Open access
Technical Papers
Mar 28, 2020

Hydrostatic Response of Deployable Hyperbolic-Paraboloid Umbrellas as Coastal Armor

Publication: Journal of Structural Engineering
Volume 146, Issue 6

Abstract

This paper introduces an innovative structural system consisting of four-sided hyperbolic-paraboloid (hypar) roof umbrellas as hard countermeasures against nearshore hazards. The umbrellas line the coast and remain upright during normal operation, providing shade and shelter along the waterfront while not limiting access to the shore. A hinge at the hypar-column interface permits tilting to form a physical barrier against surge-induced coastal inundation. Analytical equations based on idealized boundary conditions are formulated in the hydrostatic regime. The equations provide insight into optimal geometric parameters and are used to validate a decoupled numerical scheme constituting smoothed particle hydrodynamics (SPH) and the finite element method (FEM). All numerical reactions concur with the analytical solutions for water inundation matching the total deployed height. A proof-of-concept study was employed to successfully illustrate the applicability of deployable hypar umbrellas as coastal armor from a structural engineering perspective. This work ultimately demonstrates the feasibility of decoupled SPH-FEM methods in modeling fluid-structure interaction involving hypar forms, while establishing a foundation for their analysis and design for coastal hazard adaptation.

Introduction

The accelerating urbanization of the world’s coastlines coupled with rising sea levels driven by climate change accentuates the necessity of providing adequate countermeasures against the increasing likelihood of coastal hazards (Mooyaart and Jonkman 2017; Nicholls and Cazenave 2010). Shoreline erosion and flooding, storm surges, tsunamis, and extreme waves often impose significant social and economic burdens upon vulnerable and increasingly interdependent communities (Small and Nicholls 2003). In the context of hard countermeasures against hazards threatening open beaches and shorelines, coastal armoring in the form of artificial dikes and seawalls has remained a popular option for centuries (Charlier et al. 2005). While relatively low in cost, traditional forms of armor are frequently associated with numerous adverse social and environmental consequences. Seawalls are often unsightly and restrict public access to beaches while fostering a false sense of security amongst beachfront developers, amplifying habitation density within potentially hazardous zones (Pilkey and Wright 1988). Seawalls are also observed to adversely affect the natural ecosystem, with shorebirds experiencing habitat loss in addition to a reduction in viable prey adjacent to armored shore sections (Dugan et al. 2008). An exploration into alternative adaptable schemes to conventional static infrastructure is therefore desirable for the promotion of social and environmental sustainability within coastal regions.
This paper introduces an innovative structural system in the form of deployable four-sided hyperbolic-paraboloid (hypar) shells as hard countermeasures against nearshore hazards. Inspired by the works of Spanish-Mexican master builder and artist Félix Candela (1955), the proposal seeks to formalize the amalgamation between structural art and coastal hazard engineering. Commonly referred to as umbrellas, such structures arise from the merger of four straight-edged quadrants derived from a hypar manifold exhibiting negative Gaussian curvature [Figs. 1(a and b)]. Since Candela’s initial experimentation in the early 1950s (Fig. 2), architecture embodying the four-sided hypar has graced the roofs of public facilities across the Americas. Typically fabricated from reinforced concrete, the geometric efficiency inherent to the design of each umbrella facilitates its ability to cover an area exceeding 150  m2 while remaining only 40 mm thick (Garlock and Billington 2008). In addition to the popular four-sided form, eight- and even twelve-sided renditions were realized, as exemplified via the roofs of Metro Candelaria in Mexico City. However, such forms were demonstrated to experience increased stresses and deflections relative to their more modest four-sided counterparts (Levine and Garlock 2018). For a detailed architectural account of Candela’s many works involving the hypar umbrella, the reader is referred to Garlock and Billington (2008).
Fig. 1. (a) Quadrant of panel extracted from hypar manifold to create the form of a four-sided umbrella in (b) upright and (c) deployed configuration with location of hinge indicated; and (d) architectural concept of tilting mechanism and a typical panel system in an urban waterfront setting (e) during normal operation (image courtesy of Mauricio Loyola) and (f) under hazard scenarios [images (e and f) courtesy of FotosFortheFuture/Shutterstock.com].
Fig. 2. Candela’s experimental hypar at Las Aduanas. (Image courtesy of Princeton University Library.)
In this study, the typical concept of a four-sided hypar roof structure is modified to incorporate a hinge at the vertex, allowing the umbrella to tilt from its initial upright position [Fig. 1(c)]. When simultaneously deployed in a row, the panels effectively act as a physical barrier to coastal inundation during surge or extreme tidal events [Fig. 1(d)]. This dynamic mechanism introduces an element of flexibility into coastal hazard adaptation strategies not demonstrated by classical shoreline armoring initiatives (Horn 2015). The kinetic panels are intended to remain in their upright position during normal operation, providing shade and shelter along the waterfront while not limiting access to the shore [Fig. 1(e)]. The deployment of this defense system through its transition into a conventional seawall is only initiated during imminent hazard scenarios [Fig. 1(f)], reflecting a seamless conversion between structural art and coastal armor.
Evidently, the feasibility of such a system needs to be evaluated via considering fluid-structure interaction (FSI) in the hydrostatic and hydrodynamic regime. Hence, the primary objective of this paper is to develop and explore the analytical and numerical tools necessary for such assessments. First, the spatial parameterization of a deployed hypar panel as a function of various input parameters delineating its geometric properties is developed. Assuming the presence of idealized restraints at the base and vertex, analytical expressions are derived to quantify the net reactions at these locations under hydrostatic and self-weight demands. Next, the validity and implication of assumed boundary conditions are explored via imposing a condition of zero net uplift at the panel base. Finally, a decoupled numerical scheme constituting smoothed particle hydrodynamics (SPH) and the finite element method (FEM) is proposed to assess the structural response to fluid accumulation in the hydrostatic regime. Boundary reactions arising from the SPH-FEM scheme are subsequently compared against analytical results to ascertain the validity of the numerical procedure. This subsequently enables a proof-of-concept study to be carried out assessing the capabilities of a typical deployed hypar in resisting hydrostatic inundation from a structural engineering perspective. Through this process, the practical feasibility of hypar shells and its analysis via decoupled SPH-FEM numerical methods in the context of coastal armoring solutions are explored. The scope of the research is limited to hydrostatic and gravity loading, while hydrodynamic and wind loading will be addressed in future work.

Theoretical Structural Analysis

Geometric Description

Prior to the commencement of structural analysis, it is important to characterize the geometry of a tilted hypar shell via a mathematically consistent approach. The Cartesian coordinates {x,y,z}, with corresponding orthonormal ordinates {x^,y^,z^}, of any given four-sided singly symmetric hypar panel in its deployed configuration (Ωp) are parameterized via a set constituting seven fundamental geometric parameters, Si={b1,b2,c,r,h,tpv,tpe} illustrated in Fig. 3, where
b1 = longitudinal dimension of the lower panel, see Fig. 3(a), (b1>0);
b2 = longitudinal dimension of the upper panel, see Fig. 3(a), (b20);
c = transverse dimension of half the panel width, see Fig. 3(b), (c>0);
r = rise of the vertex on the concave side, see Figs. 3(a–c), (r0);
h = vertical height of the vertex on the convex side above ground, see Fig. 3(c), (0<hb1);
tpv = projected thickness of the panel at the vertex, see Figs. 3(a and b), (tpv0); and
tpe = projected thickness of the panel at the edges, see Figs. 3(a and b), (tpe0).
Fig. 3. Geometric parameters of deployed hypar panel along (a and c) longitudinal and (b) transverse spine with (d) planar projection of the shield also illustrated. Geometric parameters constituting Si are boxed.
Fig. 3 illustrates the geometry of an arbitrary tapered hypar characterized by Si with such parameters accentuated within boxes. Furthermore, the local basis {u,v,w} defines the planar projection of the surface of inclination Πuv[c,c]×[0,b1+b2] as shown in Figs. 3(a and d). The angle between Πuv and the horizontal plane is referred to as the angle of inclination and is denoted as θ1* as shown in Fig. 3(c). Various special terminologies are also introduced which are referenced throughout this paper. The concave (front in contact with fluid) and convex (rear) surfaces are referred to as the shield and soffit, respectively. Coordinates constituting such surfaces are henceforth denoted, respectively, by Πk and Πm, as indicated in Fig. 3(c). In addition, the deployed panel is divided into lower and upper sections separated by the transverse spine [Fig. 3(b)]. The term β{1,2} is thus adopted to distinguish between the two sections as highlighted in Figs. 3(a and c). In addition, Ωp is mapped via the orthogonal ratios λ1,λ2,λ3[0,1], increasing along ±u, v, and w, respectively, as shown in Figs. 3(b and d). This ensures that the full geometry can be easily represented via a 3D grid of structured cells which streamlines its application within numerical models as implemented within this paper. The reader is referred to Appendix I for a complete derivation pertaining to the geometric parameterization of Ωp by any given set Si via λ discretization.

Derivation of Boundary Reactions

Relations are derived for net reactions at the vertex and base hinges for a deployed panel with geometry delineated via Si under self-weight and hydrostatic loading on Πk. The structural system considered in the analysis is presented in Fig. 4. To consider the highest loading scenario, the relations do not consider potential fluid accumulation on Πm. To simplify the procedure, the contribution of normal force to deformation is neglected as commonly accepted in structural analysis (Hibbeler 2008). The vertex hinge is idealized as unrestrained in x to simulate a flexible support offering minimal lateral resistance, while the base is continuously pinned along y. This effectively reflects a panel supported at the vertex by a vertical column exhibiting axial stiffness much greater than that in the lateral direction. Additionally, the approach ensures that flexural failure at the column base would not result in global failure of the deployed system. The validity of a flexible vertex design philosophy is to be examined in the context of uplift suppression at the soffit base. The effective height of water submersion measured from the base, hw0 [Fig. 3(c)], is introduced as an additional parameter for the quantification of hydrostatic forces. The net reactions at the base (fb) and vertex (fv) hinge under combined hydrostatic and self-weight loading adhering to the flexible support condition are analytically derived via principles of static equilibrium. Appendixes II and III detail the derivation pertaining to the hydrostatic (fvw,fbw) and self-weight (fvg,fbg) components, respectively. The total net reactions are subsequently encapsulated via
fv=fvg+fvw,fb=fbg+fbw
(1)
Fig. 4. Simplified schematic (ignoring panel thickness) of kinetic umbrella in (a) original and (b) deployed state with idealized restraints and boundary reactions shown for (c) non-uplift and (d) uplift conditions via principles of graphic statics.
Figs. 4(a and b) depict an upright and deployed hypar, respectively, with their idealized boundary conditions shown in Figs. 4(c and d). Supplementary to Appendixes II and III, the concept of panel uplift at the base is illustrated by applying graphic statics to resolve the force polygon associated with the total hydrostatic and self-weight applied force resultant (fw+fg), and reactions at the base (fb) and vertex hinges (fv). If the applied force resultant vector intersects the vertical reaction vector below z=0, uplift is suppressed [Fig. 4(c)]. A line must be drawn between that point of intersection and the base support to maintain equilibrium. This line represents the direction of the reaction vector at the base where the vertical component fb·z^>0. Conversely, should the intersection of the resultant force be above z=0 as shown in Fig. 4(d), fb·z^<0 occurs and the presence of vertical restraints are necessary along the panel base.

Parametric Investigation

It is evident that if fb·z^<0, uplift is expected to occur at the panel base, assuming no vertical restraints are present. Hence, the relationship between panel geometry and net vertical reactions at the soffit base as a function of inundation height (hw) is subsequently explored. This effectively enables exploration into the conditions governing panel uplift at ground level, allowing for the optimization of geometric forms such that uplift can be minimally suppressed. A parametric study was implemented, the parameters of which are summarized in Table 1 and include the ratio between the projected lower-to-upper longitudinal dimension of the panel (b1/b2), along with changes to the rise-to-area ratio (r/A), with A denoting the total area of the inclined surface Πuv. In this study, the total projected area A of the hypar was held constant at 64  m2, reflecting a square with sides of 8 m. In addition, b1/b2 was varied from 1 to 4, corresponding to the lower projected length b1 increasing from 4 to 6.4 m and b2 decreasing from 4 to 1.6 m in response. Values of r*/A were selected between 0 and 0.047  m1, inferring a total vertex rise r* from 0 to 3 m, respectively. Note that Candela recommends a rise/area ratio greater than 0.015  m1 for the minimization of short- and long-term deflections of upright concrete hypar umbrellas (Garay 1994). Furthermore, three angles of inclination (θ1*) constituting 55°, 65°, and 75° were adopted in addition to three inundation heights hw for each level of tilt. hw was designated as a proportion of the maximum effective panel height (h*) with 0.5 h*, 0.75 h*, and h* considered. For simplicity, the panel thickness was neglected with tpv=tpe=0. Hence, only hydrostatic demands were examined via the parametric investigation while ignoring the influence of self-weight and shell-thickness gradients. This effectively leads to a conservative assessment of inundation-induced panel uplift. Table 1 summarizes the parameters and the range of their respective values implemented for the parametric investigation.
Table 1. Parameters adopted for parametric study involving panel uplift
Parameter variableUnitsRange of values
r/Am100.047
rm0.2n;n{0,1,2,,15}
Am264 (constant)
b1/b2Dimensionless14
b1m4+0.16n;n{0,1,2,,15}
b2m40.16n;n{0,1,2,,15}
θ1*Radians (shown as degrees){55°,65°,75°}
hwm{0.5h*,0.75h*,h*}
The net vertical reaction at the panel base fb·z^ was computed via Eq. (1) for the geometries and inundation heights considered within the parametric study. Fig. 5 depicts critical uplift lines reflecting the geometric configuration in terms of r/A and b1/b2 for a given submersion height and angle of inclination such that the net vertical base reaction is zero. This implies that the hydrostatic resultant from Fig. 4 intersects the z-axis at z=0. If one begins with a design constraint of hw and θ1*, combinations of r/A and b1/b2 to the right of any given line thus encompass the entire range of hypar geometries to which a laterally unrestrained vertex hinge would not result in panel uplift. If one begins with a design geometry of r/A and b1/b2, then critical uplift lines to the left (holding b1/b2 constant) represent hw and θ1* values that would not result in uplift. As an example, if Candela’s limit of r/A=0.015  m1 were adopted, for b1/b21.25, uplift is avoided for θ1*=55° up to hw=0.75h* or θ1*=65° up to hw=0.5h*. Furthermore, it is observed that the critical uplift lines exhibit a rightward shift with increasing inundation heights and angles of inclination, constraining the range of geometric combinations available to prevent uplift. For a constant b1/b2, increasing the r/A ratio appears to reduce the uplift potential of the panel for any given θ1* and hw. This is attributed to greater concavity of the shield maximizing the gravity component of the hydrostatic demand acting against the uplift tendency of the deployed hypar. Likewise, an increase in the b1/b2 ratio while maintaining a constant r/A also aids in the suppression of panel uplift due to the greater volume of water acting on the lower shield (β=1). Fig. 5 therefore informs the selection of geometric parameters for a squarely projected hypar pertinent to the parametric study such that hydrostatic uplift is avoided. If such conditions are violated, additional structural modifications may include the provision of lateral rigidity at the vertex (via a diagonal strut or otherwise) and/or vertical restraints along the base.
Fig. 5. (a) Critical uplift lines denoting the geometric limit to which fb·z^0 associated with a laterally unrestrained vertex hinge for angles of inclination 55°, 65°, and 75° and inundation heights of 0.5 h*, 0.75 h*, and h* for various ratios of r/A and b1/b2 pertaining to the parametric study; and (b) schematic illustrating the relationship between hw and θ1*.

Numerical Modeling and Validation

The analytical boundary reactions and transverse fluid forces obtained via Eqs. (1) and (37) are utilized to validate a decoupled SPH-FEM numerical model in the hydrostatic regime. A doubly symmetric square hypar with tpv=tpe=89  mm, exhibiting sides of 8 m and rise (r) 1.92 m with the vertex hinge 3 m above ground level, was adopted. This resulted in a rise/area ratio (r/A) of 0.03  m1, b1/b2 ratio of 1, and 68° as the angle of inclination (θ1*). The hypar geometry considered was inspired by Candela’s 1953 experimentation at Las Aduanas (Fig. 2), albeit with an increased rise at the vertex. Note that the thickness of this particular experimental hypar is larger than that of Candela’s typical designs (Garlock and Billington 2008). The open-source CUDA-enabled SPH solver DualSPHysics (Crespo et al. 2015) was implemented for the spatial discretization of fluid forces acting upon the panel, facilitating the subsequent determination of boundary reactions via a structural model executed within the Open System for Earthquake Engineering Simulation (OpenSees version 2.5.0) FEM software (Mazzoni et al. 2006). Through justifying the validity of the SPH-FEM scheme under hydrostatic conditions, a solid foundation can be established for future simulations involving hydrodynamic fluid-structure interaction of deployed hypar forms responding to a range of coastal phenomena.

Smoothed Particle Hydrodynamics

Smoothed particle hydrodynamics (SPH) is a mesh-free Lagrangian particle method commonly adopted for the simulation of FSI involving free surface fluid behavior (Liu and Liu 2010). While mesh-based techniques such as FEM have been utilized to model FSI problems in isolation (Zhu and Scott 2014), expensive considerations such as mesh rezoning are often required to accommodate excessive deformations, potentially introducing additional errors (Liu and Gu 2005). With SPH, all geometries are nonconstrained irrespective of complexity and the extent of their evolution beyond initial conditions (Crespo et al. 2008), which reinforces its applicability toward coastal engineering applications (Altomare et al. 2015).
In this study, SPH was implemented via DualSPHysics to model hydrostatic fluid accumulation on the concave surface of the deployed hypar. The fluid continuum was represented as a set of discrete particles, each exhibiting physical properties including position, velocity, density, and pressure. 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, to which a kernel function (W) was adopted to represent this contribution. The kernel function is governed by a smoothing length (hs) which delineates its zone of influence and should be greater than the initial particle separation (Barreiro et al. 2013). The weighted interpolant approximating any quantity field B(r) is thus expressed as
B(r)=ΩB(r)W(rr,hs)dr
(2)
where r = position vector; and r denotes the position of all remaining particles within the support domain Ω. Eq. (2) may subsequently be expressed in discrete form to compute any quantity B for particle i as
Bi=jΩimjρjBjWij
(3)
where mj and ρj = mass and density of particle j, respectively; and Wij=W(rirj,hs) is the weighting kernel. Wij governs the interaction between adjacent particles and must be selected in strict accordance with a prescribed set of conditions (Liu et al. 2003; Liu and Liu 2010). In this work, the quintic kernel by Wendland (1995) was adopted:
W(q)=αD(1q2)4(2q+1)for  0q2
(4)
where q = ratio between the interparticle distance and smoothing length; and αD=21/(16πhs3) for analyses in R3. Particle clumping is avoided via use of the tensile instability correction developed by Monaghan (1999). The determination of fluid forces results from changes in the velocity field stipulated by the conservation of linear momentum. This was implemented via the popular artificial viscosity scheme developed by Monaghan (1992):
dvidt=jmj(Pjρj2+Piρi2+Γij)iWij+g
(5)
where vi = particle velocity; Pi,j = pressure; and g = gravitational acceleration vector. Γij is the viscous term representing an artificial viscosity model (Monaghan 1994):
Γij={αc¯ijμijρ¯ij,vij·rij<00,vij·rij>0
(6)
with
μij=hsvij·rijrij2+ζ2
(7)
vij=vivj and rij=rirj, respectively denote the relative velocity and position of the particle; ζ2=0.01hs2; and c¯ij=0.5(ci+cj) is the average speed of sound. For CFD applications regarding water, it is suggested that α falls between 0.01 and 0.1 (Dalrymple and Rogers 2006; Monaghan 1994). The choice of α is designed to prevent numerical instability (Crespo et al. 2011), to which Altomare et al. (2015) and Barreiro et al. (2013) confirm that 0.01 is the most suitable value for the consideration of fluid interaction with coastal structures. In addition, changes in the fluid density were determined via the equation governing the conservation of mass:
dρidt=jmjvij·iWij
(8)
As such, the weakly compressible assumption for fluids was adopted such that an equation of state may be implemented to determine the pressure-density relationship. The following equation of state for water enabled the determination of fluid pressure via its density (Batchelor 2000):
P=c02ρwκ[(ρρw)κ1]
(9)
where ρw=1,000  kg/m3 is the fluid reference density; κ=7 (Monaghan 1994); and c0 is an artificial speed of sound to suppress density fluctuations within 1% of ρw (Monaghan 2012). This ensures that the fluid behavior will not deviate significantly from the incompressible approach. Through the resolution of Eq. (5), the spatial distribution of fluid forces imposed upon boundary particles constituting the shield of the deployed panel at any given timestep was determined via
fsw=mbi=1Kdvidt
(10)
where mb = mass of a boundary particle; and K = total number of particles forming the impacted surface over which fsw acts. In this work, dynamic boundary conditions proposed by Crespo et al. (2007) were adopted where boundary particles exert a repulsive force on approaching fluid particles. All boundary particles exhibit the same equations of continuity and state as their fluid counterparts but are fixed in position. The repulsive force is generated by an increase in pressure as the distance between boundary and fluid particles fall within the kernel threshold (2hs), effectively preventing fluid penetration.
The SPH computational domain in this study consists of a box 6 m long in x, 8 m wide in y, and 10 m tall in z. The concave side of the tilted panel Πk in contact with the fluid was discretized into a collection of quadrilaterals with corner coordinates derived per Eq. (23) from Appendix I and positioned 0.5 m from the front wall of the domain with the origin x0={0.5,0,0} as shown in Fig. 6. The height of fluid inundation (hw) was 7.4 m, matching the total effective height of the deployed panel (h*). An initial interparticle spacing (dp) of 0.1 m was implemented along with an artificial viscosity constant α=0.01. A total 15 s of physical time was simulated in order to provide sufficient time for initial particle oscillations to dampen (Fig. 6). The x, y, and z components of the total hydrostatic force fsw acting on each quadrilateral s constituting Πk were subsequently determined via the time averaging of fluid forces over the final 7.5 s of the simulation as demonstrated for a typical cell in Fig. 6. Time-stepping was conducted via the computationally efficient Verlet algorithm (Verlet 1967), with each time step determined according to the Courant-Friedrichs-Lewy condition (Monaghan and Kos 1999). The simulation was performed within DualSPHysics utilizing a NVIDIA GeForce GTX 1060 GPU with 1280 CUDA cores (Santa Clara, California).
Fig. 6. Illustration of SPH computational domain for deployed hypar discretized into 64 cells (with 16 cells per quadrant) and the determination of time-averaged hydrostatic force components acting on a typical cell.

Finite Element Modeling

The computation of hydrostatic demands via SPH subsequently enables the determination of reaction forces at the vertex and base supports via finite element analysis. OpenSees was implemented for evaluation of the structural response in which the hypar continuum was modeled via eight-node hexahedral stabilized single-point integration brick elements (SSPbrick), eliminating the potential for volumetric or shear locking (McGann et al. 2011). Each SPH cell was represented using four brick elements [Figs. 7(a and b)] where the eight corner nodes of each element occupied coordinates determined via Eq. (25) in Appendix I. Hence, each element exhibits four nodes on both Πk and Πm. The individual Cartesian components of the hydrostatic force on each SPH cell fsw were subsequently applied to the center node on Πk of each four-element group constituting a given cell [Fig. 7(c)]. Due to symmetry of the panel about the xz plane, only half of the hypar extending in the positive y ordinate was modeled. As such, nodes forming the longitudinal spine were translationally constrained in the y direction, effectively restricting rotation about z. The implementation of this boundary constraint was shown to be applicable for the modeling of symmetrical hypar shells as demonstrated by Levine (2018). Furthermore, the vertex node was free to translate only in x while all nodes at the base were fully translationally constrained, permitting only rotation about y.
Fig. 7. (a) Discretization of SPH cells implemented in DualSPHysics; (b) translated into OpenSees brick elements in which half a panel was modeled via FEM; (c) application of hydrostatic force vectors within the finite element model; and (d) the nodal distribution of panel self-weight. Note that this figure pertains to 16 SPH cells per quadrant.
For the determination of panel self-weight, each brick element was discretized into eight sectors, with the volume of each sector approximated via computation of the minimum volume encapsulating the interpolated coordinates of its eight corner positions [Fig. 7(d)] using the convex hull algorithm. The total computed volume of the panel arising from the summation of all sector volumes was then compared against the true theoretical hypar volume (Vp) from Eq. (46) in Appendix III. This allows for the determination of a global volume correction factor γc:
γc=Vp2i=1Cci
(11)
where ci = volume of sector i derived via the convex hull; and C = total number of element sectors constituting the entire half-hypar, equal to eight times the total number of brick elements. The self-weight force vector of a given sector is therefore γcciρsg and was assigned to the corner node of the brick element pertaining to sector ci, as shown in Fig. 7(d). The mechanical properties of the panel material matching that of typical structural concrete were adopted, with density ρs=2,400  kg/m3, Poisson’s ratio ν=0.15, and Young’s modulus E=30,000  MPa. Implementation of the ElasticIsotropic OpenSees material thus results in panel behavior constrained entirely within the elastic regime. Note that while concrete properties were adopted for the validation study, the most ideal material for such structures remains under investigation.

Validation of Numerical Scheme

The decoupled SPH-FEM scheme was validated via comparing the net reactions at the vertex and base to their analytically derived counterparts per Eq. (1). The total hydrostatic demand imposed along the positive y ordinate was also evaluated against the theoretical result given by Eq. (37) in Appendix II. The validation was carried out by means of a convergence study in which the total number of SPH cells constituting a panel quadrant was increased from 1 to 49, corresponding to an increase in the number of FEM elements from 4 to 196. In addition, the deflection at five locations on the upper panel was evaluated as a function of mesh resolution to obtain the minimum cell count necessary for adequate convergence.
Fig. 8(a) summarizes the relative error between reaction forces produced via the numerical scheme and solutions to the analytical expressions derived in the appendices as a function of modeling resolution. Particular forces constituted the total horizontal base reaction fb·x^ (Base-X), total vertical base reaction fb·z^ (Base-Z), and the vertex hinge vertical reaction fv·z^ (Vertex-Z) as shown in Fig. 4. The total transverse hydrostatic force acting on each half of the deployed hypar fhalfw·y^ (Half Panel-Y) per Eq. (37) in Appendix II was also included. As expected, Fig. 8(a) reveals that increasing the resolution of the numerical model generally enhances the overall convergence between the analytical and numerical results. When adopting 36 SPH cells per quadrant (corresponding to 144 brick elements), the forces computed via the SPH-FEM scheme were Base-X = 2,120  kN, Base-Z = 857  kN, Vertex-Z=2,143  kN, and Half Panel-Y = 310 kN. This compares well to the analytically derived values of 2,149, 832, 2,127, and 296 kN, respectively, exhibiting differences of 1.35%, 3.00%, 0.75%, and 4.73%. The assumption of small deformations pertaining to the deformed structural configuration is therefore satisfied. In addition, total OpenSees runtimes were extracted for each modeling resolution [Fig. 8(b)]. The relative increase in FEM runtimes (all less than 1 s) were seen to be negligible compared to the 30 min required to process SPH simulations across all cell discretizations. Note that increasing the number of SPH cells while maintaining a fixed initial interparticle spacing of dp=0.1  m does not alter the total number of particles, thus resulting in a constant processing time. As a further point of comparison, the delta-SPH formulation (Molteni and Colagrossi 2009) introducing a diffusive term into the continuity equation [Eq. (8)] was also investigated in addition to modifications to the artificial viscosity constant at the fluid-boundary interface. The recommended delta-SPH coefficient (δΦ) of 0.1 was considered alongside a viscosity scaling factor (fvb) of 10 to compare against the default implementation of δΦ=0 and fvb=1 for 36 SPH cells per quadrant. Fig. 9(a) reveals that under hydrostatic conditions, the incorporation of δΦ and/or changes to fvb does not have a significant effect on the synchronicity between numerical and analytical results. However, an overall reduction in relative error is produced by reducing the initial interparticle spacing as illustrated via Fig. 9(b). By adopting dp=0.06  m, the maximum error in absolute terms was only 0.89%, compared to 4.73% for dp=0.1  m. Nevertheless, this level of detail results in a runtime over seven times that of dp=0.1  m at 214 min due to the increase in overall particle count from 185,025 to 786,133.
Fig. 8. (a) Comparison between analytical and numerical results in terms of relative error for the total horizontal and vertical base reaction (Base-X and Base-Z), vertical vertex reaction (Vertex-Z), and total transverse hydrostatic demand on each half of the panel (Half Panel-Y) as a function of the number of SPH cells and FEM elements adopted per quadrant; and (b) OpenSees runtime for each resolution.
Fig. 9. (a) Relative error (expressed in absolute terms) between numerical and analytical results for various SPH cases concerning δΦ and fvb for dp=0.1  m with 36 SPH cells per quadrant adopted; and (b) relative error and DualSPHysics runtime across varying dp for δΦ=0 and fvb=1.
The implementation of 36 SPH cells per quadrant of the squarely projected hypar is also observed to result in the convergence of nodal deflections taken at five locations on the panel computed via OpenSees (Fig. 10). Because analytical or experimental results pertaining to panel deflections under hydrostatic loading are not available, the convergence study provides insight into the optimal cell resolution required to achieve an acceptable deformation profile for square hypars. Thus, the decoupled SPH-FEM numerical scheme proved capable in assessing the behavior of deployed hypar panels under complete fluid inundation in the hydrostatic regime. This validation effectively reinforces the applicability of such models for future FSI studies involving hydrostatic and hydrodynamic actions on hypar surfaces.
Fig. 10. (a) Absolute numerical nodal deflections at five locations on the deployed hypar as a function of SPH cells and FEM elements per quadrant; and (b) nodal locations are shown where P5 is always located in the center of the quadrant irrespective of the number of cells/elements.

Proof of Concept

The feasibility of tilted hypar shells as coastal armor is subsequently evaluated within a structural engineering context. The objective is to provide preliminary justification for the viability of such structures as hard countermeasures against coastal hazards. Assuming traditional concrete construction of both the hypar panel and vertex support column, reinforcement and member sizes appropriate to resist hydrostatic inundation characteristic of an extreme surge event are examined. This effectively establishes a baseline for future work involving the refinement and detailed analyses of such designs in response to a wide range of coastal loading phenomena.

Geometry and Hydrostatic Demand

The structural geometry modeled via the SPH-FEM scheme within the previous section was selected as characteristic of a typical deployed hypar. For this exercise, however, the projected thickness tpv and tpe was increased to 116 mm to obtain a true thickness of 116cosα*=105  mm along the spine [where α*=αβ|λ1=0=25.6° per Eq. (14) in Appendix I]. This modification ensures that a minimum clear cover of 1.5″ (38 mm) can be achieved when considering a single mesh composed of #4 reinforcement within the panel as stipulated by ACI 318 (ACI 2014) for exposed structures. The height of hydrostatic inundation hw was selected based on observations of historical storm surge and storm tide inundation events impacting the east coast of the United States. Fig. 11 summarizes 70 observations of the maximum storm surge or storm tide inundation heights associated with individual hurricane events from 1899 to 2012 (Needham et al. 2015). The region captured encompassed coastal zones extending from Massachusetts to Florida. Note that all recorded inundations were below the maximum deployed panel height of 7.4 m, with mean and median heights of 2.5 and 2.3 m, respectively. A conservative fluid inundation height of 0.75h* equating to 5.6 m was henceforth adopted for the application of hydrostatic loading on the tilted hypar. This was attributed to Fig. 11(b) revealing only a single event (associated with Hurricane Wilma in 2005) exceeding this threshold over the entire observational period [Fig. 11(b)]. While ongoing climate change may result in future surges of greater severity (Needham et al. 2015), the assumption of data stationarity is adopted for the proof of concept. A submersion height of 5.6 m may therefore conservatively capture the extent of coastal inundation expected from storm surges and storm tides impacting the eastern United States.
Fig. 11. (a) Histogram; and (b) time series summarizing all recorded maximum storm surge and storm tide inundation heights observed on the United States’ east coast from 1899 to 2012. (Data from Needham et al. 2015.)

Material Properties and Member Capacities

A 500-mm square reinforced concrete column was adopted to support the hypar considered for the proof of concept. The dimensions are such that slenderness effects may be neglected per ACI 318 (ACI 2014). Structural concrete with a compressive strength (fc) of 40 MPa was implemented for the column and panel, incorporating A706 Grade 80 reinforcing steel. The yield and ultimate strength of G80 reinforcement were taken as 552 and 689 MPa (80 and 100 ksi), while the strain at the onset of strain hardening and the ultimate tensile strain were 0.0074 and 0.0954, respectively (Overby et al. 2015). Longitudinal reinforcement within the column consisted of eight #7 bars evenly positioned around the perimeter confined with #4 stirrups at 200 mm spacing. The resulting 54 mm of clear cover to transverse reinforcement thus complies with durability requirements per ACI 318. The hypar shell was reinforced with a mesh of #4 bars spaced at approximately 170 mm in the longitudinal and transverse direction located about the neutral axis of the cross section.
The panel capacity was assessed in terms of the bending moment demand associated with hydrostatic loading acting on the shield. A 3D finite element model was created within OpenSees based upon the implementation of 36 SPH cells per panel quadrant. For the computation of shell-bending moments, 144 element sets were used for each quadrant, where each set was represented by 10 layers of SSPbrick elements, effectively translating to 10 integration points through the thickness of the panel [Fig. 12(b)]. This was achieved by discretizing λ3 into 10 uniformly spaced increments (Appendix I), giving a total of 144×10=1,440 elements per quadrant. The bending moment per unit width along the u and v ordinates, respectively denoted Mu and Mv, were determined for each element stack as follows:
[MuMv]=ts2ts2w[σuσv]dwtsni=1n[σiuσiv]wi
(12)
where ts = projected thickness along the integration points; w = distance along the positive w ordinate measured from the neutral axis; and σu and σv = components of stress resolved along u and v, respectively. The integral in Eq. (12) was discretized within OpenSees, with n=10 being the number of SSPbrick integration points through the thickness, wi denoting the w component of a vector from the neutral axis to integration point i, and
σiu=u^·(Tiu^),σiv=v^·(Tiv^)
(13)
in which Ti = Cauchy stress tensor evaluated at integration point i. Positive values of Mu and Mv were selected to reflect tension on the shield, hence i increases along the negative w ordinate as shown in Fig. 12(b).
Fig. 12. (a) OpenSees implementation of hypar model with (b) distribution of SSPbrick integration points through the panel thickness; and (c) maximum enveloped bending moment distribution along longitudinal and transverse ordinates for the proof-of-concept study.
As only the left half of the deployed hypar (extending along the negative u ordinate) was modeled, the same boundary conditions as those implemented for the validation case in the previous section were adopted. To account for the column, a linear spring with stiffness kv=3EIc/h3 was implemented at the vertex hinge, to which Ic represents half the second moment of area of the square column cross section [Fig. 12(a)]. This allows for the determination of axial, shear, and moment demands associated with the column due to hydrostatic inundation. Fig. 12(c) illustrates the maximum enveloped bending moment distribution of the panel along both the transverse (u) and longitudinal (v) ordinates determined per Eq. (12). Note that the spatial moment distribution computed via OpenSees was independently verified using the commercial finite element software SAP2000 via Mindlin-Reissner shell approximations. Moments associated with the area immediately adjacent to the vertex node [enclosed by dashed lines in Fig. 12(c)] were ignored because the boundary condition does not account for the tributary area associated with the vertex hinge and will be highly overestimated at this location. The design and detailing of the hinging mechanism will be investigated as part of future research and is beyond the scope of this paper. The maximum bending moment averaged over 1 m (M¯p*) was subsequently determined to be 14  kN·m/m from Fig. 12(c). In addition, the axial (Nc*=fv·z^), shear (Vc*=fv·x^), and moment (Mc*=Vc*h) demands attributed to the column were 1,013 kN, 47 kN, and 141 kN·m, respectively. The small shear relative to the axial demand thus affirms the assumption of a laterally unrestrained vertex hinge for purposes of structural analysis of the hypar shell per Fig. 4.
The nonlinear sectional analysis program for reinforced concrete, Response-2000 (Bentz and Collins 2001), was utilized to determine the theoretical capacities of the hypar shell and supporting column. The layout of reinforcement within the 500-mm square concrete column and a 1-m segment of the 105-mm-thick panel as previously described are illustrated in Figs. 13(a and b), respectively. The trilinear moment-curvature relationship characteristic of reinforced concrete is also shown, which was produced via sectional analyses implemented within Response-2000, including considerations of axial-shear interaction for the column. The yield moment (Mn) of the concrete column and hypar shell were subsequently computed to be 494 kN · m and 18.6  kN·m/m, respectively, exceeding their respective moment demands of 141 kN · m and 14  kN·m/m. Hence, the detailing of structural members adopted for the proof-of-concept study proved adequate for the hydrostatic forces exerted on the deployed hypar. The conceptual feasibility of such structures for use as coastal armor against storm surge inundation in terms of a simplified hydrostatic situation is therefore demonstrated. Future studies will involve the geometric optimization of hypar forms for pertinent FSI scenarios as well as hydrodynamic loads.
Fig. 13. (a) Moment-curvature relationship with yield capacity indicated for reinforced concrete column; and (b) hypar panel segment adopted for the proof-of-concept study.

Discussion

New analytical and computational tools were introduced for the analysis of tilted hypar shell structures under hydrostatic loading. The exploration of boundary reactions at the vertex and panel base associated with any geometric configuration defined by Si in response to fluid inundation of a given height effectively paves the way for future work into the design of appropriate support structures, including the detailing of a robust hinging mechanism. Analytical solutions to the hydrostatic problem derived within this paper also mathematically define the conditions promoting panel uplift, thus informing the implementation of appropriate countermeasures to suppress such behavior. In addition, analytical boundary reactions were adopted to successfully validate a decoupled SPH-FEM numerical scheme implemented within DualSPHysics and OpenSees. This promotes a relatively simple computational framework for the analysis of hyperbolic paraboloidal geometries in the context of FSI within an entirely open-source computing environment. The ability of SPH methods to accurately simulate complex wave and coastal phenomena (Barreiro et al. 2013) can therefore be integrated with complex hypar forms described via FEM models (Levine and Garlock 2018) to obtain detailed panel behavior for a wide range of loading and geometric scenarios.
Furthermore, a proof of concept involving the preliminary design of a typical panel and supporting column was considered to assess the structural viability of tilted hypar systems in coastal hazard engineering. A modified umbrella inspired by one of Candela’s earliest prototypes (Fig. 2) inclined at 68° was subjected to 5.6 m of hydrostatic inundation to conservatively mimic extreme surge conditions on the United States’ east coast. Reinforced concrete was assumed for both the panel and supporting column, to which bending demands in the shell were computed via tensor manipulation from OpenSees data. The linear spring approximation in representation of the column stiffness also validates the laterally unrestrained simplification at the vertex hinge as assumed in the appendices. Through the subsequent determination of panel and column capacities via Response-2000, the study successfully illustrates the preliminary feasibility of deployable hypar shells as an armoring solution against coastal inundation.
However, the study detailed within this paper exhibits several inherent limitations. The analytical relations pertaining to reactions at the vertex and base reflected via Eq. (1) were derived under the postulation of rigid body behavior for a static structure. This characteristic also extends to the spatial and temporal distribution of fluid forces extracted via SPH. Hence, the assumption of small local deformations must hold under the considered fluid loading for a given geometry. Global displacement of the panel must also remain small such that the displaced form does not significantly deviate from its initial configuration before the application of hydrostatic or hydrodynamic loads. Such assumptions appear to be valid for the hypar form investigated within the validation study as demonstrated in Fig. 10. Thus far, hydrodynamic considerations have not been incorporated into the analysis procedure. Complex phenomena associated with coastal hazards such as dynamic wave-structure interaction will almost certainly alter the demands imposed upon the panel and supporting structure relative to the purely hydrostatic case (Altomare et al. 2015). As an example, Goda’s parametric relations (Goda 2000) predict a total horizontal force of 89  kN/m on a vertical breakwater for 3 m of inundation exhibiting waves 2 m high propagating with a frequency of 0.1 Hz. This value is approximately twice that of the hydrostatic scenario (in fact equivalent to 4.26 m of static inundation), with the force resultant acting over 2.3 times higher above the base. In addition, the response of such structures to severe winds typically accompanying nearshore hazards including storm surges has also not been considered. Future research will focus on the performance of tilted hypar forms against dynamic loading scenarios arising from both wind and water sources.
From an engineering perspective, the design of an appropriate hinging mechanism at the vertex is paramount to the successful implementation of the proposed structural system as an adaptable form of coastal armor. The hinge must support a seamless transition away from an initial state of rotational fixity during normal operation, permitting immediate panel deployment prior to imminent hazard scenarios. Confirmation of the hinging mechanism would therefore enable the detailed design of the hypar shell and supporting column, including appropriate systems to resist panel uplift if required. Ideally, the design would require minimal maintenance and utilize materials exhibiting the durability necessary for long-term operation within coastal environments. While reinforced concrete was assumed for the proof of concept, the selection of an optimal material remains under investigation. Finally, an appropriate sealing solution closing the gaps between adjacent deployed panels for water retention on the shield is required. However, if a perfect watertight seal cannot be guaranteed (to be expected), pumping stations will likely be necessary to mitigate any accompanying seepage as widely adopted throughout the polder areas of the Netherlands (Hoes and Schuurmans 2006). Such challenges will be addressed as part of an ongoing endeavor advancing the concept of deployable hypar shells as hard countermeasures against coastal hazards.

Conclusions

An innovative solution to coastal armoring against nearshore hazards in the form of deployable hypar shells was introduced, enhancing the flexibility of traditional coastal hazard adaptation strategies. Relations pertaining to the spatial discretization of tilted four-sided hypar forms as a function of various geometric parameters were first established. Analytical expressions describing the boundary reactions necessary to suppress panel uplift in response to hydrostatic fluid inundation of any given height were subsequently derived. A parametric study was considered to quantify the influence of panel geometry, parameterized via the rise/area and lower/upper panel length ratios for a squarely projected hypar, and inundation height on the geometric conditions required for uplift suppression. A relatively simple decoupled numerical scheme constituting smoothed particle hydrodynamics and the finite element method for the simulation of fluid-structure interaction in the context of deployable hypar structures was next introduced. The model was successfully validated in the hydrostatic regime with numerical reactions synchronizing within 5% to their analytical counterparts. Hence, assuming small structural displacement and deformation, the decoupled model proves a useful tool in the detailed analysis of hypar surfaces for fluid-structure interaction. A proof-of-concept study was subsequently implemented to assess the performance of such systems within a structural engineering context. Preliminary designs of the panel and support column were evaluated, to which its feasibility in resisting hydrostatic inundation characteristic of extreme surge events along the United States’ Eastern Seaboard were successfully demonstrated. In leveraging the capabilities of both SPH and FEM techniques to respectively capture the fluid and structural response, future in-depth studies into the proposed armoring solution can be confidently implemented under both hydrostatic and hydrodynamic conditions. This work ultimately establishes a theoretical foundation for the analysis and design of dynamic hypar shells for coastal defense applications.

Appendix I. Geometric Parameterization of Deployed Panel

The angles of concavity pertinent to the shield along the longitudinal and transverse ordinates [Figs. 3(a and b)] are first defined:
αβ(λ1)=tan1(r{λ1}bβ),γβ(λ2)=tan1(r{λ2}βc)
(14)
to which the following notational conventions are introduced:
{A}=(1A),{A}β={A,β=1(1A),β=2
(15)
Subsequently, the angle of inclination θ1*=(v,x) can also be computed via Pythagorean considerations:
θ1*:h+Δtpcosθ1*=k1(λ1)sin[θ1*α1(λ1)]|λ1=0
(16)
where Δtp=tpvtpe is the difference between the projected panel thickness at the vertex and edges. The longitudinal dimension along the surface of the upper and lower shield (kβ) and soffit (mβ) is parameterized by λ1 and respectively determined via
kβ(λ1)=|ΔΠkβ|=bβ2+(r{λ1})2
(17)
mβ(λ1)=|ΔΠmβ|=bβ2+[r{λ1}+Δtp{λ1}]2
(18)
where |*| = Euclidean and absolute norm for R3 and R1 vector spaces, respectively; and ΔΠξβ=(Πξ2Πξ1)|λ2=β1. The continuum Πξ(λ1,λ2)=βΠξ{k,m}β{1,2}(λ1,λ2) thus represents the manifold of the shield (ξ=k) or soffit (ξ=m) as a union of the lower (β=1) and upper (β=2) surfaces. The parameterization of Πξβ by λ1 and λ2 is henceforth implied and omitted for clarity. Additionally, θβ(λ1) and ϕβ(λ1) give the angle between ΔΠkβ and ΔΠmβ, respectively, and the x ordinate [Fig. 3(c)]:
θβ(λ1)=(ΔΠkβ,x)={θ1*tan1(r{λ1}b1),β=1θ1(λ1)+θ(λ1),β=2
(19)
ϕβ(λ1)=(ΔΠmβ,x)={θ1*tan1(r{λ1}+Δtp{λ1}b1),β=1ϕ1(λ1)+ϕ(λ1),β=2
(20)
where
θ(λ1)=(ΔΠk2,ΔΠk1)=βsin1[r{λ1}kβ(λ1)]
(21)
ϕ(λ1)=(ΔΠm2,ΔΠm1)=βsin1[r{λ1}+Δtp{λ1}mβ(λ1)]
(22)
The Cartesian coordinates of all points constituting Πξβ for a deployed panel parameterized via λ1 and λ2 can thus be computed from geometric considerations as expressed via
Πkβ={[k1(λ1)λ2cosθ1(λ1)±cλ1k1(λ1)λ2sinθ1(λ1)+tpecosθ1*]+x0,β=1[k2(λ1)λ2cosθ2(λ1)+k1(λ1)cosθ1(λ1)±cλ1k2(λ1)λ2sinθ2(λ1)+k1(λ1)sinθ1(λ1)+tpecosθ1*]+x0,β=2
(23)
Πmβ={[m1(λ1)λ2cosϕ1(λ1)+tpesinθ1*±cλ1m1(λ1)λ2sinϕ1(λ1)]+x0,β=1[m2(λ1)λ2cosϕ2(λ1)+m1(λ1)cosϕ1(λ1)+tpesinθ1*±cλ1m2(λ1)λ2sinϕ2(λ1)+m1(λ1)sinϕ1(λ1)]+x0,β=2
(24)
where x0={x0,y0,z0} represents the origin. The parameter λ3 is finally introduced to map the continuum Ωp=βΩpβ, with Ωpβ[Πkβ,Πmβ] describing the global coordinates of all points constituting the deployed hypar shell between the manifolds of the shield and soffit as expressed in
Ωpβ(λ1,λ2,λ3)|Si=Πkβ+λ3(ΠmβΠkβ)
(25)

Appendix II. Derivation of Hydrostatic Boundary Reactions

Net vertical and horizontal reactions at the base and vertex hinge under hydrostatic demands are computed via integration along λ1 and λ2. The magnitude of hydrostatic pressure acting inward normal to Πkβ as a function of λ1 and λ2 is characterized as follows:
pβ(λ1,λ2)={ρwghwk1(λ1)λ2sinθ1(λ1),β=1ρwghw[hv(λ1)+k2(λ1)λ2sinθ2(λ1)],β=2
(26)
where ρw = fluid density; and g = acceleration due to gravity. hv(λ1) is the vertical height from the shield base to the plane of delineation between the lower and upper panel, Πuw|β,λ2=1 [Fig. 3(c)]:
hv(λ1)=|ΔΠk1·z^|=(Δtp+rλ1)cosθ1*+h
(27)
and * signifies the Macaulay ramp function as expressed in
A={0,A<0A,A0
(28)
The net hydrostatic reaction at the vertex hinge (fvw) is thus evaluated via consideration of moment equilibrium about the base hinge:
fvw=2cdbβ[01kβ(λ1)pβ(λ1,λ2)dβ(λ1,λ2)dλ2dλ1]z^limM,N2cMNdbβ[m=0M1n=0N1kβ(Λm)pβ(Λm,Λn)dβ(Λm,Λn)]z^
(29)
where dβ(λ1,λ2) = magnitude of the distance along the orthogonal from a given hydrostatic force vector to the base hinge on Πxz:
dβ(λ1,λ2)=|(ΠkβΠm1|λ2=0)·ΔΠ^kβ|={k1(λ1)λ2tpesinα1(λ1),β=1k2(λ1)λ2+k1(λ1)cosθ(λ1)+tpesin[θ(λ1)α1(λ1)],β=2
(30)
with ΔΠ^kβ representing the unit vector associated with ΔΠkβ. Λm and Λn, respectively translate λ1 and λ2 into discretized form, enabling the numerical evaluation of Eq. (29) via:
Λm=1+2m2M,Λn=1+2n2N
(31)
to which M,N=100 was adopted in this paper. Finally, db represents the distance between the base and vertex hinge along the x ordinate:
db=|ΔΠm1|λ1=0·x^|=m1(λ1)2h2|λ1=0
(32)
By solving Eq. (29), the horizontal and vertical components of the net hydrostatic reaction along the base hinge (fbw) may respectively be determined via consideration of force equilibrium along x and z as per Eqs. (33) and (34). The specialized result pertaining to the net horizontal reaction applicable only for fluid inundation below the total effective height of the deployed panel is also presented:
fbw·x^={2cβ[01kβ(λ1)pβ(λ1,λ2)sinθβ(λ1)dλ2dλ1],hw0ρwghw2c,0hwh*limM,N2cMNβ[m=0M1n=0N1kβ(Λm)pβ(Λm,Λn)sinθβ(Λm)]
(33)
fbw·z^=2cβ[01kβ(λ1)pβ(λ1,λ2)cosθβ(λ1)dλ2dλ1]fvw·z^limM,N2cMNβ[m=0M1n=0N1kβ(Λm)pβ(Λm,Λn)cosθβ(Λm)]fvw·z^
(34)
Evidently, fbw·y^=0 due to symmetry about Πxz|λ1=0. h* denotes the total effective height of the deployed panel [Fig. 3(c)], spanning the distance between the base and tip of Πk1 and Πk2, respectively, along the z ordinate:
h*=|(Πk2|λ2=1Πk1|λ2=0)·z^|=(b1+b2)sinθ1*
(35)
While transverse symmetry negates the presence of net reactions along the y ordinate, interactions between the hydrostatic demand and concavity of the deployed panel results in equal and opposite fluid forces acting outward normal to the plane of symmetry (Πxz|λ1=0). The total transverse hydrostatic force per unit length along the longitudinal spine is therefore the bounded integral of the hydrostatic pressure across half the panel width:
fhalfw(λ2)|β·y^=±c01pβ(λ1,λ2)tanγβ(λ2)dλ1
(36)
where γβ(λ2) is determined from Eq. (14) in Appendix I. The total hydrostatic force acting outward normal to the vertical plane of symmetry along the global y ordinate on each half of the panel thus results from the further integration of Eq. (36) on λ2:
fhalfw·y^=±β[bβ01pβ(λ1,λ2)r{λ2}βdλ1dλ2]limM,N±β[bβMNn=0N1m=0M1pβ(Λm,Λn)r{Λn}β]
(37)

Appendix III. Derivation of Self-Weight Reactions

Net vertical reactions at the base and vertex hinge from panel self-weight are determined via integration of the projected panel thickness orthogonal to Πuv along λ1 and λ2. The thickness dimension as a function of λ1 and λ2 is
tp(λ1,λ2)|β=|ΠkβΠmβ|={λ2Δtp{λ1}+tpe,β=1tpt(λ1)λ2Δtp{λ1},β=2
(38)
where tpt(λ1) = projected thickness of the panel along the transverse spine:
tpt(λ1)=|Πk1Πm1||λ2=1=|Πk2Πm2||λ2=0=tpvΔtpλ1
(39)
The angle of inclination between a vector sβ(λ1) passing through the midpoint of tp(λ1,λ2)|β,λ1 (in which λ3=0.5    λ1,λ2) to the global x ordinate may be determined via
ηβ(λ1)=[sβ(λ1),x]=sin1[hβ(λ1)sβ(λ1)]
(40)
where sβ(λ1)=|sβ(λ1)| is the magnitude of said vector pertaining to the lower or upper panel:
sβ(λ1)=|Πkβ+Πmβ2|λ2=1Πkβ+Πmβ2|λ2=0|=bβ2+(r{λ1}+Δtp{λ1}+tpe2)2
(41)
Furthermore, hβ(λ1) is the magnitude of the z ordinate projection of sβ(λ1):
hβ(λ1)=sβ(λ1)·z^=mβ(λ1)sinϕβ(λ1)+(1)(β+1)tpv(λ1)2cosθ1*
(42)
The distance from the midpoint of tp(λ2)|λ1 to the base hinge along the x ordinate may subsequently be computed:
eβ(λ1,λ2)=[sβ(λ1)λ2+(β1)s1(λ1)+Πk1Πm12|λ2=0]·x^=sβ(λ1)λ2cosηβ(λ1)+(β1)s1(λ1)cosη1(λ1)tpe2sinθ1*
(43)
The vertex reaction from self-weight (fvg) is therefore determined from the consideration of moment equilibrium about the base hinge utilizing Eqs. (38) and (43):
fvg=2cρsgdbβ[bβ01tp(λ1,λ2)eβ(λ1,λ2)dλ2dλ1]z^2cρsgdblimM,Nβ[bβMNm=0M1n=0N1tp(Λm,Λn)eβ(Λm,Λn)]z^
(44)
where ρs = density of the panel material. The total self-weight base reaction (fbg) from vertical force equilibrium is thus expressed as
fbg=fgfvg
(45)
in which fg=Vpρsg is the total weight of the panel, with Vp being the hypar volume:
Vp=c(tpv+3tpe)(b1+b2)2
(46)
and g=gz^ is the gravitational acceleration vector.

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 School of Architecture at Princeton University, for contributing to the production of Fig. 1.

References

ACI (American Concrete Institute). 2014. Building code requirements for structural concrete. ACI 318. Farmington Hills, MI: ACI.
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.
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 Mathematical Library.
Bentz, E., and M. P. Collins. 2001. Response-2000 user manual. Toronto: Univ. of Toronto.
Candela, F. 1955. “Structural applications of hyperbolic paraboloidical shells.” J. Am. Concr. Inst. 51 (1): 397–416.
Charlier, R. H., M. C. P. Chaineux, and S. Morcos. 2005. “Panorama of the history of coastal protection.” J. Coastal Res. 21 (1): 79–111. https://doi.org/10.2112/03561.1.
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 Tech. Sci. Press 5 (3): 173–184.
Crespo, A. J. C., M. Gomez-Gesteira, and R. A. Dalrymple. 2008. “Modeling dam break behavior over a wet bed by a SPH technique.” J. Waterway, Port, Coastal, Ocean Eng. 134 (6): 313–320. https://doi.org/10.1061/(ASCE)0733-950X(2008)134:6(313).
Dalrymple, R. A., and B. D. Rogers. 2006. “Numerical modeling of water waves with the SPH method.” Coastal Eng. 53 (2–3): 141–147. https://doi.org/10.1016/j.coastaleng.2005.10.004.
Dugan, J. E., D. M. Hubbard, I. F. Rodil, D. L. Revell, and S. Schroeter. 2008. “Ecological effects of coastal armoring on sandy beaches.” Supplement, Mar. Ecol. 29 (S1): 160–170. https://doi.org/10.1111/j.1439-0485.2008.00231.x.
Garay, G. 1994. Proyecto de historia oral de la Ciudad de Mexico: Testimonios de sus arquitectos (1940–1990). Mexico City: Instituto de Investigaciones Dr. José María Luis Mora.
Garlock, M. E. M., and D. P. Billington. 2008. Felix Candela: Engineer, builder, structural artist. Princeton, NJ: Princeton University Art Museum.
Goda, Y. 2000. Random seas and design of maritime structures. Singapore: World Scientific Publishing.
Hibbeler, R. C. 2008. Structural analysis. 7th ed. Toronto: Pearson Education Canada.
Hoes, O., and W. Schuurmans. 2006. “Flood standards or risk analyses for polder management in the Netherlands.” Supplement, Irrig. Drain. 55 (S1): 113–119. https://doi.org/10.1002/ird.249.
Horn, D. P. 2015. “Storm surge warning, mitigation, and adaptation.” Chap. 6 in Coastal and marine hazards, risks, and disasters. Waltham, MA: Elsevier.
Levine, A. 2018. “A structural analysis and comparison of Felix Candela’s 4-sided and 8-sided concrete umbrellas.” M.S. thesis, Dept. of Civil and Environmental Engineering, Princeton Univ.
Levine, A., and M. E. M. Garlock. 2018. “Structural comparison of Candela’s 4-sided and 8-sided concrete umbrellas.” In Proc., IASS Symp. Boston: Massachusetts Institute of Technology.
Liu, G. R., and Y. T. Gu. 2005. An introduction to meshfree methods and their programming. Dordrecht, Netherlands: Springer Netherlands.
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.
Liu, M. B., G. R. Liu, and K. Y. Lam. 2003. “Constructing smoothing functions in smoothed particle hydrodynamics with applications.” J. Comput. Appl. Math. 155 (2): 263–284. https://doi.org/10.1016/S0377-0427(02)00869-5.
Mazzoni, S., F. McKenna, M. H. Scott, and G. L. Fenves. 2006. OpenSees command language manual. Berkeley, CA: Pacific Earthquake Engineering Research Center.
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.
Molteni, D., and A. Colagrossi. 2009. “A simple procedure to improve the pressure evaluation in hydrodynamic context using the SPH.” Comput. Phys. Commun. 180 (6): 861–872. https://doi.org/10.1016/j.cpc.2008.12.004.
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. 1999. “SPH without a tensile instability.” J. Comput. Phys. 159 (2): 290–311. https://doi.org/10.1006/jcph.2000.6439.
Monaghan, J. J. 2012. “Smoothed particle hydrodynamics and its diverse applications.” Annu. Rev. Fluid Mech. 44: 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).
Mooyaart, L. F., and S. N. Jonkman. 2017. “Overview and design considerations of storm surge barriers.” J. Waterway, Port, Coastal, Ocean Eng. 143 (4): 06017001. https://doi.org/10.1061/(ASCE)WW.1943-5460.0000383.
Needham, H. F., B. D. Keim, and D. Sathiaraj. 2015. “A review of tropical cyclone-generated storm surges: Global data sources, observations, and impacts.” Rev. Geophys. 53 (2): 545–591. https://doi.org/10.1002/2014RG000477.
Nicholls, R. J., and A. Cazenave. 2010. “Sea level rise and its impact on coastal zones.” Science 328 (5985): 1517–1520. https://doi.org/10.1126/science.1185782.
Overby, D., M. Kowalsky, and R. Seracino. 2015. A706 Grade 80 reinforcement for seismic applications. Raleigh, NC: Constructed Facilities Laboratory, NC State Univ.
Pilkey, O. H., and H. L. Wright III. 1988. “Seawalls versus beaches.” J. Coastal Res. 4: 41–64.
Small, C., and R. J. Nicholls. 2003. “A global analysis of human settlement in coastal zones.” J. Coastal Res. 19 (3): 584–599.
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.
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.
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 146Issue 6June 2020

History

Received: Apr 2, 2019
Accepted: Oct 22, 2019
Published online: Mar 28, 2020
Published in print: Jun 1, 2020
Discussion open until: Aug 28, 2020

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]
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

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share