural behavior of which was evaluated in response to combined surge/wave loading consistent with Hurricane Sandy. The hinge in Figs. 9(d and e) is shown only for conceptual purposes. A possible design of the hinge mechanism was explored by Wang et al. (2021a).
Fig. 9. (a) Satellite view of the Monmouth Beach study area, with adopted beach profile marked; (b) elevation indicating the location of the proposed umbrella (vertical scale exaggerated); (c) existing seawall at Monmouth Beach; (d) architectural rendering of proposed kinetic umbrellas in the upright configuration at location in Fig. 9(c); and (e) architectural rendering of proposed kinetic umbrellas in the deployed configuration at the location in Fig. 9(c). [Base map (a) from Imagery ©2021 Google, Map data ©2021; images (c–e) by authors.]

Wave Characteristics

The idealized beach profile depicted in Fig. 9(b) was implemented for the SPH simulation [Fig. 10(a)]. A total water depth (d) of 9.1 m ensured that an inundation (dw) of 2.7 m was achieved to reflect the maximum witnessed during Hurricane Sandy at Monmouth. The depth-limited wave spectrum [Fig. 10(b)] was estimated based on the maximum sustained surface wind speed of 21.6ms1 (42 kn) recorded at Monmouth Beach approximately 15 min after landfall (Blake et al. 2013). This spectrum was generated via a simplified parametric swell and wind-sea model utilizing the Donelan et al. (1985) spectral form modified for depth-limited regimes (Wang et al. 2021b). The model produced a significant wave height (Hm0) of 3.64 m, with peak period (Tp), frequency (fp), and wavelength (Lp) of 8.33 s, 0.12  s1, and 71 m, respectively. Unlike Joint North Sea Wave Project (JONSWAP), in the Donelan spectrum the high-frequency face of the spectral energy (SD) is proportional to f4 as follows (Young 2003):
SD(f)=αDg2(2π)4fpf4exp[(ffp)4]γDexp{[(ffp)2]/(2σD2fp2)}
(7)
where
αD=0.0165νD0.55γD={6.489+6logνD,νD0.1591.7,νD<0.159σ=0.08+1.29×103νD3
(8)
and
νD=fpg[g2m06.365×106(fpg)3.3]1/0.7,m0=(Hm04)2
(9)
Fig. 10. (a) Elevation of SPH simulation domain of Monmouth Beach (exaggerated vertical scale); (b) estimated wave spectra; (c) corresponding inlet velocity time history and normalized force on obstacle over the entire storm duration; (d) 60-s excerpt of inlet velocity producing the maximum wave force; and (e) corresponding force history.
The adopted Lp of 71 m constitutes a rough approximation of the sea state, because using Stokes fifth-order yields Lp=75.9  m, and in stream function theory, Lp=76.03  m.
Prior to the 3D simulation of kinetic umbrellas, it was necessary to obtain the wave conditions yielding the maximum hydrodynamic force associated with Hurricane Sandy at the proposed umbrella location. For this purpose, irregular waves (which provide a more realistic representation of waves found in nature) were generated via the deterministic spectral simulation method. To ascertain the desired forcing conditions for structural analysis, a two-dimensional (2D) SPH simulation of the domain [Fig. 10(a)] was implemented. This enabled the force history on an arbitrary obstacle at the proposed umbrella location to be captured over a total storm duration of 5,000 s (equivalent to approximately 660 waves propagating at the mean period Tm0.9Tp), with the inlet velocity [Fig. 10(c)] defined as follows:
Vb(t)=rt(t)i=1N=100Aiωiξicos(ωit+ϕ˜i)
(10)
where Ai = amplitude associated with each discrete frequency fi{0.05,0.30}  s1 of wave spectrum comprising N=100 frequency bands
Ai=2SD(fi)Δf
(11)
where Δf = constant width of each frequency band; ωi=2πfi; and
ξi=2[cosh(2kid)1]sinh(2kid)+2kid
(12)
where ki is calculated using the dispersion relation [Eq. (3)]. To capture the stochastic nature of realistic sea states, ϕ˜iU(π,π) imposes an arbitrary phase shift randomly sampled from a uniform distribution between π and π. A 60-s excerpt of the inlet velocity [Fig. 10(d)] responsible for producing the maximum wave force on the obstacle [Fig. 10(c)] subsequently was identified and adopted as the forcing conditions for the 3D structural analysis of kinetic umbrellas. Fig. 10(e) shows the force on the obstacle when utilizing the inlet velocity sequence in Fig. 10(d).

Geometric Properties and Model Definition

Full-scale kinetic umbrellas measuring 8×8  m (S=11.3  m and A=64  m2) [Fig. 1(a)] with r/S ratios 0, 0.085, 0.170, and 0.255, all inclined at θ=65°, were adopted for the case study. This mirrored the geometries previously examined by Wang et al. (2021c) in their parametrization of the hydrostatic response. To examine the influence of waves approaching at oblique angles, umbrellas with r/S=0.170 rotated counterclockwise about the vertical axis at the base at angles θw=15°, 30°, and 45° also were implemented. A SPH domain width of 16 m was adopted for the beach profile presented in Fig. 10(a), of which the central 8 m wide umbrella was modeled using 144 SPH cells and which was flanked by two half-umbrellas, each 4 m wide, acting as the boundary condition. The umbrellas were positioned 50 m shoreward from the top of the sloping beach, and experienced 2.7 m of still water inundation. Fig. 11 illustrates the full SPH computational domain and showcases all umbrella geometries and rotations considered. For each configuration, over 14.7 million fluid particles were simulated for a total physical time of 60 s, per the inlet velocity history in Fig. 10(d), to cpature the peak hydrodynamic force occuring at approximately t=46.5  s [Fig. 10(e)]. Due to computational limitations, it would be infeasible to simulate the entire 5,000-s storm; hence only the 60-s duration yielding the maximum force was considered. Each simulation took approximately 63 h to complete.
Fig. 11. SPH simulation domain of Monmouth Beach showing bird’s-eye view of kinetic umbrellas (SPH cells highlighted) adopting various r/S ratios and θw. Arrows indicate the direction of wave propagation.
Only the peak hydrodynamic SPH force measurements generated from DualSPHysics were used for structural analysis (as opposed to temporal force data, which were used in the validation studies); thus it was simpler and more effective and efficient to use the commercial finite-element software SAP2000 v20 for the structural analysis of kinetic umbrellas, rather than OpenSees. Furthermore, the use of SAP2000 facilitated a straightforward approach for the determination of internal shell stresses. FEM models and structural detailing followed those described by Wang et al. (2021c), for which only a brief overview is presented here. All panels were 100 mm thick (modeled using 576 Mindlin–Reissner shell elements) and incorporated a standard steel 125-mm parallel flange channel framing the perimeter for additional stiffening [Fig. 12(b)]. High-strength concrete with compressive capacity (fc) and Young’s modulus (Ec) of 65 and 37,900 MPa, respectively, were used as the shell material. A mesh of 16-mm glass fiber–reinforced polymer (GFRP) bars with Young’s modulus Ef=66  GPa and tensile strength fu=1,135  MPa acted as reinforcement (Abdelkarim et al. 2019). GFRP was used to mitigate the issue of reinforcement corrosion associated with traditional steel-reinforced concrete exposed to marine environments (Medeiros and Helene 2009). In contrast to the material properties adopted for model-scale experiments (reflecting the 3D printed resin relevant only for validating the SPH-FEM modeling technique), the properties defined herein pertain to umbrellas built for real-world applications. Nonlinear compression-only gap elements were utilized along the base of each umbrella such that displacements were permitted only along the positive z-ordinate, whereas translation along the x- and y-axes was restrained [Fig. 12(a)]. This reflected the panel base nested within a groove recessed into the ground allowing only uplift to occur. To simulate the behavior of the supporting column, lateral springs with stiffness kv=3EcIc/hc3 along the x- and y-ordinates were assigned to the hinge (offset 0.3 m from the vertex), where Ic is the second moment of area for a 1-m-square concrete column [Fig. 12(c)], and hc denotes the column height. Moment–curvature relations for the umbrella shell and column computed using the RC sectional analysis program Response-2000 version 1.0.5 (Bentz and Collins 2001) are shown in Figs. 12(d and e), respectively. A detailed description of the FEM model and material characterisics were given by Wang et al. (2021c).
Fig. 12. (a) Elevation view of kinetic umbrellas (with human for scale) exhibiting r/S ratios from 0 to 0.255 along with the corresponding lateral column stiffness (kv) for a 1-m-square concrete column; (b) plan view of umbrella; (c) cross section of column with detailing consistent with Wang et al. (2021c); (d) moment–curvature relations of the hypar shell; and (e) moment–curvature relations of the concrete column.

Structural Response under Hurricane Waves

Application of the spatial forces corresponding to the peak total hydrodynamic force (computed via SPH) to the FEM structural model (constructed using SAP2000) enabled the commencement of structural analysis. Fig. 13 summarizes the maximum out-of-plane bending (M*), shear (V*), and in-plane normal (N*) forces within the 100-mm-thick hypar umbrella (averaged over a 1-m width) across the various r/S ratios (with θw=0) and θw angles (with r/S=0.170) considered at the instant of maximum wave impact occurring at approximately t=45  s. At θw=0, the moment and shear demands [Fig. 13(a)] dramatically decreased with the introduction of hypar geometry (i.e., increasing r/S beyond 0), which is consistent with the hydrostatic behavior reported by Wang et al. (2021c). The reduction in M* and V* was attributed to the marked increase in compressive (C) and tensile (T) normal stresses [Fig. 13(b)] facilitated by the double curvature of hypar geometry dramatically increasing the shell’s resistance against out-of-plane deformation (Wang et al. 2020b). This effect is observed clearly via the reduction in maximum shell deflections in Fig. 13(b), which confirms that the assumption of small deformations was valid for umbrellas exhibiting hypar warping when considering the geometries and hydrodynamic loads examined in this study. Evidently, this behavior is highly desirable because the structural design of thin plates and shells is limited primarily by out-of-plane bending and shear effects. From a design perspective, the nominal moment capacity for the 100-mm-thick shell was approximately 41  kNm/m [Fig. 12(d)]. This is significantly lower than the M*=137  kNm/m hydrodynamic moment computed for the flat plate (r/S=0) [Figs. 13(a) and 14(a)], signifying flexural failure of the shell. However, as hypar geometry was introduced commencing with r/S=0.085, the moment demand decreased sharply to 28.9  kNm/m, only 70% of the flexural capacity. As the extent of hypar warping (r/S) further increased, critical shell demands across M*, V*, and N* continued to decrease. Hence, it was demonstrated that umbrellas adopting hypar geometry are not susceptible to bending failure, unlike a flat plate with zero warping. This also applies to shear demands, as is explained in the following section. When considering oblique angles of wave attack on umbrellas with r/S=0.170, bending moment decreased modestly, accompanied by an increase in shear as θw increased [Fig. 13(d)]. Compressive normal stresses also increased initially at θw=15° but decreased at larger angles of incidence [Fig. 13(e)]. A corresponding decrease in maximum shell deflections also occurred. The 1-m-square concrete column supporting the vertex exhibited decreased moment (Mc*), and shear (Vc*) with larger θw [Fig. 13(f)]. However, tensile normal stresses in the shell and axial compression in the column (Nc*) were not affected significantly by θw.
Fig. 13. Maximum moment, shear, and normal forces within the (a) hypar umbrella for r/S between 0 and 0.255 pertaining to θw=0; (b) hypar umbrella for r/S between 0 and 0.255 pertaining to θw=0, and maximum shell displacements; (c) concrete column for r/S between 0 and 0.255 pertaining to θw=0; (d) hypar umbrella for θw between 0° and 45° pertaining to r/S=0.170; (e) hypar umbrella for θw between 0° and 45° pertaining to r/S=0.170, and maximum shell displacements; and (f) concrete column for θw between 0° and 45° pertaining to r/S=0.170. T = tension and C = compression.
Fig. 14. Distribution of moment, shear, and normal forces along u- and v-ordinates pertaining to θw=0 for a shell with (a–c) r/S=0; and (d–f) r/S=0.170. Critical locations are indicated by rectangles. T = tension and C = compression.
Fig. 14 illustrates the distribution of shell moment, shear, and normal demands along the u- and v-ordinates for panels with r/S=0 and 0.170 at θw=0. In shell theory, Mu and Mv respectively denote bending in the of u- and v-directions, with positive moments signifying tension on the shield. Locations at which critical demands manifested (Fig. 14, boxes) migrated from the hinge zone [Figs. 14(a–c)] to the lower panel [Figs. 14(d–f)] with the introduction of hypar geometry, which reflected the behavior reported by Wang et al. (2021c) for purely hydrostatic scenarios. Due to the relatively large compression forces present in the hypar, a simplified yet conservative buckling assessment was undertaken. Fig. 14(f) reveals critical compression forces along the transverse spine and lower longitudinal edge with buckling lengths (Lb) of 3.7 and 3 m, respectively. Euler’s critical buckling load (Pcr) was determined for the transverse spine via the approach proposed by Wang et al. (2020b) for square hypars, and Pcr for the longitudinal edge was obtained from the buckling capacity of the 125-mm parallel flange channel framing the perimeter edge. Comparisons with the critical force (Ncr*) revealed that the umbrella sustained only 16% and 29% of the ultimate buckling capacity at the spine and edge, respectively, which clearly affirmed Candela’s belief in the superior buckling performance of four-sided hypar umbrellas (Candela 1955). The influence of oblique waves relating to θw=45° on hypar stresses for r/S=0.170 is shown in Fig. 15. Comparing these results with those for (θw=0°) [Figs. 14(d–f)] showed that differing angles of wave incidence did not significantly affect the spatial manifestation of critical demands. The decrease in critical moment with θw [Fig. 13(d)] stemmed from a reduction in transverse bending about the longitudinal spine on the lower panel [Fig. 15(a)]. Conversely, elevated shear demands at larger θw resulted from larger shear stresses generated along the transverse base on the right side [Fig. 15(b)] as the direction of wave propagation became more orthogonal with the right half of the umbrella (Fig. 11). Likewise, an increase in the buckling force (Ncr*) along the left longitudinal edge also occurred, but was only 38% of the ultimate buckling capacity (Pcr).
Fig. 15. Distribution of forces along u- and v-ordinates for a shell with r/S=0.170 and θw=45°: (a) moment; (b) shear; and (c) normal. Critical locations are indicated by rectangles. T = tension and C = compression.

Structural Feasibility of Kinetic Umbrellas under Hurricane Waves

Wang et al. (2021c) demonstrated that a kinetic umbrella with r/S=0.170 designed in accordance with Figs. 12(b and c) successfully can resist hydrostatic inundation matching the entire deployed height (i.e., dw=h*=7.25  m). To ascertain its overall structural feasibility against surge and wave attack generated by Hurricane Sandy, all critical hydrodynamic demands pertaining to the umbrella shell and supporting column were compared with those of its hydrostatic counterpart at a given level of static inundation. The hydrodynamic demand ratio (HDR) is introduced
HDR%=Ddyn*D%*
(13)
where Ddyn* and D%* = critical demand of any component (bending, shear, and normal forces) associated with maximum wave impact and hydrostatic inundation dw [Fig. 10(a)], respectively. Note that the percent symbol % in HDR% and D%* represents the percentage of the total deployed height subject to inundation (e.g., D100* = critical hydrostatic demand for complete inundation at 100% of the deployed height). A HDR value less than 1 thus implies that the structure experiences lower forces from the peak dynamic load case than the corresponding hydrostatic scenario. Fig. 16 reports the HDR across all relevant components of the demand for r/S=0.170 compared with the hydrostatic results in Wang et al. (2021c) for inundation equal to 100% of h*, representing the most critical loading condition. All demands resulting from Hurricane Sandy were smaller than their static equivalents at the maximum inundation level (HDR100). Therefore, it is demonstrated that the 100-mm-thick kinetic umbrella reported in this study can be designed successfully to serve as an adaptable flood barrier against combined surge and wave loading attributed to Hurricane Sandy at Monmouth Beach from a structural engineering perspective. Evidently, if larger loads from potentially more-severe storms must be resisted, the shell thickness can be increased accordingly. As an example, increasing the thickness to 150 mm and adopting two layers of 16-mm GFRP bars increased flexural capacity 173%, from 41 to 112  kNm/m.
Fig. 16. Ratio of critical hydrodynamic to hydrostatic demands resulting from inundation at dw=h*=7.25  m pertaining to the shell and column for a 100-mm-thick umbrella with r/S=0.170.

Summary and Conclusions

This paper explored the hydrodynamic response of deployable four-sided hyperbolic paraboloid shells (kinetic umbrellas) via scale-model testing and numerical modeling. Symbolizing the amalgamation between structural art and coastal hazard engineering, such structures can line the coast and can be tilted readily to form a physical barrier against surge-induced coastal inundation (Wang et al. 2020a). Dam-break experiments conducted on scale-model hypar specimens coupled with model and full-scale SPH simulations successfully validated the SPH-FEM technique for the structural analysis of kinetic umbrellas under the hydrodynamic regime. This tool effectively facilitates the analysis of unconventional architectural forms subject to arbitrary hydrodynamic loading under any bathymetric or topographic conditions. the modeling approach was adopted to simulate the behavior of 100-mm-thick 8×8  m square concrete umbrellas subject to combined surge and wave attack at Monmouth Beach, New Jersey, as Hurricane Sandy made landfall in October 2012. Four levels of hypar warping from r/S=0 to 0.255 were examined in addition to oblique angles of wave attack between θw=0° and 45°. Critical demands pertaining to the umbrella shell and concrete column at the instant of maximum wave impact from irregular waves were extracted. The major conclusions are as follows:
Compared with a flat plate (r/S=0), a small warping introduced by hypar geometry (r/S=0.085) dramatically reduces critical out-of-plane bending, shear, and deflections within the shell while significantly increasing in-plane normal stresses. This is attributed to the superior out-of-plane rigidity emblematic of geometries with negative Gaussian curvature, and reflects the behavior of kinetic umbrellas under purely hydrostatic loading (Wang et al. 2021c).
As the level of hypar warping increased further from r/S=0.085 to 0.170, critical moment, shear, and normal stresses and deflections continued to decrease within the shell. Column bending moments also decreased, but exhibited higher axial compression.
For r/S >0.170, large differences in structural demands generally were not observed, except for a decrease in the shell axial force.
For umbrellas with r/S=0.170, waves approaching at larger angles of incidence increased maximum shear forces along the longitudinal base but decreased the critical shell bending moment. Normal compressive stresses initially increased for θw=15° but decreased as the incidence angle increased further. Likewise, bending and shear loads in the column were inversely related to θw over the entire range examined. Normal tensile axial stresses in the shell, and axial forces in the column were only slightly affected by angle of incidence.
All critical demands pertaining to the hypar shell and column for an umbrella with r/S=0.170 subject to hydrodynamic forcing via Hurricane Sandy were lower than their hydrostatic counterparts for static inundation matching the entire deployed height of h*=7.25  m. Building upon the prior work of Wang et al. (2021c), a 100-mm-thick kinetic umbrella therefore can be designed successfully to resist the maximum wave forces attributed to Hurricane Sandy at Monmouth Beach from a structural engineering perspective.
This study ultimately demonstrates the structural feasibility of kinetic umbrellas as an elegant and adaptable solution to conventional monolithic floodwalls and levees. Supplementary to previous work investigating purely the hydrostatic behavior (Wang et al. 2020a, 2021c), we elucidated the remarkable strength of hypar geometries in sustaining hydrodynamic wave impact from landfalling Atlantic hurricanes. Furthermore, this research encourages future studies of kinetic umbrellas as countermeasures against extreme events such as river flooding and tsunamis, thus enabling their worldwide application beyond storm surge protection. For example, kinetic umbrellas effectively could provide a socially conscious alternative to the monolithic floodwalls built throughout Japan which drastically reduced the quality of life of many coastal residents (Kimura 2016). However, additional work is needed to address potential water seepage and the geotechnical design of appropriate substructure systems. Nevertheless, we hope this research will inspire renewed interest in the adoption of thin-shell hyperbolic paraboloidal forms across the architecture and engineering disciplines to address the needs of an increasingly cosmopolitan design environment.

Appendix. Governing Equations of SPH

In SPH, a continuum is discretized into particles exhibiting physical properties including position, velocity, density, and pressure (Liu and Liu 2010). Such quantities for any given particle are computed via the integral interpolant of adjacent particles within its support domain. The extent to which each nearby particle influences a given property is dependent upon the interparticle distance, and a kernel function (W) is adopted to quantify this contribution. The weighted interpolant approximating any quantity field B(r) is
B(r)=ΩB(r)W(rr,hs)dr
(14)
where r is the position vector in R3; r = position of all remaining particles within support domain Ω; and hs = smoothing length. In discrete form, Eq. (14) is expressed as
Bi=jΩimjρjBjWij
(15)
where mj and ρj = mass and density, respectively, of particle j; and Wij=W(rirj,hs) = weighting kernel. In this work, the Wendland (1995) quintic kernel was selected
W(q)=αD(1q2)4(2q+1)for  0q2
(16)
where q = ratio of interparticle distance to smoothing length; and αD=21/(16πhs3). Particle velocities are computed from the Navier–Stokes momentum equation expressed in discrete form via the artificial viscosity scheme (Monaghan 1992)
dvidt=jmj(Pjρj2+Piρi2+Γij)iWij+g
(17)
where vi = velocity; Pi,j = pressure; and g = gravitational acceleration vector. A viscous term Γij representing an artificial viscosity model (Monaghan 1994) is introduced
Γij={αc¯ijμijρ¯ij,vij·rij<00,vij·rij>0
(18)
where
μij=hsvij·rijrij2+ζ2
(19)
vij=vivj and rij=rirj = relative velocity and position of particle, respectively; ζ2=0.01hs2; and c¯ij=0.5(ci+cj) = average speed of sound. The artificial viscosity coefficient α is designed to prevent numerical instability and spurious oscillations (Crespo et al. 2011), for which Altomare et al. (2015) and Barreiro et al. (2013) suggest a value of 0.01 for the consideration of water interaction with coastal structures. Furthermore, changes in the fluid density were obtained via discretization of the continuity equation
dρidt=jmjvij·iWij
(20)
The weakly compressible fluid formulation was adopted such that an equation of state (Batchelor 2000) may be utilized to determine the pressure–density relationship
P=c02ρwκ[(ρρw)κ1]
(21)
where ρw=1,000  kgm3 = fluid reference density; κ=7 (Monaghan 1994); and c0 = artificial speed of sound to suppress density fluctuations within 1% of ρw (Monaghan 2012). Hence, by determining the acceleration of boundary particles constituting the frontal surface of the deployed umbrella, the spatial distribution of fluid forces (fsw) imposed upon the hypar at any given time can be described via
fsw=mbi=1Kdvidt
(22)
where mb = mass of boundary particle; and K = total number of particles forming impacted surface over which fsw acts. Dynamic boundary conditions (Crespo et al. 2007) were adopted in which boundary particles exerted a repulsive force on approaching fluid particles. All boundary particles adopted the same equations of continuity and state as their fluid counterparts, but their positions remained static. Finally, time-stepping was conducted via the Verlet (1967) scheme, implementing a variable time step Δt according to the Courant–Friedrichs–Lewy (CFL) condition (Monaghan and Kos 1999)
Δt=CCFL·min(Δtf,Δtcv),{Δtf=mini(hs/|fi|)Δtcv=mini(hsci+maxj|hsvij  ·  rijrij2+ζ2|)
(23)
where |fi| = force magnitude per unit mass of particle i; and CCFL=0.2.

Data Availability Statement

Some or all data, models, or code that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

Funding for this research was partially sponsored by Princeton University through the Project X grant and the Metropolis Project of Princeton University. The authors gratefully acknowledge Mauricio Loyola, affiliated with the Department of Architecture at the University of Chile, for contributing to the production of Figs. 1 and 8. The authors also thank Daniel J. Ruth, affiliated with the Department of Mechanical and Aerospace Engineering at Princeton University, for assisting with the experimental setup.

References

Abdelkarim, O. I., E. A. Ahmed, H. M. Mohamed, and B. Benmokrane. 2019. “Flexural strength and serviceability evaluation of concrete beams reinforced with deformed GFRP bars.” Eng. Struct. 186 (May): 282–296. https://doi.org/10.1016/j.engstruct.2019.02.024.
Altomare, C., A. J. C. Crespo, J. M. Dominguez, M. Gomez-Gesteira, T. Suzuki, and T. Verwaest. 2015. “Applicability of smoothed particle hydrodynamics for estimation of sea wave impact on coastal structures.” Coastal Eng. 96 (Feb): 1–12. https://doi.org/10.1016/j.coastaleng.2014.11.001.
Altomare, C., J. M. Dominguez, A. J. C. Crespo, J. Gonzalez-Cao, T. Suzuki, M. Gomez-Gesteira, and P. Troch. 2017. “Long-crested wave generation and absorption for SPH-based DualSPHysics model.” Coastal Eng. 127 (Sep): 37–54. https://doi.org/10.1016/j.coastaleng.2017.06.004.
Aureli, F., A. Maranzoni, P. Mignosa, and C. Ziveri. 2008. “Dam-break flows: Acquisition of experimental data through an imaging technique and 2D numerical modeling.” J. Hydraul. Eng. 134 (8): 1089–1101. https://doi.org/10.1061/(ASCE)0733-9429(2008)134:8(1089).
Barreiro, A., A. J. C. Crespo, J. M. Dominguez, and M. Gomez-Gesteira. 2013. “Smoothed particle hydrodynamics for coastal engineering problems.” Comput. Struct. 120 (Apr): 96–106. https://doi.org/10.1016/j.compstruc.2013.02.010.
Batchelor, G. K. 2000. An introduction to fluid dynamics. Cambridge, UK: Cambridge University Press.
Bentz, E., and M. P. Collins. 2001. Response-2000 user manual. Toronto: Univ. of Toronto.
Biscarini, C., S. Di Francesco, and P. Manciola. 2010. “CFD modelling approach for dam break flow studies.” Hydrol. Earth Syst. Sci. 14 (4): 705–718. https://doi.org/10.5194/hess-14-705-2010.
Blake, E. S., T. B. Kimberlain, R. J. Berg, J. P. Cangialosi, and J. L. Beven. 2013. Tropical cyclone report: Hurricane Sandy. University Park, FL: National Hurricane Center.
Candela, F. 1955. “Structural applications of hyperbolic paraboloidical shells.” J. Am. Concr. Inst. 51 (1): 397–416.
Coastal Research Center. 2021. “Profile locations.” Accessed March 24, 2021. https://stockton.edu/coastal-research-center/njbpn/locations.html.
Crespo, A. J. C., J. M. Dominguez, A. Barreiro, M. Gomez-Gesteira, and B. D. Rogers. 2011. “GPUs, a new tool of acceleration in CFD: Efficiency and reliability on smoothed particle hydrodynamics methods.” PLoS One 6 (6): e20685. https://doi.org/10.1371/journal.pone.0020685.
Crespo, A. J. C., J. M. Dominguez, B. D. Rogers, M. Gomez-Gesteira, S. Longshaw, R. Canelas, R. Vacondio, A. Barreiro, and O. Garcia-Feal. 2015. “DualSPHysics: Open-source parallel CFD solver based on smoothed particle hydrodynamics (SPH).” Comput. Phys. Commun. 187 (Feb): 204–216. https://doi.org/10.1016/j.cpc.2014.10.004.
Crespo, A. J. C., M. Gomez-Gesteira, and R. A. Dalrymple. 2007. “Boundary conditions generated by dynamic particles in SPH methods.” CMC-Comput. Mater. Continua 5 (3): 173–184.
De Padova, D., M. Ben Meftah, F. De Serio, M. Mossa, and S. Sibilla. 2020. “Characteristics of breaking vorticity in spilling and plunging waves investigated numerically by SPH.” Environ. Fluid Mech. 20 (Jun): 233–260. https://doi.org/10.1007/s10652-019-09699-5.
Donelan, M. A., J. Hamilton, and W. H. Hui. 1985. “Directional spectra of wind-generated ocean waves.” Philos. Trans. R. Soc. A 315 (1534): 509–562.
Draper, P., M. Garlock, and D. P. Billington. 2010. “Structural optimization of Felix Candela’s Hypar umbrella shells.” J. Int. Assoc. Shell Spatial Struct. 51 (1): 59–66.
Dugan, J. E., D. M. Hubbard, I. F. Rodil, D. L. Revell, and S. Schroeter. 2008. “Ecological effects of coastal armoring on sandy beaches.” Mar. Ecol. 29 (Jul): 160–170. https://doi.org/10.1111/j.1439-0485.2008.00231.x.
Garlock, M. E. M., and D. P. Billington. 2008. Felix Candela: Engineer, builder, structural artist. Princeton, NJ: Princeton Univ. Art Museum.
Goda, Y. 2000. Random seas and design of maritime structures. Singapore: World Scientific.
Grenier, N., D. Le Touze, A. Colagrossi, M. Antuono, and G. Colicchio. 2013. “Viscous bubbly flows simulation with an interface SPH model.” Ocean Eng. 69 (Sep): 88–102. https://doi.org/10.1016/j.oceaneng.2013.05.010.
Horn, D. P. 2015. “Storm surge warning, mitigation, and adaptation.” In Coastal and marine hazards, risks, and disasters. Amsterdam, Netherlands: Elsevier.
Kimura, S. 2016. “When a seawall is visible: Infrastructure and obstruction in post-tsunami reconstruction in Japan.” Sci. Cult. 25 (1): 23–43. https://doi.org/10.1080/09505431.2015.1081501.
Lauber, G., and W. H. Hager. 1998. “Experiments to dambreak wave: Horizontal channel.” J. Hydraul. Res. 36 (3): 291–307. https://doi.org/10.1080/00221689809498620.
Le Mehaute, B. 1976. An introduction to hydrodynamics and water waves. New York: Springer.
Liu, M. B., and G. R. Liu. 2010. “Smoothed particle hydrodynamics (SPH): An overview and recent developments.” Arch. Comput. Methods Eng. 17 (1): 25–76. https://doi.org/10.1007/s11831-010-9040-7.
Madsen, O. S. 1971. “On the generation of long waves.” J. Geophys. Res. 76 (36): 8672–8683. https://doi.org/10.1029/JC076i036p08672.
McGann, C., P. Arduino, and P. Mackenzie-Helnwein. 2011. “SSPbrick element—Openseeswiki.” Accessed October 17, 2018. http://opensees.berkeley.edu/wiki/index.php/SSPbrick_Element.
Medeiros, M. H. F., and P. Helene. 2009. “Surface treatment of reinforced concrete in marine environment: Influence on chloride diffusion coefficient and capillary water absorption.” Constr. Build. Mater. 23 (3): 1476–1484. https://doi.org/10.1016/j.conbuildmat.2008.06.013.
Michiels, T., M. Garlock, and S. Adriaenssens. 2016. Seismic assessment of Félix Candela’s concrete shells and their behavior during the 1985 Mexico City Earthquake. A case study on the Church of Our Lady of the Miraculous Medal. Leuven, Belgium: Taylor & Francis.
Mokos, A., B. D. Rogers, P. K. Stansby, and J. M. Dominguez. 2015. “Multi-phase SPH modelling of violent hydrodynamics on GPUs.” Comput. Phys. Commun. 196 (Nov): 304–316. https://doi.org/10.1016/j.cpc.2015.06.020.
Monaghan, J. J. 1992. “Smoothed particle hydrodynamics.” Annu. Rev. Astron. Astrophys. 30 (1): 543–574. https://doi.org/10.1146/annurev.aa.30.090192.002551.
Monaghan, J. J. 1994. “Simulating free surface flows with SPH.” J. Comput. Phys. 110 (2): 399–406. https://doi.org/10.1006/jcph.1994.1034.
Monaghan, J. J. 2012. “Smoothed particle hydrodynamics and its diverse applications.” Annu. Rev. Fluid Mech. 44 (Jan): 323–346. https://doi.org/10.1146/annurev-fluid-120710-101220.
Monaghan, J. J., and A. Kos. 1999. “Solitary waves on a Cretan beach.” J. Waterway, Port, Coastal, Ocean Eng. 125 (3): 145–155. https://doi.org/10.1061/(ASCE)0733-950X(1999)125:3(145).
Ning, D., R. Wang, L. Chen, J. Li, J. Zang, L. Cheng, and S. Liu. 2017. “Extreme wave run-up and pressure on a vertical seawall.” Appl. Ocean Res. 67 (Sep): 188–200. https://doi.org/10.1016/j.apor.2017.07.015.
Nordenson, C. S., G. Nordenson, and J. Chapman. 2018. Structures of coastal resilience. Washington, DC: Island.
Ozmen-Cagatay, H., and S. Kocaman. 2011. “Dam-break flow in the presence of obstacle: Experiment and CFD simulation.” Eng. Appl. Comput. Fluid Mech. 5 (4): 541–552. https://doi.org/10.1080/19942060.2011.11015393.
Reeve, D., A. Chadwick, and C. Fleming. 2018. Coastal engineering: Processes, theory & design practice. Boca Raton, FL: Taylor and Francis.
Robertson, I. N. 2020. Tsunami loads and effects: Guide to the tsunami design provisions of ASCE 7-16. Reston, VA: ASCE.
St-Germain, P., I. Nistor, R. Townsend, and T. Shibayama. 2014. “Smoothed-particle hydrodynamics numerical modeling of structures impacted by tsunami bores.” J. Waterway, Port, Coastal, Ocean Eng. 140 (1): 66–81. https://doi.org/10.1061/(ASCE)WW.1943-5460.0000225.
Tanimoto, K., and K. Kimura. 1985. A hydraulic experimental study on trapezoidal caisson breakwaters. Technical Note 0528. Japan: Port and Airport Research Institute.
Verbrugghe, T., J. M. Dominguez, C. Altomare, A. Tafuni, R. Vacondio, P. Troch, and A. Kortenhaus. 2019. “Non-linear wave generation and absorption using open boundaries within DualSPHysics.” Comput. Phys. Commun. 240 (Jul): 46–59. https://doi.org/10.1016/j.cpc.2019.02.003.
Verlet, L. 1967. “Computer ‘experiments’ on classical fluids. I. Thermodynamical properties of Lennard-Jones molecules.” Phys. Rev. 159 (1): 98–103. https://doi.org/10.1103/PhysRev.159.98.
Wang, S., M. Garlock, and B. Glisic. 2019. “Kinetic umbrellas for coastal defense applications.” In Proc., IASS Annual Symp. 2019. Barcelona, Spain: International Center for Numerical Methods in Engineering.
Wang, S., M. Garlock, and B. Glisic. 2020a. “Hydrostatic response of deployable four-sided hyperbolic paraboloid shells as coastal armor.” J. Struct. Eng. 146 (6): 04020096. https://doi.org/10.1061/(ASCE)ST.1943-541X.0002619.
Wang, S., M. Garlock, and B. Glisic. 2021a. “Kinematics of deployable hyperbolic paraboloid umbrellas.” Eng. Struct. 244 (Oct): 112750. https://doi.org/10.1016/j.engstruct.2021.112750.
Wang, S., M. Garlock, and B. Glisic. 2021b. “Parametric modeling of depth-limited wave spectra under hurricane conditions with applications to kinetic umbrellas against storm surge inundation.” Water 13 (3): 251. https://doi.org/10.3390/w13030251.
Wang, S., A. Levine, M. Garlock, J. A. Contreras-Jimenez, and J. J. Jorquera-Lucerga. 2020b. “Structural evaluation of Felix Candela’s 8-sided hyperbolic paraboloidal umbrellas.” Eng. Struct. 222 (Nov): 111156. https://doi.org/10.1016/j.engstruct.2020.111156.
Wang, S., V. Notario, M. Garlock, and B. Glisic. 2021c. “Parameterization of hydrostatic behavior of deployable hypar umbrellas as flood barriers.” Thin-Walled Struct. 163 (Jun): 107650. https://doi.org/10.1016/j.tws.2021.107650.
Wei, Z., R. A. Dalrymple, A. Herault, G. Bilotta, E. Rustico, and H. Yeh. 2015. “SPH modeling of dynamic impact of tsunami bore on bridge piers.” Coastal Eng. 104 (Oct): 26–42. https://doi.org/10.1016/j.coastaleng.2015.06.008.
Wendland, H. 1995. “Piecewise polynomial, positive definite and compactly supported radial functions of minimal degree.” Adv. Comput. Math. 4 (1): 389–396. https://doi.org/10.1007/BF02123482.
Young, I. R. 2003. “A review of the sea state generated by hurricanes.” Mar. Struct. 16 (3): 201–218. https://doi.org/10.1016/S0951-8339(02)00054-0.
Zhu, M., and M. H. Scott. 2014. “Modeling fluid–structure interaction by the particle finite element method in OpenSees.” Comput. Struct. 132 (Feb): 12–21. https://doi.org/10.1016/j.compstruc.2013.11.002.

Information & Authors

Information

Published In

Go to Journal of Structural Engineering
Journal of Structural Engineering
Volume 148Issue 5May 2022

History

Received: May 19, 2021
Accepted: Nov 9, 2021
Published online: Mar 14, 2022
Published in print: May 1, 2022
Discussion open until: Aug 14, 2022

Authors

Affiliations

Ph.D. Candidate, Dept. of Civil and Environmental Engineering, Princeton Univ., Princeton, NJ 08544 (corresponding author). ORCID: https://orcid.org/0000-0001-9704-4752. Email: [email protected]
Maria Garlock, Ph.D., F.ASCE [email protected]
P.E.
Professor, Dept. of Civil and Environmental Engineering, Princeton Univ., Princeton, NJ 08544. Email: [email protected]
Assistant Professor, Dept. of Mechanical and Aerospace Engineering, Princeton Univ., Princeton, NJ 08544. ORCID: https://orcid.org/0000-0002-4644-9909. Email: [email protected]
Branko Glisic, Ph.D., M.ASCE [email protected]
Associate Professor, Dept. of Civil and Environmental Engineering, Princeton Univ., Princeton, NJ 08544. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

  • Analytical solutions for the dynamic analysis of a modular floating structure for urban expansion, Ocean Engineering, 10.1016/j.oceaneng.2022.112878, 266, (112878), (2022).
  • A decoupled SPH-FEM analysis of hydrodynamic wave pressure on hyperbolic-paraboloid thin-shell coastal armor and corresponding structural response, Engineering Structures, 10.1016/j.engstruct.2022.114738, 268, (114738), (2022).

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share