ArticlePDF Available

Abstract and Figures

The unique, hierarchical patterns of leaf veins have attracted extensive attention in recent years. However, it remains unclear how biological and mechanical factors influence the topology of leaf veins. In this paper, we investigate the optimization mechanisms of leaf veins through a combination of experimental measurements and numerical simulations. The topological details of three types of representative plant leaves are measured. The experimental results show that the vein patterns are insensitive to leaf shapes and curvature. The numbers of secondary veins are independent of the length of the main vein, and the total length of veins increases linearly with the leaf perimeter. By integrating biomechanical mechanisms into the topology optimization process, a transdisciplinary computational method is developed to optimize leaf structures. The numerical results show that improving the efficiency of nutrient transport plays a critical role in the morphogenesis of leaf veins. Contrary to the popular belief in the literature, this study shows that the structural performance is not a key factor in determining the venation patterns. The findings provide a deep understanding of the optimization mechanism of leaf veins, which is useful for the design of high-performance shell structures.
Content may be subject to copyright.
1
Article published in
Journal of the Mechanical Behavior of Biomedical Materials, Vol. 123, No.104788, 2021
https://doi.org/10.1016/j.jmbbm.2021.104788
Topology of leaf veins: Experimental observation and computational
morphogenesis
Jiaming Ma a, Zi-Long Zhao a, Sen Lin b, Yi Min Xie a,*
a Centre for Innovative Structures and Materials, School of Engineering, RMIT University,
Melbourne 3001, Australia
b State Key Laboratory of Advanced Design and Manufacturing for Vehicle Body, College of
Mechanical and Vehicle Engineering, Hunan University, Changsha 410082, China
ABSTRACT
The unique, hierarchical patterns of leaf veins have attracted extensive attention in recent
years. However, it remains unclear how biological and mechanical factors influence the
topology of leaf veins. In this paper, we investigate the optimization mechanisms of leaf veins
through a combination of experimental measurements and numerical simulations. The
topological details of three types of representative plant leaves are measured. The experimental
results show that the vein patterns are insensitive to leaf shapes and curvature. The numbers of
secondary veins are independent of the length of the main vein, and the total length of veins
increases linearly with the leaf perimeter. By integrating biomechanical mechanisms into the
topology optimization process, a transdisciplinary computational method is developed to
optimize leaf structures. The numerical results show that improving the efficiency of nutrient
transport plays a critical role in the morphogenesis of leaf veins. Contrary to the popular belief
in the literature, this study shows that the structural performance is not a key factor in
* Corresponding author. Tel.: +61 399253655
Email address: mike.xie@rmit.edu.au (Y.M. Xie).
2
determining the venation patterns. The findings provide a deep understanding of the
optimization mechanism of leaf veins, which is useful for the design of high-performance shell
structures.
Keywords: Leaf veins; Topology; Computational morphogenesis; Nutrient transport; Structural
stiffness; Curved shell
3
1. Introduction
Through a long history of evolution, biological materials have formed hierarchical
structures which are closely related to their functions (Allaire et al., 2002; Meyers et al., 2008;
Wang et al., 2016; Wright et al., 2004; Zhao et al., 2015, 2016). The fractal patterns in nature
such as the river networks, snowflakes, root systems, and venation systems, have attracted
considerable attention over the past few decades (Ball, 2009, 2016; Wolfram, 2002). Leaf veins
render gas exchange and fluid/nutrition transport, such that the mesophyll cells can achieve
sufficient supplies for living (Carvalho et al., 2017; Sack and Scoffoni, 2013). In return, the
photosynthesis from mesophyll provides organics and oxygen to support the plantslives and
flourishment (Dengler and Tsukaya, 2001; Efroni et al., 2008; Scarpella et al., 2010). To
maximize the sun lighted area for photosynthesis, a leaf surface requires sufficient stiffness to
support its expansion against weight and wind loads (Brodribb et al., 2007). Veins are stiffer
than the mesophyll, and they make a significant contribution to reinforce the mesophyll and
maintain the leaf shape (Ennos et al., 2000; Gibson et al., 1988; Niklas, 1999). The
biomechanical mechanisms underlying the beautiful vein patterns have attracted extensive
interests of scientists, engineers, and designers (Blonder et al., 2011; Gokmen, 2013; Md Rian
and Sassone, 2014; Runions et al., 2005).
Several numerical approaches have been established to investigate the branch-like vein
structures. The L-system method has been used to analyze the growth process of plants
(Prusinkiewicz and Lindenmayer, 1990). However, this mathematical method only models a
vein system morphologically, without considering the biomechanical requirements such as
nutrient transport and shape maintaining. An adaptive algorithm has been developed for vein
morphogenesis by generating stiffeners in leaf model in random directions (Liu et al., 2017).
Topology optimization has now become a powerful tool to explore the morphogenesis of
natural materials. In the past three decades, several optimization techniques have been
successfully established, including the solid isotropic material with penalization (SIMP)
method (Bendsøe, 1989; Bendsøe, 1995; Sigmund and Maute, 2013), the level-set method
(Allaire et al., 2002; Wang et al., 2003), the evolutionary structural optimization (ESO) method
(Xie and Steven, 1993; Xie and Steven, 1997) and the bi-directional evolutionary structural
optimization (BESO) method (Huang and Xie, 2007, 2009). Most recently, imposing
complicated constraints during the form-finding process has been realized (Chen et al., 2020;
He et al., 2020; Xiong et al., 2020; Zhao et al., 2020a). By establishing transdisciplinary
computational methods for biomechanical morphogenesis, Zhao et al. (2018, 2020b, 2020c)
4
have revealed the optimization mechanisms of, e.g., plant leaves and animal stingers. Using
topology optimization, a golden ratio distribution rule is found in venation systems (Sun et al.,
2018). A multi-objective optimization approach is developed to explore the vein patterns on a
planar plate (Lin et al., 2020). However, there is still a lack of quantitative study on the
biomechanical mechanisms of vein distributions on the shell-like mesophyll.
In this work, we investigate, both experimentally and numerically, the topology of leaf
veins. The vein patterns of three representative plant leaves are measured, including the Syringa
vulgaris L., Rosa chinensis Jacq., and Cotoneaster submultiflorus Popov. Biomechanical
functions of veins such as nutrient transport and structural stiffness are integrated into the
computational morphogenesis. The influence of the curved shape of mesophyll on the vein
distribution is examined. This study reveals the optimization mechanisms underlying the
intriguing overall layout of venation systems. The presented methodology can be applied in the
design of high-performance shell structures such as aircraft skin ribs (Song et al., 2021).
2. Methodology
2.1. Experiments
Three types of fresh plant leaves with distinctly different shapes are collected in the same
area, including Syringa vulgaris L., Rosa chinensis Jacq., and Cotoneaster submultiflorus
Popov. Five samples of each species are randomly selected. Those samples have various
surface curvatures and their main veins have different angles of inclination. Images of the
samples are taken by a high-resolution digital camera immediately after collection. The
samples are flattened and put between a piece of A4-sized paper and a transparent thin film.
Rhinoceros (Rhino 6 SR28 version) was used for feature extractions and measurements.
2.2. Computational morphogenesis
An interdisciplinary topology optimization method is developed for investigating the
biomechanical morphogenesis of plant leaves. This method is capable of dealing with different
objective functions, e.g., the stiffness maximization and the enhancement of nutrition transport.
Veins have continuously varying thickness and material properties over the leaf surface. The
new method is developed under the framework of the SIMP technique which can produce
5
results with transitional structural features (Bendsøe, 1989; Bendsøe, 1995; Sigmund and
Maute, 2013).
Denote the design domain of a leaf as Ω, which is divided into n elements for finite
element analysis (FEA). For each element i (
1, 2, 3, ...,in=
), it has a density value
i
. The
densities of vein and mesophyll elements are defined as
1
i
=
and
0
i
=
. In the beginning,
the densities for all elements are set to the target volume fraction
f
ˆ
v
. The FEA and density
update are carried out in each iteration during the optimization process. Elements with lower
sensitivities are gradually changed towards 1, while those with higher sensitivities are gradually
changed towards 0.
In the FEA, the curved leaves under self-weight is considered as shells under vertical
uniform surface traction. The leaf material is assumed to be linearly elastic. The governing
equation for maximizing the structural stiffness is expressed as below (Bendsøe and Sigmund,
2004a)
( )
dd
1
min: 2
C
=T
U K U
ρ
(1)
d
=K U F
f
1
1
ˆ
n
ii
in
i
i
v
vv
=
=
=
0 1, 1, 2, ,
iin
  =
(2)
where 𝑣𝑖 denotes the volume of element i. F, U and
d
K
are the uniformly distributed surface
traction, global displacement, and global stiffness matrix, respectively. Assume that the
Poisson’s ratios of the vein and mesophyll are the same. The Young’s modulus between two
neighboring phases is interpolated as (Bendsøe and Sigmund, 1999)
( ) ( )
m v m p
i
E E E E

= +
(3)
where p is a penalization number (default as 3) (Bendsøe and Sigmund, 2004b), and
m
E
and
v
E
are the Young’s moduli of the mesophyll and vein, respectively. Thus, the objective
function can be rewritten as
( )
v
mm
dd
1vv
1
min: 1
2
np
i i i
i
EE
CEE

=


= + −





u k u
ρ
(4)
6
where
ui
and
v
d
k
represent respectively the displacement vector and the stiffness matrix of
element i with E =
v
E
. The sensitivity is derived by the adjoint method as
( ) ( )
d1v
m
dd
v
C1
1
2p
i i i
i
E
pE
 

= = −


u k u
(5)
To maximize the nutrient transport performance of a leaf, it can be considered as a
dissipation maximization problem. The objective function of minimizing the concentration
gradients of nutrient is written as
( )
tt
1
min: 2
C
=T
T K T
ρ
(6)
𝑠. 𝑡. ∶
t
=K T Q
f
1
1
ˆ
n
ii
in
i
i
v
vv
=
=
=
0 1, 1, 2, ,
iin
  =
(7)
where Q, T, and
t
K
represent nutrient input, nutrient solution concentration, and global
conductivity matrix, respectively. The conductivity 𝜅 of each element is interpolated as
( ) ( )
m v m p
i
 
= +
(8)
The sensitivity is derived by the adjoint method as
( ) ( )
t1 T v
mt
v
11 t t
2p
t i i i
i
Cp
 


= = −


k
(9)
A weighting factor is introduced to control the effect of the two sensitivities on the final
design. The coupled sensitivity is calculated as
(1 )
i id it
  
= + −
(10)
where
( )
max id
id id
i
=
(11)
( )
max it
it it
i
=
(12)
The optimality criteria based optimizer is adopted for updating the design variables in
each iteration (Sigmund, 2001). The algorithm is implemented in the Python environment and
linked to Abaqus (Abaqus 6.20 version) for FEA. The finite element model of each leaf is
discretized into approximately 50,000 four-node S4R doubly curved shell elements. The
7
Poisson’s ratios of both veins and mesophylls are set as 0.3, and their Young’s moduli and
thickness are ascertained according to the previous study (Sun et al., 2018). For the force
displacement analysis, the surface of the leaf is subjected to a uniformly distributed force in
the downward direction, and the petiole is fixed. For the steady conduction analysis, the surface
of the leaf is applied with uniformly distributed flux input, and the petiole is set to be zero
concentration. Adiabatic condition is applied on the leaf edges.
3. Results
Experimental and numerical results are presented and quantitatively analyzed in this
section. The geometries of real leaves are measured, including the leaf perimeter, and the
lengths, angles, and numbers of the main/secondary veins. The venation layouts optimized for
nutrition transport and stiffness maximization are presented, respectively. The experimental
and numerical results are compared in Section 3.2.3.
3.1. Experimental results
Three representative plant species, i.e., the Cotoneaster submultiflorus Popov, Rosa
chinensis Jacq., and Syringa vulgaris L., are selected for experiments. Their leaves have
distinctly different shapes, as shown in Figure 1. The leaf of Cotoneaster submultiflorus Popov
has a relative higher slenderness ratio with smooth edges, the leaf of Rosa chinensis Jacq. has
an approximately round shape with serrated edges, and the leaf of Syringa vulgaris L. is drop-
shaped.
8
Figure 1. Three leaf species used in experiments: (a) Cotoneaster submultiflorus Popov, (b)
Rosa chinensis Jacq., and (c) Syringa vulgaris L.
Five samples of each species are measured for experiments. The samples for the three
plants are in different growth stages and hence have different sizes. Their surface curvatures
and inclination angles (Niklas, 1999) are also different, and therefore in the forcedisplacement
analysis, the loading conditions will be different. In most cases, the gravity direction is not
perpendicular to the leaf surfaces.
To conduct quantitative analysis on the vein patterns, the leaf samples are characterized
by their lengths and angles. The perimeters and main vein lengths are also measured. All the
three plant species are dicotyledons and have camptodromous pinnate vein. Accordingly, the
length measurement for the secondary veins starts from their growing point and ends at their
intersection with upper neighboring ones. For illustration, the secondary veins of a Rosa
chinensis Jacq. leaf are highlighted in Figure 2(b), and the method for measuring the lengths
and angles is shown in Figure 2(c).
9
Figure 2. Measurements of the length and angle of secondary veins: (a) a leaf sample, (b) the
main vein (the blue line in the middle) and the secondary veins (separated in curved blue and
red lines), and (c) the method for measuring the lengths and angles.
The angles between the secondary and the main veins are measured. The rear-end of a
camptodromous pinnate vein bends and connects to a neighboring vein. Only the base segments
of these secondary veins are used to fit the straight lines for angle measurement. A base
segment starts from the main vein and ends at the intersection with its neighboring veins. The
base segments of the secondary veins are highlighted in blue in Figure 2(b). The method for
measuring the lengths and angles is shown in Figure 2(c).
Figure 3 shows the secondary vein lengths of the three plant species. Both left and right
secondary veins are observed and measured. The secondary vein lengths of each sample and
the position of each secondary vein are normalized by the main vein length. The normalized
length for each secondary vein is calculated as its measured length divided by the main vein
length. The relative position of each secondary vein is calculated as the distance from the root
of a secondary vein to the petiole of the main vein divided by the main vein length. It can be
seen that different species have different length distributions of secondary veins. For the same
species, the discrepancies of the normalized lengths of secondary veins at similar relative
positions are small.
10
Figure 3. Measurements of the positionlength relations of the secondary veins: (a)
Cotoneaster submultiflorus Popov, (b) Rosa chinensis Jacq., and (c) Syringa vulgaris L.
Figure 4 shows the angles of both left and right secondary veins. Similarly, different
species have different angle distributions of secondary veins, which for the same species,
samples have similar angle distributions.
Figure 4. Measurements of the positionangle relations of the secondary veins: (a) Cotoneaster
submultiflorus Popov, (b) Rosa chinensis Jacq., and (c) Syringa vulgaris L.
Apart from the length and angle of the secondary veins, their numbers are measured and
plotted in Figure 5(a). The result shows that, for each leaf species, the numbers of the secondary
veins are almost the same, although the five samples have significantly different main vein
lengths.
11
Figure 5(b) shows the relation between the leaf perimeter and the total length of the main
and the secondary veins. For each plant, the total length of veins increases linearly with the
increasing leaf perimeter. It is found that the three straight fitting lines have similar slopes. For
the three species, the ratios of the total length of veins to the leaf perimeter are 2.17, 2.63, and
2.93, respectively. The biomechanical mechanisms underlying the similar ratios require further
research.
Figure 5. Comparison of vein features across different species and samples: (a) number of the
secondary veins against the main vein length, (b) variation of the total length of veins with
respect to the leaf perimeter.
3.2. Numerical results
Topology optimizations are performed to investigate the biomechanical mechanisms of
the distribution of leaf veins. Both flat and curved leaf models are considered. Two
biomechanical functions are considered, including the stiffness maximization and the
dissipation optimization (corresponding to nutrient transport performance). A representative
leaf profile is chosen for the leaf model. Curvatures for the leaf models are determined based
on typical shapes of real leaves.
Figure 6 shows four leaf models for topology optimization. The model in Figure 6(a) is
flat. The model in Figure 6(b) is folded along its middle axis. The models in Figure 6(c) and
(d) are folded perpendicular to their middle axes, where the one in Figure 6(d) has ruffles on
its edges. All leaf models have the same surface area and the objective functions of the four
cases can be compared.
12
Figure 6. Four leaf models for topology optimization: (a) a flat model, (b) a model folded along
its middle axis, (c) a model folded perpendicular to its middle axis, and (d) a model folded
perpendicular to its middle axis and with ruffles on its edges.
3.2.1. Nutrient transport
The optimized topologies of the leaves for maximizing the efficiency of nutrient transport
are shown in Figure 7. Here and in the following, the vein material with the highest conductivity
is highlighted in light-yellow, and the mesophyll materials are marked in dark green.
Intermediate colors on the models indicate the transitional change of material properties
between the two phases. Figure 7 shows that the veins have hierarchical and fractal patterns,
regardless of the curvatures of the leaf models. The optimized results are very similar to the
real leaves. The quantitative comparison between the numerical and the experimental results
are presented in Section 3.2.3. It can be seen that veins with the highest conductivity are located
near the petiole. The secondary and tertiary veins have lower conductivity and thickness. These
results are in good agreement with real leaves.
13
Figure 7. Optimized topologies for nutrient transport maximization: (a) a flat leaf, (b) a leaf
folded along the middle axis, (c) a leaf folded perpendicular to the middle axis, and (d) a leaf
folded perpendicular to the middle axis and with wrinkled edges.
3.2.2. Structural stiffness
The optimized topologies of the leaves for stiffness maximization are shown in Figure 8.
Similarly, the light-yellow color indicates the highest stiffness vein material, and the dark green
one reveals where the soft mesophylls are. The obtained topologies depend strongly on the
curvature of the models. This phenomenon is on the contrary to that of the nutrition
transportation optimization results. Although the flat model results in a branch-like pattern of
veins, the curved models result in truss-like patterns. Such difference could be attributed to the
membrane stiffness (i.e., the in-plane stiffness) of shell elements. For a curved shell, the
external load induces in-plane stresses, while for a flat plate, there exists only bending and
shear stresses. The structural compliance of the structures in Figure 7(a-d) are 7.21 N·mm, 5.22
N·mm, 7.59 N·mm and 6.90 N·mm, respectively. The compliances of the structures in Figure
8(a-d) are significantly smaller, which are 3.95 N·mm, 1.72 N·mm, 4.36 N·mm and 3.65 N·mm,
respectively. In Figure 8, veins near the petioles are very thick, which contribute to supporting
the leaf structure.
14
Figure 8. Optimized topologies for stiffness maximization: (a) a flat leaf, (b) a leaf folded along
the middle axis, (c) a leaf folded perpendicular to the middle axis, and (d) a leaf folded
perpendicular to the middle axis and with wrinkled edges.
3.2.3. Quantitative analysis
The numerical and experimental results are compared quantitatively. For the optimization
results of nutrition transportation, the skeleton lines of the main and the secondary veins are
extracted and shown in Figure 9. The lengths of the secondary veins are measured and
normalized by the length of the main vein. Figure 10(a) shows the positionlength relations of
the secondary veins.
Figure 9. Skeleton (red lines) of the main vein and the secondary veins.
15
Figure 10. (a) Positionlength and (b) positionangle relations of the secondary veins
optimized for nutrient transport.
Figure 10 (b) shows the positionangle relations of the secondary veins optimized for
nutrient transport. Despite the shape difference of the four leaf models, their veins exhibit
similar positionlength relations and positionangle relations.
The topologies optimized for stiffness maximization are distinctly different from the
patterns of real veins, which are here not quantitatively analyzed.
To investigate the influence of the main vein length on the numbers of secondary veins
through topology optimization, the leaf models should have different sizes but the same shape.
The element sizes need to be scaled to keep the element number and layout unchanged. Since
the optimized results will be unchanged if the model and its elements are enlarged
simultaneously, the computational method is not used to check the relationship between main
vein length and the numbers of secondary veins.
4. Discussion
It is seen that different leaf species exhibit different feature sizes, edge profiles, and
material properties. However, the topologies of their veins are similar. Our experimental results
reveal that the number of secondary veins is independent of the length of the main vein. The
total length of veins increases linearly with the leaf perimeter. Although the leaves have
distinctly different shapes, curvatures, inclination angles, and living environments, the
topological differences of their veins are limited.
16
From the results in Section 3.2, the vein patterns optimized for nutrition transport are
insensitive to the leaf shapes, while the patterns optimized for stiffness depend strongly on the
leaf curvature. For stiffness optimization, truss-like veins are obtained in curved models, while
feather-like veins are generated only in the flat model. It is well known that real leaves could
have various curved shapes and inclination angles (loading conditions). Therefore, the
structural stiffness is less likely to play a dominated role in determining the topology of leaf
veins.
Figure 11. Optimization results from coupling stiffness and dissipation: (a) 𝜆Stiffness = 0.0
and 𝜆Dissipation = 1.0 , (b) 𝜆Stiffness = 0.2 and 𝜆Dissipation = 0.8 , (c) 𝜆Stiffness = 0.6 and
𝜆Dissipation = 0.4, (d) 𝜆Stiffness = 1.0 and 𝜆Dissipation = 0.0.
Veins, consisting of parenchyma, sclerenchyma, and sheath, usually have higher Young’s
modulus than the mesophyll (Gibson et al., 1988; Sun et al., 2018). Vein is usually thicker than
mesophyll. Therefore, veins can be considered as reinforcing ribs of leaves. However,
according to our results in Section 3.2.2, the stiffness enhancement is not a dominated factor
that determines the vein topologies. Figure 11 shows the coupled optimization results of vein
morphologies on a curved leaf model with different weight factors, combining stiffness
enhancement with nutrient transport enhancement. It can be seen that the nutrient transport,
which is of significant importance for photosynthesis, governs the topology of veins.
This study reveals that the enhancement of nutrient transport plays a predominant role in
determining the vein patterns. The main features of the topologies optimized for nutrient
transport agree well with the real patterns. Some other factors such as biological constraints,
17
the intercellular stress during growth, and the microstructures of plant tissues, can also affect
the topology of veins, which require further research.
5. Conclusion
In this study, intricate topologies of leaf veins have been investigated through both
experimental observation and computational morphogenesis. It is found that the enhancement
of nutrient transport plays a predominant role in determining the form of venation patterns.
Contrary to popular belief in the literature, this research reveals that the structural performance
is not the key factor of leaf vein patterns. Furthermore, the experimental measurements show
that the numbers of secondary veins are independent of the length of main veins, and the total
length of veins has a linear relationship with the leaf perimeter. The numerical results of
nutrient transport show that the vein patterns are insensitive to the variation of leaf shapes. This
research provides a deep understanding of the biomechanical mechanisms underlying the
intriguing layout of leaf veins. The presented computational method can be used for designing
efficient and innovative free-form shell structures.
Acknowledgement
The authors received financial support from the Australian Research Council
(FL190100014 and DE200100887).
Conflict of interest
The authors declare that they have no conflict of interest.
18
Reference
Allaire, G., Jouve, F., Toader, A. M., 2002. A level-set method for shape optimization. Comptes rendus.
Mathématique 334, 1125-1130.
Ball, P., 2009. Nature's Patterns: A Tapestry in Three Parts. Oxford University Press, New York.
Ball, P., 2016. Patterns In Nature: Why The Natural World Looks The Way It Does. The University of
Chicago Press, Chicago.
Bendsøe, M.P., 1989. Optimal shape design as a material distribution problem. Struct. Multidiscip.
Optim. 1, 193-202.
Bendsøe, M.P., 1995. Optimization of Structural Topology, Shape, and Material. Springer-Verlag
Berlin Heidelberg, Berlin.
Bendsøe, M.P., Sigmund, O., 1999. Material interpolation schemes in topology optimization. Arch.
Appl. Mech. (1991) 69, 635-654.
Bendsøe, M.P., Sigmund, O., 2004a. Topology optimization by distribution of isotropic material,
Topology Optimization: Theory, Methods, And Applications. Springer-Verlag Berlin Heidelberg,
Berlin, pp. 1-69.
Bendsøe, M.P., Sigmund, O., 2004b. Topology Optimization: Theory, Methods, and Applications,
Second Edition, Corrected Printing. ed.: Springer-Verlag Berlin Heidelberg, Berlin.
Blonder, B., Violle, C., Bentley, L.P., Enquist, B.J., 2011. Venation networks and the origin of the leaf
economics spectrum. Ecol. Lett. 14, 91-100.
Brodribb, T.J., Feild, T.S., Jordan, G.J., 2007. Leaf maximum photosynthetic rate and venation are
linked by hydraulics. Plant Physiol. 144, 1890-1898.
Carvalho, M.R., Turgeon, R., Owens, T., Niklas, K.J., 2017. The scaling of the hydraulic architecture
in poplar leaves. New Phytol. 214, 145-157.
Chen, A.B., Cai, K., Zhao, Z.L., Zhou, Y.Y., Xia, L., Xie, Y.M., 2021. Controlling the maximum first
principal stress in topology optimization. Struct. Multidiscip. Optim. 63, 327-339
Dengler, N.G., Tsukaya, H., 2001. Leaf morphogenesis in dicotyledons: Current issues. Int. J. Plant Sci.
162, 459-464.
Efroni, I., Blum, E., Goldshmidt, A., Eshed, Y., 2008. A protracted and dynamic maturation schedule
underlies arabidopsis leaf development. Plant Cell 20, 2293-2306.
Ennos, A.R., Spatz, H.C., Speck, T., 2000. The functional morphology of the petioles of the banana,
Musa textilis. J. Exp. Bot. 51, 2085-2093.
Gibson, L.J., Ashby, M.F., Easterling, K.E., 1988. Structure and mechanics of the iris leaf. J. Mater.
Sci. 23, 3041-3048.
Gokmen, S., 2013. A morphogenetic approach for performative building envelope systems using leaf
venetian patterns. eCAADe 2013: Computation and Performance, Proceedings of the 31st
International Conference on Education and research in Computer Aided Architectural Design in
Europe, Delft University of Technology 1, 497-506.
He, Y.Z., Cai, K., Zhao, Z.L., Xie, Y.M., 2020. Stochastic approaches to generating diverse and
competitive structural designs in topology optimization. Finite Elem. Anal. Des. 173, 103399.
Huang, X., Xie, Y.M., 2007. Convergent and mesh-independent solutions for the bi-directional
evolutionary structural optimization method. Finite Elem. Anal. Des. 43, 1039-1049.
Huang, X., Xie, Y.M., 2009. Bi-directional evolutionary topology optimization of continuum structures
with one or multiple materials. Comput. Mech. 43, 393-401.
Lin, S., Chen, L., Zhang, M., Xie, Y.M., Huang, X.D., Zhou, S.W., 2020. On the interaction of
biological and mechanical factors in leaf vein formation. Adv. Eng. Softw. 149, 1-8.
Liu, H.L., Li, B.T., Yang, Z.H., Hong, J., 2017. Topology optimization of stiffened plate/shell structures
based on adaptive morphogenesis algorithm. J. Manuf. Syst. 43, 375-384.
Md Rian, I., Sassone, M., 2014. Tree-inspired dendriforms and fractal-like branching structures in
architecture: A brief historical overview. Front. Archit. Res. 3, 298-323.
Meyers, M.A., Chen, P.Y., Lin, A.Y.M., Seki, Y., 2008. Biological materials: structure and mechanical
properties. Prog. Mater. Sci. 53, 1-206.
Niklas, K.J., 1999. A mechanical perspective on foliage leaf form and function. New Phytol. 143, 19-
31.
19
Prusinkiewicz, P., Lindenmayer, A., 1990. The Algorithmic Beauty of Plants. Springer-Verlag, New
York.
Runions, A., Fuhrer, M., Lane, B., Federl, P., Rolland-Lagan, A.G., Prusinkiewicz, P., 2005. Modeling
and visualization of leaf venation patterns. Acm. T. Graphic. 24, 702-711.
Sack, L., Scoffoni, C., 2013. Leaf venation: structure, function, development, evolution, ecology and
applications in the past, present and future. New Phytol. 198, 983-1000.
Scarpella, E., Barkoulas, M., Tsiantis, M., 2010. Control of leaf and vein development by auxin. Cold
Spring Harb. Perspect. Biol. 2, a001511.
Sigmund, O., 2001. A 99 line topology optimization code written in Matlab. Struct. Multidiscip. Optim.
21, 120-127.
Sigmund, O., Maute, K., 2013. Topology optimization approaches: A comparative review. Struct.
Multidiscip. Optim. 48, 1031-1055.
Song, L., Gao, T., Tang, L., Du, X., Zhu, J., Lin, Y., Shi, G., Liu, H., Zhou, G., Zhang, W., 2021. An
all-movable rudder designed by thermo-elastic topology optimization and manufactured by
additive manufacturing. Comput. Struct. 243, 106405.
Sun, Z., Cui, T.C., Zhu, Y.C., Zhang, W.S., Shi, S.S., Tang, S., Du, Z.L., Liu, C., Cui, R.H., Chen, H.J.,
Guo, X., 2018. The mechanical principles behind the golden ratio distribution of veins in plant
leaves. Sci. Rep. 8, 13859.
Wang, M.Y., Wang, X.M., Guo, D.M., 2003. A level set method for structural topology optimization.
Comput. Methods in Appl. Mech. Eng. 192, 227-246.
Wang, X.J., Xu, S.Q., Zhou, S.W., Xu, W., Leary, M., Choong, P., Qian, M., Brandt, M., Xie, Y.M.,
2016. Topological design and additive manufacturing of porous metals for bone scaffolds and
orthopaedic implants: A review. Biomaterials 83, 127-141.
Wolfram, S., 2002. A New Kind of Science. Wolfram Media, Champaign, Illinois.
Wright, I.J., Reich, P.B., Westoby, M., Ackerly, D.D., Baruch, Z., Bongers, F., Cavender-Bares, J.,
Chapin, T., Cornelissen, J.H.C., Diemer, M., Flexas, J., Garnier, E., Groom, P.K., Gulias, J.,
Hikosaka, K., Lamont, B.B., Lee, T., Lee, W., Lusk, C., Midgley, J.J., Navas, M.L., Niinemets,
U., Oleksyn, J., Osada, N., Poorter, H., Poot, P., Prior, L., Pyankov, V.I., Roumet, C., Thomas,
S.C., Tjoelker, M.G., Veneklaas, E.J., Villar, R., 2004. The worldwide leaf economics spectrum.
Nature 428, 821-827.
Xie, Y.M., Steven, G.P., 1993. A simple evolutionary procedure for structural optimization. Comput.
Struct. 49, 885-896.
Xie, Y.M., Steven, G.P., 1997. Evolutionary Structural Optimization, Springer, London.
Xiong, Y.L., Yao, S., Zhao, Z.L., Xie, Y.M., 2020. A new approach to eliminating enclosed voids in
topology optimization for additive manufacturing. Addit. Manuf. 32, 101006.
Yang, K., Zhao, Z.L., He, Y.Z., Zhou, S.W., Zhou, Q., Huang, W.X., Xie, Y.M., 2019. Simple and
effective strategies for achieving diverse and competitive structural designs. Extreme Mech. Lett.
30, 100481.
Zhao, Z.L., Shu, T., Feng, X.Q., 2016. Study of biomechanical, anatomical, and physiological
properties of scorpion stingers for developing biomimetic materials. Mat Sci Eng C-Mater 58,
11121121.
Zhao, Z.L., Zhao, H.P., Ma, G.J., Wu, C.W., Yang, K., Feng, X.Q., 2015. Structures, properties, and
functions of the stings of honey bees and paper wasps: a comparative study. Biol. Open. 4, 921-
928.
Zhao, Z.L., Zhou, S.W., Feng, X.Q., Xie, Y.M., 2018. On the internal architecture of emergent plants.
J. Mech. Phys. Solids. 119, 224-239.
Zhao, Z.L., Zhou, S.W., Cai, K., Xie, Y.M., 2020a. A direct approach to controlling the topology in
structural optimization. Comput. Struct. 227, 106141.
Zhao, Z.L., Zhou, S.W., Feng, X.Q., Xie, Y.M., 2020b. Morphological optimization of scorpion telson.
J. Mech. Phys. Solids. 135, 103773.
Zhao, Z.L., Zhou, S.W., Feng, X.Q., Xie, Y.M., 2020c. Static and dynamic properties of pre-twisted
leaves and stalks with varying chiral morphologies. Extreme Mech. Lett. 34, 100612.
... As one of the density-based methods, the Solid Isotropic Material with Penalization (SIMP) method is often employed [10], which describes the iteratively varying model using the density variables at elements or nodes. Currently, the TO approach has been implemented in designing structural members and biological morphogenesis [12], which usually generates structures towards the prescribed load-bearing capacities [11] and the maximum structural stiffness (i.e., minimum compliance) [10]. ...
... However, the structures generated by the topology optimization method may be technically difficult to manufacture in construction practice [12,13], nor fulfill the compelling need for aesthetic patterns. These free-form designs usually comprise orderless and slim segments, exhibiting no regularized patterns and restricting modular prefabrication [14]. ...
... Herby the identification of zero-force elements is based on elemental compliancec i , i.e., the elemental strain energy induced by external loading. In Eq. (12), η c is the penalty parameter for determining zero-force components (10 − 7 is often used) with respect to c max , while c max is the maximum elemental compliance in the design domain. As long as c i is significantly lower than the maximum compliance, the element i identified as of nearly zero contribution to load-bearing could be removed. ...
Article
Topology optimization is a numerical method to find an optimal structural design governed by constraint and objective functions. This method has been enabled for free-form models pursuing the best performance or a fully periodic design in unit cells for modular prefabrication. However, an intermediate design that is neither free-form nor full-periodic is of practical significance, which could facilitate the balanced need for form and performance. In this paper, a novel multi-pattern control is developed to allow intermediate designs between these two extreme premises of topology optimization. The developed scheme divides an initial design domain into multiple groups of unit cells and enables individual pattern control for each group, which thereafter generates a variety of inspiring models fulfilling the various needs of form aesthetics, manufacturability, and structural performance. The number of pattern variations (NoV) is employed as a variable to control the pattern recurrence in unit cells, which is further equipped with the developed automated grouping using the clustering method and zero-force removal for post-processing. The inspiring models and exciting benefits from the multi-pattern control have been demonstrated with beam and building frame models, which shows how the eventually optimized designs vary as the NoV is changed.
... Examples include the detailed designs on fruit peels [1], the complex venation of leaves [2], the distinctive stripes of zebras [3], and the textured surfaces of brain corals [4]. These patterns fulfill various roles: enhancing nutrient transport efficiency [5], providing visual camouflage [6], and bolstering structural rigidity [7]. However, due to disciplinary boundaries, research often concentrates on one or two of these mechanisms, and the complexity and inherent randomness of patterns pose challenges in understanding their functionality and formation [8]. ...
Article
Full-text available
This study investigates the complex textures on walnut shells, which play a vital role in enhancing crashworthiness performance. Despite the challenges in deciphering their functionality and formation, we discovered that these textures can be described by reaction-diffusion equations. These equations capture the shell hardening mechanism and simulate texture formation based on observed lignin diffusion patterns. The texture sample sets, derived from diverse local sampling positions and scopes, were analyzed using a Convolutional Neural Network classification model to determine the most representative texture classes. The parameter combinations from the control equations, integrated with impact risk assessments and personalized needs, informed the design of a protective helmet. Physical and numerical tests confirmed the helmet's impact-absorption capabilities. These insights pave the way for the development of impact-resistant devices, such as bio-armor and shell for automotive parts.
... In response to these non-structural requirements, researchers have devised various modified topology optimization methods tailored to specific objectives. These adaptations encompass introducing detailed constraints for concept design [17], eliminating enclosed voids [18], achieving self-supporting optimized models [19], generating diverse truss structures [20], and simulating leaf morphogenesis [21]. In this snow project, the spatial structures pose distinctive challenges for topologically optimized layouts, necessitating modified BESO algorithms with specialized input parameters for material properties. ...
Conference Paper
Full-text available
This research delves into advancements in structural design and construction by employing the generative design technique Bi-directional Evolutionary Structural Optimization (BESO) and natural materials in cold climates. The study specifically focuses on harnessing snow materials to create spatial structures that are not only efficient but also elegant. Snow and ice, occurring naturally in cold climates, particularly in the Polar and Northern Hemisphere, present a unique opportunity for integration into the construction process within specified parameters. Given its crystal structure, snow is well-suited for compressive structures in sub-zero temperatures, with a density ranging between 400 and 820 kg/m3. The selection of material property values for designing snow structures is contingent upon factors such as temperature, quality, and load, necessitating a case-by-case approach. The distinctive properties of snow and ice materials can be factored into calculations and construction considerations. The research incorporates topological optimization techniques, widely employed in structural engineering and architectural form-finding, to present an innovative pavilion. This pavilion serves as a tangible outcome, illustrating collaborative efforts between architecture and engineering research groups. The central focus revolves around the application of BESO technology to generate the structure. The design features branches of varying sizes, drawing inspiration from iconic structures such as the interior of Sagrada Familia Basilica by Antoni Gaudi and Shanghai Himalaya Centre by Arata Isozaki. The authors' team manually sculpted the full-scale model of the pavilion at a challenging temperature of minus 25 degrees Celsius, utilizing minimal materials. This meticulous process resulted in a remarkable and innovative structure that captures the essence of nature within the frozen landscape. The key advantages of this novel design and construction methodology lie in the efficient utilization of natural materials and the creation of elegant structural forms.
... Nowadays, topology optimization techniques have been increasingly adopted in a wide range of engineering applications, including advanced manufacturing [1][2][3][4][5][6], architectural design [7][8][9][10][11][12], biomechanical morphogenesis [13][14][15][16][17], and civil engineering [18][19][20][21][22][23][24]. Despite their growing popularity, the optimized designs from conventional optimization techniques often feature complex geometries that pose significant manufacturing challenges [1]. ...
Article
Full-text available
In recent years, topology optimization of periodic structures has become an effective approach to generating efficient designs that meet a variety of practical considerations, including manufacturability, transportability, replaceability, and ease of assembly. Traditional periodic structural optimization typically restricts designs to a uniform assembly configuration utilizing only one type of unit cell. This study proposed a novel clustering-based approach for periodic structural optimization, which allows variable orientations of individual unit cells. A dynamic k-means clustering strategy is introduced to categorize all unit cells into distinct groups and gradually eliminate less efficient unit cells from the optimized design. Meanwhile, a novel technique is introduced to identify and select more efficient orientations of unit cells during the optimization process. Several numerical examples are presented to demonstrate the effectiveness of the proposed approach. The results show that periodic structures with clustered oriented unit cells can significantly outperform their traditional periodic counterparts. This study not only incorporates assembly flexibility into periodic topology optimization but also utilizes multiple types of unit cells in a design, thereby further enhancing its structural performance.
... In addition to the civil and architectural applications, Airbus (Krog et al., 2009) successfully applied topology optimization on its A380, A400, and A350 plane models, including the design of wing ribs, engine mount frames, and floor crossbeams. Apart from the artificial structures, Ma et al. (2021b) and Zhao et al., (2018Zhao et al., ( , 2020 utilized topology optimization to facilitate the analysis of biostructures in plants and animals that have evolved over millions of years. ...
Article
Full-text available
Free-form architectural design has gained significant interest in modern architectural practice. Due to their visually appealing nature and inherent structural efficiency, free-form shells have become increasingly popular in architectural applications. Recently, topology optimization has been extended to shell structures, aiming to generate shell designs with ultimate structural efficiency. However, despite the huge potential of topology optimization to facilitate new design for shells, its architectural applications remain limited due to complexity and lack of clear procedures. This paper presents four design strategies for optimizing free-form shells targeting architectural applications. First, we propose a topology-optimized ribbed shell system to generate free-form rib layouts possessing improved structure performance. A reusable and recyclable formwork system is developed for their effective and sustainable fabrication. Second, we demonstrate that topology optimization can be combined with funicular form-finding techniques to generate a rich variety of elegant designs, offering new design possibilities. Third, we offer cost-effective design solutions using modular components for free-form shells by combining surface planarization and periodic constraint. Finally, we integrate topology optimization with user-defined patterns on free-form shells to facilitate aesthetic expression, exemplified by the Voronoi pattern. The presented strategies can facilitate the usage of topology optimization in shell designs to achieve high-performance and innovative solutions for architectural applications.
Thesis
Full-text available
The present study deals with computer-aided design, 3D-printing, large strain numerical simulation, and experimental testing of random geometries with focus on porous materials. In particular, we attempt to assess the effect of random porous features on the mechanical response at large strain by comparing the response of well-chosen random and periodic porous geometries. We first investigate the computer-aided design process of a variety of porous geometries including random polydisperse porous materials with spherical and ellipsoidal voids, standard eroded Voronoi geometries, hexagonal honeycombs, and TPMS structures. In addition, we propose a novel computer-aided design strategy to obtain a new type of random Voronoi-type porous materials called M-Voronoi (from mechanically grown) with smooth void shapes and variable intervoid ligament sizes that can reach very low relative densities. This is achieved via a numerical, large strain, nonlinear elastic, void growth mechanical process. The proposed M-Voronoi method is general and can be applied to create both two and three-dimensional random geometries and allows the formation of isotropic or anisotropic materials. The void growth process is a consequence of mass conservation and the incompressibility of the surrounding nonlinear elastic matrix phase and the final achieved relative density may be analytically estimated in terms of the determinant of the applied deformation gradient. The extremely low densities in the M-Voronoi geometries are achieved through an intermediate remesh step in the virtual fabrication process. For this purpose, we developed a versatile and general remeshing algorithm based on the geometry reconstruction of an orphan mesh that can handle arbitrarily complex meshes, including those that contain voids or multiple phases. Moreover, the studied random geometries are general to model seamlessly a wide range of composites involving particles, multi-phase, and even polycrystals with finite interfaces under mechanical or coupled loads (e.g. magneto-electro mechanical, etc.).In the next part of the study, we fabricate the designed porous materials with a polymer 3D-printer via PolyJet technology and a UV-curable resin called TangoBlack which is a highly viscous soft polymer with brittle fracture. Meanwhile, the viscous behavior of TangoBlack is studied under uniaxial tensile, loading-unloading, and relaxation tests on a new proposed specimen geometry and is subsequently characterized by a nonlinear rubber viscoelastic model for incompressible isotropic elastomers. We then use this material to 3D-print the designed two-dimensional porous materials with square representative geometries and isotropic/anisotropic features in terms of void size and realization. The mechanical response of the fabricated porous materials is experimentally investigated by testing them under uniaxial large strain compression and low strain rates. We show that the randomness of the proposed M-Voronoi geometries and their non-uniform intervoid ligament size leads to enhanced mechanical properties at large compressive strains with no apparent peak-stress and strong hardening well before densification, while they become very close to random eroded Voronoi geometries at low densities.In the last part of this study, we investigate numerically the mechanical properties of the three-dimensional random porous geometries consisting of M-Voronoi, polydisperse porous materials with spherical voids, and classical TPMS-like geometries. The simulations are performed at large strains under compression loading while considering the matrix an elastic-perfectly plastic material without hardening. We observe enhanced plastic flow stress in the geometries with random topologies as opposed to the TPMS periodic structures. This behavior is explained by noting that deformation localizes in geometries with a periodic pattern, contrary to the random geometries which exhibit a rather diffused localization.
Article
Full-text available
Ceramic materials are widely used in engineering field because of their good physical and chemical properties. The poor mechanical properties of traditional ceramics seriously limit the development of ceramic materials and have attracted extensive attention since its birth. The development of high toughness, light weight, and functional ceramic materials has long been the pursuit of materials scientists. Since ancient times, nature has been the source of all kinds of human technological ideas, engineering principles, and major inventions. The functional structure of biology provides a good research object for the bionic design and preparation of ceramic materials. Therefore, it is necessary to summarize the functional structures and mechanisms in nature and biomimetic preparation technology. The purpose of this review is to provide guidance for the functionalized biomimetic design and efficient preparation of ceramic materials. Some typical functional examples in nature are introduced in detail, and several representative biomimetic preparation methods are listed. Finally, the performance and application status of biomimetic ceramics are summarized, and the future development direction of biomimetic ceramic materials is prospected.
Article
Full-text available
In high-speed vehicles, rudders often endure both aerodynamic pressure and thermal loads. The innovative design of rudders is of great importance for the performance of the whole vehicle. In this work, thermo-elastic topology optimization is adopted to design a typical all-movable rudder structure. The compliance of the rudder skin is considered to be a new objective and the moment of inertia of the rudder is constrained during optimization to ensure its fast response to instructions of the control system. Then sensitivity analysis of the structural compliance and the moment of inertia is carried out. Optimization results show that thermal load has a great effect on the optimized configuration and minimizing the compliance of the rudder skin gives much better design than minimizing the global compliance. Subsequently, an engineering-oriented post-processing is conducted to make the optimized design suitable for additive manufacturing. An appropriate printing direction is selected based on the layout of the optimized ribs and certain ribs are reshaped with fillets to make the rudder free of internal support structures. Besides, according to a secondary topology optimization of the ribs and the stress distribution, a set of powder-discharge holes are properly opened on the ribs so that all cavities within the rudder are connected and the metal powder inside the rudder can be discharged with little effort after manufacturing. Finally, the optimized design is successfully printed using Selective Laser Melting, demonstrating the proposed post-processing is effective for additive manufacturing.
Article
Full-text available
Previous studies on topology optimization subject to stress constraints usually considered von Mises or Drucker–Prager criterion. In some engineering applications, e.g., the design of concrete structures, the maximum first principal stress (FPS) must be controlled in order to prevent concrete from cracking under tensile stress. This paper presents an effective approach to dealing with this issue. The approach is integrated with the bi-directional evolutionary structural optimization (BESO) technique. The p-norm function is adopted to relax the local stress constraint into a global one. Numerical examples of compliance minimization problems are used to demonstrate the effectiveness of the proposed algorithm. The results show that the optimized design obtained by the method has slightly higher compliance but significantly lower stress level than the solution without considering the FPS constraint. The present methodology will be useful for designing concrete structures.
Article
Full-text available
Topology optimization techniques have been widely used in structural design. Conventional optimization techniques usually are aimed at achieving the globally optimal solution which maximizes the structural performance. In practical applications, however, designers usually desire to have multiple design options, as the single optimal design often limits their artistic intuitions and sometimes violates the functional requirements of building structures. Here we propose three stochastic approaches to generating diverse and competitive designs. These approaches include (1) penalizing elemental sensitivities, (2) changing initial designs, and (3) integrating the genetic algorithm into the bi-directional evolutionary structural optimization (BESO) technique. Numerical results demonstrate that the proposed approaches are capable of producing a series of random designs, which possess not only high structural performance, but also distinctly different topologies. These approaches can be easily implemented in different topology optimization techniques. This work is of significant practical importance in architectural engineering where multiple design options of high structural performance are required.
Article
Full-text available
Topology optimization is increasingly used in lightweight designs for additive manufacturing (AM). However, conventional optimization techniques do not fully consider manufacturing constraints. One important requirement of powder-based AM processes is that enclosed voids in the designs must be avoided in order to remove and reuse the unmelted powder. In this work, we propose a new approach to realizing the structural connectivity control based on the bi-directional evolutionary structural optimization technique. This approach eliminates enclosed voids by selectively generating tunnels that connect the voids with the structural boundary during the optimization process. The developed methodology is capable of producing highly efficient structural designs which have no enclosed voids. Furthermore, by changing the radius and the number of tunnels, competitive and diverse designs can be achieved. The effectiveness of the approach is demonstrated by two examples of three-dimensional structures. Prototypes of the obtained designs without enclosed voids have been fabricated using AM.
Article
The stiffness and heat dissipation of thin planar structures in industrial engineering are taken as the background for this research. Plant leaves with various vein patterns, such as tobacco (Nicotiana tabacum L.) and chili (Capsicum annuum L.) leaves, are treated as the research objects. Through a combination of morphological and mechanical analysis, the distribution patterns and properties of leaf veins are mathematically characterized. A topological optimization algorithm is employed to simulate the vein growth process, thus revealing the effects of the mechanical and biological properties of different leaves on their vein morphologies. Additionally, the angles between the main and secondary veins are controlled to satisfy biological constraints. This comprehensive exploration of vein morphological formation can serve as a reference for the design of bionic thin planar structures in engineering.
Article
The mechanical properties and biological functions of tissues and organs in plants are closely related to their structural forms. In this study, we have performed systematic measurements and found that the leaves and stalks of several species of emergent plants exhibit morphologies of twisting and gradient chirality. Inspired by the experimental findings, we investigate, both theoretically and numerically, the static bending and vibrational properties of these plant organs. By modeling the leaves and stalks as pre-twisted cantilever beams, the effects of the cross-sectional geometry, loading condition, handedness perversion, twisting configuration, and morphological gradient, on their mechanical behavior are evaluated. Our analysis reveals that both static and dynamic responses of the beams can be easily tuned by changing their structural parameters. For any part of the beams, its chiral morphology has more significant influences on the overall structural performance (e.g., bending stiffness and natural frequencies) if it is closer to the clamped end. This work not only deepens our understanding of the structure–property–function interrelations of chiral plants, but also holds potential applications in the bio-inspired design of innovative devices and structures.
Article
Structural shape and topology optimization has undergone tremendous developments in recent years due to its important applications in many fields. However, effectively controlling the structural complexity of the optimization result remains a challenging issue. The structural complexity is usually characterized by the distribution and geometries of interior holes. In this work, a new approach is developed based on the graph theory and the set theory to control the number and size of interior holes of the optimized structures. The minimum distance between the edges of any two neighboring holes can also be constrained. The structural performance and the effect of the structural complexity control are well balanced by using this approach. We use three typical numerical examples to verify the effectiveness of the developed approach. The optimized structures with and without constraints on the structural complexity are quantitatively compared and analyzed. The present methodology not only enables the designer to have a direct control over the topology of the optimized structures, but also provides diverse and competitive solutions.
Article
Nature provides inspirations for solving many challenging scientific and technological problems. In this study, a computational methodology is developed for the morphological optimization of three-dimensional, multi-component biological organs. The structural optimization of scorpion telson, which consists of a curved stinger and a venom container, is considered as an example by using this method. Both experimental and numerical results indicate that, through a long history of natural selection, the load-bearing capacity of the venom apparatus of a scorpion has been optimized together with its flexible segmented tail, important biological functions (e.g., venom storage and transportation), and superb sting strategy. The optimal range of the sting direction of a scorpion is theoretically determined and verified by finite element analysis. The curved scorpion stinger makes the venom container a robust design that is insensitive to the loading direction. The biomechanical mechanisms underlying the robust design are deciphered by comparing the venom apparatuses of scorpions and honey bees. This work deepens our understanding of the structure–property–function interrelations of the venomous sharp weapons of both scorpions and honey bees, and the presented methodology can also be extended to design engineering structures with optimal morphologies (e.g., curved hypodermic needles and segmented robotic arms) and explore other biological tissues and organs.
Article
Shape and topology optimization techniques are widely used to maximize the performance or minimize the weight of a structure through optimally distributing its material within a prescribed design domain. However, existing optimization techniques usually produce a single optimal solution for a given problem. In practice, it is highly desirable to obtain multiple design options which not only possess high structural performance but have distinctly different shapes and forms. Here we present five simple and effective strategies for achieving such diverse and competitive structural designs. These strategies have been successfully applied in the computational morphogenesis of various structures of practical relevance and importance. The results demonstrate that the developed methodology is capable of providing the designer with structurally efficient and topologically different solutions. The structural performance of alternative designs is only slightly lower than that of the optimal design. This work establishes a general approach to achieving diverse and competitive structural forms, which holds great potential for practical applications in architecture and engineering.