Next Article in Journal
Interfacial Microstructure Produced during Dissimilar AA6013/Ti-6Al-4V Friction Stir Lap Welding under Zero-Penetration Condition
Next Article in Special Issue
Fabrication of Uniform and Rounded Closed-Cell Aluminum Foams Using Novel Foamable Precursor Particles (FPPs)
Previous Article in Journal
Performance Assessment of Wood Ash and Bone Char for Manganese Treatment in Acid Mine Drainage
Previous Article in Special Issue
Combining 3D Printing and Electrochemical Deposition for Manufacturing Tailor-Made 3D Nickel Foams with Highly Competitive Porosity and Specific Surface Area Density
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bird’s Eye View on Lattice Structures: Design Issues and Applications for Best Practices in Mechanical Design

Dipartimento di Ingegneria Meccanica e Aerospaziale, Sapienza Università di Roma, 00184 Rome, Italy
*
Author to whom correspondence should be addressed.
Metals 2023, 13(10), 1666; https://doi.org/10.3390/met13101666
Submission received: 14 August 2023 / Revised: 22 September 2023 / Accepted: 26 September 2023 / Published: 28 September 2023

Abstract

:
Lattice structures for engineering applications are patterns of unit cells designed to make a larger functional structure. Research on lattice structures ranges in many fields, from mechanical characterization and cell and pattern designs in respect of their applications, to the manufacturing process and its final shape control. From the manufacturing point of view, some kinds of lattice structures can be infeasible when approached with traditional manufacturing methods. It may offer an inevitable limitation of their adoption. However, advancements in Additive Manufacturing (AM) have solved this manufacturing issue to a great extent, allowing to obtain major complexity of the cells that can be achieved. The topology, shape of the unit cell, and the characteristics of its replication pattern allow us to obtain many kinds of structures in respect of the different engineering requirements and manufacturing constraints. Nevertheless, the necessity of new or dedicated CAD-CAE approaches arises to manage the domains of multiscale modeling. These are some of the advantages and disadvantages that may arise while approaching the design of a component using lattice structures. The aim of this paper is to provide an overview that integrates the most recent applications of lattice structures with their related design and manufacturing issues so that, from a practical design point of view, any state-of-the-art improvements may be established in respect of the related field of applications. In this article, engineers and researchers may find a practical summary of the capabilities and processes of lattice structures that are currently available from a design and development point of view.

1. Introduction

A lattice structure is made by a unit cell that is repeated in the space, making a larger functional structure; thus, it is a peculiar kind of cellular structure. The basic idea of cellular structures evolved from natural organic structures such as bamboo, bone, cork, etc. For the very first time, the concept of cellular structures was introduced by Gibson and Ashby [1,2]. Initially, they classified cellular structures into foam (open cell and closed cell, Figure 1a,b) and honeycomb. Foam’s cell shape and pattern are usually defined by random distributions, while the honeycomb structure has a regular 2D hexagonal unit cell and the cell walls are uniformly oriented in the space (as shown in Figure 1c). In open-cell foam, the cells are not totally enclosed by their walls, while in closed-cell foam, the cells are entirely enclosed by their walls. Despite these differences, the component is always defined from a micro/mesoscale cell, replicated to fill the overall macroscale volume. However, the cellular structures were later classified [3,4] into three main categories: foam, honeycomb, and lattice.
According to the literature, in many applications, lattice structures have been found to be superior to their other cellular counterparts, such as foam and honeycomb structures. A more in detail strength-to-weight ratio and mechanical performance, such as energy absorption or compliance, is generally better due to the tailored shape of the cell and the replication pattern [6,7]. Moreover, they can support functional designs, such as heat exchangers or fluid conveyors, as demonstrated in [8]. More recently, taking inspiration from natural organic structures such as bones and plant stems, the concept of graded densities increased the spread of lattice structures, for example, being used to optimize load distributions [9,10]. It also furthers the evolution towards multigrade lattice structures, coupled with different materials. From this, the concept of multigrade and multifunction lattices arises. These are the solutions in which not only the type of unit cells but also the cell’s density and materials are distributed to gain a specific performance or function. Multigrade lattice structures have numerous applications, i.e., to enhance stiffness [11], to increase energy absorption capability [12,13,14,15,16,17], to improve thermal exchange efficiency [18,19], to tailor specific behaviors such as a negative Poisson’s ratio [20,21,22], etc.
From the manufacturing point of view, Additive Manufacturing (AM) is often considered as a specific technology to manufacture lattice structures; indeed, the capability of reproducing specific cell designs and distribution through AM increases the fields of investigation pertaining to lattice structures in the respect of foams. Nevertheless, foams obtained by blowing agents (in batch foaming, extrusion, or injection molding) may have a superior reliability concerning the mechanical characteristics since the AM interlayer interfaces may present delamination and, thus, crack propagation [23]. Besides this aspect, the advancements in AM with multigrade materials for lattice structures have gained immense popularity, providing an extraordinary performance and array of functions [24,25].
Due to these promising characteristics, lattice structures have been extensively studied and they are still undergoing rigorous research to support numerous industrial applications. Research on lattice structures ranges in many fields, such as mechanical characterization, manufacturing processes and its final shape control, and cell and pattern designs in respect of their applications.
This paper aims to provide a bird’s eye view concerning the mechanical research of lattice structures, linking their geometrical characteristics with the type of applications and related design practices, considering both manufacturing and the necessity of new or dedicated CAD-CAE approaches to manage the domains of multiscale modeling.
The final goal is to highlight a general perspective for addressing the design workflow of lattice structures and support engineering best practices in the design of these kinds of solutions.
For this, in Section 2, the lattice structure classification is reviewed and documented in terms of cell, replication pattern, and mechanical characteristics. In Section 3, applications related to the main engineering fields with respect to the types of lattice structures are presented. Then, in Section 4, the CAD-CAE design approach is discussed, together with the geometrical characteristics of lattice structures and their major manufacturing issues. Finally, in Section 5, basic conclusions and perspectives are suggested.

2. Lattice Classification

According to [23], lattice structures may be classified based on:
  • The unit cell’s characteristics, which include topology, cell’s element geometry, and cell’s size.
  • The characteristics of the cell’s replication pattern.
However, in [26,27], the classification of the lattice structure has been discussed in more detail; therefore, the latest classification of lattice structure can be shown as Figure 2.

2.1. Unit Cell Characteristics

A topology-based classification defines whether the cells are closed or open and how many cells’ faces are open, while a cell’s element geometry, according to [2], distinguishes between cells made of struts and cells made of plates or shells, introducing a structural behavior classification of the cells. A cell’s element geometry includes a geometrical definition of the structural elements (sections, number of edges, …) [6,7,23,28], and the size, including all the dimensional elements that define a cell (e.g., section size and axial length for struts, length, width, and thickness for plate/shell). The cell’s element geometry, indeed, may also constrain the topology of the cell (e.g., strut cells are always open, shells and plates may be open with different geometrical shapes). Concerning the cell’s size, the distinction between homogeneous and heterogeneous defines whether the structural elements of the cell are constant or not. Figure 3 and Figure 4 show an overview of strut-based and surface-based lattice cells, respectively.
Concerning strut-based unit cells, Simple Cubic (SC), Body-Centered (BC), and Body-Centered Cubic (BCC) are derived from the same cubic cell, just with an increasing number of beams [23,29,30,31]. The Octet Truss (OT) cell comes from the face-centered cell [32]. Other frequently used cells are the modified Gibson–Ashby (GA) and the modified Wallach–Gibson (WG) cells [6,7,23,28]. Figure 3h shows an example of a re-entrant cell.
It is an auxetic cell designed to present a negative Poisson’s ratio, which means that the cell enlarges when stretched and contracts when compressed [33,34,35,36].
Concerning surface-based unit cells, we can distinguish between plate and shell cells, and, more generally, surface-based cells like Triply Periodic Minimal Surface (TPMS). In [37], according to Gibson, the plate cells were investigated connecting some struts of Figure 3 with plates, as shown in Figure 4a–c. These solutions, although may be replicated as periodic and regular, are often considered a type of closed foam, as commented in [38]. A honeycomb lattice [28], shown in Figure 4d–f, may have unit cells that are hexagonal or triangular; recently, the re-entrant configurations have also been studied.
Triply Periodic Minimal Surfaces (TPMSs) are being widely investigated. They are bio-inspired cells with boundary surfaces with a zero-mean curvature at every point [39,40]. TPMSs are relevant as functionally graded structures since their geometric characteristics allow them to reach different surface-related properties (e.g., manufacturability, fluid permeability, electrical and thermal conductivity). According to the procedure adopted to model them, they can be classified as skeletal and sheet TPMSs. In the first case, a volume is trimmed by the TPMS and the remaining part is thinned (skeletonized) with an assigned thickness. In the second case, the TPMS represents the shell sheet that is subsequently thickened.

TPMS Formulation

From a mathematical point of view, a TPMS is an infinite surface, periodic along three mutually orthogonal directions, which satisfies the following conditions:
H x , y , x = 0 H = k 1 + k 2 2
where k1 and k2 stand for the principal curvatures of the surface at a general point P. Many authors studied the theoretical formulations of TPMS, starting from different mathematical points of view. In [39], the level set method, an approach applied in mechanical engineering to optimize topology in accordance with structural requirements, is adopted to study five TPMSs, namely Primitive, Gyroid, IWP, Diamond, and Fisher–Koch. Assuming that a TPMS may be described by a sum of Fourier terms and that its leading term is the basis function of the TPMS, i.e., φ(x,y,z), the TPMS can be achieved as iso-level curves:
φ x , y , z = c o n s t
Subjected to Equation (1). Equations (3)–(5) show three formulations achieved according to such reasoning, reported in Figure 4g–i.
Schwartz primitive:
φ x , y , z = c o s x + c o s y + c o s z
Schwartz diamond:
φ x , y , z = c o s x c o s y c o s z s i n x s i n y s i n z
Gyroid:
φ x , y , z = s i n x c o s y + s i n y c o s z + s i n z c o s x
where:
x = 2 π X L x , y = 2 π Y L y , and z = 2 π Z L z .
The L x , L y , and L z are the unit cell size in the X, Y, and Z directions. x, y, and z are the periodicities, while X, Y, and Z are the cell’s repetitions.

2.2. Characteristics of the Cell Replication Pattern

Characteristics of the replication pattern are concerned with the cell’s connection, orientation, and density in the space, as discussed in [7,41,42]. It also pertains to the cell size in the sense that the cell’s length may also change along the replication pattern randomly or according to a gradient as a request of the design intent (e.g., functionally graded lattice structures) [43,44,45,46].
Replication patterns can be:
  • Regular.
  • Pseudo-regular.
    Hybrid,
    Warped (or gradient) by cell size,
    Warped (or gradient) by thickness,
    Conformal.
  • Stochastic.
    Random by cell size,
    Random also by thickness.
In a regular replication pattern, no changes regarding the unit cell characteristics occur at all (Figure 5). Hybrid patterns include different unit cells that may also have different geometric characteristics (Figure 6). They can be considered as part of pseudo-regular patterns that provide several replications based on the variation in the unit cell’s size.
Pseudo-regular patterns also include warped (gradient) and conformal structures (Figure 7). The gradient by cell size is a replication pattern that provides a structure in which the unit cell’s size varies gradually along a specified direction, while the thickness of the elements remains the same (Figure 7a). In the case of the gradient by thickness, the structure maintains a constant cell size, but its thickness is changed (Figure 7b). Conformal lattice structures consist of cells changing in length and shapes non-homogeneously, so that the replication pattern may follow the boundaries of the part (Figure 7c) [47,48].
In stochastic patterns, the cell size varies randomly, while the thickness remains the same. However, in stochastic random patterns, the thickness varies randomly or along a gradient throughout the structure (Figure 8).

2.3. Mechanical Behavior

The mechanical behavior of cellular materials is affected by the micro/mesoscale characteristics (unit cell and replication pattern), in addition to the macroscale topology and shape of the overall component [49,50]. This means that the local behaviors determine the global mechanical response and its related performance. Based on their local mechanical response, the lattice structure is generally classified into [51]:
  • Bending dominated.
  • Stretch dominated.
Bending-dominated structures react locally with a bending deformation. This makes them extremely useful (such as compliant structures) for applications related to mechanical energy absorption, especially in crashworthiness design [52,53]. Stretch-dominated structures experience a stretch deformation under the action of a uniaxial compression or tension within their material linear elastic limit (e.g., yielding strength, ultimate strength). This makes them useful for a lightweight design without losing the global stiffness of the component.
The Maxwell stability criterion may help to determine the local behavior of a truss-based cell by its Maxwell number (M):
M = s − 3n + 6
where s is the number of struts, n is the number of joints in the unit cell, and 6 is the d.o.f. in cases of three-dimensional structures. When M equals zero or is greater than zero, it means that the isostatic or over-constrained solutions undertake a stretch-dominated category. The condition of M as less than zero means that the structure is a mechanism assumed to weld the joints (as it will be in a real lattice cell) and it undertakes a bending-dominated category.
According to this reasoning, the bending-dominated cells are BCC and Kelvin structures, while honeycomb (if axially loaded) and Octet cells are stretch dominated (concerning the application of M to the honeycomb, it is worthwhile to remember that shell-based cells are derived from the truss-based cells, assuming the cells are closed by their walls).
In [54], the classification between stretch- and bending-dominated structures is enriched by proposing a third class, named “programmable active”, able to actively merge the two categories. Active structures are functionally graded materials, such as auxetic cells, which have the ability to switch their behavior from stretched dominated to bending dominated. This is extremely suitable for crashworthiness designs when the defined impact loads may activate the plastic bending.
According to the Gibson and Ashby power criterion [51], the elastic modulus of the unit cell can be calculated with the help of the following equations:
For a bending-dominated lattice,
E * E s = C ρ * ρ s 2
For a stretch-dominated lattice,
E * E s = C ρ * ρ s
where E * is the elastic modulus of the cellular structure and E s is the elastic modulus of the equivalent solid. Similarly, ρ * is the density of the cellular structure and ρ s is the density of the equivalent solid. C is the Gibson and Ashby constant which depends on the unit cell’s topology and is determined by experimental tests. From Equations (8) and (9), the modulus density chart of the lattice cells can be defined to assess performance of the cells in respect to the ideal behaviors. In addition, similar equations are available for finding the strength density chart.
Lattice structures made by AM are currently studied from the resistance point of view to validate both the structures and the technological process. More in detail data are necessary to assess the repeatability of the stress–strain behavior to be used in the integrated product–process design. The AM discontinuities at a small scale may penalize the fatigue behavior, as discussed in [55]. In the literature, many other examples of mechanical characterization on lattice structures are provided. Ref [56] studies the compression and three-point bending resistance of 3D Kagome lattices, while [57] provides a comparative study of the auxetic cells that are systematically compared in terms of the Poisson ratio, maximum volume or area reductions, and the equivalent Young’s modulus by CAD-CAE tools. In [40], five TPMS structures are characterized from a dynamic point of view with the Hopkinson bar, finding the strain-rate effect on the stress–strain response. This result is also discussed in terms of the strain-rate sensitivity of the base material and the accuracy of the AM process.

3. Applications

Querying the Scopus database for research fields and applications related to lattice structures published in journals in the last decade, an exponential trend can be clearly seen, which is also confirmed by provisional data from 2023. In detail, Figure 9a shows relative percentage bars for five main topics related to specific technological and design fields (additive manufacturing, lightweight design, energy absorption, biomedical, and heat dissipations) and two emerging keywords related to functionally graded materials that fit the lattice structure classification we made, TPMS and Auxetic. Figure 9b shows the overall number of papers per topic found by the queries. As predicted, TPMS and auxetic are still limited in respect of AM (the most populated topic), lightweight, and energy absorption design. In the middle of the rank, there are biomedical applications and studies, partially related to TPMS, and heat dissipation, which also includes heat exchangers and sinks.
Being part of the class of cellular materials, lattice structures may be adopted to design functional structures. In the respect of foams, lattice structures have more characteristics to be tailored to both the meso and macroscale level, according to the description made in Section 2. A design optimization suitable to fine tune these characteristics enables better performances, as highlighted, for example, in [44], discussing the lightweight design. According to this, one of the most recent overviews concerning lattice structure applications, presented in [58], classifies the applications in respect of the relevant design properties that the lattice structure may help to achieve.
The ability to tailor the stiffness of a structure via the cell topology and pattern [59] enables control of the stress and strain for energy absorption, both in static and dynamic conditions at a high strain rate [60]. This makes them appropriate for crashworthiness designs and for controlling shock absorption, vibration, and acoustic noise, all of which are useful for the automotive, aerospace, ergonomics, ballistics, packaging, robotics, and tool manufacturing fields [12,13,16,17,20,38,61]. According to [33,62], Kelvin cells must be cited as structures derived from basic beam cells with the best performance for dampening vibrations.
A clear example of how the lattice structure may be tailored in respect of the assigned design requirements is provided by auxetic lattice structures. They can realize structures with a negative Poisson’s ratio, which could be utilized in the aerospace sector for the design of morphing wings for next generation aircrafts [63,64,65]. Similarly, in biomedical applications, such as bones based on TPMS cells, they can tailor structures, making them suitable to match bone stiffness, surface roughness, and material compatibility for bone growth, as demonstrated in [66,67].
Another class of applications is related to their capability of achieving a high surface/volume ratio or proper thermal conduction characteristics that make the lattice structures suitable for thermal insulation, heat pipe wicks, exchangers, and sinks. In [68], applications related to thermal systems are reviewed and encouraged, thanks to AM advancements. They involve both metallic and non-metallic lattice structures. Heat transfer media made of lattice structure increases the efficiency of the process, as demonstrated in [69].
Concerning the correlation between applications and the characteristics of the lattice structures according to unit cells and replication patterns, Table 1 provides an overview of references in terms of the industrial sectors and related applications.

4. Design and Manufacturing Issues

In this section, the manufacturing issues, together with design approaches and tools, are highlighted in order to offer an overview of how to build the lattice structure components in respect of the application’s requirements.
From a design point of view, unit cell selection is strongly affected by the functional requirements that need to be accomplished; but it is also constrained by the manufacturing feasibility. The manufacturing feasibility may be related to the type of structure (according to the classifications of Section 2) or to process limits, such as the range of accuracy, lengths, necessity of post-processing, etc. According to this, Table 2 and Table 3 show the manufacturing methods versus the base materials and major limitations, together with references where such limitations are exploited. In detail, Table 2 refers to the traditional manufacturing processes and Table 3 refers to AM processes.
Generally speaking, traditional manufacturing methods have severe limitations concerning the type of cells that can be built. In particular, a limited set of cells may be obtained through joining the manufactured details achieved on relevant plates (e.g., honeycombs are manufactured by joining hexagonal cells between plates). In some cases, the replication patterns may be limited (e.g., direct foaming). It is clearly shown how AM processes may overcome the limit on the unit cell characteristics. Traditional manufacturing processes are mostly limited to base materials (mostly available in metals), few types of unit cells, mandatory post processing, high production times, and the unavailability of fine structures. However, AM solutions are limited to mandatory post processing (in the most cases), high production times, weak performance against fatigue loading, low aesthetic, build direction, residual stresses, support/sacrificial structures, etc. Particularly, the build direction/orientation during the AM should be meticulously defined to avoid volumetric expansion/geometrical errors and to minimize the residual stresses and the support volume/sacrificial supports [148,149].
According to the literature [150,151,152,153,154,155], researchers have mostly defined development workflows for the design and manufacturing of lattice structures separately. Most of these development workflows are based on very few basic requirements that pertain only to the lattice structure’s design and the manufacturing process, often without considering the overall optimization requirements of weight reduction, assembly efforts, reliability, post processing issues regarding mechanical performance (in general), support structure removal in the case of AM, and so on.
Pattern replications may be defined by topological optimization [156] to improve the component performances, also considering other manufacturing constraints, including:
For AM: working volume, support and infill design, post processing.
For traditional manufacturing processes: die design, extra and complex tooling/attachments, multiple machining/manufacturing processes, precise manufacturing process control, complex assembly/bonding process, post processing and/or set up by other design criteria such as aesthetics, assembly, etc.
In [156], AM lattices for biomedical applications related to implant design are studied and three design workflows are discussed to optimize this kind of design.
In any case, the micro/meso scale of the cell must be replicated at the macro scale to fulfill the component volumes.
From the CAD modeling point of view, it is computationally expensive, often requiring the adoption of dedicated tools and/or a change in the modeling approach. Table 4 shows a summary of current tools available to support the CAD-CAE development of known lattice structures. Only a few of them provide the opportunity for obtaining a diversity of lattice shapes and design freedom.
In cases of cells defined by shell elements, many CAD solutions (such as nTopology, Rhinoceros, Surface Design in Catia, …) adopted Implicit CAD Modelling (ICM) to vary the cell characteristics and the pattern distribution analytically. This enables the driving of results via optimization criteria since, through ICM, the geometry is directly described by a mathematical equation. Although ICM may produce formulas not physically realistic (e.g., non-manifold solutions), in cases of lattice structures, it can use the data points to influence the design with instant changes and, therefore, assists the design with any change in the geometry [104,175]. Through a field-driven design, ICM may simplify the pattern definitions for conformal, stochastic, and hybrid structures by utilizing a scalar field that represents the unit cell replication as a weight of the ICM set of data points. Changes of the field drive the geometry and alter the characteristics of the pattern. Commonly, it helps to optimize the lattice structure based on the stress field [104].
In [176], a hybrid geometric modeling method is proposed in which the selected unit cell is replicated using a voxelization approach for the complex volume being fulfilled, through geometrical functions which are suitable for describing the cell.
Generally, the computation loop includes one or more steps for filling the inner volume with cells and additional steps to constrain the boundary surface. It enables good transferability to the CAE analysis, closing the information loop between CAD modeling and FEA analysis which often limits the actions of the designer when she/he adopts commercial software. Moreover, it can provide both regular and stochastic replications in conformal configurations thanks to a Voronoi-based approach [177,178]. In [179], a NURBS-Free Form Deformation (FFD) approach is proposed and validated by FEA to build and optimize conformal truss-based lattice structures. Other methods to build conformal lattices are the Twin Curve Division Method (TCD), Arc Division Method (AD), Curvature Division Method (CD), and NURBS Free Form Deformation Method (NFFD).

5. Conclusions

This paper provides some insights into the design capabilities of lattice structures, with subsequent achievable properties in the context of engineering requirements and applications, with a wide range of materials from metal to plastic. For a few decades, researchers have extensively worked to unfold the unique characteristics of lattice structures, such as auxetic structures, negative thermal expansions, active stiffness, etc. These functional properties are derived from lattice cell characteristics; for this reason, the paper linked design and manufacturing issues, such as replication patterns, cell mechanical response, and manufacturability, to the most relevant engineering applications found in the research literature. Moreover, to provide the design practice point of view, CAD-CAE issues related to their modeling and optimization were also discussed.
This review also highlighted that the lattice structures benefit many engineering sectors in the context of engineering requirements, including lightweight structures, energy absorption structures, conduction/convection heat management, sound/vibration energy minimization, electromagnetic wave shielding, biomedical bone implants, etc. However, based on the above discussions, the following few recommendations are furnished for future research work:
  • In order to investigate the mechanical properties of a structure derived from the multiscale properties of the component(s) and from mesoscale cell characteristics, proper functional behavior may be tailored. As such, a CAD-CAE approach may massively help this study by virtual testing.
  • CAD-CAE approaches are also affected by the multiscale domains involved in the problem, with increasing computational costs. This provides impetus to adopt new modeling approaches, such as Implicit Geometric Modeling and/or homogenization techniques for CAE analyses.
  • In the case of AM, the mechanical characteristics related to stress–strain curves must be investigated and validated to distinguish between the lattice structure properties and AM setup conditions.
  • Metallic AM lattice structures suffer with fatigue loads due to the limited process conditions and/or setup. In addition, the post processing treatments may be infeasible due to the complexity of the structures. Therefore, a suitable workflow engulfing all of the design, manufacturing, and post processing requirements may be defined in order to minimize these concerns and facilitate the functional requirements of the lattice structures.

Author Contributions

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

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

List of Symbols and Acronyms

ϕ x , y , z Basis Function of TPMS k Principal Curvature at a General Point ‘P’
L Unit Cell Size x , y , z Periodicities
X , Y , Z Cell’s Repetition M Maxwell Number
s Number of Struts n Number of Joints in the Unit Cell
E * Elastic Modulus of the Cellular Structure E s Elastic Modulus of Equivalent Solid
ρ * Density of the Cellular Structure ρ s Density of the Equivalent Solid
C Gibson & Ashby Constant A M Additive Manufacturing
C A D Computer Aided Design C A E Computer Aided Engineering
S C Simple Cubic C B Body-Centered
B C C Body-Centered Cubic F C C Face-Centered Cubic
O T Octet Truss G A Gibson–Ashby
T P M S Triply Periodic Minimal Surface 3 D 3-Dimensional
I C M Implicit CAD Modeling C A E Computer-Aided Engineering
F E A Finite Element Analysis F D D Free Form Deformation
T C D Twin Curve Division Method A D Arc Division Method
C D Curvature Division Method N U R B S Non-Uniform Rational B-Splines
N F F D NURBS Free Form Deformation MethodB-SplinesBasis Splines

References

  1. Gibson, L.J. Modelling the mechanical behavior of cellular materials. Mater. Sci. Eng. A 1989, 110, 1–36. [Google Scholar] [CrossRef]
  2. Gibson, L.J.; Ashby, M.F. Cellular Solids: Structure and Properties, 2nd ed.; Cambridge University Press: Cambridge, UK, 1997; pp. 16–21. [Google Scholar] [CrossRef]
  3. Bhate, D. Four Questions in Cellular Material Design. Materials 2019, 12, 1060. [Google Scholar] [CrossRef]
  4. Tao, W.; Leu, M.C. Design of lattice structure for additive manufacturing. In Proceedings of the 2016 International Symposium on Flexible Automation Cleveland (ISFA 2016), Cleveland, OH, USA, 1–3 August 2016; pp. 326–332. [Google Scholar] [CrossRef]
  5. Mancini, E.; Campana, F.; Pilone, D.; Amodio, D.; Sasso, M. Definition of a unified material model for cellular materials with high morphological and topological dispersion: Application to an AA7075-T6 aluminium foam. Mater. Sci. Eng. A 2022, 833, 142346. [Google Scholar] [CrossRef]
  6. Pan, C.; Han, Y.; Lu, J. Design and optimization of lattice structures: A review. Appl. Sci. 2020, 10, 6374. [Google Scholar] [CrossRef]
  7. Suethao, S.; Shah, D.U.; Smitthipong, W. Recent progress in processing functionally graded polymer foams. Materials 2020, 13, 4060. [Google Scholar] [CrossRef] [PubMed]
  8. Bici, M.; Brischetto, S.; Campana, F.; Ferro, C.G.; Seclì, C.; Varetti, S.; Mazza, A. Development of a multifunctional panel for aerospace use through SLM additive manufacturing. Procedia CIRP 2018, 67, 215–220. [Google Scholar] [CrossRef]
  9. Ghouse, S.; Reznikov, N.; Boughton, O.R.; Babu, S.; Ng, K.G.; Blunn, G.; Jeffers, J.R. The design and in vivo testing of a locally stiffness-matched porous scaffold. Appl. Mater. Today 2019, 15, 377–388. [Google Scholar] [CrossRef] [PubMed]
  10. de Wild, M.; Schumacher, R.; Mayer, K.; Schkommodau, E.; Thoma, D.; Bredell, M.; Weber, F.E. Bone regeneration by the osteoconductivity of porous titanium implants manufactured by selective laser melting: A histological and micro computed tomography study in the rabbit. Tissue Eng. Part A 2013, 19, 2645–2654. [Google Scholar] [CrossRef] [PubMed]
  11. Saxena, K.K.; Das, R.; Calius, E.P. 3D printable multimaterial cellular auxetics with tunable stiffness. arXiv 2017, arXiv:1707.04486. [Google Scholar]
  12. Zhang, J.; Huang, H.; Liu, G.; Zong, H.; Zhang, C. Stiffness and energy absorption of additive manufactured hybrid lattice structures. Virtual Phys. Prototyp. 2021, 16, 428–443. [Google Scholar] [CrossRef]
  13. Yin, H.; Zhang, W.; Zhu, L.; Meng, F.; Liu, J.; Wen, G. Review on lattice structures for energy absorption properties. Compos. Struct. 2022, 304, 116397. [Google Scholar] [CrossRef]
  14. Shan, S.; Kang, S.H.; Raney, J.R.; Wang, P.; Fang, L.; Candido, F.; Bertoldi, K. Multistable architected materials for trapping elastic strain energy. Adv. Mater. 2015, 27, 4296–4301. [Google Scholar] [CrossRef]
  15. Brennan-Craddock, J.; Brackett, D.; Wildman, R.; Hague, R. The design of impact absorbing structures for additive manufacture. J. Phys. Conf. Ser. 2012, 382, 012042. [Google Scholar] [CrossRef]
  16. Ozdemir, Z.; Tyas, A.; Goodall, R.; Askes, H. Energy absorption in lattice structures in dynamics: Nonlinear FE simulations. Int. J. Impact Eng. 2017, 102, 1–15. [Google Scholar] [CrossRef]
  17. Habib, F.N.; Iovenitti, P.; Masood, S.H.; Nikzad, M. Fabrication of polymeric lattice structures for optimum energy absorption using Multi Jet Fusion technology. Mater. Des. 2018, 155, 86–98. [Google Scholar] [CrossRef]
  18. Maloney, K.J.; Fink, K.D.; Schaedler, T.A.; Kolodziejska, J.A.; Jacobsen, A.J.; Roper, C.S. Multifunctional heat exchangers derived from three-dimensional micro-lattice structures. Int. J. Heat Mass Transf. 2012, 55, 2486–2493. [Google Scholar] [CrossRef]
  19. Son, K.N.; Weibel, J.A.; Kumaresan, V.; Garimella, S.V. Design of multifunctional lattice-frame materials for compact heat exchangers. Int. J. Heat Mass Transf. 2017, 115, 619–629. [Google Scholar] [CrossRef]
  20. Chen, Y.; Li, T.; Scarpa, F.; Wang, L. Lattice metamaterials with mechanically tunable Poisson’s ratio for vibration control. Phys. Rev. Appl. 2017, 7, 024012. [Google Scholar] [CrossRef]
  21. Chen, D.; Zheng, X. Multi-material additive manufacturing of metamaterials with giant, tailorable negative Poisson’s ratios. Sci. Rep. 2018, 8, 9139. [Google Scholar] [CrossRef]
  22. Yuan, S.; Shen, F.; Bai, J.; Chua, C.K.; Wei, J.; Zhou, K. 3D soft auxetic lattice structures fabricated by selective laser sintering: TPU powder evaluation and process optimization. Mater. Des. 2017, 120, 317–327. [Google Scholar] [CrossRef]
  23. Savio, G.; Rosso, S.; Meneghello, R.; Concheri, G. Geometric modeling of cellular materials for additive manufacturing in biomedical field: A review. Appl. Bionics Biomech. 2018, 2018, 1654782. [Google Scholar] [CrossRef]
  24. Chan, Y.C.; Shintani, K.; Chen, W. Robust topology optimization of multi-material lattice structures under material and load uncertainties. Front. Mech. Eng. 2019, 14, 141–152. [Google Scholar] [CrossRef]
  25. Stanković, T.; Mueller, J.; Egan, P.; Shea, K. A generalized optimality criteria method for optimization of additively manufactured multimaterial lattice structures. J. Mech. Des. 2015, 137, 111405. [Google Scholar] [CrossRef]
  26. Ramirez-Chavez, I.E.; Anderson, D.; Sharma, R.; Lee, C.; Bhate, D. A classification of aperiodic architected cellular materials. Designs 2022, 6, 63. [Google Scholar] [CrossRef]
  27. Álvarez-Trejo, A.; Cuan-Urquizo, E.; Bhate, D.; Roman-Flores, A. Mechanical metamaterials with topologies based on curved elements: An overview of design, additive manufacturing and mechanical properties. Mater. Des. 2023, 233, 112190. [Google Scholar] [CrossRef]
  28. Zhang, Q.; Yang, X.; Li, P.; Huang, G.; Feng, S.; Shen, C.; Lu, T.J. Bioinspired engineering of honeycomb structure-Using nature to inspire human innovation. Prog. Mater. Sci. 2015, 74, 332–400. [Google Scholar] [CrossRef]
  29. Yan, C.; Hao, L.; Hussein, A.; Bubb, S.L.; Young, P.; Raymont, D. Evaluation of light-weight AlSi10Mg periodic cellular lattice structures fabricated via direct metal laser sintering. J. Mater. Process. Technol. 2014, 214, 856–864. [Google Scholar] [CrossRef]
  30. Hunt, C.J.; Morabito, F.; Grace, C.; Zhao, Y.; Woods, B.K. A review of composite lattice structures. Compos. Struct. 2022, 284, 115120. [Google Scholar] [CrossRef]
  31. Luxner, M.H.; Stampfl, J.; Pettermann, H.E. Finite element modeling concepts and linear analyses of 3D regular open cell structures. J. Mater. Sci. 2005, 40, 5859–5866. [Google Scholar] [CrossRef]
  32. Deshpande, V.S.; Fleck, N.A.; Ashby, M.F. Effective properties of the octet-truss lattice material. J. Mech. Phys. Solids 2001, 49, 1747–1769. [Google Scholar] [CrossRef]
  33. Lakes, R. Foam structures with a negative Poisson’s ratio. Science 1987, 235, 1038–1040. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, F. Systematic design of 3D auxetic lattice materials with programmable Poisson’s ratio for finite strains. J. Mech. Phys. Solids 2018, 114, 303–318. [Google Scholar] [CrossRef]
  35. Kumar, S.; Vyavahare, S.; Teraiya, S.; Kootikuppala, J.; Bogala, H. A state of the art review of additively manufactured auxetic structures. In Recent Advances in Manufacturing Processes and Systems: Select Proceedings of RAM 2021; Springer: Singapore, 2022; pp. 69–84. [Google Scholar] [CrossRef]
  36. Xue, Y.; Gao, P.; Zhou, L.; Han, F. An enhanced three-dimensional auxetic lattice structure with improved property. Materials 2020, 13, 1008. [Google Scholar] [CrossRef]
  37. Berger, J.B.; Wadley, H.N.G.; McMeeking, R.M. Mechanical metamaterials at the theoretical limit of isotropic elastic stiffness. Nature 2017, 543, 533–537. [Google Scholar] [CrossRef]
  38. Xue, R.; Cui, X.; Zhang, P.; Liu, K.; Li, Y.; Wu, W.; Liao, H. Mechanical design and energy absorption performances of novel dual scale hybrid plate-lattice mechanical metamaterials. Extrem. Mech. Lett. 2020, 40, 100918. [Google Scholar] [CrossRef]
  39. Al-Ketan, O.; Abu Al-Rub, R.K. Multifunctional mechanical metamaterials based on triply periodic minimal surface lattices. Adv. Eng. Mater. 2019, 21, 1900524. [Google Scholar] [CrossRef]
  40. AlMahri, S.; Santiago, R.; Lee, D.W.; Ramos, H.; Alabdouli, H.; Alteneiji, M.; Alves, M. Evaluation of the dynamic response of triply periodic minimal surfaces subjected to high strain-rate compression. Addit. Manuf. 2021, 46, 102220. [Google Scholar] [CrossRef]
  41. Song, J.; Tang, Q.; Feng, Q.; Ma, S.; Guo, F.; Han, Q. Investigation on the modelling approach for variable-density lattice structures fabricated using selective laser melting. Mater. Des. 2021, 212, 110236. [Google Scholar] [CrossRef]
  42. Al Khalil, M.; Lebaal, N.; Demoly, F.; Roth, S. A design and optimization framework of variable-density lattice structures for additive manufacturing. Mech. Adv. Mater. Struct. 2022, 29, 4711–4725. [Google Scholar] [CrossRef]
  43. Sienkiewicz, J.; Płatek, P.; Jiang, F.; Sun, X.; Rusinek, A. Investigations on the mechanical response of gradient lattice structures manufactured via SLM. Metals 2020, 10, 213. [Google Scholar] [CrossRef]
  44. Seharing, A.; Azman, A.H.; Abdullah, S. A review on integration of lightweight gradient lattice structures in additive manufacturing parts. Adv. Mech. Eng. 2020, 12, 1687814020916951. [Google Scholar] [CrossRef]
  45. Li, D.; Liao, W.; Dai, N.; Xie, Y.M. Anisotropic design and optimization of conformal gradient lattice structures. Comput.-Aided Des. 2020, 119, 102787. [Google Scholar] [CrossRef]
  46. Seharing, A.; Azman, A.H.; Abdullah, S. Finite element analysis of gradient lattice structure patterns for bone implant design. Int. J. Struct. Integr. 2020, 11, 535–545. [Google Scholar] [CrossRef]
  47. Fabbro, P.D.; Rosso, S.; Ceruti, A.; Meneghello, R.; Concheri, G.; Savio, G. Conformal Lattice Structures: Modeling and Optimization. In Design Tools and Methods in Industrial Engineering II, Proceedings of the International Conference on Design, Simulation, Manufacturing: The Innovation Exchange, ADM 2021, Rome, Italy, 9–10 September 2021; Springer International Publishing: Cham, Switzerland, 2021; pp. 474–485. [Google Scholar] [CrossRef]
  48. Nguyen, J.; Park, S.I.; Rosen, D.W.; Folgar, L.; Williams, J. Conformal lattice structure design and fabrication. In Proceedings of the 2012 International Solid Freeform Fabrication Symposium, Austin, TX, USA, 6–8 August 2012; University of Texas at Austin: Austin, TX, USA, 2012. [Google Scholar]
  49. D’Angelo, C.; Ortona, A.; Colombo, P. Influence of the loading direction on the mechanical behavior of ceramic foams and lattices under compression. Acta Mater. 2013, 61, 5525–5534. [Google Scholar] [CrossRef]
  50. Singamaneni, S.; Bertoldi, K.; Chang, S.; Jang, J.H.; Young, S.L.; Thomas, E.L.; Tsukruk, V.V. Bifurcated mechanical behavior of deformed periodic porous solids. Adv. Funct. Mater. 2009, 19, 1426–1436. [Google Scholar] [CrossRef]
  51. Ashby, M.F.; Gibson, L.J. Cellular Solids: Structure and Properties; Press Syndicate of the University of Cambridge: Cambridge, UK, 1997; pp. 175–231. [Google Scholar]
  52. Tumino, D.; Alaimo, A.; Orlando, C.; Valvano, S. A Preliminary Study on the Effect of Strut Waviness on the Mechanical Properties of BCC Lattice Unit Cells. In Design Tools and Methods in Industrial Engineering II, Proceedings of the International Conference on Design, Simulation, Manufacturing: The Innovation Exchange, ADM 2021, Rome, Italy, 9–10 September 2021; Springer: Cham, Switzerland, 2021; pp. 431–441. [Google Scholar] [CrossRef]
  53. Graziosi, S.; Ballo, F.M.; Libonati, F.; Senna, S. 3D printing of bending-dominated soft lattices: Numerical and experimental assessment. Rapid Prototyp. J. 2022, 28, 51–64. [Google Scholar]
  54. Wagner, M.A.; Lumpe, T.S.; Chen, T.; Shea, K. Programmable, active lattice structures: Unifying stretch-dominated and bending-dominated topologies. Extrem. Mech. Lett. 2019, 29, 100461. [Google Scholar] [CrossRef]
  55. Benedetti, M.; Du Plessis, A.; Ritchie, R.O.; Dallago, M.; Razavi, S.M.J.; Berto, F. Architected cellular materials: A review on their mechanical properties towards fatigue-tolerant design and fabrication. Mater. Sci. Eng. R Rep. 2021, 144, 100606. [Google Scholar] [CrossRef]
  56. Wei, K.; Yang, Q.; Ling, B.; Xie, H.; Qu, Z.; Fang, D. Mechanical responses of titanium 3D kagome lattice structure manufactured by selective laser melting. Extrem. Mech. Lett. 2018, 23, 41–48. [Google Scholar] [CrossRef]
  57. Elipe, J.C.Á.; Lantada, A.D. Comparative study of auxetic geometries by means of computer-aided design and engineering. Smart Mater. Struct. 2012, 21, 105004. [Google Scholar] [CrossRef]
  58. du Plessis, A.; Razavi, N.; Benedetti, M.; Murchio, S.; Leary, M.; Watson, M.; Berto, F. Properties and applications of additively manufactured metallic cellular materials: A review. Prog. Mater. Sci. 2022, 125, 100918. [Google Scholar] [CrossRef]
  59. Wu, G.; Cho, Y.; Choi, I.S.; Ge, D.; Li, J.; Han, H.N.; Yang, S. Directing the deformation paths of soft metamaterials with prescribed asymmetric units. Adv. Mater. 2015, 27, 2747–2752. [Google Scholar] [CrossRef] [PubMed]
  60. Niknam, H.; Akbarzadeh, A.H. Graded lattice structures: Simultaneous enhancement in stiffness and energy absorption. Mater. Des. 2020, 196, 109129. [Google Scholar] [CrossRef]
  61. Ji, J.C.; Luo, Q.; Ye, K. Vibration control based metamaterials and origami structures: A state-of-the-art review. Mech. Syst. Signal Process. 2021, 161, 107945. [Google Scholar] [CrossRef]
  62. Wei, Y.; Yu, B.; Yang, Q.; Gao, P.; Miao, Z.; Cheng, J.; Sun, X. Damping behaviors of steel-based Kelvin lattice structures fabricated by indirect additive manufacture combining investment casting. Smart Mater. Struct. 2020, 29, 055001. [Google Scholar] [CrossRef]
  63. Spadoni, A.; Ruzzene, M. Static aeroelastic response of chiral-core airfoils. J. Intell. Mater. Syst. Struct. 2007, 18, 1067–1075. [Google Scholar] [CrossRef]
  64. Alderson, A.; Alderson, K.L. Auxetic materials. Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng. 2007, 221, 565–575. [Google Scholar] [CrossRef]
  65. Xia, Y.; Friswell, M.I.; Flores, E.S. Equivalent models of corrugated panels. Int. J. Solids Struct. 2012, 49, 1453–1462. [Google Scholar] [CrossRef]
  66. El-Sayed, M.A.; Essa, K.; Ghazy, M.; Hassanin, H. Design optimization of additively manufactured titanium lattice structures for biomedical implants. Int. J. Adv. Manuf. Technol. 2020, 110, 2257–2268. [Google Scholar] [CrossRef]
  67. Murr, L.E.; Gaytan, S.M.; Medina, F.; Lopez, H.; Martinez, E.; Machado, B.I.; Bracke, J. Next-generation biomedical implants using additive manufacturing of complex, cellular and functional mesh arrays. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2010, 368, 1999–2032. [Google Scholar] [CrossRef] [PubMed]
  68. Sajjad, U.; Rehman, T.U.; Ali, M.; Park, C.W.; Yan, W.M. Manufacturing and potential applications of lattice structures in thermal systems: A comprehensive review of recent advances. Int. J. Heat Mass Transf. 2022, 198, 123352. [Google Scholar] [CrossRef]
  69. Shabgard, H.; Allen, M.J.; Sharifi, N.; Benn, S.P.; Faghri, A.; Bergman, T.L. Heat pipe heat exchangers and heat sinks: Opportunities, challenges, applications, analysis, and state of the art. Int. J. Heat Mass Transf. 2015, 89, 138–158. [Google Scholar] [CrossRef]
  70. Moon, S.K.; Tan, Y.E.; Hwang, J.; Yoon, Y.J. Application of 3D printing technology for designing light-weight unmanned aerial vehicle wing structures. Int. J. Precis. Eng. Manuf.-Green Technol. 2014, 1, 223–228. [Google Scholar] [CrossRef]
  71. Magerramova, L.; Volkov, M.; Afonin, A.; Svinareva, M.; Kalinin, D. Application of light lattice structures for gas turbine engine fan blades. In Proceedings of the 31st Congress of the International Council of the Aeronautical Sciences, ICAS, Belo Horizonte, Brazil, 9–14 September 2018. [Google Scholar]
  72. Spadoni, A.; Ruzzene, M. Numerical and experimental analysis of the static compliance of chiral truss-core airfoils. J. Mech. Mater. Struct. 2007, 2, 965–981. [Google Scholar] [CrossRef]
  73. Aabid, A.; Parveez, B.; Parveen, N.; Khan, S.A.; Zayan, J.M.; Shabbir, O. Reviews on design and development of unmanned aerial vehicle (drone) for different applications. J. Mech. Eng. Res. Dev. 2022, 45, 53–69. [Google Scholar]
  74. Zhang, X.; Zhou, H.; Shi, W.; Zeng, F.; Zeng, H.; Chen, G. Vibration tests of 3D printed satellite structure made of lattice sandwich panels. AIAA J. 2018, 56, 4213–4217. [Google Scholar] [CrossRef]
  75. Vasiliev, V.V.; Barynin, V.A.; Razin, A.F. Anisogrid composite lattice structures-Development and aerospace applications. Compos. Struct. 2012, 94, 1117–1127. [Google Scholar] [CrossRef]
  76. Khorrami, M.R.; Humphreys, W.M.; Lockard, D.P. An assessment of flap and main landing gear noise abatement concepts. In Proceedings of the 21st AIAA/CEAS Aeroacoustics Conference, Dallas, TX, USA, 22–26 June 2015; p. 2987. [Google Scholar] [CrossRef]
  77. Jetté, B.; Brailovski, V.; Dumas, M.; Simoneau, C.; Terriault, P. Femoral stem incorporating a diamond cubic lattice structure: Design, manufacture and testing. J. Mech. Behav. Biomed. Mater. 2018, 77, 58–72. [Google Scholar] [CrossRef] [PubMed]
  78. Heinl, P.; Müller, L.; Körner, C.; Singer, R.F.; Müller, F.A. Cellular Ti-6Al-4V structures with interconnected macro porosity for bone implants fabricated by selective electron beam melting. Acta Biomater. 2008, 4, 1536–1544. [Google Scholar] [CrossRef]
  79. Challapalli, A.; Ju, J. Continuum model for effective properties of orthotropic octet-truss lattice materials. In Proceedings of the ASME International Mechanical Engineering Congress and Exposition, Montreal, QC, Canada, 14–20 November 2014; American Society of Mechanical Engineers: New York, NY, USA, 2015; Volume 46583, p. V009T12A051. [Google Scholar] [CrossRef]
  80. Dumas, M.; Terriault, P.; Brailovski, V. Modelling and characterization of a porosity graded lattice structure for additively manufactured biomaterials. Mater. Des. 2017, 121, 383–392. [Google Scholar] [CrossRef]
  81. Stevenson, G.; Rehman, S.; Draper, E.; Hernández-Nava, E.; Hunt, J.; Haycock, J.W. Combining 3D human in vitro methods for a 3Rs evaluation of novel titanium surfaces in orthopaedic applications. Biotechnol. Bioeng. 2016, 113, 1586–1599. [Google Scholar] [CrossRef] [PubMed]
  82. Chen, W.M.; Lee, S.J.; Lee, P.V.S. Failure analysis of an additive manufactured porous titanium structure for orthopedic implant applications. Mater. Sci. Forum 2016, 863, 45–49. [Google Scholar] [CrossRef]
  83. Gorguluarslan, R.M.; Choi, S.K.; Saldana, C.J. Uncertainty quantification and validation of 3D lattice scaffolds for computer-aided biomedical applications. J. Mech. Behav. Biomed. Mater. 2017, 71, 428–440. [Google Scholar] [CrossRef] [PubMed]
  84. Soro, N.; Attar, H.; Brodie, E.; Veidt, M.; Molotnikov, A.; Dargusch, M.S. Evaluation of the mechanical compatibility of additively manufactured porous Ti-25Ta alloy for load-bearing implant applications. J. Mech. Behav. Biomed. Mater. 2019, 97, 149–158. [Google Scholar] [CrossRef] [PubMed]
  85. Rahmani, R.; Antonov, M.; Kollo, L.; Holovenko, Y.; Prashanth, K.G. Mechanical behavior of Ti6Al4V scaffolds filled with CaSiO3 for implant applications. Appl. Sci. 2019, 9, 3844. [Google Scholar] [CrossRef]
  86. Jia, D.; Li, F.; Zhang, C.; Liu, K.; Zhang, Y. Design and simulation analysis of Lattice bone plate based on finite element method. Mech. Adv. Mater. Struct. 2021, 28, 1311–1321. [Google Scholar] [CrossRef]
  87. Wang, C.; Li, Y.; Zhao, W.; Zou, S.; Zhou, G.; Wang, Y. Structure design and multi-objective optimization of a novel crash box based on biomimetic structure. Int. J. Mech. Sci. 2018, 138, 489–501. [Google Scholar] [CrossRef]
  88. Wang, C.; Wang, W.; Zhao, W.; Wang, Y.; Zhou, G. Structure design and multi-objective optimization of a novel NPR bumper system. Compos. Part B Eng. 2018, 153, 78–96. [Google Scholar] [CrossRef]
  89. Yin, S.; Chen, H.; Wu, Y.; Li, Y.; Xu, J. Introducing composite lattice core sandwich structure as an alternative proposal for engine hood. Compos. Struct. 2018, 201, 131–140. [Google Scholar] [CrossRef]
  90. Niutta, C.B.; Ciardiello, R.; Tridello, A. Experimental and Numerical Investigation of a Lattice Structure for Energy Absorption: Application to the Design of an Automotive Crash Absorber. Polymers 2022, 14, 1116. [Google Scholar] [CrossRef]
  91. Mantovani, S.; Campo, G.A.; Ferrari, A.; Cavazzuti, M. Optimization methodology for automotive chassis design by truss frame: A preliminary investigation using the lattice approach. In Transdisciplinary Engineering Methods for Social Innovation of Industry 4.0; IOS Press: Amsterdam, The Netherlands, 2018; pp. 984–992. [Google Scholar]
  92. Kumar, V.; Manogharan, G.; Cormier, D.R. Design of periodic cellular structures for heat exchanger applications. In 2009 International Solid Freeform Fabrication Symposium; University of Texas at Austin: Austin, TX, USA, 2009. [Google Scholar]
  93. Pannikottu, A.; Bandaru, S. Flexible Honeycomb Composite Vehicle Armor; SAE Technical Paper No. 2009-01-0601; SAE International: Warrendale, PA, USA, 2009. [Google Scholar] [CrossRef]
  94. Harris, J.A. Additively Manufactured Metallic Cellular Materials for Blast and Impact Mitigation. Ph.D. Thesis, University of Cambridge, Cambridge, UK, 2018. [Google Scholar]
  95. Acanfora, V.; Corvino, C.; Saputo, S.; Sellitto, A.; Riccio, A. Application of an additive manufactured hybrid metal/composite shock absorber panel to a military seat ejection system. Appl. Sci. 2021, 11, 6473. [Google Scholar] [CrossRef]
  96. Luo, J.W.; Chen, L.; Xia, Y.; Zheng, X.; Tao, W.Q. Topology optimization of natural convection using porous metal foam based on the adjoint lattice Boltzmann method and level set method. Comput. Fluids 2023, 265, 106007. [Google Scholar] [CrossRef]
  97. Baobaid, N.; Ali, M.I.; Khan, K.A.; Al-Rub, R.K.A. Fluid flow and heat transfer of porous TPMS architected heat sinks in free convection environment. Case Stud. Therm. Eng. 2022, 33, 101944. [Google Scholar] [CrossRef]
  98. Hussein, A.; Hao, L.; Yan, C.; Everson, R.; Young, P. Advanced lattice support structures for metal additive manufacturing. J. Mater. Process. Technol. 2013, 213, 1019–1026. [Google Scholar] [CrossRef]
  99. Vaissier, B.; Pernot, J.P.; Chougrani, L.; Véron, P. Genetic-algorithm based framework for lattice support structure optimization in additive manufacturing. Comput.-Aided Des. 2019, 110, 11–23. [Google Scholar] [CrossRef]
  100. Cheng, L.; Liang, X.; Bai, J.; Chen, Q.; Lemon, J.; To, A. On utilizing topology optimization to design support structure to prevent residual stress induced build failure in laser powder bed metal additive manufacturing. Addit. Manuf. 2019, 27, 290–304. [Google Scholar] [CrossRef]
  101. Alejandrino, J.D.; Concepcion, R.S., II; Lauguico, S.C.; Tobias, R.R.; Venancio, L.; Macasaet, D.; Dadios, E.P. A machine learning approach of lattice infill pattern for increasing material efficiency in additive manufacturing processes. Int. J. Mech. Eng. Robot. Res. 2020, 9, 1253–1263. [Google Scholar] [CrossRef]
  102. Ntintakis, I.; Stavroulakis, G.E. Infill Microstructures for Additive Manufacturing. Appl. Sci. 2022, 12, 7386. [Google Scholar] [CrossRef]
  103. Dong, G.; Tang, Y.; Li, D.; Zhao, Y.F. Design and optimization of solid lattice hybrid structures fabricated by additive manufacturing. Addit. Manuf. 2020, 33, 101116. [Google Scholar] [CrossRef]
  104. Ahmad, A.; Elamana, S.; Adam, K.; Bici, M.; Campana, F. Lightweight Horse Saddletree Through Reverse Engineering and Lattice Structure Design. Comput. Des. Appl. 2023, 20, 923–935. [Google Scholar] [CrossRef]
  105. Hanna, B.; Adams, R.; Townsend, S.; Robinson, M.; Soe, S.; Stewart, M.; Theobald, P. Auxetic metamaterial optimisation for head impact mitigation in American football. Int. J. Impact Eng. 2021, 157, 103991. [Google Scholar] [CrossRef]
  106. Ren, X.; Das, R.; Tran, P.; Ngo, T.D.; Xie, Y.M. Auxetic metamaterials and structures: A review. Smart Mater. Struct. 2018, 27, 023001. [Google Scholar] [CrossRef]
  107. ARUP. Canton Tower. Available online: https://www.arup.com/projects/guangzhou-tv-tower (accessed on 27 June 2023).
  108. ARUP. Beijing National Stadium (Bird’s Nest). Available online: https://www.arup.com/services/buildings/structural-engineering (accessed on 27 June 2023).
  109. Atomium. The Atomium (The Symbol of Brussels and Belgium). Available online: https://atomium.be/home/Index (accessed on 27 June 2023).
  110. Huang, R.; Ma, S.; Zhang, M.; Xu, J.; Wang, Z. Dynamic deformation and failure process of quasi-closed-cell aluminum foam manufactured by direct foaming technique. Mater. Sci. Eng. A 2019, 756, 302–311. [Google Scholar] [CrossRef]
  111. Luukkonen, T.; Yliniemi, J.; Sreenivasan, H.; Ohenoja, K.; Finnilä, M.; Franchin, G.; Colombo, P. Ag-or Cu-modified geopolymer filters for water treatment manufactured by 3D printing, direct foaming, or granulation. Sci. Rep. 2020, 10, 7233. [Google Scholar] [CrossRef] [PubMed]
  112. Singh, S.; Singh, R. Precision investment casting: A state of art review and future trends. Proc. Inst. Mech. Eng. Part B J. Eng. Manuf. 2016, 230, 2143–2164. [Google Scholar] [CrossRef]
  113. Barnett, S.O. Investment Casting—The Multi-Process Technology. Foundry Trade J. Int. 1988, 11, 33. [Google Scholar]
  114. Jones, S.; Yuan, C. Advances in shell moulding for investment casting. J. Mater. Process. Technol. 2003, 135, 258–265. [Google Scholar] [CrossRef]
  115. Wadley, H.N. Multifunctional periodic cellular metals. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2006, 364, 31–68. [Google Scholar] [CrossRef] [PubMed]
  116. Ablat, M.A.; Qattawi, A. Investigating the design and process parameters of folded perforated sheet metal. Int. J. Adv. Manuf. Technol. 2019, 102, 615–633. [Google Scholar] [CrossRef]
  117. Cheung, K.C.; Gershenfeld, N. Reversibly assembled cellular composite materials. Science 2013, 341, 1219–1221. [Google Scholar] [CrossRef] [PubMed]
  118. Jiang, S.; Sun, F.; Zhang, X.; Fan, H. Interlocking orthogrid: An efficient way to construct lightweight lattice-core sandwich composite structure. Compos. Struct. 2017, 176, 55–71. [Google Scholar] [CrossRef]
  119. Cheung, K.C.W. Digital Cellular Solids: Reconfigurable Composite Materials. Ph.D. Thesis, Massachusetts Institute of Technology, Cambridge, MA, USA, 2012. [Google Scholar]
  120. Queheillalt, D.T.; Murty, Y.; Wadley, H.N. Mechanical properties of an extruded pyramidal lattice truss sandwich structure. Scr. Mater. 2008, 58, 76–79. [Google Scholar] [CrossRef]
  121. Kang, K.J. Wire-woven cellular metals: The present and future. Prog. Mater. Sci. 2015, 69, 213–307. [Google Scholar] [CrossRef]
  122. Simovski, C.R.; Belov, P.A.; Atrashchenko, A.V.; Kivshar, Y.S. Wire metamaterials: Physics and applications. Adv. Mater. 2012, 24, 4229–4248. [Google Scholar] [CrossRef]
  123. Khoda, B.; Ahsan, A.N. A novel rapid manufacturing process for metal lattice structure. 3d Print. Addit. Manuf. 2021, 8, 111–125. [Google Scholar] [CrossRef]
  124. Li, M.; Du, W.; Elwany, A.; Pei, Z.; Ma, C. Metal binder jetting additive manufacturing: A literature review. J. Manuf. Sci. Eng. 2020, 142, 090801. [Google Scholar] [CrossRef]
  125. Du, W.; Ren, X.; Pei, Z.; Ma, C. Ceramic binder jetting additive manufacturing: A literature review on density. J. Manuf. Sci. Eng. 2020, 142, 040801. [Google Scholar] [CrossRef]
  126. Mirzababaei, S.; Pasebani, S. A review on binder jet additive manufacturing of 316L stainless steel. J. Manuf. Mater. Process. 2019, 3, 82. [Google Scholar] [CrossRef]
  127. Prashar, G.; Vasudev, H. A comprehensive review on sustainable cold spray additive manufacturing: State of the art, challenges and future challenges. J. Clean. Prod. 2021, 310, 127606. [Google Scholar] [CrossRef]
  128. Yin, S.; Cavaliere, P.; Aldwell, B.; Jenkins, R.; Liao, H.; Li, W.; Lupoi, R. Cold spray additive manufacturing and repair: Fundamentals and applications. Addit. Manuf. 2018, 21, 628–650. [Google Scholar] [CrossRef]
  129. Pathak, S.; Saha, G.C. Development of sustainable cold spray coatings and 3D additive manufacturing components for repair/manufacturing applications: A critical review. Coatings 2017, 7, 122. [Google Scholar] [CrossRef]
  130. Dass, A.; Moridi, A. State of the art in directed energy deposition: From additive manufacturing to materials design. Coatings 2019, 9, 418. [Google Scholar] [CrossRef]
  131. Yao, X.X.; Ge, P.; Li, J.Y.; Wang, Y.F.; Li, T.; Liu, W.W.; Zhang, Z. Controlling the solidification process parameters of direct energy deposition additive manufacturing considering laser and powder properties. Comput. Mater. Sci. 2020, 182, 109788. [Google Scholar] [CrossRef]
  132. Heigel, J.C.; Michaleris, P.; Reutzel, E.W. Thermo-mechanical model development and validation of directed energy deposition additive manufacturing of Ti-6Al-4V. Addit. Manuf. 2015, 5, 9–19. [Google Scholar] [CrossRef]
  133. Behera, D.; Cullinan, M. Current challenges and potential directions towards precision microscale additive manufacturing-Part I: Direct ink writing/jetting processes. Precis. Eng. 2021, 68, 326–337. [Google Scholar] [CrossRef]
  134. Costakis, W.J., Jr.; Rueschhoff, L.M.; Diaz-Cano, A.I.; Youngblood, J.P.; Trice, R.W. Additive manufacturing of boron carbide via continuous filament direct ink writing of aqueous ceramic suspensions. J. Eur. Ceram. Soc. 2016, 36, 3249–3256. [Google Scholar] [CrossRef]
  135. Tu, R.; Sodano, H.A. Additive manufacturing of high-performance vinyl ester resin via direct ink writing with UV-thermal dual curing. Addit. Manuf. 2021, 46, 102180. [Google Scholar] [CrossRef]
  136. Ning, F.; Cong, W.; Hu, Y.; Wang, H. Additive manufacturing of carbon fiber-reinforced plastic composites using fused deposition modeling: Effects of process parameters on tensile properties. J. Compos. Mater. 2017, 51, 451–462. [Google Scholar] [CrossRef]
  137. Singh, S.; Ramakrishna, S.; Singh, R. Material issues in additive manufacturing: A review. J. Manuf. Process. 2017, 25, 185–200. [Google Scholar] [CrossRef]
  138. Chacón, J.M.; Caminero, M.A.; García-Plaza, E.; Núnez, P.J. Additive manufacturing of PLA structures using fused deposition modelling: Effect of process parameters on mechanical properties and their optimal selection. Mater. Des. 2017, 124, 143–157. [Google Scholar] [CrossRef]
  139. Sukhotskiy, V.; Vishnoi, P.; Karampelas, I.H.; Vader, S.; Vader, Z.; Furlani, E.P. Magnetohydrodynamic drop-on-demand liquid metal additive manufacturing: System overview and modelling. In Proceedings of the 5th International Conference of Fluid Flow, Heat and Mass Transfer (FFHMT’18), Niagra Falls, ON, Canada, 7–9 June 2018. [Google Scholar] [CrossRef]
  140. Gan, Z.; Yu, G.; He, X.; Li, S. Surface-active element transport and its effect on liquid metal flow in laser-assisted additive manufacturing. Int. Commun. Heat Mass Transf. 2017, 86, 206–214. [Google Scholar] [CrossRef]
  141. Li, A.; Liu, X.; Wan, X.; Yang, Y. Thermal behaviors and fluid flow controlling the geometry of 7075 aluminum alloy single tracks during liquid metal flow rapid cooling additive manufacturing. Int. Commun. Heat Mass Transf. 2020, 116, 104664. [Google Scholar] [CrossRef]
  142. Dermeik, B.; Travitzky, N. Laminated object manufacturing of ceramic-based materials. Adv. Eng. Mater. 2020, 22, 2000256. [Google Scholar] [CrossRef]
  143. Tao, Y.; Yin, Q.; Li, P. An additive manufacturing method using large-scale wood inspired by laminated object manufacturing and plywood technology. Polymers 2020, 13, 144. [Google Scholar] [CrossRef] [PubMed]
  144. Luong, D.X.; Subramanian, A.K.; Silva, G.A.L.; Yoon, J.; Cofer, S.; Yang, K.; Tour, J.M. Laminated object manufacturing of 3D-printed laser-induced graphene foams. Adv. Mater. 2018, 30, 1707416. [Google Scholar] [CrossRef] [PubMed]
  145. King, W.E.; Anderson, A.T.; Ferencz, R.M.; Hodge, N.E.; Kamath, C.; Khairallah, S.A.; Rubenchik, A.M. Laser powder bed fusion additive manufacturing of metals; physics, computational, and materials challenges. Appl. Phys. Rev. 2015, 2, 041304. [Google Scholar] [CrossRef]
  146. Snow, Z.; Nassar, A.R.; Reutzel, E.W. Invited Review Article: Review of the formation and impact of flaws in powder bed fusion additive manufacturing. Addit. Manuf. 2020, 36, 101457. [Google Scholar] [CrossRef]
  147. Ladani, L.; Sadeghilaridjani, M. Review of powder bed fusion additive manufacturing for metals. Metals 2021, 11, 1391. [Google Scholar] [CrossRef]
  148. Cheng, L.; To, A. Part-scale build orientation optimization for minimizing residual stress and support volume for metal additive manufacturing: Theory and experimental validation. Comput.-Aided Des. 2019, 113, 1–23. [Google Scholar] [CrossRef]
  149. Zanini, F.; Sorgato, M.; Savio, E.; Carmignato, S. Uncertainty of CT dimensional measurements performed on metal additively manufactured lattice structures. In Proceedings of the 10th Conference on Industrial Computed Tomography, Wels, Austria, 4–7 February 2020. [Google Scholar] [CrossRef]
  150. Tamburrino, F.; Graziosi, S.; Bordegoni, M. The design process of additively manufactured mesoscale lattice structures: A review. J. Comput. Inf. Sci. Eng. 2018, 18, 040801. [Google Scholar] [CrossRef]
  151. Koch, P.; Korn, H.; Kordass, R.; Holtzhausen, S.; Schoene, C.; Mueller, B.; Stelzer, R. A CAD-based workflow and mechanical characterization for additive manufacturing of tailored lattice structures. In 2018 International Solid Freeform Fabrication Symposium; University of Texas at Austin: Austin, TX, USA, 2018. [Google Scholar]
  152. Daynes, S.; Feih, S.; Lu, W.F.; Wei, J. Design concepts for generating optimised lattice structures aligned with strain trajectories. Comput. Methods Appl. Mech. Eng. 2019, 354, 689–705. [Google Scholar] [CrossRef]
  153. Rosso, S.; Uriati, F.; Grigolato, L.; Meneghello, R.; Concheri, G.; Savio, G. An optimization workflow in design for additive manufacturing. Appl. Sci. 2021, 11, 2572. [Google Scholar] [CrossRef]
  154. Panagiotidou, V.; Koerner, A.; Cruz, M.; Parker, B.; Beyer, B.; Giannakopoulos, S. 3D extrusion of multi-biomaterial lattices using an environmentally informed workflow. Front. Archit. Res. 2022, 11, 691–708. [Google Scholar] [CrossRef]
  155. Hudak, Y.F.; Li, J.S.; Cullum, S.; Strzelecki, B.M.; Richburg, C.; Kaufman, G.E.; Aubin, P.M. A novel workflow to fabricate a patient-specific 3D printed accommodative foot orthosis with personalized latticed metamaterial. Med. Eng. Phys. 2022, 104, 103802. [Google Scholar] [CrossRef] [PubMed]
  156. Al-Tamimi, A.A.; Almeida, H.; Bartolo, P. Structural optimisation for medical implants through additive manufacturing. Prog. Addit. Manuf. 2020, 5, 95–110. [Google Scholar] [CrossRef]
  157. Nguyen, D.S.; Vignat, F. A method to generate lattice structure for additive manufacturing. In Proceedings of the 2016 IEEE International Conference on Industrial Engineering and Engineering Management (IEEM), Bali, Indonesia, 4–7 December 2016; IEEE: Piscataway, NJ, USA, 2016; pp. 966–970. [Google Scholar] [CrossRef]
  158. Jafferson, J.M.; Sharma, H. Design of 3D printable airless tyres using NTopology. Mater. Today Proc. 2021, 46, 1147–1160. [Google Scholar] [CrossRef]
  159. Nguyen, D.S. Design of lattice structure for additive manufacturing in CAD environment. J. Adv. Mech. Des. Syst. Manuf. 2019, 13, JAMDSM0057. [Google Scholar] [CrossRef]
  160. Moreno Nieto, D.; Moreno Sánchez, D. Design for additive manufacturing: Tool review and a case study. Appl. Sci. 2021, 11, 1571. [Google Scholar] [CrossRef]
  161. Dragoni, E.; Ciace, V.A. Mechanical design and modelling of lightweight additively manufactured lattice structures evolved from regular three-dimensional tessellations. Proc. Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 2021, 235, 1759–1773. [Google Scholar] [CrossRef]
  162. Abdulhadi, H.S.; Mian, A. Effect of strut length and orientation on elastic mechanical response of modified body-centered cubic lattice structures. Proc. Inst. Mech. Eng. Part L J. Mater. Des. Appl. 2019, 233, 2219–2233. [Google Scholar] [CrossRef]
  163. Al Rifaie, M.; Mian, A.; Srinivasan, R. Compression behavior of three-dimensional printed polymer lattice structures. Proc. Inst. Mech. Eng. Part L J. Mater. Des. Appl. 2019, 233, 1574–1584. [Google Scholar] [CrossRef]
  164. Lu, Y.; Birol, E.B.; Johnson, C.; Hernandez, C.; Sabin, J.E. A Method for Load-responsive Inhomogeneity and Anisotropy in 3D Lattice Generation Based on Ellipsoid Packing. In Proceedings of the 25th International Conference of the Association for Computer-Aided Architectural Design Research in Asia (CAADRIA), Bangkok, Thailand, 5–8 August 2020; pp. 395–404. [Google Scholar] [CrossRef]
  165. Robles-Linares, J.A.; Ramírez-Cedillo, E.; Siller, H.R.; Rodríguez, C.A.; Martínez-López, J.I. Parametric modeling of biomimetic cortical bone microstructure for additive manufacturing. Materials 2019, 12, 913. [Google Scholar] [CrossRef] [PubMed]
  166. Fantini, M.; Curto, M. Interactive design and manufacturing of a Voronoi-based biomimetic bone scaffold for morphological characterization. Int. J. Interact. Des. Manuf. (IJIDeM) 2018, 12, 585–596. [Google Scholar] [CrossRef]
  167. Nazir, A.; Arshad, A.B.; Jeng, J.Y. Buckling and post-buckling behavior of uniform and variable-density lattice columns fabricated using additive manufacturing. Materials 2019, 12, 3539. [Google Scholar] [CrossRef]
  168. Jin, X.; Li, G.; Gong, J. Optimal design and modeling of 3D variable-density lattice structures. In Proceedings of the 2017 8th International Conference on Mechanical and Aerospace Engineering (ICMAE), Prague, Czech Republic, 22–25 July 2017; IEEE: Piscataway, NJ, USA, 2017; pp. 320–325. [Google Scholar] [CrossRef]
  169. Fragassa, C. Electric city buses with modular platform: A design proposition for sustainable mobility. In Sustainable Design and Manufacturing 2017: Selected Papers on Sustainable Design and Manufacturing 4; Springer International Publishing: Cham, Switzerland, 2017; pp. 789–800. [Google Scholar] [CrossRef]
  170. Abate, K.M.; Nazir, A.; Yeh, Y.P.; Chen, J.E.; Jeng, J.Y. Design, optimization, and validation of mechanical properties of different cellular structures for biomedical application. Int. J. Adv. Manuf. Technol. 2020, 106, 1253–1265. [Google Scholar] [CrossRef]
  171. Dakshnamoorthy, V. Automated Lattice Optimization of Hinge Fitting with Displacement Constraint. Ph.D. Thesis, The University of Texas, Austin, TX, USA, 2016. [Google Scholar]
  172. He, Y.; Burkhalter, D.; Durocher, D.; Gilbert, J.M. Solid-lattice hip prosthesis design: Applying topology and lattice optimization to reduce stress shielding from hip implants. In Proceedings of the Frontiers in Biomedical Devices, 2018 Design of Medical Devices Conference, Minneapolis, MN, USA, 9–12 April 2018; American Society of Mechanical Engineers: New York, NY, USA, 2018; Volume 40789, p. V001T03A001. [Google Scholar] [CrossRef]
  173. Al-Ketan, O.; Abu Al-Rub, R.K. MSLattice: A free software for generating uniform and graded lattices based on triply periodic minimal surfaces. Mater. Des. Process. Commun. 2021, 3, e205. [Google Scholar] [CrossRef]
  174. Maskery, I.; Parry, L.A.; Padrão, D.; Hague, R.J.M.; Ashcroft, I.A. FLatt Pack: A research-focussed lattice design program. Addit. Manuf. 2022, 49, 102510. [Google Scholar] [CrossRef]
  175. Li, Q.; Hong, Q.; Qi, Q.; Ma, X.; Han, X.; Tian, J. Towards additive manufacturing oriented geometric modeling using implicit functions. Vis. Comput. Ind. Biomed. Art 2018, 1, 9. [Google Scholar] [CrossRef]
  176. Tang, Y.; Dong, G.; Zhao, Y.F. A hybrid geometric modeling method for lattice structures fabricated by additive manufacturing. Int. J. Adv. Manuf. Technol. 2019, 102, 4011–4030. [Google Scholar] [CrossRef]
  177. Lei, H.Y.; Li, J.R.; Xu, Z.J.; Wang, Q.H. Parametric design of Voronoi-based lattice porous structures. Mater. Des. 2020, 191, 108607. [Google Scholar] [CrossRef]
  178. Bici, M.; Campana, F.; De Michelis, M. Mesoscale geometric modeling of cellular materials for finite element analysis. Comput.-Aided Des. Appl. 2017, 14, 760–769. [Google Scholar] [CrossRef]
  179. Dal Fabbro, P.; Rosso, S.; Ceruti, A.; Boscolo Bozza, D.; Meneghello, R.; Concheri, G.; Savio, G. Analysis of a Preliminary Design Approach for Conformal Lattice Structures. Appl. Sci. 2021, 11, 11449. [Google Scholar] [CrossRef]
Figure 1. Cellular structures reproduced or adapted from [1,2,5], with permission from publishers Cambridge University Press and ELSEVIER, 2023: (a) open-cell foam; (b) closed-cell foam; (c) honeycomb.
Figure 1. Cellular structures reproduced or adapted from [1,2,5], with permission from publishers Cambridge University Press and ELSEVIER, 2023: (a) open-cell foam; (b) closed-cell foam; (c) honeycomb.
Metals 13 01666 g001
Figure 2. Classification of lattice structure.
Figure 2. Classification of lattice structure.
Metals 13 01666 g002
Figure 3. Beam Lattice (generated with nTop and the colored lines show the 3D reference system useful for comparison): (a) Simple Cubic Cell; (b) Body Centered Cubic Cell; (c) Face Centered Cubic Cell; (d) Diamond Cell; (e) Fluorite Cell; (f) Octet Cell; (g) Truncated Octahedron Cell; (h) Re-entrant Cell; (i) Truncated Cube Cell; (j) Kelvin Cell; (k) Iso Truss Cell; (l) Weaire–Phelan Cell.
Figure 3. Beam Lattice (generated with nTop and the colored lines show the 3D reference system useful for comparison): (a) Simple Cubic Cell; (b) Body Centered Cubic Cell; (c) Face Centered Cubic Cell; (d) Diamond Cell; (e) Fluorite Cell; (f) Octet Cell; (g) Truncated Octahedron Cell; (h) Re-entrant Cell; (i) Truncated Cube Cell; (j) Kelvin Cell; (k) Iso Truss Cell; (l) Weaire–Phelan Cell.
Metals 13 01666 g003
Figure 4. Surface-based lattice (generated with nTop and the colored lines show the 3D reference system useful for comparison): (a) Simple Cubic Foam Cell; (b) Body Centered Cubic Foam Cell; (c) Face Centered Cubic Foam Cell; (d) Hexagonal Honeycomb Cell; (e) Triangular Honeycomb Cell; (f) Re-entrant Honeycomb Cell; (g) Schwartz Cell; (h) Gyroid Cell; (i) Diamond Cell; (j) Split-P Cell; (k) Lidinoid Cell.
Figure 4. Surface-based lattice (generated with nTop and the colored lines show the 3D reference system useful for comparison): (a) Simple Cubic Foam Cell; (b) Body Centered Cubic Foam Cell; (c) Face Centered Cubic Foam Cell; (d) Hexagonal Honeycomb Cell; (e) Triangular Honeycomb Cell; (f) Re-entrant Honeycomb Cell; (g) Schwartz Cell; (h) Gyroid Cell; (i) Diamond Cell; (j) Split-P Cell; (k) Lidinoid Cell.
Metals 13 01666 g004
Figure 5. Examples of regular replication patterns (generated with nTop): (a) Kelvin Cell; (b) Re-entrant Cell; (c) Re-entrant Honeycomb Cell; (d) Split-p Cell.
Figure 5. Examples of regular replication patterns (generated with nTop): (a) Kelvin Cell; (b) Re-entrant Cell; (c) Re-entrant Honeycomb Cell; (d) Split-p Cell.
Metals 13 01666 g005
Figure 6. Examples of hybrid replication patterns achieved by changing the unit cells (generated with nTop): (a) BCC Lattice + Diamond Lattice + FCC Lattice; (b) BCC lattice + Gyroid TMPS.
Figure 6. Examples of hybrid replication patterns achieved by changing the unit cells (generated with nTop): (a) BCC Lattice + Diamond Lattice + FCC Lattice; (b) BCC lattice + Gyroid TMPS.
Metals 13 01666 g006
Figure 7. Examples of pseudo-regular pattern replications (generated with nTop): (a) BCC cells with gradient by thickness; (b) BCC cells with gradient by cell size; (c) Simple cubic cells conformal.
Figure 7. Examples of pseudo-regular pattern replications (generated with nTop): (a) BCC cells with gradient by thickness; (b) BCC cells with gradient by cell size; (c) Simple cubic cells conformal.
Metals 13 01666 g007aMetals 13 01666 g007b
Figure 8. Examples of stochastic replication patterns (generated with nTop): (a) cell’s size variation; (b) thickness variation by gradient.
Figure 8. Examples of stochastic replication patterns (generated with nTop): (a) cell’s size variation; (b) thickness variation by gradient.
Metals 13 01666 g008
Figure 9. Query results from Scopus: (a) timeline relative percentage distribution of journal papers in Scopus sorted by topic (subject limit to “Engineering”), (b) amount of journal papers per topic (subject limit to “Engineering”).
Figure 9. Query results from Scopus: (a) timeline relative percentage distribution of journal papers in Scopus sorted by topic (subject limit to “Engineering”), (b) amount of journal papers per topic (subject limit to “Engineering”).
Metals 13 01666 g009
Table 1. Applications: overview by field of expertise, lattice structure classification, and useful references.
Table 1. Applications: overview by field of expertise, lattice structure classification, and useful references.
SectorsUnit Cell/Replication PatternApplicationsReferences
Aerospace Engineering
  • BCC
  • Auxetic
  • Rhombic
  • Chiral
  • Honeycomb
  • Pyramidal
  • Heat Exchanger
  • Wings
  • Gas Turbine Fan Blades
  • Airfoil
  • Drone Structure
  • Satellite Structure
  • Rocket Body
  • Landing Gear
[18,70,71,72,73,74,75,76]
Biomedical Engineering
  • Diamond Cubic
  • Stochastic
  • Octet-truss
  • Tetrahedral
  • BCC
  • TPMS
  • Honeycomb Lattice
  • Hip Implant
  • Orthopedic Implant
  • Bone Scaffold
  • Femoral Stem
[77,78,79,80,81,82,83,84,85,86]
Mechanical Engineering:
Automotive
  • Honeycomb
  • Truss-based
  • TPMS
  • Auxetic
  • Crash Box
  • Bumper
  • Engine Hood
  • Crash Absorber
  • Chassis Frame
  • Heat Exchanger
[87,88,89,90,91,92]
Mechanical Engineering: Defense
  • BCC
  • Honeycomb
  • Auxetic
  • Military Armor Body
  • Helmet Liner
  • Military Seat Shock Panel
[93,94,95]
Mechanical Engineering: Heat Exchanger
  • Truss-based
  • TPMS
  • Natural Convection
[96,97]
Mechanical Engineering
  • BCC
  • Cubic
  • FCC
  • TPMS
  • AM Supporting Structure
  • AM Volume Infills
[98,99,100,101,102,103]
Product Design: Sports
  • Auxetic
  • Honeycomb
  • Truss-base Random
  • Football Helmet
  • Sports Shoes
  • Horse Saddle
[104,105,106]
Civil Engineering: Building Construction
  • BCC
  • Conformal
  • Stochastic
  • BCC
  • Canton Tower
  • Beijing National Stadium (Bird’s Nest)
  • The Atomium
[107,108,109]
Table 2. Correlation table among traditional manufacturing methods, base material, limitations, and related references.
Table 2. Correlation table among traditional manufacturing methods, base material, limitations, and related references.
Manufacturing MethodBase MaterialIssuesReferences
Direct Foaming
Metals
  • Cell Replication Pattern: Only stochastic structure can be obtained.
[110,111]
Investment Casting
Metals
  • Low accuracy for fine structures.
  • Long production time.
[112,113,114]
Stamping Forming
Metals
  • Cell’s characteristics.
  • Post processing is required.
  • Difficult to produce fine structures.
[115,116]
Interlocking Grid
Assembly
Metals
Composites
Polymer Fibers
  • Cell’s characteristics.
  • Difficult to produce fine structures.
[117,118,119]
Extrusion Wire Cutting
Metals
  • Cell’s characteristics.
  • Costly processes.
[120]
Lap Assembly
Metals
  • Cell’s characteristics.
  • Post-processing is required.
[115]
Wire-Woven Method
Metals
  • Cell’s characteristics.
  • Post processing is required.
[121,122,123]
Table 3. Correlation table among Additive Manufacturing (AM) methods, base material, limitations, and related references.
Table 3. Correlation table among Additive Manufacturing (AM) methods, base material, limitations, and related references.
AM MethodBase MaterialIssuesReferences
Binder jetting (BJ)
Metals
Ceramics
Polymers
  • Part shrinkage 2–3%
  • Low accuracy and tolerance
[124,125,126]
Cold Spray Additive Manufacturing (CSAM)
Metals
Ceramics
  • Loss of ductility due to plastic deformation
  • Post processing
[127,128,129]
Direct Energy Deposition
(DED)
Metals
Ceramics
  • Low resolution
  • Poor surface finish
  • Costly
[130,131,132]
Direct Ink Writing (DIW)
Ceramics
Metals
Slurries
Polymers
Sol-gel inks
  • Fine surface finish, in cases of ceramics
[133,134,135]
Fused Deposition Modeling (FDM)
Metals
Thermoplastics
Wax
  • Slow Speed
  • Low Accuracy
[136,137,138]
Liquid Metal
Additive
Manufacturing
Metals
  • Melting and cooling of the printing material influencing the final geometry
[139,140,141]
Laminated Object Manufacturing (LOM)
Papers
Plastics
Metals
  • Poor surface finish
[142,143,144]
Powder Bed
Fusion (PBF)
Metals
Ceramics
Thermoplastics
Wax
  • Volumetric Expansion
  • Low accuracy and tolerance
[145,146,147]
Table 4. Three-dimensional modeling of lattice structures.
Table 4. Three-dimensional modeling of lattice structures.
Design ToolsSoftwareLattice TypeReferences
CAD Software/
Numerical Solvers
Solid Works
nTopology
Optistruct
Inspire
Rhinoceros Grasshopper
Fusion 360
PTC Creo
Netfabb
Materialize
Free CAD
Ansys
Beam Lattice
TPMS
Honeycomb
Voronoi
Auxetic
Conformal
Gradient Lattice
[152,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172]
Special Tools
MS Lattice
FLatt Pack
TPMS
[173,174]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ahmad, A.; Belluomo, L.; Bici, M.; Campana, F. Bird’s Eye View on Lattice Structures: Design Issues and Applications for Best Practices in Mechanical Design. Metals 2023, 13, 1666. https://doi.org/10.3390/met13101666

AMA Style

Ahmad A, Belluomo L, Bici M, Campana F. Bird’s Eye View on Lattice Structures: Design Issues and Applications for Best Practices in Mechanical Design. Metals. 2023; 13(10):1666. https://doi.org/10.3390/met13101666

Chicago/Turabian Style

Ahmad, Abas, Luca Belluomo, Michele Bici, and Francesca Campana. 2023. "Bird’s Eye View on Lattice Structures: Design Issues and Applications for Best Practices in Mechanical Design" Metals 13, no. 10: 1666. https://doi.org/10.3390/met13101666

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

Article Metrics

Back to TopTop