Next Article in Journal
A Review of Mechanical Fine-Pointing Actuators for Free-Space Optical Communication
Previous Article in Journal
Superhydrophobic Coatings for Corrosion Protection of Stainless Steel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanical Performance of a Node-Reinforced Body-Centered Cubic Lattice Structure: An Equal-Strength Concept Design

1
Hebei Key Laboratory of Mechanical Reliability for Heavy Equipments and Large Structures, Yanshan University, Qinhuangdao 066004, China
2
School of Civil Engineering and Mechanics, Yanshan University, Qinhuangdao 066004, China
3
State Key Laboratory of Metastable Materials Science and Technology, Yanshan University, Qinhuangdao 066004, China
*
Author to whom correspondence should be addressed.
Aerospace 2024, 11(1), 4; https://doi.org/10.3390/aerospace11010004
Submission received: 21 November 2023 / Revised: 15 December 2023 / Accepted: 18 December 2023 / Published: 19 December 2023

Abstract

:
Lattice structures are characterized by a light weight, high strength, and high stiffness, and have a wide range of applications in the aerospace field. Node stress concentration is a key factor affecting the mechanical performance of lattice structures. In this paper, a new equal-strength body-centered cubic (ES-BCC) lattice structure was additively manufactured using 316L stainless steel via selective laser melting (SLM). The results of a mechanical compression test and finite element analysis revealed that the failure location of the ES-BCC structure changed from the nodes to the center of the struts. At the same density, the energy absorption, elastic modulus, and yield strength of the ES-BCC structure increased by 11.89%, 61.80%, and 53.72% compared to the BCC structure, respectively. Furthermore, the change in angle of the ES-BCC structure achieves significant changes in strength, stiffness, and energy absorption to meet different design requirements and engineering applications. The equal-strength concept design can be applied as a general design method to the design of other lightweight energy-absorbing lattice structures.

1. Introduction

Lattice structures are structures constructed by combining one or more structural cytons in a periodic array. Due to their superior characteristics of light weight, high strength, high stiffness, energy absorption, vibration damping, and heat and noise insulation, they are widely used in lightweight design, aerospace, biomedicine, and automotive ships [1,2,3]. The advancement of additive manufacturing technology has advantages in the design and processing of lattice structures, which has greatly promoted the development of such structures and made them a hotspot for lightweight structure research in the aerospace and aviation fields [4].
The current reported lattice structures include body-centered cubic (BCC), face-centered, simple cubic, octet, diamond, rhombic dodecahedron, pyramidal, octet-truss, triply periodic minimal surface, and hybrid structures [5,6,7,8,9,10,11]. Besides the materials used for manufacturing structures, the shape and geometrical parameters of the lattice structure determine its mechanical properties, i.e., the number, orientation, and diameter of the struts [12]. Lattice structures have been classified into tensile-dominated and bending-dominated types [13,14,15]. Therefore, the deformation mechanism of the struts under load has been considered as determining the mechanical performance of the structure [16,17]. However, almost all of the currently reported stress concentration is often initiated at the nodes, which leads to faster crack initiation and failure [18,19,20].
Some studies improve the mechanical properties of lattice structures through the node design. A novel BCC lattice structure with tapered struts was designed to reduce the stress concentration at the nodes [6]. The results show that the structure fracture location changed from the nodes to the tapered strut center, and the elastic modulus was increased by 67%. A new graded-strut body-centered cubic structure was obtained by replacing node right angles with node fillets [21,22]. In the filleted structure, strut failure initiated near the end of the fillet, and the compressive modulus and yield strength were increased by at least 38.20% and 34.12%, respectively. Taking inspiration from a crystal structure, a BCC structure with atoms as reinforced nodes and atomic bonds as connected struts has been manufactured [23], and the yield strength and collapse strength were increased by 12.31% and 31.19%, respectively. Although experience-based node reinforcement design improves the mechanical performance of structures, it lacks theoretical support and is difficult to widely apply.
In the present work, a new equal-strength body-centered cubic (ES-BCC) lattice structure was proposed. The material is additively manufactured with 316L stainless steel via the SLM process. The deformation mode, elastic modulus, yield strength, and energy absorption of the ES-BCC structure were investigated using numerical simulation and compression tests and compared with the BCC structure. The equal-strength design of the struts can effectively improve the mechanical properties of the structure, which provides a new idea for the design of lightweight energy-absorbing structures in the aerospace field.

2. Materials and Methods

2.1. Design of ES-BCC Structure

The ES-BCC structure was designed based on the equal-strength concept of the BCC structure. Figure 1 shows the schematic diagram of the BCC structure under axial compression: the shaded portion indicates that the bottom of the structure is a fixed constraint. The BCC lattice structure, as shown in Figure 1a, contains unit cells with bottom dimensions of L × L and the height H. The eight struts surrounding the node are taken as the analysis module, as shown in Figure 1b. Assuming that the compression load is F, since the structure is geometrically symmetric, any strut near the node can be taken for force analysis, which is simplified into a cantilever beam structure. The strut AB is taken for force analysis, as shown in Figure 1c.
Decomposing the load F, the axial and tangential forces on the strut can be obtained as:
F x = F sin θ F y = F cos θ
where θ is the angle between the bottom surface and the structural strut.
Under combined compression-bending deformation, the stress can be represented by:
σ x = F x A x + M x W z x
For circular cross-section struts, A x = π d 2 x 4 is the cross-sectional area at x, M x = F y x is the bending moment at x, and W x = π d 3 x 32 is the bending section coefficient.
Letting the stress at any x be equal to the allowable stress of the base material, we have:
σ max = σ = 4 F sin θ π d 0 2 x = 0 4 F sin θ π d 1 2 + 32 F cos θ l π d 1 3 x = l
where d0 is the initial diameter (diameter at the center of struts), and d1 is the maximum diameter (diameter at nodes), l = 2 L 2 + H 2 / 4 . From Equation (3), we have:
F = π d 0 2 σ 4 sin θ
Substituting Equation (4) into σ x = σ , the function of the strut diameter can be given as:
d 3 x d 0 2 d x = 8 cot θ d 0 2 x
Equation (5) is the design formula for the ES-BCC structure. Obviously, the diameter of the strut is related to d0 and θ .
As shown in Figure 1b, ϕ is the angle between the struts in the front view of the structure. From the geometric relationship, the relationship between ϕ and θ can be obtained:
sin θ = tan ϕ 2 2 + tan 2 ϕ 2
ES-BCC unit cell structures with d0 = 0.3 mm are shown in Table 1. Since a non-smooth transition at the center of the strut would result in stress concentration, a rounded corner transition with R = 3 mm was performed in the actual modeling. As shown in Table 1, the shapes of the strut profiles determined using Equation (5) are similar, with differences only in the size of the diameters at each location. For structures with ϕ of 75°, 90°, 105°, and 120°, the minimum diameters of the struts after adding rounded corners are 0.86, 0.76, 0.69, and 0.63 mm when d0 is 0.30 mm, respectively.

2.2. Relative Density

The ratio of the solid volume of the structure to the volume of the quadrilateral filled by the structure is the relative density of the lattice structure. When filling the lattice structure into the cube, the ends of each strut are partially truncated in the x, y, and z directions, and a schematic diagram of the intercepted structural strut dimensions is shown in Figure 2.
The distance between the nodes of the force can be expressed by 2( l Δ l ). If the strut is uniform, then Δ l = a + b 2 = d 4 tan θ + 1 tan θ , where d is the strut diameter. For the new design structures, the diameter d is a function of x, and d is approximated as d 0 + d 1 / 2 . Thus, we obtain:
Δ l = d 0 + d 1 8 tan θ + 1 tan θ
here, d0 is the minimum diameter after adding rounded corners and d1 is the maximum diameter.
As shown in Figure 2, the volume of the variable-section strut of the 2l segment is V 1 = 2 0 l π r 2 x d x .
From the above analysis, the volume of the 2( l Δ l )-segment strut can be obtained as:
V 2 = 2 0 l Δ l π r 2 x d x
Since each strut of the structure is partially truncated, the volume of the Δ l -segment strut is approximated as V Δ l = V 1 V 2 / 8 . The unit cell volume can be obtained as follows:
V = 8 V 2 + 2 V Δ l
The relative density can be obtained with rectification as follows:
ρ ¯ = V L 2 H

2.3. Specimen Fabrication

The bottom surface dimensions of the ES-BCC unit cell structure are L × L, taking L = 6 mm, and the height H is related to ϕ . The ES-BCC structure is named according to ϕ to facilitate later analysis. For example, a lattice structure with ϕ = 75° is named an A075 structure.
The specimens were fabricated using a selective laser melting (SLM) process, and 316L stainless steel was selected as the fabrication material. To eliminate the influence of the specimen printing direction, all specimens were fabricated along the direction of the experimental load. The overall dimensions of the structures were 30 mm × 30 mm × 30 mm. To compare the mechanical properties of the ES-BCC structures with those of the original BCC lattice structure, the conventional BCC lattice structure (strut diameter d = 1.025 mm) was fabricated with the same dimensions and relative density as the A090 structure. In order to avoid errors in the experimental process, two specimens were prepared for all lattice structures. The specimens are shown in Figure 3. Table 2 shows 316L stainless steel’s chemical composition [24].

2.4. Experimental Procedure

Quasi-static tests were performed on the specimens using an Instron 5982 universal material testing machine, including tensile and compressive. The quasi-static compression test setup is shown in Figure 4a. The loading speed was set to 2 mm/min. For compression testing, the loading was stopped at the densification stage or 25 mm displacement. For tensile testing, the loading was stopped when the specimen fractured. The force and displacement are automatically recorded by the data acquisition system. A DSLR camera Canon 80 D was used to record the entire process of the experiment in real time. The tensile stress–strain curves of the additive manufacturing 316L stainless steel specimens are shown in Figure 4b. The mechanical properties shown in Table 3 are similar to those in reference [25,26], which were used in the subsequent numerical simulations.

2.5. Finite Element Modeling

The modeling, meshing, and solving were performed using the CAD, HyperMesh 2021, and LS-DYNA R11.1.0 software. The finite element model is shown in Figure 5a. The effect of the strain rate is not considered in the quasi-static compression tests. To simulate the upper and lower indenters in the compression test, rigid plates are placed and meshed with shell elements, and the lattice structure is meshed with solid elements (C3D8R). The top rigid plate was loaded with displacement at a constant rate of 1 m/s [27,28], which was to improve the computational efficiency of the numerical simulation [29,30,31,32], with a compression displacement of 25 mm, and the bottom rigid plate was completely fixed. The static friction coefficient between the lattice structure and indenters is 0.3 [33,34].
The mesh convergence test was conducted to reduce the influence of the mesh element size on the stress–strain curve, as shown in Figure 5b. The element sizes were 0.2, 0.3, 0.5, and 0.8 mm, respectively. The stress–strain curve of the finite element model with an element size of 0.8 mm had large deviations from those of the other sizes, and the error between the results of element sizes of 0.2, 0.3, and 0.5 mm was very small. Therefore, elements with a grid size of 0.3 mm were used for subsequent simulation calculations.

3. Results and Discussion

3.1. Morphological Characteristics

The masses of all specimens were obtained by weighing them, and the relative densities were calculated. Figure 6 shows the design values, theoretical values, and measured values of the relative density of the ES-BCC structures. The relative density of all specimens was smaller than the design value, the maximum error was 6.2%, and the relative density error between the design and theoretical values was less than 6.9%. The main reason for errors would have been rounded transitions, approximately calculated volumes, defects, and residual powder particles.
Scanning electron microscopy (SEM) was used to assess the manufacturing quality of the experimental specimens. As shown in Figure 7, the surface of each strut was rough. The adhesion of partially melted powder particles can be observed, especially at the nodes and below the struts. Moreover, porosity defects were observed on the struts. At the intersection of the struts, small-angle structures had more residual powder particles than large-angle structures. Furthermore, porosity defects might also exist inside the struts of the lattice structure.

3.2. Mechanical Properties

The compressive stress–strain curves of the A090 structure and the BCC lattice structure were compared, as shown in Figure 8. The simulation curves are in good agreement with the experimental results, especially in the initial phases. Under quasi-static compression, the lattice structure exhibited the following deformation responses: an initial linear elastic phase, a plastic plateau phase where the stress remained constant, and a densification phase where the stress sharply grew. The pentagram in the Figure 8 represents the starting point of the structure into the densification phase. The compressive stress–strain curves of the A090 and BCC lattice structures had the same trend, but at the same strain, the former’s stress exceeded the latter.
The deformation behavior can directly reflect the damage failure mode. Figure 9 depicts the deformation process of the lattice structure in compression tests and simulations. When the strain reached 40%, both the A090 and BCC lattice structures had significant deformation on the diagonal struts, and the center of the structure showed local densification. Struts in various areas exhibited different deformation patterns, with the external struts producing tensile deformation and the internal struts producing buckling deformation. The BCC lattice structure was deformed uniformly in the compression process. The overall stress distribution of the structure was not uniform, and at the nodes of the struts, stress concentration was observed. However, the A090 structure was the first to show bending deformation and compaction at the center of the strut where it made contact with the upper and lower plates. The rest of the layers deformed uniformly. During the compression process, the stress distribution of the A090 structure was more uniform, and the damage failure location was at the center of the strut.
To quantitatively describe the mechanical properties and energy absorption characteristics of the ES-BCC structures, the elastic modulus (E), yield strength ( σ y ), and energy absorption (EA) were compared. The elastic modulus was defined as the slope of the initial linear elastic segment (about 1–3% of the strain). The yield strength was defined as the stress at the point where the stress–strain curve intersected a straight line whose slope equaled the elastic modulus and passed through the 0.2% strain [35]. The energy absorption of a lattice structure can be represented by the energy absorption per unit volume as the area under the compressive stress–strain curve:
E A = 0 ε d σ ε d ε
where σ ε and ε are the stress and strain during compression, respectively, and ε d is the maximum strain when the structure is compressed to the densification stage [36]. Table 4 shows the densification strains of the experimental and simulated compression stress–strain curves. It can be observed that the densification strain increases with ϕ . The ε d of the BCC lattice structure exceeded that of the A090 structure, consistent with the compressive stress–strain curves shown in Figure 9.
Table 5 compares the mechanical properties of the structures under study in compression tests and simulations. The discrepancies among the elastic moduli, yield strengths, and energy absorptions were within 10.5%, 1.2%, and 2.6%, respectively, proving the reliability of the numerical simulation. The elastic modulus, yield strength, and EA of the A090 structure outperformed those of the BCC lattice structure by 61.80%, 53.72%, and 11.89%, respectively.

3.3. Influence of ϕ

At the initial diameter d0 = 0.3 mm, the effect of ϕ on the mechanical properties of the ES-BCC structures was studied, as shown in Figure 10. The stress was initially concentrated at the center of the struts of the top and bottom cell layers and propagated to the adjacent layers with the ongoing crush. At a strain of 40%, significant densification occurred at the center of the structure. The smaller ϕ was, with more crystal cells and a higher relative density resulting in faster densification of the structure during compression, the more uniform the stress distribution was. It was also noticed that the collapse began at the top and bottom cell layers, followed by the middle layer, and finally the adjacent two layers began to collapse. The specimens did not produce fracture damage during compression.
The compressive stress–strain curves of ES-BCC structures with different ϕ are presented in Figure 11. It demonstrates the typical stress–strain behavior of lattice structures with three characteristic regions consisting of a linear elastic region, plateau region, and densification region. The change in ϕ did not affect the general trend of the stress–strain curve but at the same strain, the stress decreased with an increasing ϕ . The smaller the angle, the greater the elastic modulus, yield strength, and energy absorption, but the influence of mass is not taken into account.
The mechanical properties of ES-BCC structures at different ϕ values were further evaluated by calculating their relative elastic modulus ( E ρ ¯ ), relative yield strength ( σ ρ ¯ ), and specific energy absorption (SEA) [37,38]. The calculation formula is as follows:
E ρ ¯ = E ρ ¯
σ ρ ¯ = σ y ρ ¯
S E A = E A ρ = 0 ε d σ ε d ε ρ
where ρ is the density of the lattice structure.
Figure 12 shows the mechanical properties of ES-BCC structures with different ϕ . The larger ϕ , the greater the relative elastic modulus, relative yield strength, and specific energy absorption. When the ϕ of the structure increased from 75° to 120°, the relative elastic modulus grew from 2.19 to 3.64 GPa (i.e., by 66.21%), the relative yield strength grew from 44.34 to 65.87 MPa (i.e., by 48.56%), and the SEA increased from 3.89 to 5.59 J/g (i.e., by 43.70%). Therefore, the A120 structure had a superior elastic modulus, yield strength, and EA at the same relative density. Mainly due to the increase in angle, the compression force of the structure increases and the bending force decreases.

3.4. Gibson–Ashby Model Prediction

The Gibson–Ashby (GA) model [39] is a classical model theory based on cubic porous structures simplified by cubic struts. It uses the geometric mechanics method to link porous materials’ mechanical properties and can effectively predict their tensile and compressive properties, such as the elastic modulus, tensile and compressive strength, and energy absorption. The normalized elastic modulus and normalized yield strength of the lattice structure can be characterized by power functions of the relative density via the GA equations as follows:
E E s = C 1 ρ ¯ n 1
σ y σ s = C 2 ρ ¯ n 2
where E and σ y are the compressive modulus and yield strength of the lattice structure, respectively, E s and σ s are the elastic modulus and yield strength of the matrix material, respectively, while the coefficient C and the exponent n characterize the geometry of the unit cell. The exponent reflects the mechanical properties of the lattice structure better than the coefficient because the functional relationship follows power function characteristics. Structures with different relative densities were analyzed with a d0 of 0.15, 0.20, 0.25, and 0.30 mm. The fitted curves of the GA model for the E and σ y of the lattice structures are shown in Figure 13.
As shown in Figure 13, the E and σ y of the lattice structures increased as ρ ¯ increased. The mechanical properties improved significantly with increasing ϕ , while the E and σ y of the A120 structure significantly exceeded those of other lattice structures with the same relative density. This substantiates the design concept of high-strength and high-stiffness lattice structures.
The fitted parameters of the GA model for the mechanical properties of ES-BCC structures are listed in Table 6. Due to the large difference in the exponents of lattice structures with various geometric configurations [40], according to the Maxwell criterion [41], lattice structures are classified as tensile- or bending-dominated based on their deformation forms. Bending- and tensile-dominant deformation structures have a normalized elastic modulus index n1 close to 2 and 1, respectively, and a normalized yield strength index n2 close to 1.5 and 1, respectively. Notably, the indices n1 and n2 of the ES-BCC structures were close to 2 and 1.5, respectively. Therefore, the ES-BCC structures under study exhibited bending-dominated deformation under quasi-static compression, which were more stable in the plateau phase and had higher energy absorption [42].

3.5. Theoretical Analysis

3.5.1. Equivalent Elastic Modulus

Assume that the lattice structure is only subjected to an axial compression load F, as shown in Figure 14. The load F is decomposed into the axial force Fx and tangential force Fy, and the strut displacement can be decomposed into axial displacement δ x and tangential displacement δ y .
The axial displacement δ x of the strut is:
δ x = 0 l F x E A x d x = 0 l 4 F sin θ E π d 2 x d x
where E is the elastic modulus of the structural material.
Tangential displacement δ y is caused by the combined action of the bending moment and shear. The tangential displacement can be obtained using the unit load method as:
δ y = 0 l M ¯ M x E I x d x + 0 l K s F y G A x d x
where M ¯ = x is the bending moment under unit load; Ks = 10/9 is the shear coefficient for a circular section; G = E / 2 ( 1 + μ ) is the shear modulus; and I x = π d 4 x 64 = π r 4 x 4 is the moment of inertia at x.
Let the displacement of the unit cell in the compression direction be δ . Projecting the axial and tangential displacements onto the force direction, we can obtain:
δ = 4 δ x sin θ + δ y cos θ
The equivalent elastic modulus can be expressed as:
E e = σ e ε e
where σ e = 4 F L 2 is the equivalent stress, and ε e = δ H is the equivalent strain, collated to obtain:
E e = 4 F H L 2 δ

3.5.2. Equivalent Yield Strength

When the lattice structure is subjected to axial compression, the combined action of the axial force and bending moment is applied to the variable-section struts, resulting in combined compression-bending deformation, eventually causing the damage failure of the structure. Struts are prone to buckling under compressive loading, but the struts in the ES-BCC structure studied in this paper do not satisfy the slender rod condition, so the buckling effect is not considered. According to the compression test, it is found that the plastic yielding damage of the structure was first produced on the tensile side, so taking the tensile stress surface as the plastic damage failure criterion surface, the stress expression at x is as follows:
σ x = M x W x F x A x
At a maximum x, i.e., when x = l, the maximum value of stress is observed. Making σ x max equal to the yield strength of the substrate, we obtain:
F max = σ l 32 cos θ l π d 3 l 4 sin θ π d 2 l
Then, the plastic damage strength is:
σ y = 4 F max L 2
When the lattice structure undergoes plastic failure, it can be regarded as reaching the yield state, and the plastic damage strength at this time is the equivalent yield strength.

3.5.3. Elastic Modulus and Yield Strength Analysis

The parameters E and σ y of ES-BCC structures with different ϕ can be obtained by applying the theoretical prediction equation. As shown in Figure 15, the theoretical values of the A075 and A090 structures are smaller than the experimental and simulated values because in the theoretical prediction, the calculation object is the unit cell structure with direct contact between the top layer and the indenter, but in the actual compression process, the top and bottom layers of the structure are the first to produce damage failure, while the inner unit cell is more constrained and bears part of the load. The theoretical values of the A105 and A120 structures are larger than the experimental and simulated values, probably because the struts are longer. The struts are prone to buckling and combined bending and torsional deformation at the nodes, while the influence of strut buckling is not considered in the theoretical calculations.
Table 7 shows the E and σ y values of ES-BCC structures with different ϕ values. The discrepancies in E between the experimental and simulated results ranged from −7.52% in the A090 structure to 11.73% in the A105 structure, while those in σ y ranged from −4.13% for the A075 structure to 14.26% for the A105 structure. The deviations between the experimental and simulated or theoretical values of the A105 structure were the largest because of the unavoidable processing defects in the specimens. However, its simulated results were more consistent with the theoretical ones. The discrepancies between the experimental and theoretical values of E and σ y for the ES-BCC structures were within 18.93 and 13.29%, proving that the theoretical model could predict E and σ y of the ES-BCC structures effectively.

4. Conclusions

In this study, an equal-strength concept was proposed for building novel lattice structures, and a new BCC lattice structure, equal-strength body-centered cubic (ES-BCC), is designed, which can be applied as a lightweight structure in the aerospace field. The mechanical properties of the designed structure were analyzed using experiments and finite element analysis, and the main conclusions are as follows:
(1) ES-BCC lattice structures based on the equal-strength concept were designed to reduce the node stress concentration effectively. The yield position of the BCC lattice structures changed from the node region to the center of strut. The elastic modulus, yield strength, and energy absorption of the A090 structure were improved by 61.80%, 53.72%, and 11.89% compared to the BCC structure, respectively.
(2) The ES-BCC lattice structures all exhibited bending-dominated deformation. A larger ϕ corresponded to a higher relative elastic modulus, relative yield strength, and SEA. Those of the A120 structure reached 3.64 GPa, 65.87 MPa, and 5.59 J/g, respectively, outperforming those of the A075 structure by 66.21, 48.56, and 43.70%, respectively.
(3) By changing ϕ from 75° to 120°, substantial variations in strength, stiffness, and energy absorption were achieved to meet different design requirements and engineering applications. Thus, small-angle ES-BCC lattice structures can be used in energy-buffering devices, and large-angle ones can meet high-strength requirements.

Author Contributions

Z.L.: conceptualization, methodology, funding acquisition, investigation, writing—review and editing. R.Z.: writing—original draft, software, visualization, validation. C.T.: writing—original draft, software, validation. Y.W.: writing—original draft, validation. X.L.: project administration, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (11902287).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of 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.

References

  1. Maconachie, T.; Leary, M.; Lozanovski, B.; Zhang, X.; Qian, M.; Faruque, O.; Brandt, M. SLM lattice structures: Properties, performance, applications and challenges. Mater. Des. 2019, 183, 108137. [Google Scholar] [CrossRef]
  2. Tao, C.; Wang, Z.; Liu, Z.; Wang, Y.; Zhou, X.; Liang, X.; Li, H. Crashworthiness of Additively Manufactured Lattice Reinforced Thin-Walled Tube Hybrid Structures. Aerospace 2023, 10, 524. [Google Scholar] [CrossRef]
  3. Alkhatib, S.E.; Karrech, A.; Sercombe, T.B. Isotropic energy absorption of topology optimized lattice structure. Thin-Walled Struct. 2023, 182, 110220. [Google Scholar] [CrossRef]
  4. Kumar, A.; Collini, L.; Daurel, A.; Jeng, J.-Y. Design and additive manufacturing of closed cells from supportless lattice structure. Addit. Manuf. 2020, 33, 101168. [Google Scholar] [CrossRef]
  5. 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]
  6. Zhao, M.; Zhang, D.Z.; Li, Z.; Zhang, T.; Zhou, H.; Ren, Z. Design, mechanical properties, and optimization of BCC lattice structures with taper struts. Compos. Struct. 2022, 295, 115830. [Google Scholar] [CrossRef]
  7. Liu, R.; Chen, W.; Zhao, J. A review on factors affecting the mechanical properties of additively-manufactured lattice structures. J. Mater. Eng. Perform. 2023, 1–27. [Google Scholar] [CrossRef]
  8. Park, J.H.; Park, K. Compressive behavior of soft lattice structures and their application to functional compliance control. Addit. Manuf. 2020, 33, 101148. [Google Scholar] [CrossRef]
  9. Cao, X.; Xiao, D.; Li, Y.; Wen, W.; Zhao, T.; Chen, Z.; Jiang, Y.; Fang, D. Dynamic compressive behavior of a modified additively manufactured rhombic dodecahedron 316L stainless steel lattice structure. Thin-Walled Struct. 2020, 148, 106586. [Google Scholar] [CrossRef]
  10. Li, Z.; Nie, Y.; Liu, B.; Kuai, Z.; Zhao, M.; Liu, F. Mechanical properties of AlSi10Mg lattice structures fabricated by selective laser melting. Mater. Des. 2020, 192, 108709. [Google Scholar] [CrossRef]
  11. Guo, X.; Ding, J.; Li, X.; Qu, S.; Hsi Fuh, J.Y.; Lu, W.F.; Song, X.; Zhai, W. Interpenetrating phase composites with 3D printed triply periodic minimal surface (TPMS) lattice structures. Compos. B Eng. 2023, 248, 110351. [Google Scholar] [CrossRef]
  12. Lee, K.W.; Lee, S.H.; Noh, K.H.; Park, J.Y.; Cho, Y.J.; Kim, S.H. Theoretical and numerical analysis of the mechanical responses of BCC and FCC lattice structures. J. Mech. Sci. Technol. 2019, 33, 2259–2266. [Google Scholar] [CrossRef]
  13. Li, S.J.; Xu, Q.S.; Wang, Z.; Hou, W.T.; Hao, Y.L.; Yang, R.; Murr, L. Influence of cell shape on mechanical properties of Ti–6Al–4V meshes fabricated by electron beam melting method. Acta Biomater. 2014, 10, 4537–4547. [Google Scholar] [CrossRef] [PubMed]
  14. Eckel, Z.C.; Zhou, C.; Martin, J.H.; Jacobsen, A.J.; Carter, W.B.; Schaedler, T.A. Additive manufacturing of polymer-derived ceramics. Science 2016, 351, 58–62. [Google Scholar] [CrossRef]
  15. Jens, B.; Stefan, H.; Iwiza, T.; Ruth, S.; Oliver, K. High-strength cellular ceramic composites with 3D microarchitecture. Proc. Natl. Acad. Sci. USA 2014, 111, 2453–2458. [Google Scholar]
  16. Ozdemir, Z.; Hernandez-Nava, E.; Tyas, A.; Warren, J.A.; Fay, S.D.; Goodall, R.; Todd, I.; Askes, H. Energy absorption in lattice structures in dynamics: Experiments. Int. J. Impact. Eng. 2016, 89, 49–61. [Google Scholar] [CrossRef]
  17. Liu, Y.J.; Li, S.J.; Wang, H.L.; Hou, W.T.; Hao, Y.L.; Yang, R.; Sercombe, T.B.; Zhang, L.C. Microstructure, defects and mechanical behavior of beta-type titanium porous structures manufactured by electron beam melting and selective laser melting. Acta Mater. 2016, 113, 56–67. [Google Scholar] [CrossRef]
  18. Qiu, C.; Yue, S.; Adkins, N.J.E.; Ward, M.; Hassanin, H.; Lee, P.D.; Withers, P.J.; Attallah, M.M. Influence of processing conditions on strut structure and compressive properties of cellular lattice structures fabricated by selective laser melting. Mater. Sci. Eng. A 2015, 628, 188–197. [Google Scholar] [CrossRef]
  19. Yu, G.; Li, Z.; Li, S.; Zhang, Q.; Hua, Y.; Liu, H.; Zhao, X.; Dhaidhai, D.T.; Li, W.; Wang, X. The select of internal architecture for porous Ti alloy scaffold: A compromise between mechanical properties and permeability. Mater. Des. 2020, 192, 108754. [Google Scholar] [CrossRef]
  20. Dallago, M.; Raghavendra, S.; Luchin, V.; Zappini, G.; Pasini, D.; Benedetti, M. The role of node fillet, unit-cell size and strut orientation on the fatigue strength of Ti-6Al-4V lattice materials additively manufactured via laser powder bed fusion. Int. J. Fatigue 2021, 142, 105946. [Google Scholar] [CrossRef]
  21. Latture, R.M.; Rodriguez, R.X.; Holmes, L.R.; Zok, F.W. Effects of nodal fillets and external boundaries on compressive response of an octet truss. Acta Mater. 2018, 149, 78–87. [Google Scholar] [CrossRef]
  22. Bai, L.; Yi, C.; Chen, X.; Sun, Y.; Zhang, J. Effective design of the graded strut of BCC lattice structure for improving mechanical properties. Materials 2019, 12, 2192. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, Y.; Zhang, J.; Gu, X.; Zhou, Y.; Yin, Y.; Tan, Q.; Li, M.; Zhang, M.-X. Mechanical performance of a node reinforced body-centred cubic lattice structure manufactured via selective laser melting. Scr. Mater. 2020, 189, 95–100. [Google Scholar] [CrossRef]
  24. Zhang, Y.; Xu, X.; Fang, J.; Huang, W.; Wang, J. Load characteristics of triangular honeycomb structures with self-similar hierarchical features. Eng. Struct. 2022, 257, 114114. [Google Scholar] [CrossRef]
  25. Roach, A.M.; White, B.C.; Garland, A.; Jared, B.H.; Carroll, J.D.; Boyce, B.L. Size-dependent stochastic tensile properties in additively manufactured 316L stainless steel. Addit. Manuf. 2020, 32, 101090. [Google Scholar] [CrossRef]
  26. Li, P.; Wang, Z.; Petrinic, N.; Siviour, C.R. Deformation behaviour of stainless steel microlattice structures by selective laser melting. Mater. Sci. Eng. A 2014, 614, 116–121. [Google Scholar] [CrossRef]
  27. Jiang, F.; Yang, S.; Qi, C. Quasi-static crushing response of a novel 3D re-entrant circular auxetic metamaterial. Compos. Struct. 2022, 300, 116066. [Google Scholar] [CrossRef]
  28. Zhu, G.; Wen, D.; Wei, L.; Wang, Z.; Zhao, X. Mechanical performances of novel cosine function cell-based metallic lattice structures under quasi-static compressive loading. Compos. Struct. 2023, 314, 116962. [Google Scholar] [CrossRef]
  29. Zhu, G.; Zhao, Z.; Cao, X.; Wei, L.; Wang, Z.; Zhao, X. On crushing behavior of square aluminum/CFRP hybrid structures subjected to quasi-static loading. Thin-Walled Struct. 2023, 184, 110486. [Google Scholar] [CrossRef]
  30. Li, P.; Yang, F.; Bian, Y.; Zhang, S.; Wang, L. Designable mechanical properties of modified body-centered cubic lattice materials. Compos. Struct. 2023, 317, 117060. [Google Scholar] [CrossRef]
  31. Yao, S.; Zhou, Y.; Li, Z.; Zhang, P.; Cao, Y.; Xu, P. Energy absorption characteristics of square frustum lattice structure. Compos. Struct. 2021, 275, 114492. [Google Scholar] [CrossRef]
  32. Ha, N.S.; Pham, T.M.; Tran, T.T.; Hao, H.; Lu, G. Mechanical properties and energy absorption of bio-inspired hierarchical circular honeycomb. Compos. B Eng. 2022, 236, 109818. [Google Scholar] [CrossRef]
  33. Tabacu, S. Axial crushing of circular structures with rectangular multi-cell insert. Thin-Walled Struct. 2015, 95, 297–309. [Google Scholar] [CrossRef]
  34. Li, Z.; Yao, S.; Ma, W.; Xu, P.; Che, Q. Energy-absorption characteristics of a circumferentially corrugated square tube with a cosine profile. Thin-Walled Struct. 2019, 135, 385–399. [Google Scholar] [CrossRef]
  35. Yang, J.; Chen, X.; Sun, Y.; Zhang, J.; Feng, C.; Wang, Y.; Wang, K.; Bai, L. Compressive properties of bidirectionally graded lattice structures. Mater. Des. 2022, 218, 110683. [Google Scholar] [CrossRef]
  36. Sharma, D.; Hiremath, S.S. Bio-inspired repeatable lattice structures for energy absorption: Experimental and finite element study. Compos. Struct. 2022, 283, 115102. [Google Scholar] [CrossRef]
  37. Jia, H.; Lei, H.; Wang, P.; Meng, J.; Li, C.; Zhou, H.; Zhang, X.; Fang, D. An experimental and numerical investigation of compressive response of designed Schwarz Primitive triply periodic minimal surface with non-uniform shell thickness. Extreme Mech. Lett. 2020, 37, 100671. [Google Scholar] [CrossRef]
  38. Rodrigo, C.; Xu, S.; Durandet, Y.; Fraser, D.; Ruan, D. Quasi-static and dynamic compression of additively manufactured functionally graded lattices: Experiments and simulations. Eng. Struct. 2023, 284, 115909. [Google Scholar] [CrossRef]
  39. Gibson, L.J.; Ashby, M.F. Cellular Solids: Structure and Properties; Cambridge University Press: Cambridge, UK, 1997. [Google Scholar]
  40. Bai, L.; Gong, C.; Chen, X.; Zheng, J.; Yang, J.; Li, K.; Sun, Y. Heterogeneous compressive responses of additively manufactured Ti-6Al-4V lattice structures by varying geometric parameters of cells. Int. J. Mech. Sci. 2022, 214, 106922. [Google Scholar] [CrossRef]
  41. Li, C.; Lei, H.; Zhang, Z.; Zhang, X.; Zhou, H.; Wang, P.; Fang, D. Architecture design of periodic truss-lattice cells for additive manufacturing. Addit. Manuf. 2020, 34, 101172. [Google Scholar] [CrossRef]
  42. Beharic, A.; Rodriguez Egui, R.; Yang, L. Drop-weight impact characteristics of additively manufactured sandwich structures with different cellular designs. Mater. Des. 2018, 145, 122–134. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the BCC structure under axial compression. (a) The BCC lattice structure; (b) analysis module; (c) simplified analysis of the force on the struts.
Figure 1. Schematic diagram of the BCC structure under axial compression. (a) The BCC lattice structure; (b) analysis module; (c) simplified analysis of the force on the struts.
Aerospace 11 00004 g001
Figure 2. Schematic diagram of the geometric dimensions of the strut.
Figure 2. Schematic diagram of the geometric dimensions of the strut.
Aerospace 11 00004 g002
Figure 3. Fabricated lattice structure specimens.
Figure 3. Fabricated lattice structure specimens.
Aerospace 11 00004 g003
Figure 4. The quasi-static tests. (a) Instron testing machine; (b) stress–strain curves of the additive manufacturing 316L stainless steel specimens.
Figure 4. The quasi-static tests. (a) Instron testing machine; (b) stress–strain curves of the additive manufacturing 316L stainless steel specimens.
Aerospace 11 00004 g004
Figure 5. Numerical simulation. (a) Finite element model setup; (b) mesh convergence test.
Figure 5. Numerical simulation. (a) Finite element model setup; (b) mesh convergence test.
Aerospace 11 00004 g005
Figure 6. Comparison of relative density of specimens, design models, and theoretical values.
Figure 6. Comparison of relative density of specimens, design models, and theoretical values.
Aerospace 11 00004 g006
Figure 7. SEM images of ES-BCC structures. (ac) A090; (df) A120.
Figure 7. SEM images of ES-BCC structures. (ac) A090; (df) A120.
Aerospace 11 00004 g007
Figure 8. Compressive stress–strain curves: A090 and BCC lattice structures.
Figure 8. Compressive stress–strain curves: A090 and BCC lattice structures.
Aerospace 11 00004 g008
Figure 9. The compression deformation process of the lattice structures.
Figure 9. The compression deformation process of the lattice structures.
Aerospace 11 00004 g009
Figure 10. Stress-strain curve of 316L stainless steel material.
Figure 10. Stress-strain curve of 316L stainless steel material.
Aerospace 11 00004 g010
Figure 11. The stress–strain curves of ES-BCC structures with different ϕ .
Figure 11. The stress–strain curves of ES-BCC structures with different ϕ .
Aerospace 11 00004 g011
Figure 12. The mechanical properties of ES-BCC structures with different ϕ : (a) relative elastic modulus; (b) relative yield strength; (c) specific energy absorption.
Figure 12. The mechanical properties of ES-BCC structures with different ϕ : (a) relative elastic modulus; (b) relative yield strength; (c) specific energy absorption.
Aerospace 11 00004 g012
Figure 13. Mechanical properties–relative density fitting curves of lattice structures. (a) E/Es; (b) σy/σs.
Figure 13. Mechanical properties–relative density fitting curves of lattice structures. (a) E/Es; (b) σy/σs.
Aerospace 11 00004 g013
Figure 14. The force schematic of the 1/2 variable-section strut.
Figure 14. The force schematic of the 1/2 variable-section strut.
Aerospace 11 00004 g014
Figure 15. Comparison of the mechanical properties of ES-BCC structures. (a) E; (b) σ y .
Figure 15. Comparison of the mechanical properties of ES-BCC structures. (a) E; (b) σ y .
Aerospace 11 00004 g015
Table 1. The ES-BCC unit cell structure with d0 = 0.3 mm.
Table 1. The ES-BCC unit cell structure with d0 = 0.3 mm.
ϕ θ d (x)Strut ProfileUnit Cell Model
75°27.37° d 3 x 0.09 d x = 18 2 25 3 1 x Aerospace 11 00004 i001Aerospace 11 00004 i002
90°35.26° d 3 x 0.09 d x = 18 2 25 x Aerospace 11 00004 i003Aerospace 11 00004 i004
105°42.66° d 3 x 0.09 d x = 18 2 25 2 + 6 2 3 Aerospace 11 00004 i005Aerospace 11 00004 i006
120°50.77° d 3 x 0.09 d x = 18 6 75 x Aerospace 11 00004 i007Aerospace 11 00004 i008
Table 2. 316L stainless steel’s chemical composition (%) [24].
Table 2. 316L stainless steel’s chemical composition (%) [24].
CSiMnSPCrNiMoFe
0~0.030~1.00~2.00~0.030~0.04516.5~18.510.0~13.02.0~3.0Balance
Table 3. The mechanical property parameters of 316L stainless steel.
Table 3. The mechanical property parameters of 316L stainless steel.
Elastic Modulus (GPa)Density (g/cm3)Poisson’s RatioYield Stress (MPa)Ultimate Stress (MPa)
677.590.3443.63551.63
Table 4. Structural parameters of lattice structures.
Table 4. Structural parameters of lattice structures.
Lattice TypeBCCA075A090A105A120
Relative densityExp.0.12650.17570.13070.10230.0841
Sim.0.13440.18420.13440.10700.0880
Densification strainExp.0.62880.44920.48480.53560.5864
Sim.0.60750.43270.51420.52830.5708
Table 5. The elastic modulus, yield strength, and energy absorption of lattice structures.
Table 5. The elastic modulus, yield strength, and energy absorption of lattice structures.
Lattice TypeElastic Modulus (MPa)Yield Strength (MPa)Energy Absorption (J/cm3)
Exp.Sim.DEVESExp.Sim.DEVESExp.Sim.DEVES
A090373.58345.46−7.53%6.696.61−1.20%4.684.802.56%
BCC238.34213.51−10.42%4.284.300.47%4.214.291.90%
Note: Deviation DEVES = (Sim. − Exp.)/Exp. × 100%.
Table 6. The fitting parameters of the GA model for the mechanical properties of ES-BCC structures.
Table 6. The fitting parameters of the GA model for the mechanical properties of ES-BCC structures.
Lattice TypeNormalized Elastic ModulusNormalized Yield Strength
C1n1R2C2n2R2
A0750.162.000.9990.321.870.996
A0900.332.060.9860.581.930.999
A1050.502.080.9991.012.030.999
A1200.682.030.9981.231.940.999
Table 7. Mechanical properties of ES-BCC structures: experimental, simulated, and theoretical.
Table 7. Mechanical properties of ES-BCC structures: experimental, simulated, and theoretical.
Lattice TypeElastic Modulus (MPa)Yield Strength (MPa)
Exp.Sim.Theor.DEVESDEVETExp.Sim.Theor.DEVESDEVET
A075384.84360.05342.66−6.44%−10.96%7.757.437.17−4.13%−7.48%
A090373.58345.46340.66−7.52%−8.81%6.696.616.47−1.20%−3.29%
A105285.85319.38339.9511.73%18.93%5.195.935.8814.26%13.29%
A120306.21323.60340.095.68%11.06%5.556.075.399.37%−2.88%
Note: DEVES = (Sim. − Exp.)/Exp. × 100%; DEVET = (Theor. − Exp.)/Exp. × 100%.
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

Liu, Z.; Zhao, R.; Tao, C.; Wang, Y.; Liang, X. Mechanical Performance of a Node-Reinforced Body-Centered Cubic Lattice Structure: An Equal-Strength Concept Design. Aerospace 2024, 11, 4. https://doi.org/10.3390/aerospace11010004

AMA Style

Liu Z, Zhao R, Tao C, Wang Y, Liang X. Mechanical Performance of a Node-Reinforced Body-Centered Cubic Lattice Structure: An Equal-Strength Concept Design. Aerospace. 2024; 11(1):4. https://doi.org/10.3390/aerospace11010004

Chicago/Turabian Style

Liu, Zeliang, Rui Zhao, Chenglin Tao, Yuan Wang, and Xi Liang. 2024. "Mechanical Performance of a Node-Reinforced Body-Centered Cubic Lattice Structure: An Equal-Strength Concept Design" Aerospace 11, no. 1: 4. https://doi.org/10.3390/aerospace11010004

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