You are on page 1of 48

Accepted Manuscript

A geometry projection method for continuum-based topology


optimization with discrete elements

J. A. Norato, B. K. Bell, D. A. Tortorelli

PII: S0045-7825(15)00171-1
DOI: http://dx.doi.org/10.1016/j.cma.2015.05.005
Reference: CMA 10618

To appear in: Comput. Methods Appl. Mech. Engrg.

Received date: 19 December 2014


Revised date: 26 April 2015
Accepted date: 6 May 2015

Please cite this article as: J.A. Norato, B.K. Bell, D.A. Tortorelli, A geometry projection
method for continuum-based topology optimization with discrete elements, Comput. Methods
Appl. Mech. Engrg. (2015), http://dx.doi.org/10.1016/j.cma.2015.05.005

This is a PDF file of an unedited manuscript that has been accepted for publication. As a
service to our customers we are providing this early version of the manuscript. The manuscript
will undergo copyediting, typesetting, and review of the resulting proof before it is published in
its final form. Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
*Manuscript
Click here to download Manuscript: main.pdf Click here to view linked References

1
2
3
4
5
6
7
8
9 A geometry projection method for continuum-based
10
11
topology optimization with discrete elements
12
13 J. A. Noratoa,∗, B. K. Bellb , D. A. Tortorellic
14
a Department of Mechanical Engineering, The University of Connecticut, 191 Auditorium
15
16 Road, U-3139, Storrs, CT 06269
b Ricardo, Chicago Technical Campus, 7850 S. Grant Street, Burr Ridge, IL 60527, United
17 States
18 c Department of Mechanical Science and Engineering, University of Illinois at
19 Urbana-Champaign, 1206 West Green Street, MC-244, Urbana, IL 61801, United States
20
21
22
23
24 Abstract
25
26 This article describes a method for the continuum-based topology optimization
27
28 of structures made of discrete elements. In particular, we examine the opti-
29 mization of linearly elastic planar structures made of bars of fixed width and
30
31 semicircular ends. The design space for the optimization consists of the end-
32
point locations of the bar’s medial axes and their out-of-plane thicknesses. To
33
34 circumvent re-meshing upon design changes, we project the design onto a fixed
35
36 analysis grid using a differentiable geometry projection that results in a density
37 field indicating the fraction of solid material anywhere in the design space, as
38
39 in density-based topology optimization methods. The out-of-plane thickness is
40
penalized so that the optimizer is capable of removing bars during the opti-
41
42 mization. The differentiability of the projection allows for the computation via
43
44 the chain rule of design sensitivities of responses of interest, and therefore it
45 allows for the use of robust and efficient gradient-based optimization methods.
46
47 Notably, this approach makes it easier to fabricate optimal designs by using off-
48 the-shelf stock material. Furthermore, the method considers the case where bars
49
50 overlap at their joints (i.e. their thicknesses are added at the joint) and when
51
they do not. Finally, our proposed method naturally accommodates the impo-
52
53
54 ∗ Corresponding author. Phone: +1 (860)486-2345. Fax: (860)486-5088. E-mail address:
55 norato@engr.uconn.edu.
56
57
58
59 Preprint submitted to Computer Methods in Applied Mechanics and EngineeringApril 26, 2015
60
61
62
63
64
65
1
2
3
4
5
6 2
7
8
9 sition of several fixed length scales. We demonstrate the proposed approach on
10
classical problems of compliance-based topology optimization and identify its
11
12 benefits as well as research directions to be addressed in the future.
13
14 Keywords: Topology optimization, geometry projection, manufacturability,
15 multiscale topology optimization.
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 3
7
8
9 1. Introduction
10
11 Topology optimization (TO) methods have matured to the point where the
12
13 use of commercial software to explore structural designs is commonplace in in-
14
dustry. This maturity rests on the density-based approach, in which the design
15
16 space is parameterized by a (discretized) density field and the local material
17
18 properties are defined as differentiable functions of the density. Its brilliance
19 lies in its natural ability to accommodate changes in topology, circumvent re-
20
21 meshing upon design changes, and allow for response sensitivities to be com-
22 puted so that robust and efficient gradient-based nonlinear programming al-
23
24 gorithms can be used for the optimization. The ability to produce a mostly
25
26 solid-void design is given by the Solid Isotropic Material Penalty (SIMP) tech-
27 nique ([1, 2]), which penalizes the presence of intermediate density regions by
28
29 making them inefficient (in terms of, e.g., stiffness). The more recent level set
30 TO methods ([3, 4, 5]) represent the structural design implicitly via the zero
31
32 level set of a function and retain the aforementioned advantages of density-based
33
methods; however, they do not require penalization of intermediate densities,
34
35 since the level set representation naturally induces a solid-void design almost
36
37 everywhere. This advantage comes at the expense of other challenges, such as
38 relatively poor convergence (cf. [6]).
39
40 The aforementioned approaches have thrived because of their ability to pro-
41 duce complex, organic shapes. In some cases, like in the fabrication of planar
42
43 microelectromechanical systems (MEMS), designs can be manufactured that
44
45 closely resemble the optimal topologies ([7]). Additive manufacturing tech-
46 niques hold the promise to extend that realm. However, in most applications
47
48 where a particular manufacturing process is employed or when, due to economic
49 and joining considerations, one wishes to employ stock material (such as tubes,
50
51 squares, rounds, plates, etc.), it is difficult to directly use the TO results, because
52
the resulting design cannot possibly be manufactured with the available process
53
54 or stock material. As a result of this, the designer has to spend a significant
55
56 amount of time creating a manufacturable design that captures the intent of the
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 4
7
8
9 TO; as (s)he does that, structural responses of interest may significantly dete-
10
riorate (for example, by introducing stress concentrations), which will prompt
11
12 additional design iterations and potentially add weight, thereby reducing the
13
14 effectiveness of the optimization. It should be noted that ground structure TO
15 approaches ([8, 9]), which work by removing discrete elements from an initial
16
17 network (the ground structure) and by moving the end points of those elements,
18
naturally accommodate the use of one-dimensional stock material (for instance,
19
20 I-beams). However, the analysis that is used to predict the performance of these
21
22 designs contains numerous simplifications versus the more accurate continuum
23 theory. Therefore, we restrict our attention to TO on continua.1
24
25 The idea of using an explicit geometric design representation was first ex-
26 plored in the bubble method ([10]), where splines are used to represent the
27
28 structural boundaries, with the control points of the splines being the design
29
parameters. A conforming grid is used for the analysis, requiring re-meshing
30
31 when the shape of existing boundaries changes or a new hole is introduced. The
32
33 locations where holes are nucleated are determined by a characteristic function,
34 which was later shown to correspond to the topological derivative of compliance
35
36 ([11]). In density-based TO methods, the earliest geometric consideration is
37
perhaps the use of filtering techniques; while originally intended to make the
38
39 optimal topology independent from the mesh size, they showed the capability of
40
41 introducing a minimum member size for the compliance problem, either by fil-
42 tering the sensitivities ([12]) or directly the densities [13, 14])—although it was
43
44 soon after identified that they were not capable of removing so-called hinges (i.e.
45 one-node connections) in the design of compliant mechanisms ([12]). A mini-
46
47 mum length scale is also imposed via filtering and thresholding in [15] whereby
48
an additional regularized Heaviside thresholding of the filtered density renders
49
50 sharper boundaries (i.e. less intermediate densities). Not surprisingly (that is,
51
52 because this is a significant industry need), more advances in so-called manufac-
53
54 1 In this work, however, we restrict attention to compliance-based topology optimization
55
56 and defer the subject of stress-based topology optimization to future work.
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 5
7
8
9 turing constraints came from the commercial TO software industry, where side
10
constraints are imposed on the design variables so that the resulting topologies
11
12 require no cores if the structure is cast or have a constant section if it is ex-
13
14 truded ([16]). Density-based morphology filters (such as erosion and dilation)
15 inspired by digital image processing techniques were investigated in [17] in order
16
17 to overcome the onset of hinges in compliant mechanism design. A maximum
18
length scale was introduced in [18] by imposing element-wise constraints on the
19
20 minimum void volume in a circle (sphere) of given size centered at each element
21
22 centroid; this approach produces topologies that satisfy the specified maximum
23 length scale, albeit at the expense of a large number of constraints in the op-
24
25 timization problem. In [19] casting constraints are introduced by employing
26 a regularized Heaviside function that provides the location of the structural
27
28 boundary along a column of elements along the casting draw direction. In [20]
29
the circular neighborhood used in [15] for the definition of the regularized Heav-
30
31 iside projection is extended to the access surface in order to impose casting or
32
33 milling constraints.
34 Incorporation of parameterized primitives as features in density-based topol-
35
36 ogy optimization was reported in [21]. In this work, holes in the design that
37
have a fixed, parameterized shape but variable location and orientation are im-
38
39 posed by constructing an implicit function with regularized Heaviside functions.
40
41 Boolean operations among features are supported by the use of R-functions. As
42 in the bubble method, holes are introduced based on the topological derivative of
43
44 the objective function, but using a fixed grid for the analysis. The shape of the
45 nucleated hole is chosen from a library of primitives using morphing techniques,
46
47 which is amenable to subsequent shape optimization. Upon completion of the
48
optimization, shape matching techniques are employed to convert the optimal
49
50 topology into geometric primitives. In [22], the void region in the design space
51
52 is represented by the union of circular or rectangular masks of variable size and
53 location. Elements inside the mask are assigned a density close to zero, whereas
54
55 elements outside the mask are assigned a density of 1, hence this method pro-
56
duces 0-1 designs; consequently, however, it requires a zero-order optimization
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 6
7
8
9 algorithm to determine the optimal size and location of the masks. This method
10
is extended in [23] for the design of photonics crystal waveguides with circular
11
12 holes, whereby regularized Heaviside functions are used to map the circular
13
14 holes onto the fixed grid, hence supporting sensitivities and allowing for the use
15 of gradient-based optimizers. The embedding of fixed shape solid components
16
17 in [24] is attained by approximating the shape of these components with the
18
union of a set of circles (the finite circle method). Contrary to the methods in
19
20 [21] and [23], however, this method requires re-meshing in a region around the
21
22 fixed component, and hence does not provide a continuous and differentiable
23 relationship between the component’s geometry and the analysis.
24
25 In level set TO methods, several strategies have been employed to incorpo-
26 rate geometric constraints. In [25], as in [21], free-form shapes and parame-
27
28 terized primitives are combined into a single implicit representation whose zero
29
level set defines the structural boundary, and boolean operations among these
30
31 primitives are supported via R-functions. Unlike the aforementioned density-
32
33 based methods, however, the analysis is not performed on a fixed finite element
34 grid, but using a meshfree method with distance fields previously developed by
35
36 the same group ([26]). This method differs from most level set TO methods,
37
which use the solution of the Hamilton-Jacobi equation to update the design.
38
39 Rather, it uses a time-stepping scheme that is based on shape sensitivities; there
40
41 is no connection between the level set function and density-based methods. In
42 [27] a constraint, enforced via a penalty method, limits the interactions between
43
44 points on the level set so as to produce designs with constant member size. The
45 method in [28], like [25], uses a composite level set function to accommodate
46
47 both free shape and parameterized primitives. Instead of using R-functions to
48
support Boolean operations among the primitives, however, a product of Heav-
49
50 iside functions is used. The analysis is performed using the extended finite
51
52 element (XFEM), which explicitly represents the structural boundary and en-
53 ables the computation of the shape sensitivities (provided the boundary does
54
55 not pass through a node of the fixed mesh), at the expense of re-meshing and
56
the additional cost of XFEM. Finally, in the recent work in [29], the authors at-
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 7
7
8
9 tempt to produce topologies of constant member size by introducing a constraint
10
(via the augmented Lagrangian) on the average distance between the structure’s
11
12 medial surface and the zero level set. However, their reported topologies are not
13
14 as smooth and do not attain a constant member size as well as those of [28].
15 Recently, the researchers developed a method for the optimization of struc-
16
17 tures comprised of rectangular bars (similar to the geometry considered in this
18
paper) whereby, in the same vein of some of the aforementioned approaches,
19
20 a level set function that implicitly represents the Boolean union of all bars is
21
22 employed to model the structural geometry [30]; like [29], it explicitly repre-
23 sents the boundary via the XFEM. This method resembles ours in the use of a
24
25 parametric representation of the bars; however, ours is in effect a density-based
26 method, and hence it can be used with standard finite element analysis and
27
28 nonlinear programming algorithms. Moreover, we note that the method in [30]
29
does not remove bars from the initial design, does not accommodate overlapping
30
31 bars (i.e. with increased joint thickness), and, as in the results in [29], and does
32
33 not produce designs with smooth boundaries, which can be a challenge for, e.g.,
34 stress-based TO.
35
36 The aforementioned methods are capable of generating designs with certain
37
geometric features by imposing various innovative constraints on the design
38
39 space (i.e. the density field or the level set function). However, with the ex-
40
41 ception of [31], none of them produces designs exclusively comprised of discrete
42 elements. In this paper we produce such designs in the context of density-based
43
44 TO. In particular, we examine the optimization of linearly elastic planar struc-
45 tures made of bars of fixed width and semicircular ends. The design space for
46
47 the optimization consists of the endpoint locations of the bar’s medial axes and
48
their out-of-plane thicknesses. As we wish to perform the analysis on the same
49
50 fixed grid throughout the optimization, and use efficient gradient-based opti-
51
52 mizers, we project the bar design onto the analysis grid using a differentiable
53 mapping from the design space, e.g. the endpoint positions, to a continuously
54
55 varying density field, which indicates the fraction of solid material anywhere
56
in the design space as in density-based topology optimization methods. The
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 8
7
8
9 out-of-plane bar thickness is penalized giving the optimizer the capability to
10
remove bars. Our method also accommodates the cases where bars overlap
11
12 at their joints (i.e. their thicknesses are added at the joint) or when they do
13
14 not. Finally, our proposed method naturally accommodates the imposition of
15 several fixed length scales, i.e. different bar widths. The method extends our
16
17 earlier work in [32] by using bars with semicircular ends instead of rectangles,
18
which simplifies the geometry projection and the sensitivity analysis and leads
19
20 to improved robustness and designs.
21
22 The remainder of the paper is organized as follows. In Section 2 we formalize
23 the description of the geometry of the bars and describe the geometry projection
24
25 on the analysis grid. In Section 3 we detail the sensitivity analysis for responses
26 of interest. We then formulate the optimization problem and provide details of
27
28 the computer implementation in Section 4. In Section 5 we demonstrate the
29
proposed approach on classical problems of compliance-based TO. Finally, we
30
31 identify the benefits of the proposed approach and identify future directions of
32
33 research in Section 6.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 9
7
8
9 2. Geometry Projection
10
11 The design is herein defined by a set of bars of constant width w and with
12
13 semicircular ends of radius w/2. Each bar q is parameterized by the location
14 of the end points of its medial axis xqo and xqf (cf. Figure 1) and by its out-
15
16 of-plane thickness t. We assume all bars have the same thickness and width for
17
simplicity, but note that the method accommodates bars of different thickness
18
19 and width. The signed distance from any point in space p to a bar q, which will
20
21 be used in the sequel to define the geometry projection, is given by
22 w
23 φq (dq (xqo , xqf , p), w) := dq (xqo , xqf , p) − , (1)
2
24
25 where dq above is the minimum distance from p to the bar’s medial axis, given
26
by ([33])
27 
28 
 ||b|| if a · b ≤ 0


29 
30 dq (xqo , xqf , p) := ||g|| if 0 < a · b < a · a (2)


31 

32 ||e|| if a · b ≥ a · a
33
34 with
35
36 a := xqf − xqo (3)
37
38 b := p − xqo (4)
39
e := p − xqf (5)
40  
41 1
42 g := I − a ⊗ a b =: P⊥
a b. (6)
||a||2
43
44 In the above, || · || denotes the Euclidean 2-norm, ⊗ denotes the dyadic tensor
45 product, and P⊥
a is the perpendicular projection operator. From hereon, and
46
47 for the sake of brevity, we omit the argument w from the previous equations,
48
because it is considered fixed. We note that this is a mater of choice so that
49
50 we obtain constant width bars; however, nothing in our method would prevent
51
52 w from being a design variable or from designing with bars of multiple fixed
53 width. The use of round ends allows us to obtain the distance from any point
54
55 to a bar with a single equation (i.e. Equation 1), as opposed to having rectan-
56
gular bars as in [32], where we have to find the minimum distance among the
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 10
7
8
9
10 w
11 xf
12
13
14
15
16
17 g
e
18 xo
19
20 b
21 p
22
23 Figure 1: Bar geometry
24
25
26 minimum distance to each individual edge (which, additionally, incurs problems
27 with differentiability of the signed distance function due to the non-smooth bar
28
29 boundaries). Not only is the formulation simpler, but the computation of the
30
geometry projection and the sensitivity analysis are also more efficient since it
31
32 is not necessary to loop over all four edges of the rectangular bars.
33
34 To perform the analysis on a fixed grid, we project the geometry onto a
35 density field. To this end, we employ the geometry projection algorithm of [34],
36
37 which can be described as follows: given a sampling window, i.e. the ball Brp of
38 radius r centered at p, the density ρ(p, r) is given by the fraction of the window
39
40 that intersects the solid structure (cf. Figure 2), i.e.
41
42 |Brp ∩ ω|
ρ(p, r) := , (7)
43 |Brp |
44
45 where ω denotes the region of space occupied by the structure and | · | is a
46
47 measure of the volume (or area). If the window size is negligible in relation to
48 the curvature of the structural boundary ∂ω inside the window, then we can
49
50 approximate the portion of ∂ω that intersects the window with a straight line
51 whose mid point is the point on ∂ω closest to p (cf. Figure 2). In that case,
52
53 the numerator in Equation 7 is given by the volume (area) of the spherical cap
54
(circular segment) on the solid side. In the case of our bar problem, the density
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 11
7
8
9 is expressed in terms of the signed distance from p to the bar q:
10 
11 
0 if φq > r


12 
 h  


13  1 2 r2 arccos φq (dq ) −
πr r
14 ρq (dq (xqo , xqf , p), r) := p i (8)


15 
 φq (dq ) r2 − φq (dq )2 if − r ≤ φq ≤ r
16 



17 1 if φ < −r,
q
18
19 where the arguments of dq , i.e. (xqo , xqf , p), have been omitted on the right-
20
21 hand side of the equation for conciseness. Also, since r is fixed in the proposed
22
method, it will be omitted as an argument hereafter.
23
24
25
26 ∂ω
27
ω
28
29
30
31
32
33 φq
34
r p
35
36
37
38
39 Figure 2: Geometry projection
40
41
42 This projection was employed in [34] for shape optimization on fixed grids,
43
where the structural boundary was given by a piecewise approximation of an
44
45 ellipse parameterized by its radii in one example, and by the level set of an
46
47 implicit linear combination of radial basis functions in another. It was also
48 used in [35] to perform topology optimization; instead of using an independent
49
50 geometry model, however, the structural boundary in that work was given by a
51
level set of the topological derivative of the structural compliance.
52
53 Since we are performing the analysis on a fixed grid, the points in space
54
55 whose sampling window does not intersect the structure will have a zero density,
56 which will make the analysis problem ill-posed (the way in which the geometry
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 12
7
8
9 projection modifies the analysis problem will be defined in the following). In
10
order to avoid this, we introduce a lower bound ρmin on the density, and define
11
12
13 ρ̃q (ρq (dq (xqo , xqf , p))) := ρmin + ρq (dq (xqo , xqf , p))(1 − ρmin ). (9)
14
15 In the case when only one bar intersects Brp , we modify the material’s elasticity
16
17 tensor C as a function of the elasticity tensor of the solid material Co (scaled by
18
the bar’s out-of-plane thickness t) via
19
20
21 C(ρ̂q (ρ̃q (ρq (dq (xqo , xqf , p))), αq , ηc )) := ρ̂q (ρ̃q (ρq (dq (xqo , xqf , p))), αq , ηc )Co ,
22
(10)
23
24 where
25
26
ρ̂q (ρ̃q (ρq (dq (xqo , xqf , p))), αq , η) := ρ̃q (ρq (dq (xqo , xqf , p)))αqη , (11)
27
28
29 is an effective density, wherein αq ∈ [0, 1] can be seen as a weighting of the
30
out-of-plane thickness t, and ηc is a penalization power that has the same role
31
32 as in SIMP methods, i.e. it penalizes intermediate thickness values so that the
33
34 resulting design is made of bars of a specified thickness t, and of bars with zero
35 thickness, which can be considered as non-existent. However, αq need not be
36
37 associated with a physical dimension in general, and can be thought of as a
38 mechanism to push an entire discrete member towards being fully effective or
39
40 completely disappearing.
41
One advantage of having an independent explicit geometry model is that
42
43 we can consider the two separate cases when bars overlap and merge. In the
44
45 first case when bars overlap, we assume their thicknesses add and define the
46 composite density as
47
48 N
X
49 ρ(z, p, η) := ρ̂q (ρ̃q (ρq (dq (xqo , xqf , p))), αq , η) (overlap), (12)
50 q=1
51
52 where N is the number of bars, and z is the vector of design variables given by
53 z := [zT1 , zT2 . . . , zTN ]T , with zq = [xTqo , xTqf , αq ]T . Assuming at least one bar has
54
55 a non-zero thickness, i.e. αmax := maxq {αq } > 0, which is always the case when
56
the optimization is driven by structural criteria, then it can be shown that
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 13
7
8 PN
9 ρ ∈ [ρmin αmax
η
, q αq ]. As the optimization progresses and the penalization
10
pushes all αq towards 0 or 1, then ρ ∈ [ρmin , N ].
11
12 In the second case, i.e. when bars merge at their intersections, we assume
13
14 that the effective thickness is the maximum thickness among all bars in the
15 intersection (this would be the equivalent to the existing density-based and level
16
17 set methods). Since the maximum function is not differentiable, we approximate
18
it using a p-norm and define the composite density as
19
20 " N
# p1
21 X p
ρ(z, p, η) := ρ̂q (ρ̃q (ρq (dq (xqo , xqf , p))), αq , η) (merge), (13)
22
q=1
23
24 Assuming again that αmax > 0, then for finite p we have that ρ ≥ ρmin αmax
η
,
25
26 with the equality holding if and only if all bars except one have αq = 0. On the
27
other hand, ρ ≥ maxq αqη , with the equality holding if and only if all bars but
28
29 one have αq = 0. If all bars have either αq = 0 or αq = 1, then as p → ∞ the
30
31 lower bound on ρ tends to ρmin from above, and the upper bound on ρ tends
32 to 1 also from above. Hence, for finite p we will always exceed the maximum
33
34 thickness, as we show in the examples of Section 5.
35 To illustrate the difference between merged and overlapped bar designs, con-
36
37 sider the case where all bars have the same thickness t. Joints of bars in merged
38
39 designs will have an effective thickness approximately equal to the maximum
40 thickness of all bars at the joint, in this case t; hence, merged designs can be
41
42 thought of bars being cut and then welded to each other at the joint, much like
43 the miter joint in a picture frame. Joints of bars in overlapped designs, on the
44
45 other hand, will have an effective thickness that equals the sum of thicknesses
46
of the bars that form the joint. Therefore, overlapped designs can be thought
47
48 of bars being cut and then welded (or bolted) to a gusset of thickness t(Nb − 1),
49
50 where Nb denotes the number of bars intersecting at the joint. Because the joints
51 in either case are fully effective (i.e. they constitute solid regions in space), they
52
53 are capable of transferring moments and hence the structures produced with
54 our method can be interpreted as frames. Our method differs from the variable
55
56 thickness method in that the structure is made of bars of fixed thickness, and
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 14
7
8
9 the joint thicknesses are integer multiples of the bar thicknesses that comprise
10
the joint. Considerations for joint strength are not made here and are deferred
11
12 to future investigation.
13
14 In terms of the composite density for all bars, the elasticity tensor is then
15 modified as
16
17 C(ρ(z, p, ηc )) := ρ(z, p, ηc )Co . (14)
18
19 As evidenced by this equation, the entire geometry projection is encapsulated
20
21 in the calculation of the composite density ρ; from thereon, material properties
22 can be modified as functions of the effective density in various ways, akin to ma-
23
24 terial distribution methods.2 In other words, the geometry projection bridges
25 an explicit geometry representation and the density-based representation, and
26
27 hence it enjoys the benefits of density-based TO methods, such as the ability
28
to employ a fixed grid and standard finite element methods for the analysis,
29
30 efficient and robust off-the-shelf nonlinear programming algorithms for the op-
31
32 timization, and, as will be shown in the following section, a simplified sensitivity
33 analysis. We note that the method can be readily extended to other geometric
34
35 primitives, granted that a suitable signed distance (exact or approximated) that
36 is differentiable with respect to the design parameters exists.
37
38 Following Equation 14, we define the analysis problem as follows: find u ∈ U
39
40 such that
41 a(u(ρ(z, p, ηc )), v) = l(v), (15)
42
43 where
44
Z
45
46 a(u, v) := ∇v · C(ρ(z, p, ηc ))∇u dv (16)
47 ZΩ
48 l(v) := v · t ds (17)
49 Γt
50
for all v ∈ U. In the above, a(u, v) and l(v) are the bilinear and linear forms
51
52 respectively, Ω is the fixed design domain, Γt is the region on which the applied
53
54
55 2 It should be noted that in SIMP methods the penalization takes place on the density,
56 whereas in the proposed method it takes place on the weights αq .
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 15
7
8
9 traction t is applied, and U := {u ∈ H 1 (Ω) : u = 0 on ΓD } is the set of
10
admissible displacements, with ΓD the region on which homogeneous Dirichlet
11
12 boundary conditions are imposed. We assume no body loads are applied, and
13
14 both the applied tractions t and the regions Γt and ΓD are design independent.
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 16
7
8
9 3. Sensitivity Analysis
10
11 To employ efficient gradient-based optimization algorithms to solve the TO
12
13 problems, design sensitivities of the objective and constraint functions are re-
14
quired. These functions will in general have the form θ(u(z), z), where u is
15
16 the solution to the primal analysis, which here is that of Equation 15. To
17
18 efficiently compute the sensitivities of these responses, the adjoint sensitivity
19 analysis method is used (cf. [36]); this method is widely employed and docu-
20
21 mented in TO and will not be described here (cf., for example, [9]). What is
22 different with respect to density-based TO, is that here the design variables
23
24 are the bars’ locations and weights αq (i.e. zq ) instead of the discretized den-
25
26 sity field; therefore, the design sensitivities of the modified elasticity tensor are
27 computed via the chain rule as
28
29 dC(ρ(z, p, ηc )) ∂C(ρ(z, p, ηc )) ∂ρ(z, p, ηc )
30 = , (18)
dz ∂ρ ∂z
31
32 where, from Equation 14,
33 ∂C 1
34 = Co = C. (19)
∂ρ ρ
35
36 For the sake of conciseness, we omit function arguments in the rest of this
37 section, but they should be clear from the preceding one. From Equations 12
38
39 and 13 we find
40 N
41 ∂ρ X ∂ ρ̂q
= (overlap) (20)
42 ∂zj ∂zj
q=1
43
XN  p−1
44 ∂ρ ρ̂q ∂ ρ̂q
45 = (merge) (21)
∂zj q=1
ρ ∂zj
46
47
It is clear from the definition of the vector of design variables z and from the
48
49 definition of the bar effective density that ∂ ρ̃q /∂zj 6= 0 only when j ∈ {5(q −
50
51 1) + 1, . . . , 5(q − 1) + 4}, and ∂αq /∂zj 6= 0 only when j = 5q. Consequently,
52 from Equation 11 we have
53
 
54 ∂ ρ̂q ∂ ρ̃q ∂ρq ∂dq η
55 = δqdj/5e α (1 − δ(5q)j ) + η ρ˜q αq δ(5q)j ,
η−1
(22)
∂zj ∂ρq ∂dq ∂zj q
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 17
7
8
9 where δij is the Kronecker delta and d·e is the ceiling function. In the computer
10
implementation, of course, this means that these sensitivities are ‘assembled’
11
12 bar by bar. Continuing with the terms in Equation 22, from Equation 10, we
13
14 have
15 ∂ ρ̃q
= 1 − ρmin . (23)
16 ∂ρq
17 The sensitivity of the geometry projection of Equation 8 with respect to the
18
19 minimum distance is given by
20 
21  p
∂ρq − 2 2 (r2 − d2q ) if − r ≤ dq ≤ r
πr
22 = (24)
23 ∂dq 
0 otherwise,
24
25
where we considered from Equation 1 that ∂φq /∂zj = ∂dq /∂zj . It is worth em-
26
27 phasizing that we require that r be less than half the bar width, i.e. r < w/2 so
28
29 that ∂ω ∩ Brp cannot have two disjoint branches of ∂ω inside the sampling
30 window, in which case the geometry projection is incorrect and dq is non-
31
32 differentiable.
33
Up to this point, the calculation of the sensitivities is independent of the
34
35 parametric representation of the design. In fact, we posit that if one considers
36
37 the situation in which each bar is of zero length (so that it collapses to a circle
38 of radius w/2), and there is one collapsed bar centered at each element cen-
39
40 troid (or node) in the grid with its location fixed, and further that w/2 is, say,
41 the radius of the circle that circumscribes each corresponding finite element,
42
43 then we recover the density-based method without filtering.3 The geometry
44
representation appears only in the relationship between the minimum distance
45
46 dq and the design parameters z. Therefore, we must either choose a geometry
47
48 representation such that this relationship is differentiable, or remedy (e.g. via
49 regularization) non-differentiability issues (such as, for example, when there are
50
51 corners in ∂ω).
52
Design sensitivities of the minimum distance function of Equation 2 are given
53
54
55 3 One could of course modify the composite density function to effectively render a filter.
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 18
7
8
9 by
10 
11 
 − b if a · b ≤ 0


12
∂dq  ||b||
h i

13 = 1 1
(a ⊗ b)T + (a · b)I − I g if 0 < a · b < a · a (25)
∂xqo  ||g|| ||a||2
14 


0
15 if a · b ≥ a · a,
16
17 and
18
19 

 0 if a · b ≤ 0
20 

∂dq 
21 
= − (26)
||g|| ||a||2 (a ⊗ b) + (a · b)I g if 0 < a · b < a · a
1 T
22 ∂xqf 


 e
23  − ||e|| if a · b ≥ a · a.
24
25
26 Two things are worth noticing in these equations. The first is that the sensitiv-
27 ities are continuous across the branches of the function; indeed, when a · b = 0,
28
29 b = g, and we have ∂||g||/∂xqo = ∂||b||/∂xqo = b/||b|| and ∂||g||/∂xqf = 0;
30 likewise, when a · b = a · a, g = b − a = e, and we have ∂||g||/∂xqo = 0 and
31
32 ∂||g||/∂xqf = ∂||e||/∂xqf = e/||e||. The second is that when the point p from
33
which the minimum distance dq is being evaluated lies on the medial axis of
34
35 the bar, the second branch of the above equations applies; since the term ||g||
36
37 in the denominator equals zero, this sensitivity is not defined. However, since
38 we require that r < w/2, then by Equation 24 the sensitivity of the geometry
39
40 projection must necessarily be zero. This is expected since an infinitesimal mo-
41
tion of the bar will not alter the density (ρq = 1) of a sample window smaller
42
43 than the bar width and centered on the bar’s medial axis. Another situation
44
45 arises if the bar length collapses to zero, and the ||a|| term in the denominator
46 of the second branch of both expressions causes the sensitivity to be undefined;
47
48 in this case, however, only the first and third branches apply, in which case the
49 sensitivity is zero either by virtue of considering the branch with zero sensitivity,
50
51 or by considering once again that r < w/2 and applying Equation 24. Hence,
52
the sensitivities for the proposed geometry model and geometry projection exist
53
54 with respect to any design provided r < w/2.
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 19
7
8
9 4. Optimization Problem and Computer Implementation
10
11 We consider the classical TO problem of minimizing compliance subject to
12
13 a volume constraint:
14 R
15 minz θ(u(ρ(z, p, ηc ))) := Γt
u(ρ(z, p, ηc )) · t ds
16 subject to
17 V (z,ηv )
18 |Ω| − vf∗ ≤ 0
(27)
19 a(u(ρ(z, p, ηc )), v) = l(v)
20
21 xqo , xqf ∈ Ω
22
0 ≤ αq ≤ 1
23
24 where V (z, ηv ) is the volume of the design, |Ω| is the volume of the fixed design
25
26 domain, ηv is a volume penalization power, and vf∗ is a specified volume fraction.
27
In the case of overlapping bars, we utilize the explicit geometry representation
28
29 and compute the design volume as
30  
N
X N
X
31 πw2
V (z, ηv ) = vq := αqηv tq lq w + (overlap), (28)
32
q=1 q=1
4
33
34 wehre lq := ||a|| (cf. Equation 3) is the bar length. For the penalization to work
35
36 (i.e. for the optimizer to penalize intermediate values of αq ), it is necessary
37 that ηc > ηv , which is consistent with SIMP methods. Since the design volume
38
39 expressed in this way is not a function of the geometry projection and hence ρ,
40
41 we need not use the expressions of the previous section to obtain its sensitivities,
42 and instead obtain
43
44 ∂V ∂V wαqηv tq ∂V vq
=− =− a, = ηv . (29)
45 ∂xqo ∂xqf li ∂αq αq
46
47 We could also impose a constraint on the 2-dimensional perimeter
48 N
X N
X
49 P (z) = pq := αqηv (2lq + πw) (overlap), (30)
50 q=1 q=1
51
52 where the factor αq is introduced so that the perimeter of bars with zero thick-
53 ness is not counted towards the total perimeter. Perimeter constraints were
54
55 introduced in density-based TO ([37]) as a means to make the problem well-
56
posed and produce mesh-independent designs. Here the perimeter constraint
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 20
7
8
9 would have the same result—namely, to limit the number and length of the
10
bars in the optimal design, however the perimeter is very easily computed here
11
12 as opposed to methodologies to compute it from the parameterized density field.
13
14 Moreover, the complexity of the design is controlled by the number of bars.
15 In the case of merging bars, one could in principle also compute the design
16
17 volume as a function of z directly. However, computing the volume of inter-
18
sections between bars is far more complicated than the expression of Equation
19
20 28. Therefore, we choose to employ the usual way of computing volume in
21
22 density-based TO, i.e.
23 Z
24 V (ρ(z, p, ηv )) = ρ(z, p, ηv ) dv (merge). (31)
25 Ω
26 This is another significant difference with the work presented in [32], where the
27
28 use of the ηc = ηv caused significant difficulties in attaining a 0-1 design. The
29
sensitivities of the merged volume are computed from Equations 20–26.
30
31 Using adjoint sensitivity analysis (cf., for example, [9]), we find that the
32
33 compliance sensitivity with respect to ρ is given by
34
∂θ 1
35 = −u · Co u = − u · Cu (32)
36 ∂ρ ρ
37
One point worth making is that while the compliance decreases monotonically
38
39 with respect to the composite density ρ (since C is positive definite and hence
40
41 the right-hand side in the above equation is always less than zero), the design
42 sensitivities ∂θ/∂zj can have any sign, which precludes the use of optimality
43
44 criterion methods for the optimization.
45 The equilibrium constraint in Equation 27 is not passed as a constraint to
46
47 the optimizer, but enforced exactly at each iteration by performing a finite
48
element analysis to obtain u as customary in TO. The constraint on xqo and
49
50 xqf in Equation 27 is intended to guarantee that the bar endpoints stay inside
51
52 the design domain. In the case where Ω is a rectangular domain aligned with
53 the coordinate system, as in the examples presented in the following section,
54
55 then these constraints can be imposed as bounds on the individual components
56
of xqo and xqf . It should be noted, however, that in the case of overlapping
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 21
7
8
9 bars, this constraint can be removed from the optimization problem, since a
10
portion of a bar outside of the design domain will add to the volume but not
11
12 decrease the compliance. This might not be true, however, for other structural
13
14 criteria such as stress. In the case where Ω is not convex, a strategy is needed
15 to keep the bars inside the design domain; this is outside of the scope of this
16
17 work and is left for future investigation. However, a constraint based on the bar
18
volume outside of the design domain can be enforced, e.g., |ω ∩ Ωc | = 0, where
19
20 Ωc := R2 \ Ω.
21
22 The computer implementation of the method was done in Matlab
R
. The
23 finite element analysis is performed on a regular grid of square bilinear elements
24
25 of size h. To perform the geometry projection, we employ a sampling window
26 √
centered at each element centroid and with radius r = 2h/2 (i.e. the window
27
28 that circumscribes the element) and obtain an element composite density ρe . In
29
accordance with Equation 14, the fully effective element stiffness matrix keo is
30
31 multiplied by ρe . The compliance design sensitivity is then given by
32
33 X N
el X N
el
∂θ ∂θ ∂ρe ∂ρe
34 = = uTe keo ue , (33)
∂zj ∂ρe ∂zj ∂zj
35 e=1 e=1
36
37 where Nel is the number of elements in the mesh and ue is the vector of nodal
38 displacements for element e. A similar expression ensues for the volume of
39
40 merging bars:
41 N N
42 ∂V X el
∂V ∂ρe X el
∂ρe
= = ve (merge), (34)
43 ∂zj ∂ρe ∂zj ∂zj
e=1 e=1
44
45 where ve is the volume of element e and we recall from Equation 31 that we use
46
47 η = ηv for the volume calculation.
48 The optimization is solved using the method of moving asymptotes (MMA)
49
50 [38]. To improve convergence, we impose move limits on the design variables at
51
every iteration. To facilitate this, we scale the design variables xqo and xqf by
52
53 the corresponding dimensions of Ω,4 so that 0 ≤ ẑj ≤ 1, where ẑj denotes the
54
55
56 4 The αq do not require scaling.
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 22
7
8
9 scaled design variable. The move limits
10
11 (k−1)
max(0, ẑj
(k)
− m) ≤ ẑj
(k−1)
≤ min(1, ẑj + m) (35)
12
13
14 are considered with the other side constraints in Equation 27, where the super-
15
script denotes the current optimization iteration and 0 < m ≤ 1 is the move
16
17 limit.
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 23
7
8
9 5. Results
10
11 In this section we present numerical examples to demonstrate the proposed
12
13 method. In all examples, we use an isotropic elastic material with Young’s
14
modulus of E = 1E5 and Poisson’s ratio of ν = 0.3. The sample window
15 √
16 radius is r = 2h/2, where h is the element size; hence, the sample window
17
18 circumscribes the element. To impose the penalization we use ηc = 3 and ηv = 1,
19 consistent with SIMP methods. The density lower bound is ρmin = 1E−4, and
20
21 the p-norm power for computing the composite density in the case of merging
22 bars is p = 8. All bars have a thickness of t = 1, and initial αq = 0.5, which
23
24 facilitates the early removal of bars by the optimizer. Unless specified, we use
25
26 a move limit of m = 0.1. Finally, all examples are run until the relative change
27 in the objective function in subsequent iterations is less than 1E−4. The MMA
28
29 parameters we used are those employed for the numerical results in [39].
30
31 5.1. Two-bar Cantilever Beam
32
33 The first example is a simple one for which the optimal solution is known: a
34
35 square cantilever beam with a transverse load on the midpoint of the right edge
36 (cf. Figure 3a). The initial design is made of two bars of width w = 0.2 placed
37
38 inside the design envelope as shown in Figure 3b. For this example we assume
39
40 the bars can merge. The design domain mesh is 40 × 40.
41 The optimal design, which is what we expect for this problem, appears in
42
43 Figure 3c, and its corresponding composite density ρ appears in Figure 3d.
44 The bars in the optimal design are not filled so as to clearly illustrate their
45
46 boundaries (this will be more important in the subsequent examples). The
47
number of iterations to convergence, optimal design compliance and volume
48
49 fraction, the volume fraction constraint and move limit for all examples are
50
51 listed in Table 1. Also listed is the maximum composite density in the design
52 envelope; in the case of overlapping bars, it corresponds to the maximum number
53
54 of bars intersecting at any point in the design envelope (up to roundoff error).
55 In the case of merging bars, it should ideally correspond to 1.0 (since all bars
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 24
7
8
9 have t = 1.0). However, the p-norm of Equation 13 incurs an error, and hence
10
the maximum composite density is greater than 1.0. Increasing the value of
11
12 p brings the maximum density closer to 1.0, but it also introduces significant
13
14 ill-conditioning into the optimization problem, which makes convergence more
15 difficult; we found that reasonable results are obtained with p = 8.
16
17 Two aspects of this result are worth mentioning. The first is that, unlike
18
ground structure methods, the ends of the bars need not be connected to other
19
20 bar ends or to the loads or displacement boundary conditions, and in fact they
21
22 can ‘float’ inside the design envelope. This provides more freedom to the opti-
23 mizer to move and size the bars, and as we will show in subsequent examples,
24
25 allows the optimizer to lock into a reasonable topology in a few iterations. The
26 second aspect worth noting is that the volume fraction constraint is not active.
27
28 This is expected, since the design space is restricted to two bars of constant
29
width. We note this is different from free-form compliance-based TO, where the
30
31 volume fraction constraint is always active.
32
33 The compliance and volume fraction of this result can be compared to ana-
34 lytical results, corresponding to two bars connected to the each of the corners of
35
36 the left-hand side edge and to the midpoint of the right-hand side edge, respec-
37
tively. The compliance of this design is obtained using the expression for strain
38
39 energy of a bar under axial load (see, for example, [40]), U = P 2 E/(2AL), where
40
41 P is the axial load, L the bar length, and A its cross-sectional area. Since the
42 compliance equals twice the total strain energy, for the two bars we have θ = 4U .
43 √
44 From the bar geometry A = 0.2 and L = 5, and from equilibrium of forces
45 √ √
P = (F/2) 5 = 5 5. This amounts to θanalytical = 0.02795, and a correspond-
46
47 ing volume fraction vf = 2AL/4 = 0.2236. These values correspond closely to
48
those obtained with our method and listed in Table 1, namely θ = 0.0283 and
49
50 vf = 0.2114. The analytical compliance is slightly lower and volume fraction
51
52 slightly higher. We attribute these discrepancies to our method’s continuum
53 model versus the analytical uniaxial model, and to our bars with semicircular
54
55 ends versus the analytical rectangular bars.
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 25
7
8
9 1

10 0.9

11 0.8

12 0.7

13 0.6

14 0.5

15 0.4

16 0.3

17 0.2

18 0.1

19 0

20
21
22
23 (a) Problem definition (b) Initial design (the color denotes αq )
24
1
25 0.9 1

26 0.8

27 0.7
0.8

28 0.6
0.6

29 0.5

0.4
30 0.3
0.4

31 0.2
0.2

32 0.1

33 0 0

34
35
(c) Optimal design (color denotes αq ) (d) Optimal design (color denotes ρ)
36
37
38 Figure 3: Square beam example
39
40 Table 1: Optimal design values for all examples
41
42
43
44 Fig. Joints It. θ V (z)/|Ω| vf∗ m maxx∈Ω ρ(x)
45 3 merge 25 0.0283 0.2114 0.30 0.10 1.0653
46
47 5 merge 82 0.5405 0.2869 0.30 0.10 1.1892
48 11 merge 80 0.5313 0.2998 0.30 0.10 1.1710
49
50 13 overlap 148 0.5933 0.3003 0.30 0.10 2.0019
51
15 merge 94 0.5298 0.2997 0.30 0.10 1.1892
52
53 16 merge 55 0.5562 0.2802 0.30 0.05 1.1484
54
55 18 merge 54 0.4311 0.3470 0.35 0.05 1.1490
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 26
7
8
9 5.2. Long Cantilever Beam with Right Edge Top Load
10
11 The second example is more interesting in that it has more bars. It consists
12 of a cantilever beam clamped on its left edge, and with a transverse tip load on
13
14 the top right corner as shown in Figure 4a. The design space is made of 28 bars
15
16 of width w = 0.5, placed as shown in Figure 4b, and the design domain mesh
17 is 160 × 40. To illustrate how the optimization proceeds, we plot the design
18
19 at several iterations during the optimization in Figure 5. Bars that have been
20 removed (i.e. those with αq ≤ 0.01) or that are shorter than 0.1 are plotted with
21
22 a light gray color. The composite density for the optimal design is plotted in
23
Figure 6 and the iteration history of the objective and constraint functions is
24
25 plotted in Figure 7.
26
27 Several things are worth noting regarding this run. One is that despite the
28 complexity of the design space, the optimizer is able to lock into the optimal
29
30 topology in less than 30 iterations, and another that it is able to effectively
31 remove bars by making their αq ≈ 0 and/or by making their length close to
32
33 zero (i.e. collapsing them to a circle). This is an indication that the sensitivities
34
35 are guiding the optimizer quickly to a good solution. Of course, convergence
36 is dampened by the move limits, which are necessary for the stability of the
37
38 algorithm.5
39 Something else worth noting is that the members are for the most part
40
41 connected—i.e. there are no gaps between intersecting bars. This is driven by
42
the structural requirements, since we do not impose any geometric constraint
43
44 that enforces this. We encountered in our experiments, however, that in order
45
46 to attain this quality, the void region material must be sufficiently soft, i.e. ρmin
47 must not be too large, as otherwise gaps will be stiff enough to transfer load; on
48
49 the other hand, we recall that ρmin cannot be arbitrarily small lest the analysis
50 become ill-posed, and this is particularly important in early iterations, when
51
52
53 5 This strategy is common in density-based methods (cf. for example [9]). In level set
54
methods the need to satisfy the Courant-Friedrichs-Lewy (CFL) condition in the solution of
55
56 the Hamilton-Jacobi equation acts in effect as a move limit.
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 27
7
8
9 there may not be a connected path of bars between the loads and the supports.
10
Our experience indicates that the value ρmin = 1E−4 is appropriate. As shown
11
12 in the sequel, the design obtained with our method attains a compliance value
13
14 comparable to those obtained with the SIMP method, which has a much smaller
15 bound since ρmin is penalized and hence effective densities of, e.g., (1E−3)3 =
16
17 1E−9 are attained. On the other hand, we note that other methods (for instance,
18
the level set method in [4]) employ an effective lower bound of 1E−3 and produce
19
20 comparable results to those of SIMP.
21
22 Another observation that must be made is that the optimizer has the freedom
23 to offset two (or more) bars to increase the effective width and increase the
24
25 bending stiffness (see Figures 8a and 8b). This can be seen, for example, in the
26 bottom members on the left edge of the last iteration in Figure 5. Similarly,
27
28 in order to produce a longer member or a curved load path, the optimizer
29
‘concatenates’ bars (in what looks like sausage string, see Figures 8c and 8d).
30
31 It should be possible to circumvent these features by imposing constraints on
32
33 the intersections between, length of, and angles between bars; this is outside of
34 the scope of this paper, however, and is deferred to future work.
35
36 It should be noted that the design produced by our method in the merged
37
bars model is not unique, as it can render one bar that engulfs another, cf.
38
39 Figures 8e and 8f. In such cases, the effective thickness is the maximum thickness
40
41 amongst the two bars, and the sensitivity of the composite density with respect
42 to a change in the position of the engulfed bar’s endpoint along its medial axis
43
44 is zero.6 This poses no problem for the interpretation, however, as one would
45 simply choose the larger bar in this case; one may also combine two bars into a
46
47 single bar in the case where they are collinear but one is not entirely contained
48
in the other (indeed, this was done to create the beam models presented in the
49
50 sequel). These joining operations could be done in an automated manner, but
51
52 this is outside of the scope of this paper and deferred to future work.
53
54 6 In practice, the sensitivity is not exactly zero because of the approximation of the maxi-
55
56 mum thickness via the p-norm.
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 28
7
8
9
10
11
12
13
14
15
16
17
18
19
20 (a) Problem definition
21
22 1
23
24
0.5
25
26
27 0
28
29
30 (b) Initial design (the color denotes αq )
31
32 Figure 4: Cantilever beam example
33
34
35 Finally, it is worth noting that the convergence behavior of this and all other
36 examples is good. In particular, the number of iterations for most runs is less
37
38 than 100, which is on par with density-based methods, and is significantly less
39
40 than most level set-based methods.
41 We compare the results of this optimization to those obtained using element
42
43 density variables by using the code in [41] is employed to produce these results,
44 wherein a filter radius of r = w/2 = 0.125 is imposed to produce a design of
45
46 similar complexity to that obtained with our method. For the 60 × 40 mesh,
47
however, this radius exactly inscribes the element, and therefore the correspond-
48
49 ing result suffers from extensive checkerboarding. Therefore, we use a slightly
50
51 larger radius of 0.1375 (i.e. 1.1 element sizes) to produce the design appearing
52 in Figure 9a. The compliance for this design is greater than the one produced in
53
54 our method, possibly because our approximation of the maximum bar thickness
55 via the p-norm renders an effective density greater than 1.0, and possibly from
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 29
7
8
9
10
11 1 1

12
0.5 0.5
13
14 0 0

15
16 (a) Iteration 1 (b) Iteration 5
17
1 1
18
19 0.5 0.5

20
21 0 0

22
23 (c) Iteration 10 (d) Iteration 15
24 1 1

25 0.8

0.6
26 0.4
0.5

27 0.2

0 0
28
29 (e) Iteration 20 (f) Iteration 30
30
31 1 1

32 0.5 0.5

33
34 0 0

35
36 (g) Iteration 40 (h) Iteration 82
37
38 Figure 5: Cantilever beam optimization history (the color denotes αq )
39
40
41
42
43
44
45
1
46
47
48 0.5

49
50 0

51
52
53
54 Figure 6: Cantilever beam composite density for optimal design
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 30
7
8
9 3
10
10
11 Objective Function 10
2

12 1
10
13
14 10
0

15 −1
10
16 0 10 20 30 40 50 60 70 80
Iteration
17
18
19 0.3
Volume Fraction

20 0.25

21 0.2

22 0.15

23 0.1

24 0.05
0 10 20 30 40 50 60 70 80
Iteration
25
26
27
Figure 7: Cantilever beam objective and constraint history
28
29
30
the increased filter radius in the SIMP design.
31
32 We also create an ‘interpretation’ of our design using beam elements. The
33
34 locations of the nodes are averaged at some joints; at other joints the location
35 is approximated in an ad hoc manner (for example, bar centers near the edges
36
37 are simply moved to the edge); the beam model for this example is shown in
38 Figure 9c. The purpose of this analysis is to demonstrate that a beam model
39
40 can be readily created from our design that renders comparable compliance and
41
volume. The analysis is done using beam elements of constant cross section (of
42
43 in-plane width w = 0.2 and out-of-plane thickness t = 1.0) using the commercial
44
45 finite element code ABAQUS R
. The beam model has a compliance and vol-
46 ume fraction that are 12% and 11% greater than our design respectively. The
47
48 increased volume is due to the facts that the beam model does not have the
49
semicircular ends, that the size variable αq for some of the bars in our design is
50
51 less than 1.0, and possibly from interpretation errors. The increased compliance
52
53 is attributed to the error in the approximation of the maximum density via the
54 p-norm in the geometry projection design, and to interpretation errors.
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 31
7
8
9
10
11
12
13
14 1

15 0.8 1
0.6
16 0.4 0.5

17 0.2
0
18 0

19
(a) (b)
20 1

21 0.8 1

22 0.6

0.5
0.4
23 0.2
0
24 0

25
26 (c) (d)
1
27 0.8 1
28 0.6

29 0.4 0.5

0.2
30 0
0

31
32 (e) (f)
33 1

0.8
34 0.6
1

35 0.4 0.5

36 0.2
0
0
37
38 (g) (h)
39 1
2
40 0.8
1.5
0.6
41 0.4
1

0.5
42 0.2
0
43 0

44
(i) (j)
45
46
47 Figure 8: Size αq and composite density ρ for different types of bar intersections: a) and b)
48 merged bars with in-plane width offset; c) and d) concatenated merged bars; e) and f)
49 collinear merged bars; g) and h) in-plane width stacking of merged bars; and j) and k)
50 out-of-plane thickness stacking of overlapped bars
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 32
7
8
9 1

10
0.5
11
12 0

13
14 (a) (b)
15
16
17
18
19
20 (c)
21
22
Figure 9: Comparison of cantilever beam designs: a) element density result: θ = 0.622,
23
24 vf = 0.303; b) geometry projection method result: θ = 0.541, vf = 0.287; and c) beam
25 model interpretation: θ = 0.606, vf = 0.333
26
27
28
29
30
31
32
33
34
35
36 Figure 10: MBB beam problem definition
37
38
39 5.3. MBB Beam
40
41 Our following example corresponds to the well-known Messerschmitt-Bölkow-
42
43 Blohm (MBB) beam. The loading and boundary conditions are depicted in
44 Figure 10. Since the analysis is symmetric with respect to the center plane of
45
46 the beam (perpendicular to the paper, indicated by a dashed line in Figure 10),
47
we model only the right hand side of the beam. The design space and mesh for
48
49 the half-beam are the same as the previous example. The goal of this example
50
51 is to show the difference between using merged and overlapped bars, and also
52 to exemplify the existence of local minima.
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 33
7
8
9
10 1
11
12
0.5
13
14
15 0

16
17 (a) Composite density
18
19 1
20
21
22 0.5

23
24
0
25
26
27 (b) Bar layout (the color denotes αq )
28
29 Figure 11: Optimal design for MBB beam example (merge)
30
31
32 5.3.1. Merged Bars
33
34 We start by running the problem with merged beams. The optimized com-
35 posite density and layout of bars are shown in Figures 11a and 11b respectively.
36
37 This result is consistent with known results for the MBB beam (cf., for example,
38
[9]). We observe that the in-plane stacking of the widths of the bars (such as
39
40 the one depicted in Figures 8g and 8h) and the concatenation of bars is more
41
42 pronounced than in the previous example. As before, we compare the results of
43 this example with those obtained with the SIMP method, and with an interpre-
44
45 tation made of 1-dimensional beam elements; the corresponding results appear
46 in Figure 12.
47
48
49 5.3.2. Overlapped Bars
50
51 We now run the MBB problem using overlapping bars, and recall that in this
52 case the volume is computed directly from the design parameters via Equation
53
54 28. The composite density and bar layout for the optimized design are shown
55
in Figures 13a and 13b respectively. The first observation we make is that, as
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 34
7
8
9 1

10 0.5

11
12 0

13
14 (a) (b)
15
16
17
18
19 (c)
20
21
Figure 12: Comparison of MBB beam designs: a) element density result: θ = 0.622,
22
23 vf = 0.303; b) geometry projection method result: θ = 0.531, vf = 0.3; and c) beam model
24 interpretation: θ = 0.634, vf = 0.324
25
26
27 can be seen from Table 1, the merged bars design is about 12% stiffer than this
28
29 design. That means that, at least in this example, the optimizer can find a stiffer
30 design by taking the extra material in the joints of the overlapped bar design
31
32 and distributing it more effectively. Qualitatively, we see that the ends of the
33
bars in this example do not overlap other bars as is the case in the merged bar
34
35 design, and we see some gaps between bars. This is due to the added stiffness
36
37 arising from the addition of the effective densities for all bars (cf. Equation 12).
38 Indeed, as seen in the composite density plot of Figure 13a, there are higher
39
40 composite density regions (ρ ≈ 2) near the ends of concatenated bars. Also, we
41 note that the optimizer does not stack bars in-plane for increased width as in
42
43 the previous example, but in this case it gets the added stiffness by stacking the
44
bars out-of-plane (i.e. by stacking their thicknesses, as shown in Figures 8i and
45
46 8j), which it does on the bottom of the beam. Although the bending stiffness
47
48 of a bar, which is proportional to its area moment of inertia I = tw3 /12, favors
49 an increase in w, the bending stiffness of the entire MBB beam increases by
50
51 placing more material along the top and bottom edges, and therefore out-of-
52
plane thickness stacking of bars along the bottom edge is more desirable than
53
54 in-plane width stacking.
55
56 Finally, another interesting observation is that in the merged bar design some
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 35
7
8
9
2
10
11 1.5
12 1
13
0.5
14
15 0

16
17 (a) Composite density
18
19 1
20
21
22 0.5

23
24
0
25
26
27 (b) Bar layout (the color denotes αq )
28
29 Figure 13: Optimal design for MBB beam example (overlap)
30
31
32 of the bars go slightly outside the design envelope, whereas in the overlapped bar
33
34 design they stay inside. This is due to the fact that the volume of the overlapped
35 bars is computed from the geometry of the bars (cf. Equation 28) and so any
36
37 portion of the bar outside of the design domain contributes to the volume but
38 not to the stiffness and hence is inefficient, which causes the optimizer to bring
39
40 the bar inside the design domain. In the case of merged bars, on the other
41
42 hand, the volume is computed from the densities (cf. Equation 31), and hence
43 portions of the bars outside the design domain do not contribute to the volume
44
45 and consequently they incur a penalty. Note that the endpoints of the bars
46 themselves do stay inside the design envelope because of the side constraints.
47
48 To the best of the authors’ knowledge, the problem with overlapping bars cannot
49
be solved with existing methods. Figure 14 depicts a beam model of this result in
50
51 which the joints with double thickness are modeled using shorter beam elements
52
53 whose out-of-plane thickness is 2.
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 36
7
8
9
10
11
12
13
14
15
16 (a)
17
18 (b)
19
20 Figure 14: Comparison of MBB beam designs (the element density design appears in Figure
21
12a): a) geometry projection method result: θ = 0.593, vf = 0.3; and b) beam model
22
23 interpretation: θ = 0.726, vf = 0.338
24
25
26 5.3.3. Initial Design Dependency
27
Since the initial design in our method is a ground structure, it is natural
28
29 to expect that there is dependence of the optimal design on the initial design.
30
31 We start by noting that in ground structure methods even if the joint locations
32 are allowed to change during the optimization, the initial connectivity among
33
34 bars is maintained throughout the optimization. And as aforementioned, that
35 is not the case in our method since the bars move independently in space, and
36
37 hence it is fair to expect that our method does not depend on the initial design
38
39 as much as ground structure methods. To illustrate this, we repeat the MBB
40 beam example with a different initial design, depicted in Figure 15a, which has
41
42 the same number of bars as that of Figure 4b.
43 The optimal design, shown in Figure 15b, is very similar to that of Figure
44
45 11b, and from Table 1 we see that their compliances are within 0.2%, which
46
confirms the fact that they are local minima. However, we also see that it took
47
48 23 more iterations for the optimizer to converge for this design, which indicates
49
50 this initial design is ’farther’ from the optimal design than that in Figure 4b.
51 But more importantly, this shows our method is robust, as it does not appear
52
53 to be dependent on the initial design as one may initially think.
54 We are, however, able to obtain a different optimal design from the ones
55
56 obtained so far by doing something different: decreasing the move limit from
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 37
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 1
22 0.8
23
24 0.6
25
26 0.4
27 0.2
28
29 0
30
31 (a) Initial design
32 1
33
34 0.8
35
0.6
36
37 0.4
38
39 0.2
40
0
41
42 (b) Optimal design bar layout
43
44 Figure 15: MBB beam example with different initial design (merged; the color denotes αq )
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 38
7
8
9 1
10
11 0.8
12 0.6
13
14 0.4
15
16 0.2
17 0
18
19 Figure 16: MBB beam example with move limit m = 0.05 (merged; the color denotes αq )
20
21
22
23
24
25
26
27
28
29 Figure 17: Beam model interpretation of the optimization result in Figure 16 (the element
30 density design appears in Figure 12a): θ = 0.653, vf = 0.347
31
32
33 0.1 to 0.05. By doing this, we effectively impose a different design trajectory for
34
35 the optimizer which leads to a different design. Starting again from the initial
36
design of Figure 4b, we obtain the design shown in Figure 16. This design is
37
38 richer than the ones previously obtained, and also matches well known results
39
40 in the literature. It has a slightly greater compliance than the previous results
41 (about 5% greater than the design with m = 0.1), and some of the bars have
42
43 intermediate values of αq . We think the orientation and location of some of the
44 removed bars indicate the optimization is converging to a design with a slightly
45
46 greater volume fraction constraint. It is worth noting that, as seen in Table
47
1, this example shows that decreasing the move limit does not necessary lead
48
49 to more iterations, possibly because controlling the extent of each design step
50
51 might lead the optimizer towards a better design earlier in the optimization
52 thereby decreasing the total number of iterations. A beam model of this result
53
54 is shown in Figure 17.
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 39
7
8
9 5.4. Multiple Length Scales
10
11 Our final example concerns the design of a structure with multiple fixed
12 length scales. While different length scales naturally arise in topology optimiza-
13
14 tion (see for example [9]), to our knowledge it is not possible to enforce multiple
15
16 constant size for all members in existing methods. Here, we design a cantilever
17 beam with two sets of merging bars of widths w1 = 0.25 and w2 = 0.75 in
18
19 the initial design, as depicted in Figure 18a. The problem is similar to that
20 described in Figure 4a, however the load is applied on the midpoint of the right
21
22 edge. We use a 160 × 40 mesh, and we enforce a move limit of m = 0.05 and a
23
volume fraction constraint of vf∗ = 0.35. The optimal design appears in Figure
24
25 18b. Two element density designs are provided, resulting from filters of radii
26
27 0.125 (corresponding to the thinner bar width of 0.25), and 0.375 (correspond-
28 ing to the thicker bars width of 0.75), as well as a beam model interpretation;
29
30 all results are shown in Figure 19. As is expected, our method produces a de-
31 sign with lower compliance than that of the element density design with filter
32
33 size of r = 0.375. However, our design is also better than the element density
34
35 design with r = 0.125, possibly due to the reasons mentioned in Section 5.2.
36 Our design is not symmetric, as we expect for this problem; this is possibly due
37
38 to locking into a local minimum during the optimization.
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 40
7
8
9
10
11 1
12
13
14 0.5
15
16
17
0
18
19
20 (a) Initial design
21
22 1
23
24 0.8
25 0.6
26
0.4
27
28 0.2
29
0
30
31
32 (b) Optimal design bar layout
33
34 Figure 18: Cantilever beam with two member sizes (merge; the color denotes αq )
35
36
37
38
39
40
41
42
43 (a) (b)
44
45
46
47
48
49 (c)
50
51 Figure 19: Comparison of cantilever beam designs: a) element density design with filter
52
radius r = 0.125: θ = 0.479, vf = 0.341; b) element density design with filter radius
53
54 r = 0.375: θ = 0.643, vf = 0.348; and c) two-scale beam model interpretation: θ = 0.484,
55 vf = 0.387
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 41
7
8
9 6. Conclusions
10
11 We have demonstrated a method for continuum-based topology optimiza-
12
13 tion of structures made of discrete elements. It uses a differentiable projection
14
of the geometry of the structure onto a fixed analysis grid to support simplified
15
16 analysis for any design and provide continuous sensitivities for the optimization,
17
18 thereby inheriting the main advantages of density-based topology optimization.
19 By using the chain rule, sensitivities of structural responses with respect to the
20
21 explicit geometry parameterization are readily derived. This means the geome-
22 try projection can be used in conjunction with standard nonlinear programming
23
24 algorithms, which accommodates multiple constraints. Another advantage of
25
26 our approach is its ability to modify material properties and penalize interme-
27 diate size members much like in SIMP methods. In fact, as discussed in Section
28
29 3, the SIMP method with filtered densities may be seen as a special case of
30 the method proposed here. This means that the geometry projection can be
31
32 coded in a highly modular way, and it can be readily combined with existing
33
methodologies for density-based topology optimization.
34
35 By having an independent explicit geometry, our method is capable of pro-
36
37 ducing designs with members of predefined cross-sections. Moreover, it can
38 accommodate the case where the thicknesses of the bars are added at their in-
39
40 tersections (overlapping bars), and when it does so, it can compute the volume
41 fraction and its sensitivities directly from the bar’s geometry and not from the
42
43 composite density. Unlike ground structure methods, the discrete elements in
44
45 our method need not be connected during the optimization, hence giving more
46 freedom to the optimizer to optimally place and size the bars. These capabili-
47
48 ties, to the best of the authors’ knowledge, do not exist in current density-based
49 TO methodologies. In level set TO methodologies the previously cited work in
50
51 [30] can attain some of these characteristics such as constant thickness mem-
52
bers,7 but not the ability to consider overlapping bars, and more importantly,
53
54
55 7 In [30] the width of the bars is a variable and so the members in the structure are not of
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 42
7
8
9 to remove bars from the design space. Moreover, as mentioned in Section 1,
10
the boundaries for the designs appearing in [30] are not smooth, which may be
11
12 a challenge if stress constraints were to be considered. We believe our method
13
14 can be in principle extended to incorporate stress constraints by using existing
15 density-based techniques for stress-based topology optimization (for example,
16
17 [42]); however, if the bars in any given design iterate do not provide at least
18
one connected load path between the loads and the supports, the design will
19
20 likely exhibit high localized stress regions that will adversely affect the opti-
21
22 mization convergence. A similar challenge will arise in design problems with
23 large deformations. These will be subjects for future investigation.
24
25 The presented examples show that the method produces designs consistent
26 with well known topologies in the literature, and it does so without the need for
27
28 tuning of numerical parameters (with the exception perhaps of the move limit,
29
which is also often modified in density-based methods). The convergence of our
30
31 method is comparable to that of density-based methods for the case of merged
32
33 bars, which is substantially better than most level set methods.
34 The most immediate and necessary direction for future work is the imposi-
35
36 tion of geometric constraints to achieve desired characteristics. These include
37
constraints on the bar’s lengths, constraints to eliminate the stacking of bars
38
39 (in- or out-of-plane), the formation of strings of short concatenated bars that
40
41 produce curved load paths, and bars encroaching a non-designable regions when
42 the design envelope is not convex, etc. Other extensions are the consideration
43
44 of 3-dimensional problems, the incorporation of structural responses other than
45 compliance, and the exploration of other geometry models.
46
47
48
49
50
51
52
53
54
55 equal size, but it is possible to fix the bars width to attain constant size members as we do
56
here. On the other hand, we can add the bar width w as a design variable as in [30].
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 43
7
8
9 References
10
11 [1] M. P. Bendsøe, Optimal shape design as a material distribution problem,
12
13 Structural Optimization 1 (4) (1989) 193–202.
14
15 [2] M. Zhou, G. Rozvany, The COC algorithm, part II: topological, geomet-
16
17 rical and generalized shape optimization, Computer Methods in Applied
18 Mechanics and Engineering 89 (1) (1991) 309–336.
19
20 [3] G. Allaire, F. Jouve, A.-M. Toader, A level-set method for shape optimiza-
21
22 tion, Comptes Rendus Mathématique 334 (12) (2002) 1125–1130.
23
24 [4] G. Allaire, F. Jouve, A.-M. Toader, Structural optimization using sensi-
25
26 tivity analysis and a level-set method, Journal of Computational Physics
27 194 (1) (2004) 363–393.
28
29
30 [5] M. Y. Wang, X. Wang, D. Guo, A level set method for structural topology
31 optimization, Computer Methods in Applied Mechanics and Engineering
32
33 192 (1) (2003) 227–246.
34
35 [6] N. P. van Dijk, K. Maute, M. Langelaar, F. Van Keulen, Level-set methods
36
for structural topology optimization: a review, Structural and Multidisci-
37
38 plinary Optimization 48 (3) (2013) 437–472.
39
40 [7] O. Sigmund, Design of multiphysics actuators using topology optimization–
41
42 part I: One-material structures, Computer Methods in Applied Mechanics
43
and Engineering 190 (49) (2001) 6577–6604.
44
45
46 [8] G. Rozvany, M. Bendsøe, U. Kirsch, Layout optimization of structures,
47 Applied Mechanics Reviews 48 (2) (1995) 41–119.
48
49
[9] M. P. Bendsøe, O. Sigmund, Topology optimization: theory, methods and
50
51 applications, Springer, 2003.
52
53 [10] H. A. Eschenauer, V. V. Kobelev, A. Schumacher, Bubble method for topol-
54
55 ogy and shape optimization of structures, Structural Optimization 8 (1)
56
(1994) 42–51.
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 44
7
8
9 [11] T. Lewiński, J. Sokolowski, Energy change due to the appearance of cavities
10
in elastic solids, International Journal of Solids and Structures 40 (7) (2003)
11
12 1765–1803.
13
14 [12] O. Sigmund, On the design of compliant mechanisms using topology opti-
15
16 mization, Journal of Structural Mechanics 25 (4) (1997) 493–524.
17
18 [13] T. E. Bruns, D. A. Tortorelli, Topology optimization of non-linear elas-
19
20 tic structures and compliant mechanisms, Computer Methods in Applied
21
Mechanics and Engineering 190 (26) (2001) 3443–3459.
22
23
24 [14] B. Bourdin, Filters in topology optimization, International Journal for Nu-
25 merical Methods in Engineering 50 (9) (2001) 2143–2158.
26
27
28 [15] J. K. Guest, J. Prévost, T. Belytschko, Achieving minimum length scale
29 in topology optimization using nodal design variables and projection func-
30
31 tions, International Journal for Numerical Methods in Engineering 61 (2)
32 (2004) 238–254.
33
34
35 [16] M. Zhou, R. Fleury, Y.-K. Shyy, H. Thomas, J. Brennan, Progress in topol-
36 ogy optimization with manufacturing constraints, in: Proceedings of the
37
38 9th AIAA MDO conference AIAA-2002-4901, 2002.
39
40 [17] O. Sigmund, Manufacturing tolerant topology optimization, Acta Mechan-
41
42 ica Sinica 25 (2) (2009) 227–239.
43
44 [18] J. K. Guest, Imposing maximum length scale in topology optimization,
45
Structural and Multidisciplinary Optimization 37 (5) (2009) 463–473.
46
47
48 [19] A. R. Gersborg, C. S. Andreasen, An explicit parameterization for cast-
49 ing constraints in gradient driven topology optimization, Structural and
50
51 Multidisciplinary Optimization 44 (6) (2011) 875–881.
52
53 [20] J. K. Guest, M. Zhu, Casting and milling restrictions in topology optimiza-
54
55 tion via projection-based algorithms, in: ASME 2012 International Design
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 45
7
8
9 Engineering Technical Conferences and Computers and Information in En-
10
gineering Conference, American Society of Mechanical Engineers, 2012, pp.
11
12 913–920.
13
14 [21] G. Cheng, Y. Mei, X. Wang, A feature-based structural topology optimiza-
15
16 tion method, in: IUTAM Symposium on Topological Design Optimization
17 of Structures, Machines and Materials, Springer, 2006, pp. 505–514.
18
19
[22] A. Saxena, Are circular shaped masks adequate in adaptive mask overlay
20
21 topology synthesis method?, Journal of Mechanical Design 133 (1) (2011)
22
23 011001.
24
25 [23] F. Wang, J. S. Jensen, O. Sigmund, High-performance slow light photonic
26 crystal waveguides with topology optimized or circular-hole based mate-
27
28 rial layouts, Photonics and Nanostructures-Fundamentals and Applications
29
10 (4) (2012) 378–388.
30
31
32 [24] W. Zhang, L. Xia, J. Zhu, Q. Zhang, Some recent advances in the integrated
33 layout design of multicomponent systems, Journal of Mechanical Design
34
35 133 (10) (2011) 104503.
36
37 [25] J. Chen, V. Shapiro, K. Suresh, I. Tsukanov, Shape optimization with topo-
38
logical changes and parametric control, International Journal for Numerical
39
40 Methods in Engineering 71 (3) (2007) 313–346.
41
42 [26] M. Freytag, V. Shapiro, I. Tsukanov, Field modeling with sampled dis-
43
44 tances, Computer-Aided Design 38 (2) (2006) 87–100.
45
46 [27] S. Chen, M. Y. Wang, A. Q. Liu, Shape feature control in structural topol-
47
ogy optimization, Computer-Aided Design 40 (9) (2008) 951–962.
48
49
50 [28] M. Zhou, M. Y. Wang, Engineering feature design for level set based struc-
51 tural optimization, Computer-Aided Design 45 (12) (2013) 1524–1537.
52
53 [29] X. Guo, W. Zhang, W. Zhong, Explicit feature control in structural topol-
54
55 ogy optimization via level set method, Computer Methods in Applied Me-
56
chanics and Engineering 272 (2014) 354–378.
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 46
7
8
9 [30] X. Guo, W. Zhang, W. Zhong, Doing topology optimization explicitly and
10
geometrically—a new moving morphable components based framework,
11
12 Journal of Applied Mechanics 81 (8) (2014) 081009.
13
14 [31] X. Guo, W. Zhang, W. Zhong, Topology optimization based on moving
15
16 deformable components: A new computational framework, arXiv preprint
17
18 arXiv:1404.4820.
19
20 [32] B. Bell, J. Norato, D. Tortorelli, A geometry projection method for
21
continuum-based topology optimization of structures, in: 12th AIAA Avia-
22
23 tion Technology, Integration, and Operations (ATIO) Conference and 14th
24
25 AIAA/ISSMO Multidisciplinary Analysis and Optimization Conference,
26 2012.
27
28
[33] P.-O. Persson, G. Strang, A simple mesh generator in Matlab, SIAM
29
30 Review 46 (2) (2004) 329–345.
31
32 [34] J. Norato, R. Haber, D. Tortorelli, M. P. Bendsøe, A geometry projec-
33
34 tion method for shape optimization, International Journal for Numerical
35
Methods in Engineering 60 (14) (2004) 2289–2312.
36
37
38 [35] J. A. Norato, M. P. Bendsøe, R. B. Haber, D. A. Tortorelli, A topological
39 derivative method for topology optimization, Structural and Multidisci-
40
41 plinary Optimization 33 (4-5) (2007) 375–386.
42
43 [36] D. A. Tortorelli, P. Michaleris, Design sensitivity analysis: overview and
44
45 review, Inverse problems in Engineering 1 (1) (1994) 71–105.
46
47 [37] R. Haber, C. Jog, M. P. Bendsøe, A new approach to variable-topology
48
49 shape design using a constraint on perimeter, Structural Optimization
50 11 (1-2) (1996) 1–12.
51
52
53 [38] K. Svanberg, The method of moving asymptotes—a new method for struc-
54 tural optimization, International Journal for Numerical Methods in Engi-
55
56 neering 24 (2) (1987) 359–373.
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6 47
7
8
9 [39] K. Svanberg, A class of globally convergent optimization methods based
10
on conservative convex separable approximations, SIAM Journal on Opti-
11
12 mization 12 (2) (2002) 555–573.
13
14 [40] F. P. Beer, E. R. Johnston, J. T. DeWolf, Statics and mechanics of mate-
15
16 rials, McGraw-Hill, 2011.
17
18 [41] O. Sigmund, A 99 line topology optimization code written in matlab, Struc-
19
20 tural and Multidisciplinary Optimization 21 (2) (2001) 120–127.
21
22 [42] C. Le, J. Norato, T. Bruns, C. Ha, D. Tortorelli, Stress-based topology
23
24 optimization for continua, Structural and Multidisciplinary Optimization
25 41 (4) (2010) 605–620.
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65

You might also like