You are on page 1of 28

Marine Structures 79 (2021) 103055

Contents lists available at ScienceDirect

Marine Structures
journal homepage: www.elsevier.com/locate/marstruc

A multiscale material-structure-hydroelasticity coupled analytical


model for floating sandwich structures with hierarchical cores
Hui Fang a, *, Huadan Zhu a, Aijun Li a, Huichao Liu b, Senlin Luo b, Yilun Liu b, **,
Yong Liu a, Huajun Li a
a
Shandong Province Key Laboratory for Ocean Engineering, School of Engineering, Ocean University of China, Qingdao, 266100, China
b
State Key Laboratory for Strength and Vibration of Mechanical Structures, Xi’an Jiaotong University, Xi’an, 710049, China

A R T I C L E I N F O A B S T R A C T

Keywords: A novel material-structure-hydroelasticity coupling analytical model is proposed for marine


Multiscale coupling structures, which is utilized for the calculation and optimization of a very large floating sandwich
Floating sandwich structure structure (VLFSS) with a hierarchical ultrahigh-performance concrete (UHPC) core in this study.
Hierarchical core
For the coupling material and structure analysis, three-dimensional representative volume
element and self-consistent methods are developed to reveal the physical relations between the
UHPC core’s macroscale mechanical properties (e.g., modulus and density) and mesoscale hier­
archical characteristics (e.g., aggregate and porosity) and to obtain the corresponding parame­
terized formulas. For the coupling material-structure-hydroelasticity analysis, a sixth-order
dynamical equation for the potential flow model of the VLFSS, in which the hierarchical core’s
parameters are introduced through the material-structure coupling formulas, is developed. The
hydroelasticity equations containing multiscale parameters are solved, and the mechanical re­
sponses are calculated. Using this coupled multiscale method, the shear force in the representa­
tive VLFSS is optimized for a smaller amplitude, which relies on the interactivity of the
hierarchical structural parameters and wave conditions. These results demonstrate the potential
of the multiscale coupling methodology to achieve the physically significant optimization of a
floating composite structure in ocean engineering.

1. Introduction

Very large floating structure (VLFS for short) plays an important role in ocean space utilization, solving the scarcity of land in
islands and coastal countries with high rise of population and urban development. For example, Japan established the Mega-Float
airports in the Tokyo bay [1], the Yumeshima-Maishima floating bridge in Osaka, the floating emergency rescue bases in Yoko­
homa, Tokyo and Osaka, and the floating oil storage systems in Shirashima and Kamigoto [2]. Canada built a floating heliport in
Vancouver and the Kelowna floating bridge on Lake On in British Columbia. In past several decades, researchers have made
considerable efforts in the development of analysis for feasibility investigations of VFLSs.
Early studies on the hydroelastic analysis mainly focused on water wave interaction with large floating plates, such as ice sheets,
which have similar characteristics with the VLFSs. The floating sheets and VLFSs were often modeled as large thin elastic plates. By

* Corresponding author.
** Corresponding author.
E-mail addresses: fanghui@ouc.edu.cn (H. Fang), yilunliu@mail.xjtu.edu.cn (Y. Liu).

https://doi.org/10.1016/j.marstruc.2021.103055
Received 28 October 2020; Received in revised form 7 June 2021; Accepted 8 June 2021
Available online 6 July 2021
0951-8339/© 2021 Elsevier Ltd. All rights reserved.
H. Fang et al. Marine Structures 79 (2021) 103055

using an error function method, Fox and Squire [3] developed an analytical solution for the interaction of surface waves with a
semi-infinite floating ice sheet. The effect of incident angle on wave scattering by a semi-infinite floating ice sheet was further
examined by Fox and Squire [4], and the results showed that there was a critical angle of incident waves beyond which the waves were

Fig. 1. (a) Illustrations of the VLFSS and (b) hierarchical UHPC core, (c) illustration of three-stage multi-scale analysis of the VLFSS with UHPC
hierarchical core.

2
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. 1. (continued).

fully reflected. Subsequently, Squire [5] reviewed the investigations on the hydroelastic response of ice sheets subjected to ocean
waves. Teng et al. [6] studied the hydroelastic response of a semi-infinite floating plate with a finite draft. They showed that the zero
draft assumption was feasible for low wave frequencies, and the effect of the plate draft became noticeable for high wave frequencies.
Recently, Guo et al. [7] developed an analytical solution for interactions between obliquely incident waves and a semi-infinite floating

3
H. Fang et al. Marine Structures 79 (2021) 103055

plate on a step topography, and presented effects of the draft and the step height on the deflection and bending moment of the plate.
Based on potential flow theory and thin plate theory, Mohapatra and Guedes Soares [8] developed a general three-dimensional
mathematical model for wave interaction with a floating and submerged flexible plate. Different from the assumption of thin
plates, Zhao et al. [9] used the Viener-Hopf technique to solve the problem of surface wave diffraction by floating elastic plates, which
was modeled as Mindlin thick plates. In addition to horizontal flat structures, water wave interaction with the VLFSs having other
different types were also studied. Watanabe et al. [2] detailed a literature survey of investigations on the hydroelasticity analysis of
pontoon-type VLFSs. By adopting the coupling method of higher-order boundary element and finite element, Li et al. [10] solved the
hydroelastic response problem of a pontoon-type VLFS subjected to short-crested irregular waves. In their solution procedure, the VLFS
was modeled by a Mindlin plate, and the mode superposition method was adopted to reduce the dimension of dynamic equations.
Different from the above VLFSs, Santos et al. [11] performed the hydroelasticity analysis for wave-induced dynamic response of a fast
patrol boat, and their results indicated some limitations of the linear fluid-structure interaction in the case of planning conditions with
relatively large Froude numbers. Hirdaris and Temarel [12] reported some applications of hyroelasticity theory in ship field, and
discussed the long-term potential use of nonlinear fluid-structure interaction for improving dynamic response models of ships.
It is noted that the VLFSs in all above studies were regarded as single-layer homogeneous plates. In the practical designs and
applications of VLFSs, their deadweight and the mechanics performance should be considered more carefully. To reduce the structural
weight and improve the performance without sacrificing the stiffness and strength, composite materials and structures have been
widely used in advanced marine, civil and other structures [13–16]. As a representative engineering composites,
ultrahigh-performance concrete (UHPC) with the necessary stiffness, toughness and durability [17] is a high-quality base material for
the hierarchical core that can be used for the development of sandwich-type floating structures [18,19]. As a representative engi­
neering structure, a very large floating sandwich structure (VLFSS) with a hierarchical UHPC core is proposed and shown in Fig. 1, and
multiple structural and material characteristics at different geometric scales must be modeled in the calculation and design. As
illustrated in Fig. 1(a), the floating body as a whole has a very large scale, with a length that can reach hundreds or even thousands of
meters. As shown in Fig. 1(b), the sandwich core’s hierarchical structure has a medium scale, in which the diameter unit of a minimum
hole is on the order of centimeters and the UHPC’s grading structure is on the order of millimeters and microns. As a whole, the
hydroelasticity of VLFSS is a complex multiscale problem, related to the microstructure of UHPC, hierarchical structure of the core,
geometrical features of the floating body and the wave actions, ranging from micrometer to hundreds of meter scale. And a multiscale
material-structure-hydroelasticity coupled analysis should be developed for the design and optimization of this marine structure.
In terms of the material scale, concrete is a class of hierarchical materials. They behave like a homogeneous material at macroscale,
but the heterogeneity manifests itself when the scale goes down [20]. To study its property, this kind of materials is often broken down
under several scales, and a certain homogenization method is employed to bridge neighbor scales. For example, Borges et al. [21] build
a two-scale model to study the mechanics phenomena during the loading process of concrete structure. At macroscale, a concrete beam
is studied, while at mesoscale, concrete is regarded as a three-phase composite, consisting of matrix, aggregate and the interface
transition zone (ITZ) of the two. Finite element method (FEM) is utilized to simulate the behavior at the two scales. The response of
every element at macroscale is predicted by the representative volume element (RVE) at mesoscale. Abbès et al. [22] consider High
Performance Concrete as a three-level material to study its elasto-viscoplastic properties. The macroscale concrete is a homogenized
material of matrix, aggregate and large air pores at mesoscale. And the above-mentioned matrix is also a homogenized material of a
four phase RVE at microscale. UHPC is more complex and thus demands more levels of the model. Chen et al. [23] use a three-level
homogenization model to take account of the four scales of UHPC. The first level is a gel matrix RVE, consisting of solid gel and gel
pores. Its effective properties are predicted by Mori-Tanaka method [24]. The second level is the cement paste RVE, which includes the
homogenized gel matrix, pores, silica fume, portlandite crystals, cement clinker and so on. The effective modulus is calculated by a
stochastic micromechanical model [25]. The third level is a concrete RVE, with homogenized cement paste as matrix and sand, air
voids and fiber, etc. as inclusion. A two-step approach is adopted for this level. The first step is to calculate the concrete without fiber
using the same approach as the second level calculation, while an iteration process [26] is employed to take the effect of fiber into
account in the second step. To give a thorough understanding of the imperfect interface effect on the property of UHPC, an even more
complex multiscale model, including five scales, as well as a four-level homogenization model, is employed. The first level is all the
way down to calcium silicate hydrate (CSH) scale, which includes low density CSH as matrix and high density CSH as inclusion. The
cement paste is considered as a composite with homogenized calcium as matrix and portlandite crystals, clinker and capillary pore as
inclusion, which is regarded as the homogenized matrix of mortar, together with sand and ITZ as the other two constituents. Similarly,
the concrete RVE is also considered as a three-phase composite with homogenized mortar as matrix and coarse aggregate and ITZ as
inclusions. A modified Mori-Tanaka scheme is developed by taking the imperfect interface into consideration, to bridge all the
aforementioned scales. However, since this paper intends to present a design framework, we calculate the properties of UHPC only
from mesoscale for the sake of simplicity. As presented above, UHPC is consisted of aggregates and fibers that are bonded together by
mastic [27]. The aggregate gradation and content are two important parameters to determine the modulus and density of UHPC, which
can be analyzed by a micromechanical model with the representative volume element (RVE) homogenization [28].
At the structure scale, the design of the hierarchical core and sandwich structure with the required density and bending stiffness is
crucial for VLFSS, so that the density of VLFSS is smaller than sea water and the mechanical responses of VLFSS are in an optimal range
under wave actuation. In order to get density and stiffness of the hierarchical structure, the multiscale RVE method is employed for the
design and optimization of the hierarchical core. At the hydroelasticity analysis scale, the hydroelastic response of sandwich structure
under wave actuation with potential flow model could be solved using the homogeneous mechanical parameters inputted from the
material and structure level models. And the hydroelastic boundary conditions (the dynamic equation of the sandwich structure)
should be obtained. Currently, the simplified model for uniform plates is based on Euler beam theory [29,30]. Considering the

4
H. Fang et al. Marine Structures 79 (2021) 103055

sandwich structure, the stiffness of the core is very weak compared to the steel face panels, so that the bending moment is mostly born
by the upper and lower panels, while shear stress is generated in the core [28]. Thus, the specific bending behaviors of sandwich
structure should be considered in the hydroelasticity analysis. To sum up, the hydroelastic responses of the VLFSS with hierarchical
UPHC core are conducted by multiscale material-structure-hydroelasticity coupled analysis that combines the material parameters of
UHPC, structural parameters of core, geometry of VLFSS and wave conditions in an analytical model, so that one can analyze and
optimize VLFSS from material level to hydroelasticity level.
In this study, a novel material gradation - honeycomb structure - hydroelasticity coupling analysis methodology is developed and
performed for a VLFSS with a hierarchical UHPC core. RVEs are constructed to obtain the relations between the macroscale mechanical
properties and the mesoscale characteristics of the hierarchical UHPC core, which is verified by the self-consistent theory. A high-order
hydroelastic equation is established for the potential flow model of a sandwich-type very large floating body, in which the relationship
between the meso- and macroscale characteristics obtained above is substituted into the high-order dynamical equation. The
hydroelasticity equations containing multi-scale parameters are solved, and the mechanical responses are calculated. As an illustra­
tion, a representative VLFSS is analyzed, and the response characteristics are optimized.

Fig. 2. FE model of the UHPC hierarchical core’s RVEs of (a) 2-level aggregate and (b) 3-level aggregate UHPC, (c) simple cube foam and (d) body
center cube UHPC foam, and (e) square and (f) hexagonal lattices of the through holes.

5
H. Fang et al. Marine Structures 79 (2021) 103055

2. A VLFSS with a hierarchical UHPC core

Fig. 1 (a) is the illustration of VLFSS with UHPC hierarchical core. Fig. 1 (b) presents the sandwich structure of VLFSS, i.e. the upper
and lower steel face panels of thickness h1 = h3, and a UHPC hierarchical core of thickness h2. In principle, the UHPC hierarchical core
can be designed as three level structures. First, in the material scale UHPC is consisted of cement and multilevel aggregates, in which
the mechanical property is modulated by the number of aggregate levels and volume fraction of aggregate. Second, for the lightweight
design of VLFSS, the UHPC core can be constructed with UHPC foam, i.e. small voids are randomly distributed in UHPC, thus the
corresponding properties depend on the void ratio and the distribution of void. Third, the UHPC core with through holes, in which new
design factors, viz., ratio of holes and distribution of holes are added in this level. Indeed, the second and third level structures can be
used independently in VLFSS depending on the specific requirements of VLFSS in practical applications.
The structural analysis of the VLFSS with hierarchical UHPC core is of certain difficulty because it has to consider both the analysis
of the internal micro and meso-structure responsible of the desired properties of UHPC core, and the analysis of the macro-structure
itself. A profitable technique is the multi-scale analysis, traditionally used, for example, multi-scale FEM in the mechanics of composite
materials. Herein, an efficient procedure for the multiscale FEM through a continuum mechanics bridge, which is to get the mechanics
properties of the UHPC core at different scales (Fig. 1(b)) from Virtual Tests of a RVE of FEM, derives the mechanics behaviour of the
micro and meso-structure of the UHPC core under different deformation modes. Each RVE is built and analyzed in ABAQUS. And a
novel hydroelastic model for wave interaction with flexible VLFSS, in which the material property of the UHPC core obtained by the
multi-scale RVE are substituted, is developed to analyze the structural response. Fig. 1(c) illustrates the three-stage multi-scale analysis
of the VLFSS with UHPC hierarchical core: at first stage, the elastic properties of the matrix UHPC material with different aggregate
gradations are calculated by RVE models including three levels aggregates, and are unified into continuous equivalent bulk; at second
stage, the elastic properties of the lightweight UHPC core with holes are calculated by two-levels RVE model, in which the matrix
UHPC’s mechanics parameters obtained in the last stage are utilized as input parameters of the RVE models. And the UHPC core’s
mechanics parameters are calculated and further unified into continuous equivalent bulk that will be utilized in the next stage of
hydroelastic analysis; at third stage, the hydroelastic model of the sandwich structure is adopted, in which the lightweight UHPC core’s
mechanics parameters obtained in the last stage are utilized to derive the dynamic equations, and the vibration of the VLFSS subjected
to the action of small-amplitude surface gravity waves are calculated by a novel approach of frequency-domain analysis.

2.1. Mechanical properties of the hierarchical UHPC core

The representative volume element (RVE) method is employed to describe the mechanical behaviors of hierarchical UHPC core, i.
e., the UHPC with different aggregate gradations, UHPC foam and UHPC core with through hole, as shown in Figs. 1(b) and Fig. 2.
Here, the RVE means a typical volume over which a measurement can be made to represent the average value of the whole. Each RVE is
built and analyzed in ABAQUS to obtain the mechanical properties. Considering that the elastic property of a material can be expressed
by a 6 × 6 elastic matrix C using Voigt notation, 6 load cases, where only one strain component is prescribed to be nonzero (but small in
the infinitesimal deformation region) are analyzed. After the RVE is constructed and meshed, the calculation is automatically carried
out in ABAQUS MicroMechanics Plugin [31], which directly outputs the desired elastic parameters. Furthermore, we also employ the
widely used Mori-Tanaka method [32] in theory of composite materials to predict the elastic properties of UHPC. The Young’s modulus
and shear modulus of composite materials with spherical inclusion are expressed by Young’s modulus and shear modulus of the matrix
E0 , G0 , those of the inclusion Ef , Gf and the volume ratio of the inclusion C1 as
9Kf G
E= (1)
3Kf + G

C1 G 0
G = G0 − (2)
1 − βf (1 − C1 )

where,
C1 (K1 − K0 )K0
Kf = K0 +
K0 + αf (1 − C1 )(K1 − K0 )

C1 (G1 − G0 )G0
G = G0 +
G0 + βf (1 − C1 )(G1 − G0 )

E0 G 0
K0 =
3(3G0 − E0 )

Ef Gf
K1 = ( )
3 3Gf − Ef

1 E0
αf =
3 4G0 − E0

6
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. 3. (a) Category for set of the nodes on the surfaces of a cubic RVE. (b) Deformation of the UHPC’s RVEs of 3-level aggregates: (b.1) x tension,
(b.2) y tension, (b.3) z tension, (b.4) xy shear, (b.5) xz shear and (b.6) yz shear. (c.1) Stress σ11 and (c.2) strain ε11 of the corresponding nodes
located on two opposite surfaces FaceRight and FaceLeft of the base UHPC material’s RVEs of 3-level aggregates (Periodic boundary condition ε =
[ε, 0, 0, 0, 0, 0]T ).

7
H. Fang et al. Marine Structures 79 (2021) 103055

2 18G0 − 5E0
βf =
15 4G0 − E0
The results of FEM-RVE are fitted to a polynomial as the representative model of each level. Based on this, the shear modulus and
density of core material are expressed by design variables, and the optimal solution is obtained and discussed.
Fig. 2 gives the illustrations of RVEs of the hierarchical UHPC core. Fig. 2 (a) and (b) are examples of RVEs for 2-level aggregate and
3-level aggregate UHPC. Fig. 2 (c) and (d) are the RVEs of Simple Cube and Body Center Cube UHPC foam. Fig. 2 (e) and (f) are the
RVEs of UHPC core with square and hexagonal lattice of through holes. No appreciable difference has been observed between the
results obtained from linear elements (e.g., C3D4) or quadratic elements (e.g., C3D10 M). Meshes generated from seed controls of 1/8
RVE edge or 1/20 RVE edge are also tested, and the results are insensitive to the seed size.

2.1.1. Simulating principle


The RVE based on finite element homogenization method is a proven method to predict the effective mechanical properties of
composites, and thus it is adopted to evaluate the UHPC core with the complicated micro-structures in this study. The effective me­
chanical properties are intrinsic to the materials and thus should not depend on the external effects, such as presence of body force and
boundary condition. Although the boundary condition will not affect the effective mechanical properties of composites, it is necessary
and must satisfy Hill’s energy law [33]. The existing investigations have demonstrated that the better approximations for the effective
mechanical properties of composites could be obtained when the periodic boundary condition instead of the uniform tensile or linear
displacement boundary condition is adopted [34–36]. Thus, the periodic boundary condition will be used here. In the ABAQUS, the
periodic boundary condition can be applied on the surface nodes of the RVE (Fig. 3(a)) of composites through a linear multi-point
constraint, which requires that a linear combination of nodal displacements equals zero.
To validate the RVE models, six kinds of monotonic loading cases involving the pure x, y, z tension and xy, xz, yz shear are applied
on the surfaces of the UHPC hierarchical core’s RVEs of aggregate UHPC (3-level), simple cube foam and body center cube UHPC foam,
and square and hexagonal lattices of the through holes. Fig. 3(b.1-b.6) give the corresponding deformation behaviors of the UHPC’s
RVEs of 3-level aggregates under these six loading cases, respectively. The stress and strain continuity conditions on the boundaries of
the RVE models are checked for verifying the periodic boundary condition. The simulation details of other RVE validation are pre­
sented in the Appendix: Figs. A1, A3, A5 and A7 give the corresponding deformation behaviors of the RVE models (of simple cube foam
and body center cube, foam RVEs of square and hexagonal lattices of the through holes, respectively) under these six loading cases, and
Figs. A2, A4 and A6 show the stress σ 11 and strain ε11 of the corresponding nodes on two opposite surfaces FaceRight and FaceLeft of
the RVE models (see Fig. 3(a)) under the x tension loading case; as a supplement, Fig. A9 gives the stress σ 12 and strain ε12 of the
corresponding nodes on two opposite surfaces FaceRight and FaceLeft of the UHPC core’s RVEs of hexagonal lattices of the through
holes under the xy shear loading case.The elastic deformation of the RVE models are amplified with several times for descripting the
loading modes. It should be mentioned that the relative deviations result from the numerical errors due to the extrapolation of the
stresses and strains at the respective integration points and from the mesh density of the surfaces of the RVE. As shown above, the
predefined periodic boundary condition guarantees the stress and strain continuity conditions of the corresponding nodes on the
boundaries of the RVE and the mechanics properties of RVEs satisfy Hill’s energy law.
Effective media theories have been used in continuum micromechanics. Such an approach was pioneered by Mori and Tanaka [32,
37]. Mori-Tanaka method can calculate the average internal stress in the matrix of a material containing inclusions with trans­
formation strain. The average stress in the matrix is uniform throughout the material and in-dependent of the position of the domain
where the average treatment is carried out. The actual stress in the matrix is average stress plus the locally fluctuation stress, the
average of which vanished in the matrix. And this Mori-Tanaka method has been used to predict the elastic properties of a wide range
of composites (unidirectional, short-fiber and particulate). The basic formulae of the Mori-Tanaka method were reformulated by
Benveniste [38]. The effective stiffness tensor of a two-phase composite material with an isotropic matrix phase is given. As an

Fig. 4. The relation between the modulus of UHPC and total aggregate volume proportion, indicating the elastic modulus is determined by the total
aggregate volume proportion instead of aggregate levels.

8
H. Fang et al. Marine Structures 79 (2021) 103055

application, the elastic property of the UHPC hierarchical core can be obtained following the formulations presented by Shen et al.
[28].

2.1.2. Simulation results


Limited by the 3D random aggregate algorithm, the volume ratio of aggregate in UHPC considered herein can only reach 0.4. As
shown in Fig. 4, the calculated values of Young’s modulus or shear modulus of UHPC with two-level aggregate and three-level
aggregate lie on the same line, which indicates the elastic modulus of UHPC is mainly influenced by total volume proportion of
aggregate (or average of density, Fig. 4), instead of aggregate level. The relationships of Young’s modulus and shear modulus to
aggregate fraction can be fitted by two polynomials (shown in Fig. 5). In addition, the average density of UHPC is given according to
the rule of mixture,
( )
ρUHPC = ρagg · cagg + ρcem · 1 − cagg (3)

where ρagg and ρcem are the densities of the aggregate and cement, respectively; and cagg is the total volume proportion of the aggregate.
For the second level, spherical voids arranged in simple cubes and body center cubes are considered (Fig. 2 (c) and (d), respec­
tively). Structural integrity requires the maximum void ratios of the simple cube and body center cube to be approximately 0.9651 and
0.9945, respectively. Nine samples of each arrangement are calculated. The resultant moduli are normalized by the corresponding
moduli of the matrix (e.g., UHPC) to rule out the influence of the specific matrix.
As shown in Fig. 6, the normalized moduli decrease when the volume ratio of the void increases, clearly because of the reduction in
the material bearing loads. This finding is similar to the results predicted by the Mori-Tanaka method. The difference between the
simple cube and body center cube is quite small, so we omit the influence of the void arrangement in the following.
Notably, when the void ratio goes beyond 0.5, the void begins to overlap with each other, and the spherical shape is no longer
maintained. Besides, even before the void starts to contact with each other, the interaction among them grows to be strong when they
are pretty close. These breach the assumptions of Mori-Tanaka method. Hence, appreciable difference is observed when void ratio
becomes large, and we need to put more value on the results of FE-RVE. As mentioned above, the void ratio is the main factor that have
effects on the mechanical properties of UHPC foam. The normalized relationship is given by fitting (Fig. 6). In addition, the mean
density of this level
ρfom = ρUHPC ·(1 − cvod ) (4)

where cvod is the void ratio.


For the third level, we only considered the square lattice and hexagonal lattice of the holes (Fig. 2 (e) and (f)). The maximum ratios
of the holes are approximately 0.78 for the square arrangement and approximately 0.92 for the hexagonal arrangement due to the
structural integrity. The following sections show that only the transverse shear modulus or G13 of the core material plays a large role in
the response of the sandwiched VLFSS. To show the shear modulus reduction in the holes, the calculated shear modulus G13 is
normalized by that of the matrix (e.g., UHPC foam).
The normalized shear modulus of both the square arrangement and hexagonal arrangement decrease as the ratio of holes increases,
apparently due to the loss of the material bearing loads. The hexagonal arrangement suffers less shear modulus loss compared with the
square arrangement, if the ratio of holes is large. Thus, it is reasonable to believe that the hexagonal lattice outperforms the square
lattice (shown in Fig. 7). Hence, only the normalized out-of-plane shear modulus G13 of the hexagonal lattice is fitted,
G13
= − 0.707c3hol + 1.57c2hol − 1.93chol + 1 (5)
Gfom

Fig. 5. The average modulus of UHPC against the volume proportion of the aggregate.

9
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. 6. Normalized moduli of the UHPC foam against void the ratio.

Fig. 7. Normalized moduli of UHPC with through holes against the ratio of holes.

where chole is the ratio of holes. The density is


ρhol = ρfom ·(1 − chol ) (6)

Finally, the expression of the out-of-plane shear modulus of the core material with a three-level structure is obtained from the
abovementioned equations in the form
( )
G = G13 = G chol , cvod , cagg (7)

and its density is expressed as


( )
ρ = ρhol = ρ chol , cvod , cagg (8)

The expressions of a two-level core material design can be obtained by omitting the modulus reduction and density reduction of the
second or third level. The optimized design can be obtained by maximizing the function G under necessary constraint conditions. For
instance, we obtained the maximum out-of-plane shear modulus of the different level core materials under the constraints
( ( )) /
ρUHPC = ρagg · cagg + ρmot · 1 − cagg ·(1 − cvod )(1 − chol ) ≤ 400kg m3 (9a)

0 ≤ cagg ≤ 0.4 (9b)

0 ≤ cvod ≤ 0.965 (9c)

0 ≤ chol ≤ 0.92 (9d)


As shown in Table 2, the optima of design ➀ and design ➂ are close and much better than design ➁. Furthermore, when design ➀, i.
e., the three-level design, reaches the optimum, cvod is near zero. This finding indicates that the second level fails to contribute to the
optimum. Therefore, we should take design ➂ as the best, for the sake of simplicity.
In the three-level modelling above, the size of microstructural constituent of the RVE is much smaller than that of the sandwich
structure adopted in the hydroelastic analysis below (subsection 2.2). But the share of bending in structural bulk of the third RVE

10
H. Fang et al. Marine Structures 79 (2021) 103055

modelling of the UPHC core with through holes in comparison to membrane shearing would increase when the diameter of the hole
approximates the size of the bulk; that is to say, the size effect will occurs when the size of structural component in comparison to that
of the structure is large and then more advanced homogenization is needed, i.e. multiscale theory studied by Kouznetsova [39], in
which a second-order computational homogenization procedure based on a proper incorporation of the gradient of the macroscopic
deformation tensor into the kinematical micro-macro framework is proposed and verified. Because the high-order theory of the ho­
mogenization is very complex, the structural components (through holes) are designed with a much smaller size than that of the
sandwich structure (the size of the generalized one is 1000 × 100 × 4.9 m3) and the RVE modelling presented above could be utilized
availably in this study.

2.2. Elastic flexure of the sandwich structure with an UHPC hierarchical core

In Fig. 8, the concept of the floating body with three layers is presented. The floating body is likely to be used in the construction of
maritime airports or other structures. To realize the very large floating structure, the model of the sandwich beam [40] is adopted in
this hydroelastic analysis. The analysis model of this sandwich beam is shown in Fig. 8.
As shown in Fig. 8(a), the sandwich beam of the unit width is considered. It has panels of thickness h1 and h3 and core material of
thickness h2. The panel is purely elastic, with Young’s modules of E1 and E3. The core is linearly viscoelastic with a shear modulus G. In
Fig. 8(b), the displacement is expressed in terms of w, and the longitudinal X-direction displacement of the mid-planes in the panel is u1
and u3. The longitudinal displacement component of any core point is u (without suffix). At the same time, the direct transverse strain
of the core material and panel is ignored so that the transverse displacement w of each point on the section is equal.
It is assumed that both the shear strain of the panel and the longitudinal direct stress of the core are negligible. The shear strain γ in
3
the core exists. Therefore, we can see obviously in Fig. 8(c) that the shear force consists of three parts: (a) a shear force S1 = D1 ∂∂xw3 ,
E1 h31 3
where the flexural rigidity D1 is expressed by D1 = (b) a shear force S3 = D3 ∂∂xw3 , where the flexural rigidity D3 is expressed by D3 =
12 ;
[ ]
3
E3 h3 d ∂w u1 − u3
12 ; and (c) a shear force S2 = − τd, where the shear stress τ = Gγ = G h2 ∂x + h2 ,d = h2 + (h1 + h3 )/2. Hence, the shear force can
be presented by
( )
∂3 w d ∂w u1 − u3
S = S1 + S2 + S3 = Dt − Gd + (10)
∂x3 h2 ∂x h2

where Dt = D1 + D3 .
Next, the transverse loading, p, on the beam is related to S by
∂S
p= (11)
∂x
In Fig. 8(d), consider the longitudinal equilibrium of a lengthwise element of the lower panel. Clearly,
δP3 = − τδx (12a)

Fig. 8. (a) Dimensions and coordinate system. (b) Displacement system. (c) Forces and moments acting on a section. (d) Longitudinal forces on the
lower face-plate element.

11
H. Fang et al. Marine Structures 79 (2021) 103055

or
∂P3
=− τ (12b)
∂x

where P3 = E3 h3 ∂∂ux3 .
According to the expression Eqs. (11) and (12), the simplest pair of differential equations is formed that relate the displacement w
and u3 to the applied loading p. We can eliminate u3 from the equations according to the expression, and a single sixth-order differential
equation in w is obtained
( )
∂6 w ∂4 w 1 ∂2 p
− G t (1 + Y) = − Gt p (13)
∂x6 ∂x4 Dt ∂x2

where
( ) ( )
∂2 w G 1 1 E1 h31 E3 h33 d2 E1 h1 E3 h3
p = − ms + q(x, t), Gt = + , Dt = + , and Y = .
∂t2 h2 E1 h1 E3 h3 12 12 Dt E1 h1 + E3 h3

where ms is the mass per unit area of the beam, and q(x, t) is an time-dependent loading.
The structure of the sandwich beam has the possible boundary conditions for w as follows. The longitudinal forces at both ends of
the sandwich beam are
[ ]
Dt ∂ 4 w ∂2 w p
P1 = − P3 = − G t Y − (14a)
Gt d ∂x4 ∂x2 Dt
The bending moment applied on the panels including upper and lower is

∂2 w ∂2 w (14b)
M1 = D1 2
, M3 = D3 2
∂x ∂x

and the bending moment applied on the core is


[ ]
Dt ∂4 w ∂2 w p
M2 = − 4
+ Gt Y 2 + . (14c)
Gt ∂x ∂x Dt
Then, the total bending moment applied to the sandwich beam is
[ ]
Dt ∂4 w ∂2 w p
M= − 4
+ Gt (1 + Y) 2 + , (14d)
Gt ∂x ∂x Dt

and the shear force applied on the sandwich beam is


[ ]
Dt ∂5 w ∂3 w 1 ∂p
S= − 5
+ Gt (1 + Y) 3 + . (14e)
Gt ∂x ∂x Dt ∂x

2.3. Hydroelastic analysis of the VLFSS with a hierarchical UHPC core

As shown in Fig. 9, the problem considered in this study is two-dimensional in nature. A sandwich beam of thickness c floats on the
equilibrium surface of a fluid, whose depth is assumed to be h. Rectangular Cartesian coordinates are adopted. The origin is in the
middle of the plate at the intersection of the still water level, and the z-axis is vertically upward. The coordinate of the left edge of the
beam is (− b, 0). Meanwhile, we consider the vibration of the sandwich beam when it is subjected to a train of small-amplitude surface

Fig. 9. Analytical model of wave scattering by a sandwich beam floating on the surface of a fluid and the definition of coordinate system.

12
H. Fang et al. Marine Structures 79 (2021) 103055

gravity waves propagating from left to right in the direction of the positive x-axis. Fig. 8(a) shows the transverse vibration model of the
sandwich beam. The length of the beam is B (B = 2b). Under wave action, the differential equation of the beam motion is given in Eq.
(13).
The present physical problem is solved by using the potential theory [41], where the fluid is assumed to be inviscid and incom­
pressible, and its motion is irrotational. Then, the fluid motion can be described by a velocity potential Φ(x, z, t). Further considering
time-harmonic waves with angular frequency ω, the velocity potential is written as
[ ]
Φ(x, z, t) = Re φ(x, z)e− iωt (15)

where Re denotes the real part of the argument, φ(x, z) is the spatial velocity potential.
When the beam makes harmonic motion with a small amplitude, we have
[ ]
q(x, t) = Re Q(x)e− iωt (16)
[ ]
w(x, t) = Re W(x)e− iωt
(17)
[ ]
S(x, t) = Re S(x)e− iωt
(18)

[ ]
Mj (x, t) = Re M j (x)e− iωt
, j = 1, 2, 3 (19)

where Q(x) is the spatial time-independent wave load; W(x) and S(x) are the spatial displacement and shear force of the sandwich
beam, respectively; and Mj (x) the spatial bending moment with the subscripts j = 1, 2 and 3 denoting the j-th layer of the sandwich
beam.
For convenience of study, the whole fluid domain is divided into three regions: the left open region Ω1, x ≤ − b and –h ≤ z ≤ 0; the
region below the floating beam Ω2, − b ≤ x ≤ b and –h ≤ z ≤ 0; the right open region Ω3, x ≥ b and –h ≤ z ≤ 0. The spatial velocity
potentials satisfy the Laplace equation

∂2 φj ∂2 φj
+ 2 = 0, j = 1, 2, 3 (20)
∂x2 ∂z

where the subscript j denotes region j. The velocity potentials also satisfy the boundary conditions associated with free surface, water
bottom and far fields

∂φj ω2
= φj , z = 0, j = 1, 3 (21)
∂z g

∂φj
= 0, z = − h, j = 1, 2, 3 (22)
∂z
( )

lim + ik0 (φ1 − φІ ) = 0 (23a)
x→− ∞ ∂x
( )

lim − ik0 φ3 = 0 (23b)
x→∞ ∂x

where g is the acceleration of gravity, k0 is the wave number, and φI is the velocity potential of incident waves.
Under the assumption that there is no gap between the fluid and the structure, the kinematic and dynamic conditions on the
boundary of the beam are, respectively, as follows:
∂φ2
= − iωW, z = 0 (24)
∂z

Q
= iωφ2 − gW, z = 0 (25)
ρ

where ρ is the density of seawater. By applying Eqs. 13 and 15 − (17), (24) and (25), the elastic boundary condition on the floating
sandwich beam is derived as follows:
[ 6 ( )]
∂ ∂4 ρg − ms ω2 ∂2
− Gt (1 + Y) 4 + − Gt φ2z |z=0
∂x 6 ∂x Dt ∂x 2
( ) (26)
ρω2 ∂2
− − Gt φ2 | z=0 = 0
Dt ∂x2

13
H. Fang et al. Marine Structures 79 (2021) 103055

The motion of the elastic beam must satisfy suitable edge conditions according to its support on two ends. Here we consider typical
edge conditions with zero deflection, zero bending moment applied on the core and panels. The mathematical expressions of these
edge conditions on (x, z) = (0, –b) are given as follows.

(i) Zero deflection at the two ends gives



i ⃒
W(x) ≡ φ2z ⃒⃒ = 0, x = ±b (27a)
ω z=0

(ii) Zero bending moment at ends of the panels gives


⃒ 2
⃒∂
M 1 (x)| =|D1 ⃒⃒ 2 φ2z |z=0 = |0
∂x
, x = ±b (27b)
⃒ 2
⃒∂
M 3 (x)| =|D3 ⃒⃒ 2 φ2z |z=0 = |0
∂x

(iii) Zero bending moment at ends of the core gives


[ ]
Dt ∂4 ∂2 ρg − ms ω2 Dt ρω2
M 2 (x) = − Gt Y + φ2z |z=0 − φ | = 0, x = ±b#27c (27c)
Gt ∂x4 ∂x2 Dt Gt Dt 2 z=0
Meanwhile, the shear force applied on the sandwich beam is calculated by
[ ]
Dt ∂5 ∂3 ρg − ms ω2 ∂ ρω2
S(x) = − + Gt (1 + Y) − φ | + φ | (28)
Gt ∂x5 ∂x3 Dt ∂x 2z z=0 Dt 2x z=0
In the left open region Ω1, by using the separation of the variables method, the series expression of the velocity potential, satisfying
the governing equation (20) and the boundary conditions (21), (22) and (23a), is written as
[ ]
igH ( ik0 (x+b) ) ∑∞
φ1 = − e + R0 e− ik0 (x+b)
Z0 (z) + Rn ekn (x+b)
Zn (z) (29)
2ω n=1

where H is the height of the incident wave, Rn (n ≥ 0) are unknown complex expansion coefficients, and the vertical eigenfunctions Zn
(z) are given by
cosh k0 (z + h) coskn (z + h)
Z0 (z) = , Zn (z) = (n ≥ 1) (30)
cosh k0 h coskn h
In Eqs. (29) and (30), the wave numbers kn are the positive real roots of the dispersion relation

ω2 = gk0 tanhk0 h = − gkn tankn h(n ≥ 1) (31)

In the floating sandwich beam region Ω2, the velocity potential satisfying the governing equation (20) and the boundary conditions
(22) and (26) is written as
∞ ( )
igH ∑ cosh λn x sinh λn x
φ2 = − An + Bn Yn (z) (32)
2ω n=− 2 cosh λn b cosh λn b

where An and Bn (n ≥ − 2) are unknown complex expansion coefficients, and the vertical eigenfunctions Yn(z) are given by
cos λn (z + h)
Yn (z) = (n ≥ − 2) (33)
cos λn h
In Eqs. (32) and (33), the wavenumbers λn satisfy the following dispersion relation:
[ ]
ρg − ms ω2 ( 2 ) ρω2 ( 2 )
λ6n − Gt (1 + Y)λ4n + λn − Gt λn tan λn h = λ − Gt (34)
Dt Dt n
It is noted that λ0 is the negative pure imaginary root of Eq. (34); λ− 2 and λ− 1 are a pair of conjugate complex roots with positive real
parts; and λn (n ≥ 1) are the positive real roots.
In the right open region Ω3, the velocity potential satisfying Eq. (20) − (22) and (23b) is written as
[ ]
igH ∑∞
φ3 = − T0 eik0 (x− b) Z0 (z) + Tn e− kn (x− b) Zn (z) (35)
2ω n=1

14
H. Fang et al. Marine Structures 79 (2021) 103055

where Tn (n ≥ 0) are unknown complex expansion coefficients.


The expansion coefficients in Eqs. (29), (32) and (35) need to be determined by the sandwich beam edge condition (27) and the
following matching conditions on the common boundaries between adjacent regions
φ1 = φ2 , x = − b, − h ≤ z ≤ 0 (36a)

∂φ1 ∂φ2
= , x = − b, − h ≤ z ≤ 0 (3b)
∂x ∂x

φ3 = φ2 , x = b, − h ≤ z ≤ 0 (37a)

∂φ3 ∂φ2
= , x = b, − h ≤ z ≤ 0 (37b)
∂x ∂x
Substituting Eqs. (29), (32) and (35) into Eqs. (36) and (37), multiplying both sides of the resulting equations by Zm (z), and
integrating with respect to z from − h to 0, which yields


Rm I m − (An − Bn tanhλn b)Λnm = − δm0 I0 , m = 0, 1, 2, ⋯ (38a)
n=− 2



̃
km Rm Im − λn ( − An tanhλn b + Bn )Λnm = ̃
k0 δm0 I0 , m = 0, 1, 2, ⋯ (38b)
n=− 2



Tm Im − (An + Bn tanhλn b)Λnm = 0, m = 0, 1, 2, ⋯ (39a)
n=− 2



̃
km Tm Im + λn (An tanhλn b + Bn )Λnm = 0, m = 0, 1, 2, ⋯ (39b)
n=− 2

where
∫0 ∫0
Λnm = Yn (z)Zm (z)dz, Im = Zm2 (z)dz.
− h − h

= 0), ̃
δ00 = 1, δm0 = 0(m ∕ k0 = − ik0 , ̃
km = km (m ∕
= 0)
Substituting Eq. (32) into the edge condition (27) gives the following equations:


(An − Bn tanh λn b)λn tanλn h = 0 (40a)
n=− 2



(An − Bn tanh λn b)λ3n tanλn h = 0 (40b)
n=− 2


∞ [ ( ) ]
ρg − m s ω 2 ρω2
(An − Bn tanh λn b) λn tanλn h λ4n − Gt Yλ2n + + =0 (40c)
n=− 2
Dt Dt



(An + Bn tanh λn b)λn tanλn h = 0 (41a)
n=− 2

Table 1
Physical parameters of the VLFSS.
Model h B First/Third layer Second layer Gradation parameter (hierarchical core)

ρ1 h1 = h3 E1 = E3 ρ2 h2 G cagg chol

1 100 1000 7800 0.05 192 852 4.9 2.57 0.16 0.65
2 0.10 708 4.8 2.10 0.18 0.71
3 0.20 400 4.6 1.24 0.4 0.84

*h: the depth of water (unit: m), B: the length of the floating body (unit: m), ρ2 : the density of hierarchical UHPC core (unit: kg/m3), h1: the thickness
of the first layer (unit: m), h3: the thickness of the third layer (unit: m), h2: the thickness of the second layer (unit: m), E1: the elastic modulus of the first
layer and the third layer (unit: GPa), G: the shear modulus of the second layer (unit: GPa), Cagg: the first gradation of the UHPC, Chol: the third
gradation of the UHPC.

15
H. Fang et al. Marine Structures 79 (2021) 103055

Table 2
Optimal designs of hierarchical UHPC core.
Hierarchical UHPC core cagg cvod chol G (GPa)

➀ three levels 0.4 8.0163 × 10 − 6 0.8395 1.2303


➁ first + second level 0.4 0.8395 – 0.7124
➂ first + third level 0.4 – 0.8395 1.2453

Table 3
Convergence examination of the calculation results for model 3.
Truncated number N k0 h = 1 k0 h = 5 k0 h = 10

Kr K2r + K2t Kr K2r + K2t Kr K2r + K2t

5 0.791 1.0000 0.986 1.0000 0.996 1.0000


10 0.802 1.0000 0.988 1.0000 0.997 1.0000
15 0.805 1.0000 0.988 1.0000 0.997 1.0000
20 0.807 1.0000 0.989 1.0000 0.997 1.0000
25 0.809 1.0000 0.989 1.0000 0.997 1.0000
30 0.810 1.0000 0.989 1.0000 0.997 1.0000
35 0.810 1.0000 0.989 1.0000 0.997 1.0000

Fig. 10. Variations of the reflection coefficient Kr and the transmission coefficient Kt versus wave number k0h for model 3 (Table 1).



(An + tanh λn b)λ3n tanλn h = 0 (41b)
n=− 2


∞ [ ( ) ]
ρg − m s ω 2 ρω2
(An + Bn tanh λn b) λn tanλn h λ4n − Gt Yλ2n + + =0 (41c)
n=− 2
Dt Dt

After truncating n in Eq. (38) − (41) to N + 1 terms, and m in Eq. (38) to N terms, we obtain a system of linear equations including 4
N + 10 unknowns. Solving this linear system gives all the expansion coefficients in the velocity potentials.
The reflection coefficient Kr and the transmission coefficient Kt of the sandwich structure are calculated by
Kr = |R0 | (42a)

Kt = |T0 | (42b)
The present water wave motion satisfies the energy conservation relation

Kr2 + Kt2 = 1 (43)

The deflection of the sandwich structure W(x) is calculated by

16
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. 11. Hydroelastic properties of the VLFSS models and physical parameters in model 3.

N +1 ( )
gH ∑ cosh λn x sinh λn x
W(x) = − 2
An + Bn λn tanλn h (44)
2ω n=− 2 cosh λn b cosh λn b

The bending moment on the sandwich beam is calculated by


Dt gH
M(x) =
Gt 2ω2
(45)
N +1 (
∑ )[ ( ) ]
cosh λn x sinh λn x ρg − ms ω2 ρω2
An + Bn λn tanλn h λ4n − Gt (1 + Y)λ2n − +
n=− 2
cosh λn b cosh λn b Dt Dt

The shear force on the sandwich beam is calculated by


Dt gH
S(x) =
Gt 2ω2
(46)
N +1 (
∑ ) [ ( ) ]
sinh λn x cosh λn x ρg − ms ω2 ρω2
An + Bn λn λn tanλn h λ4n − Gt (1 + Y)λ2n − +
n=− 2
cosh λn b cosh λn b Dt Dt

The normalized deflection, bending moment and shear force are, respectively, defined as:
⃒ ⃒
⃒2W ⃒
CW = ⃒⃒ ⃒⃒ (47a)
H

17
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. 12. Deflection, bending moment and shear force for models (a) 1, (b) 2 and (c) 3 in Table 1.

⃒ ⃒
⃒ 2M ⃒
⃒ ⃒
CM = ⃒ ⃒ (47b)
⃒ρgHh2 ⃒

⃒ ⃒
⃒ 2S ⃒
⃒ ⃒
CS = ⃒ ⃒ (47c)
⃒ρgHh⃒

Here three different models of sandwich beam are considered, and the detailed physical parameters are listed in Table 1. The length
of all three models is 1000 m, and the water depth is 100 m. In the first model, the thickness of the panel is set as to 0.05 m, and the
thickness of the core material is 4.9 m. In the other two models (see Table 1), the total thickness of the sandwich structure is kept
constant, and the thicknesses of the three layers are changed. Then, the differences of the hydrodynamic quantities among the three
models are examined and analyzed.
To examine the convergence of the calculation results, the values of the reflection coefficient Kr at different truncated numbers N
are listed in Table 3. It is observed from the table that the results with three-figure accuracy are obtained when the truncated number N
reaches to 35, and the energy conservation relation (43) is well satisfied. Thus, the truncated number N = 35 is adopted in the sub­
sequent calculations.
In Fig. 10, the variations of the reflection coefficient Kr and the transmission coefficient Kt versus wavenumber k0 h are plotted for
model 3. The results in Fig. 10 show that the variations of the reflection and transmission coefficients with the increasing wavenumber
are oscillating, and zero wave reflection occurs to some specific wave frequencies, that is to say, the incident wave energy propagates

18
H. Fang et al. Marine Structures 79 (2021) 103055

completely towards the leeside of the floating structure.


The calculation results of the normalized deflection, bending moment and shear force on the VLFSS model 3 (see Table 1) are
presented in Fig. 11, where three different incident wavenumbers (frequencies) are considered. It is observed from Fig. 11(a) that there
are multiple peaks of the normalized deflection, and the peak increases with the increasing incident wavenumber. The variations of the
normalized bending moment and shear force along the beam length are similar to that of the normalized deflection, as shown in Fig. 11
(b) and (c). Actually, the number of peaks highly depends on the physical parameters of the sandwich beam. The locations corre­
sponding to these peaks should be the focus of attention in the practical designs of floating sandwich structures.

3. Mechanical optimizations of the VLFSS with a hierarchical UHPC core

As shown in section 2, the hydroelasticity characteristics, honeycomb core and multilevel aggregate of UHPC are coupled in an
analytic model. For engineering applications, this model can be utilized to calculate the hydroelastic responses and optimize the
hierarchical structures of the VFLSS. As an illustration, the three models presented in Table 1 are considered, in which the depth of
water is 100 m and the length and height of the floating body are 1000 m and 5 m, respectively. The average density (= 990 kg/m3) and
total thickness of the floating body are set as limiting factors and remain unchanged in this study, and the floating capacities of these
three models are all the same.
Considering model 3, the thickness of the surface panel is 0.05 m, and the thickness of the UHPC core is 4.9 m. In conformity with
the optimization method of the UHPC hierarchical core presented in subsection 2.1, other physical parameters are calculated and
shown in Table 1. According to the coupled analytic model above, the deflection, bending moment and shear force along the beam
length and k0 h are shown in Fig. 12 (a). k0 h ranges from 0.01 to 6 [42]. By virtue of the edge condition, the surface deflection
magnitude will vanish at x/b = − 1 and x/b = 1. As expected, the hydroelastic and mechanics responses are consistent with the wave
propagation characteristics that are presented in section 2.3. The shear forces (S) significantly exist at some cross-sectional areas of the
VLFSS. According to the theory of bending beams, the shear force is in equilibrium with the section bending moment and external
applied loads. It is complex to optimize the shear force using the traditional methodology based on the independent theory of
hydroelasticity and composite structure because the material, structural and wave characteristics interact with each other. Utilizing
the coupled analytical methodology presented in section 2, the parameters of the UHPC core aggregate, hole and thickness are syn­
chronously embedded and adjusted in the hydroelasticity and mechanics calculations. As an initial optimization, model 2 (Table 1) is
designed with a thicker core (h2 = 4.8 m), a smaller proportion of aggregates (cagg = 0.18) and a smaller porosity of the core (chol =
0.71). The equivalent Young’s modulus of the UHPC core is G = 2.1 GPa. Fig. 12 (b) shows the hydroelastic response of model 2. The
deflection and moment of model 2 are different from those of model 3, but the maximal values of these responses are similar to those of
model 3. The shear force amplitudes (Fig. 12 (b3)) are much smaller than those of model 3. As a further optimization, model 1 (Table 1)
is designed with a thicker core (h2 = 4.9 m), a smaller proportion of aggregates (cagg = 0.16) and a smaller porosity of the core (chol =
0.65). The equivalent Young’s modulus of the UHPC core is G = 2.57 GPa. Fig. 12 (c) shows the hydroelastic response of model1. The
maximal values of the deflection and moment responses are similar to those of models 3 and 2. The shear force amplitudes (Fig. 12
(c3)) are smaller than those of model 2. In this optimization procedure above, the shear force is set as a design objective that provides a
demonstration of the multiscale material-structural-hydroelastic coupled analytical model and optimization methodology. In engi­
neering applications, the adaptation to specific objectives would be performed.

4. Conclusion

In this study, high-performance composite materials and structures are effectively utilized to enhance mechanical performance of
marine structures. Considering the multiple material, structure and hydroelasticity characteristics at different geometric scales, a novel
model of the hydroelasticity equations containing multi-scale material, structural and wave parameters are innovated and utilized for
the optimization of a VLFSS with a hierarchical UHPC core. 3-D RVE, sixth-order dynamic motion and potential flow models are
deduced and coupled as governing equations containing multi-scale parameters, which are solved to reveal the physical relations
between the macroscale hydroelasticity and mesoscale hierarchical characteristics. As an illustration, the mechanics characteristics of
a representative VLFSS are effectively optimized, which demonstrate the potential of the multiscale coupling model to analyze the
interactivity of hierarchical structural parameters and wave conditions of a composite structure in ocean engineering. In the further,
the multiscale hydro-elastoplastic model of floating structures [21] have been studied, in which the nonlinear coupling model is the
key issue.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgments

This work was supported by the Key Engineering Innovation Project of Shandong Province (2019JZZY010301), NSFC (51725903)
and the Taishan Scholar Project of Shandong Province under Grant No. ts20190915. We thank PHD student Jibo Sun for numerical
simulations to the revised manuscript.

19
H. Fang et al. Marine Structures 79 (2021) 103055

Appendix. Details of simulating principle

Fig. A1. Deformation of the UHPC core’s RVEs of simple cube foam: (a) x tension, (b) y tension, (c) z tension, (d) xy shear, (e) xz shear and (f)
yz shear.

20
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. A2. (a) Stress and (b) strain ε11 of the corresponding nodes located on two opposite surfaces FaceRight and FaceLeft of the UHPC core’s RVEs of
simple cube foam (Periodic boundary condition ε = [ε, 0, 0, 0, 0, 0]T ).

21
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. A3. Deformation of the UHPC core’s RVEs of body center cube foam: (a) x tension, (b) y tension, (c) z tension, (d) xy shear, (e) xz shear and (f)
yz shear.

22
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. A4. (a) Stress σ 11 and (b) strain ε11 of the corresponding nodes located on two opposite surfaces FaceRight and FaceLeft of the UHPC core’s
RVEs of body center cube foam (Periodic boundary condition ε = [ε, 0, 0, 0, 0, 0]T ).

23
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. A5. Deformation of the UHPC core’s RVEs of square lattices of the through holes: (a) x tension, (b) y tension, (c) z tension, (d) xy shear, (e) xz
shear and (f) yz shear.

24
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. A6. (a) Stress σ 11 and (b) strain ε11 of the corresponding nodes located on two opposite surfaces FaceRight and FaceLeft of the UHPC core’s
RVEs of square lattices of the through holes (Periodic boundary condition ε = [ε, 0, 0, 0, 0, 0]T ).

25
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. A7. Deformation of the UHPC core’s RVEs of hexagonal lattices of the through holes: (a) x tension, (b) y tension, (c) z tension, (d) xy shear, (e)
xz shear and (f) yz shear.

26
H. Fang et al. Marine Structures 79 (2021) 103055

Fig. A8. (a) Stress σ 11 and (b) strain ε11 of the corresponding nodes located on two opposite surfaces FaceRight and FaceLeft of the UHPC core’s
RVEs of hexagonal lattices of the through holes (Periodic boundary condition ε = [0, 0, 0, γ, 0, 0]T ).

References

[1] Isobe E. Research and development of mega-float. In: Ertekin RC, Kim JW, editors. Proc 3rd int WkspVery large floating structures, vol. I; 1999. p. 7–13.
Honolulu, Hawaii, USA: September 22–24.
[2] Watanabe E, Utsunomiya T, Wang CM. Hydroelastic analysis of pontoon-type VLFS: a literature survey. Eng Struct 2004;26:245–56. https://doi.org/10.1016/j.
engstruct.2003.10.001.
[3] Fox C, Squire VA. Reflection and transmission characteristics at the edge of shore fast sea ice. J Geophys Res 1990;95(C7):11629–39. https://doi.org/10.1029/
JC095iC07p11629.
[4] Fox C, Squire VA. On the oblique reflection and transmission of ocean waves at shore fast sea ice. Philos T R Royal Soc A 1994;347(1682):185–218. https://doi.
org/10.1098/rsta.1994.0044.
[5] Squire VA. Of ocean waves and sea-ice revisited. Cold Reg Sci Technol 2007;49:110–33. https://doi.org/10.1016/j.coldregions.2007.04.007.
[6] Teng B, Gou Y, Cheng L, Liu S. Draft effect on wave action with a semi-infinite elastic plate. Acta Oceanol Sin 2006;25(6):116–27. https://doi.org/10.1016/j.
marchem.2005.09.003.
[7] Guo Y, Liu Y, Meng X. Oblique wave scattering by a semi-infinite elastic plate with finite draft floating on a step topography. Acta Oceanol Sin 2016;35(7):
113–21. https://doi.org/10.1007/s13131-015-0760-2.
[8] Mohapatra SC, Guedes Soares G. Interaction of ocean waves with floating and submerged horizontal flexible structures in three-dimensions. Appl Ocean Res
2019;83:136–54. https://doi.org/10.1016/j.apor.2018.10.009.
[9] Zhao G, Hu C, Wei Y, Zhang J, Huang W. Diffraction of surface waves by floating elastic plates. J Fluid Struct 2008;24(2):231–49. https://doi.org/10.1016/j.
jfluidstructs.2007.07.006.
[10] Li HT, Wang QB, Zong Z, Sun L, Liang H. Stochastic hydroelastic analysis of a very large floating structure using pseudo-excitation method. Appl Ocean Res
2014;48:202–13. https://doi.org/10.1016/j.apor.2014.08.003.
[11] Santos FM, Temarel P, Guedes Soares C. On the limitations of two- and three-dimensional linear hydroelasticity analyses applied to a fast patrol boat. P I Mech
Eng M-J Eng 2009;223:457–78. https://doi.org/10.1243/14750902JEME152.
[12] Hirdaris SE, Temarel P. Hydroelasticity of ships: recent advances and future trends. P I Mech Eng M-J Eng 2009;223:305–30. https://doi.org/10.1243/
14750902JEME160.
[13] Crupi V, Epasto G, Guglielmino E. Comparison of aluminium sandwiches for lightweight ship structures: honeycomb vs. foam. Mar Struct 2013;30:74–96.
https://doi.org/10.1016/j.marstruc.2012.11.002.

27
H. Fang et al. Marine Structures 79 (2021) 103055

[14] Fang H, Li JG, Chen HB, Liu B, Huang W, Liu YL, Wang TJ. Radiation induced degradation of silica reinforced silicone foam: experiments and modeling. Mech
Mater 2017;105:148–56. https://doi.org/10.1016/j.mechmat.2016.11.006.
[15] Hoo Fatt MS, Sirivolu D. Marine composite sandwich plates under air and water blasts. Mar Struct 2017;56:163–85. https://doi.org/10.1016/j.
marstruc.2017.08.004.
[16] Zhang ZJ, Zhang QC, Zhang DZ, Li Y, Jin F, Fang DN. Enhanced mechanical performance of brazed sandwich panels with high density square honeycomb-
corrugation hybrid cores. Thin-Walled Struct 2020;151:106757. https://doi.org/10.1016/j.tws.2020.106757.
[17] Luo J, Shao X, Fan W, Cao J, Deng S. Flexural cracking behavior and crack width predictions of composite (steel + UHPC) lightweight deck system. Eng Struct
2019;194:120–37. https://doi.org/10.1016/j.engstruct.2019.05.018.
[18] Fujikubo M, Yao T. Structural modeling for global response analysis of VLFS. Mar Struct 2001;14:295–310. https://doi.org/10.1016/S0951-8339(00)00062-9.
[19] Fujikubo M. Structural analysis for the design of VLFS. Mar Struct 2005;18:201–26. https://doi.org/10.1016/j.marstruc.2005.07.005.
[20] Bernard O, Ulm FJ, Lemarchand E. A multiscale micromechanics-hydration model for the early-age elastic properties of cement-based materials. Cement Concr
Res 2003;33(9):1293–309. https://doi.org/10.1016/S0008-8846(03)00039-5.
[21] Borges DC, JJC Pituba. Analysis of quasi-brittle materials at mesoscopic level using homogenization model. Adv Concr Constr 2017;5(3):221–40. https://doi.
org/10.12989/acc.2017.5.3.221.
[22] Abbès Fazilay, Abbès Boussad, Benkabou Rim, Asroun Aïssa. A FEM multiscale homogenization procedure using nanoindentation for high performance
concrete. Journal of Applied and Computational Mechanics 2020;6(3):493–504. https://doi.org/10.22055/JACM.2019.29832.1640.
[23] Chen Q, Wang H, Li HX, Jiang ZW, Zhu HH, Ju JW, Yan ZG. Multiscale modelling for the ultra-high performance concrete: from hydration kinetics to
macroscopic elastic moduli. Construct Build Mater 2020;247:118541. https://doi.org/10.1016/j.conbuildmat.2020.118541.
[24] Pichler C, Lackner R, Mang HA. Multiscale model for creep of shotcrete-from logarithmic-type viscous behavior of CSH at the μm-scale to macroscopic tunnel
analysis. J Adv Concr Technol 2008;6(1):91–110. https://doi.org/10.3151/jact.6.91.
[25] Chen Q, Zhu HH, Ju JW, Guo F, Wang LB, Yan ZG, Deng T, Zhou S. A stochastic micromechanical model for multiphase composites containing spherical
inhomogeneities. Acta Mech 2015;226(6):1861. https://doi.org/10.1007/s00707-014-1278-y.
[26] Garboczi EJ, Berryman JG. Elastic moduli of a material containing composite inclusions: effective medium theory and finite element computations. Mech Mater
2001;33(8):455–70. https://doi.org/10.1016/S0167-6636(01)00067-9.
[27] Guo W, Fan W, Shao XD, Shen DJ, Chen BS. Constitutive model of ultra-high-performance fiber reinforced concrete for low velocity impact simulations. Compos
Struct 2018;185:307–26. https://doi.org/10.1016/j.compstruct.2017.11.022.
[28] Shen GL, Hu GK, Liu B. Mechanics of composite materials. Beijing: Tsinghua University Press; 2013.
[29] Suzuki H. Overview of Megafloat: concept, design criteria, analysis, and design. Mar Struct 2005;18:111–32. https://doi.org/10.1016/j.marstruc.2005.07.006.
[30] Teng B, Cheng L, Liu SX, Li FJ. Modified eigenfunction expansion methods for interaction of water waves with a semi-infinite elastic plate. Appl Ocean Res 2001;
23:357–68. https://doi.org/10.1016/S0141-1187(02)00005-6.
[31] ABAQUS 6.14. Analysis user’s manual.
[32] Toshio M. Micromechanics of defects in solids. Mechanics of elastic and inelastic solids 3. 2nd. Dordrecht, Netherlands; Boston: Hingham, MA, USA: Martinus
Nijhoff Publishers; 1987.
[33] Hill R. Elastic properties of reinforced solids: some theoretical principles. J Mech Phys Solid 1963;11(5):357–72. https://doi.org/10.1016/0022-5096(63)
90036-X.
[34] Hazanov S, Huet C. Order relationships for boundary conditions effect in heterogeneous bodies smaller than the representative volume. J Mech Phys Solid 1994;
42(12):1995–2011. https://doi.org/10.1016/0022-5096(94)90022-1.
[35] Hori M, Nemat-Nasser S. On two micromechanics theories for determining micromacro relations in heterogeneous solids. Mech Mater 1999;31:667–82. https://
doi.org/10.1016/S0167-6636(99)00020-4.
[36] Bouaoune L, Brunet Y, Moumen AE, Kanit T, Mazouz H. Random versus periodic microstructures for elasticity of fibers reinforced composites. Compos B Eng
2016;103:68–73. https://doi.org/10.1016/j.compositesb.2016.08.026.
[37] Mori T, Tanaka K. Average stress in matrix and average elastic energy of materials with misfitting inclusions. Acta Metall 1973;21(5):571–4. https://doi.org/
10.1016/0001-6160(73)90064-3.
[38] Benveniste YA. A new approach to the application of Mori-Tanaka’s theory in composite materials. Mech Mater 1987;6(2):147–57. https://doi.org/10.1016/
0167-6636(87)90005-6.
[39] Kouznetsova VG. Computational homogenization for the multi-scale analysis of multi-phase materials. Eindhoven: Technische Universiteit Eindhoven; 2002.
[40] Mead J, Markus S. The forced vibration of a three-layer, damped sandwich beam with arbitrary boundary conditions. J Sound Vib 1969;10(2):163–75. https://
doi.org/10.1016/0022-460X(69)90193-X.
[41] Mei CC, Stiassnie M, Yue DKP. Theory and applications of ocean surface waves: linear aspects. World Scientific Publishing; 2005.
[42] Meylan M, Squire VA. The response of ice floes to ocean waves. J Geophys Res 1994;99:891–900. https://doi.org/10.1029/93JC02695.

28

You might also like