AN EFFECTIVE METHOD OF SYNTHESIZING ... - Semantic Scholar

1 downloads 0 Views 666KB Size Report
University of Michigan, Ann Arbor, MI 48109-2125, USA ... G.G. Brown, 2350 Hayward, Ann Arbor, MI 48105-2125, USA. ...... end (Howell and Midha, 1995).
Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

AN EFFECTIVE METHOD OF SYNTHESIZING COMPLIANT ADAPTIVE STRUCTURES USING LOAD PATH REPRESENTATION Kerr-Jia Lu* and Sridhar Kota Department of Mechanical Engineering University of Michigan, Ann Arbor, MI 48109-2125, USA

ABSTRACT Synthesis of shape morphing compliant mechanism is inherently different from typical single output design problems, due to the multiple output points along the morphing boundary. We have previously developed a genetic algorithm (GA) based synthesis approaches, incorporating a binary ground structure parameterization, to systematically design shape morphing compliant mechanisms. However, the approach is ineffective due to issues such as the generation of disconnected structures and the need to choose an initial mesh. In this paper, we will present the ‘load path representation,’ which is developed to overcome the issues encountered using the binary ground structure parameterization. The performance of the load path approach over the binary ground structure approach will be demonstrated through several design examples. The results have shown that the load path approach offers several advantages, such as (a) eliminating the need of an initial ground structure, (b) ensuring structural connectivity, and (c) yielding solutions that generate desired shape change efficiently. Key Words: compliant mechanism, adaptive shape morphing, load path representation, binary ground structure, genetic algorithm

*

[email protected]; tel: 734-763-4916; fax: 734-647-3170; Department of Mechanical Engineering, 2231 G.G. Brown, 2350 Hayward, Ann Arbor, MI 48105-2125, USA.

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

INTRODUCTION

A compliant mechanism is a single-piece flexible structure that is designed to be flexible to transmit motions, yet stiff enough to withstand external loads. Various synthesis approaches (Ananthasuresh, Kota and Kikuchi, 1994; Frecker et al., 1997; Hetrick and Kota, 1999; Larsen, Sigmund and Bouwstra, 1997; Saggere and Kota, 1999; Tai and Chee, 2000), utilizing structural optimization techniques, have been developed in the past decade to design practical devices such as motion amplifiers (Hetrick and Kota, 1999) in MEMS and compliant grippers (Joo, Kota and Kikuchi, 2000). One unique feature of compliant mechanisms is its smooth deformation field, due to the distributed compliance, that provides a novel means to morph structural shape. However, unlike the single output problems studied in most of the previous research, shape morphing involves deforming the structural boundary, where every point is an output point, from its initial profile to a desired target shape. Development of a new synthesis approach is, therefore, required to address the multiple output points presented in such type of problem (Lu and Kota, 2003).

In a previous paper (Lu and Kota, 2003), we have developed a genetic algorithm (GA) based synthesis approach for shape morphing compliant mechanisms, using a binary ground structure parameterization. As shown in Figure 1 and Figure 2, the ‘binary ground structure approach’ starts with a preprocessor to estimate problem feasibility and identify several candidate locations along the morphing boundary to be the ‘output points’ of the compliant mechanism. The design domain is then parameterized using a ground structure

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

of beam elements (Figure 2(b)). Each beam element is described by two design variables: one binary variable determining the presence of an element (topology), and one real variable defining the cross section area (dimension). The binary topology variable distinguishes itself from traditional ground structure optimization approach, hence the name ‘binary ground structure.’ Although the results have demonstrated feasibility of the synthesis approach, it is unclear how the initial ground structure is selected (the mesh configuration and resolution in Figure 2(b)). Furthermore, the representation using binary variables actually includes invalid structures that are disconnected from input or ground support in the design space, as shown in Figure 3. An additional verifying algorithm is thus required to ‘search for’ and penalize the disconnected designs, leading to inefficiency in the overall synthesis process. Moreover, due to the direct correspondence between the design variables and the FE elements, the number of design variables will increase dramatically when a more refined initial mesh is used.

Figure 1. Flowchart for both GA based synthesis approaches using binary ground structure and load path representation.

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

Figure 2. Compliant mechanism synthesis using the binary ground structure shown in Figure 1; (a) problem specification, (b) design domain parameterization, and (c) the hypothetical design obtained from the GA.

Figure 3. Three types of invalid structures that can be generated using the binary ground structure shown in Figure 2(b).

In this paper, we have developed a ‘load path representation’ to address the issues associated with the binary ground structure. The load path representation renders a novel parameterization to replace the binary ground structure parameterization in Figure 1. The overall synthesis approach is, therefore, referred to as the ‘load path approach.’ It ensures structural connectivity and allows variable mesh configurations without the need to choose an initial discretization mesh. It is largely inspired by the morphological representation (Tai and Chee, 2000) tailored to design structures and single output compliant mechanisms. Interested readers can refer to Tai and Chee (2000) for more

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

details on the morphological representation. The load path representation will be presented in this paper, followed by several design examples to study the improvements over the binary ground structure approach.

LOAD PATH REPRESENTATION

In every compliant mechanism, three classes of points, termed as ‘essential ports,’ must exist to ensure proper functionality of the system: (1) input points for actuation; (2) fixed points to ground the structure; and (3) output points to interact with the external environment. In the shape morphing problem, there are multiple output points, while the number of input and fixed points are specified by the designer. In this research, only one actuator is employed. As can be seen in Figure 4, these essential ports should be connected directly or indirectly to each other to form a valid compliant mechanism. The physical connections are the ‘load paths’ in a structure to transfer energy between the input actuator and the output points. In the load path representation, it is assumed that “every essential port is connected to every other essential port of a different class via one load path.” Thus, three types of load paths will be generated: (1) paths from Input Æ Output, (2) paths from Input Æ Fixed points, and (3) paths from Fixed point Æ Output. In addition, a set of intermediate connection ports (interconnect ports) are introduced to allow connections between different load paths, thus achieving ‘indirect’ connections between input/output/Fixed points.

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

Figure 4. The essential ports should be connected directly or indirectly by load paths (non-directional).

Topology and Connectivity Based on the layout shown in Figure 4, each design can be represented by a graph with vertices as the essential and interconnect ports. Figure 5 shows an example design that is ‘fully connected’ (defined later). As seen in Table 1, each load path is represented in terms of the ‘path sequence’ (pathSeq) containing vertices along the path. For each load path, there is also a corresponding binary variable, pTop, which indicates the presence or elimination of this path: when all pTop = 1, the graph is considered ‘fully connected’ with all paths present; when some pTop = 0 in a graph, the pTop = 0 paths are eliminated from the design. For example, the design in Figure 6 has the same path sequences as Figure 5, but different pTop values create a different topology. The pTop values for Figure 6 are shown in the 5th column in Table 1, while the pathSeq’s (3rd column) are identical to those in Figure 5. The pathSeq and pTop are, therefore, considered the topology design variables. By changing the pathSeq and pTop, various topologies can be created.

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

Figure 5. An example design with 6 intermediate connection ports. This design is considered fully connected because pTopi = 1 for all paths (i = 1 ~ 7).

Figure 6. An example design with identical pathSeq as in Figure 5, but different pTopi values (5th column, Table 1) lead to a different topology.

Table 1. The pathSeq and pTop for the designs in Figure 5 and Figure 6. Note that the pathSeq’s are identical for the two designs, and the pDim is only for Figure 6.

Path Path type No. In 1 2

Path sequence {1,6,3} {1,7,11,4}

pTopi Figure 5 1 1

pTopi Figure 6 1 0

pDimi Figure 6 {1,0.75} {1,1,2}

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

↓ Out

3

In ↓ Fix

4

Fix ↓ Out

5 6 7

{1,7,11,5}

{4.5,0.75,3}

1

1

{1,6,2}

1

1

{2.5,5}

{2,3} {2,6,7,4} {2,7,11,5}

1 1 1

1 1 0

{0.75} {5,1.5,8} {1.8,1,1}

To ensure all designs are well-connected, each design must satisfy two connectivity requirements: (1) the structure must be grounded at one or more fixed points, and (2) the input must be connected to the rest of the structure. These rules can easily be incorporated in the load path approach by monitoring the pTop values of paths from input and fixed point(s) to the output points. At least one of the pTop values in each path type has to be 1 to satisfy the connectivity requirement, thus the structural connectivity is now shown explicitly in the design variables. This is a huge advantage over the binary ground structure approach where a topology variable only determines the existence of one element and the connectivity has to be ‘searched for.’

Size and Geometry Aspects It is assumed that the load paths are comprised of rectangular beam elements, thus the size design variables are then defined as the cross section dimensions of each beam section between two vertices. In this research, the out-of-plane beam dimension is prescribed as a constant, thus only in-plane dimensions are considered design variables. Take Figure 6 for example, each path has a corresponding sequence, pDim, to describe the section dimensions, shown in the 6th column in Table 1. When two load paths have overlapping sections, such as in path #1 and #4 between vertices 1 and 6 (bold in Table 1),

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

one of the two available pDim values (1 and 2.5 from pDim1 and pDim4) is randomly selected to describe the section dimension. Since only one value is required to define one section, this random selection can be seen as the selected value dominating the recessive one. An additional variable, hBoundary, is also included in each design to define the inplane dimension of the morphing boundary. Finally, to control the geometry of the compliant mechanism, the locations of the interconnect ports are allowed to move within the design domain and are considered the geometry design variables (portLocation).

Using the design variables defined in the load path representation, the topology, size, and geometry of the compliant mechanism can be explored simultaneously. This parameterization scheme is incorporated into the same synthesis procedure shown in Figure 1 in place of the binary ground structure, and is referred to as the ‘load path approach’ in this paper. One crucial component in this GA based synthesis approach is the reproduction scheme, including selection, crossover, and mutation. Due to the difference in data structure from the binary ground structure, a new set of genetic operation strategies are required to work with load path representation. The genetic operations need to maintain sufficient diversity in each generation to prevent pre-matured convergence to a dominant design. In the following section, we will give a brief overview of the GA synthesis approach (Lu and Kota, 2003) and introduce the genetic operation schemes incorporated in the load path approach. Additional information on genetic algorithms, in general, can be found in Holland (1975) and Goldberg (1989).

GENETIC OPERATION STRATEGIES IN GA

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

GA simulates the selection scheme seen in nature and evolves a population of designs based on the principle of ‘survival of the fittest.’ The ‘fitness’ of each design is defined based on the objective function. In the shape morphing problem, the objective is to minimize the ‘shape deviation’ between the deformed shape and the desired target shape. Two shape deviation objective functions have been developed based on least square errors and Fourier Transformation respectively (Lu and Kota, 2003). The smaller the deformed shape deviates from the target shape, the better the shape morphing is achieved, hence the fitter the design is. As shown in Figure 1, in the GA, the designs in the first generation are randomly generated to somewhat sample the whole solution space. This parent generation then produces a new generation from a reproduction process. The load path approach incorporates a ‘roulette wheel’ selection scheme, where fitter individuals have higher chances to be selected for reproduction, and inferior ones have lower probability to reproduce. The selected parent designs then produce new offspring designs through the genetic operations: crossover and mutation. The genetic operations create diversity within each generation, which, in fact, provides the power behind GA to improve designs as generations evolve. The crossover and mutation strategies are, therefore, essential to the performance of GA.

Crossover Strategies The crossover strategy in the load path approach is to ‘exchange’ randomly selected paths between two parent designs. More than one path can be selected for crossover. The pathSeq and pTop of the parent designs are exchanged during this process while preserving the original pDim and portLocation. Since new pathSeq leads to new

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

connectivity in the offspring design, the number of segments (length of pDim) along each path may be different from that in the parent design. The pDim must be modified by inserting additional values or removing extra ones to maintain compatibility with the pathSeq. In addition, the boundary dimension (hBoundary) of the two parent designs can be exchanged according to the crossover probability.

Figure 7 shows two example designs (P1 and P2) with their load path information listed in Table 2. Path #1 and #5 are selected, for example, to illustrate the crossover operation. The two new designs are shown in Figure 8 and Table 3 as K1 and K2. As shown in Table 3, the original pathSeq and pTop of Path #1 and #5 in P1 and P2 are replaced with the values from the other parent, while pDim and interconnect port locations remain the same. However, the length of pDim changes after the crossover. Therefore, a random value (within the pDim bounds) is inserted into pDim if the new pathSeq is longer; a randomly selected ‘bit’ is removed from pDim if shorter. In addition, the boundary information is exchanged during crossover; since the connectivity of the morphing boundary is invariant, only the cross-section dimension is changed (hBoundary).

Figure 7. Two example parent designs with load path information listed in Table 2.

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004) Table 2. The load path information for the parent designs shown in Figure 7.

Path type

Path #

InÆOut

1 2 3 4 5 6 7

InÆFix FixÆOut hBoundary

Figure 7(a) P1 PathSeq pTop {1,6,3} 1 {1,7,11,4} 0 {1,7,11,5} 1 {1,6,2} 1 {2,3} 1 {2,6,7,4} 1 {2,7,11,5} 0 {3,4,5} 1

pDim {1,0.75} {1,1,2} {4.5,0.75,3} {2.5,5} {0.75} {5,1.5,8} {1.8,1,1} {5}

Figure 7(b) P2 PathSeq pTop {1,8,11,3} 0 {1,6,4} 1 {1,5} 1 {1,10,2} 0 {2,9,3} 1 {2,8,6,4} 1 {2,8,9,5} 0 {3,4,5} 1

pDim {2,5,3} {3,1.5} {4} {2,4} {1,2} {1,5,2} {2,1.2,3} {2}

Figure 8. Two offspring designs obtained from the parent designs in Figure 7 by exchanging path #1 and #5. Their load path information is listed in Table 3.

Table 3. The load path information of the offspring designs shown in Figure 8.

Path type

Path #

InÆOut

1* 2 3 4 5* 6 7

InÆFix FixÆOut hBoundary

Mutation Strategies

Figure 8(a) K1 PathSeq pTop {1,8,11,3} 0 {1,7,11,4} 0 {1,7,11,5} 1 {1,6,2} 1 {2,9,3} 1 {2,6,7,4} 1 {2,7,11,5} 0 {3,4,5} 1

pDim Æ{1,2,0.75} {1,1,2} {4.5,0.75,3} {2.5,5} Æ{0.75,3} {5,1.5,8} {1.8,1,1} {2}

Figure 8(b) K2 PathSeq pTop {1,6,3} 1 {1,6,4} 1 {1,5} 1 {1,10,2} 0 {2,3} 1 {2,8,6,4} 1 {2,8,9,5} 0 {3,4,5} 1

pDim Æ{5,3} {3,1.5} {4} {2,4} Æ{1} {1,5,2} {2,1.2,3} {5}

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

Four options are offered in the mutation process: (1) mutation of hBoundary, (2) mutation of path destination, (3) mutation of pTop, and (4) mutation of portLocations. In hBoundary mutation, the boundary dimension is replaced by a randomly generated value within the upper and lower bounds. To mutate the path destination, the end vertices of some randomly selected paths can mutate to a different one within the same class. For example, a path originally connecting the input to one of the output points can be mutated into a path connecting the input to another output point, simply by changing the last vertex in the pathSeq. The binary topology variable (pTop) is also allowed to mutate from 0 to 1 and vice versa for one randomly selected load path, thus changing the topology. The connection port location can also be mutated to a different location within the design domain. Figure 9 shows an example design mutated from K1 in Figure 8(a). The mutation in hBoundary, the destination change in Path #6, and mutation of pTop1 and pTop5 are shown in Table 4, while the location change of interconnect port 7 can be seen in Figure 9.

Figure 9. The new K1 (Figure 8(a)) after mutation shown in Table 4. Note that interconnect port 7 is also mutated to a different location.

Table 4. The load paths for the original K1 in Figure 8(a) and its mutated version in Figure 9.

Path type

Path #

Figure 8(a) K1 – original PathSeq pTop pDim

Figure 9 K1 – mutated PathSeq pTop pDim

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

InÆOut

1* 2 3 InÆFix 4 FixÆOut 5* 6* 7 hBoundary *

{1,8,11,3} {1,7,11,4} {1,7,11,5} {1,6,2} {2,9,3} {2,6,7,4} {2,7,11,5} {3,4,5}

0 0 1 1 1 1 0 1

{1,2,0.75} {1,1,2} {4.5,0.75,3} {2.5,5} {0.75,3} {5,1.5,8} {1.8,1,1} {2}

{1,8,11,3} {1,7,11,4} {1,7,11,5} {1,6,2} {2,9,3} {2,6,7,3} {2,7,11,5} {3,4,5}

Æ1 0 1 1 Æ0 1 0 1

{1,2,0.75} {1,1,2} {4.5,0.75,3} {2.5,5} {0.75,3} {5,1.5,8} {1.8,1,1} {4.2}

It is noted that when the two parent designs are identical, the ‘exchanging paths’ (crossover) strategy fails to produce any new design. In fact, the offspring designs will be identical to the parent designs. The mutation probability is, therefore, higher in this approach to enhance diversity in each generation. Higher mutation also helps improve the crossover performance, because the more diverse a generation is, the less likely it is to select two identical parent designs.

CONVERGENCE TO LOCAL OPTIMUM

Due to the heuristic nature of GA, the algorithm is capable of searching the whole solution space more extensively without being trapped in a local region. Although GA is more efficient in locating a region close to a local optimum, finding the exact location may be quite difficult. If the GA can indeed explore the entire solution space thoroughly, performing a local search following the GA can accelerate the convergence to the nearest local optimum, which is very likely to be the global optimum. However, there is no guarantee that GA can explore or sample the solution space evenly, so adding a local search after GA can only lead to a local optimum. In order to enhance the chance of finding the global optimum, a global search, DIRECT optimization algorithm (Jones,

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

Perttunen and Stuckman, 1993), is adopted to help investigate the global optimality. DIRECT optimization algorithm is a sampling algorithm that requires no knowledge of the objective function gradient. The algorithm samples points in the solution space and uses the information it has obtained to decide where to search next. It operates at both the global and local level. Once the global part of the algorithm finds the basin of convergence of the optimum, the local part of the algorithm quickly and automatically exploits it (Jones, Perttunen and Stuckman, 1993). The topology of the optimal solution obtained from GA is, therefore, used as a basic layout for the DIRECT algorithm to perform additional iterations on the connection port locations and beam section dimensions. However, the sampling nature of DIRECT algorithm implies that the obtained solution depends greatly on the number of iterations (sampling points). Therefore, a local search algorithm is utilized to accelerate the convergence to the nearby local optimum, after a prescribed number of iterations are carried out using the DIRECT algorithm. In this paper, the optimization toolbox in Matlab is used to perform the local search. More information on the optimization toolbox can be found in the Matlab documentation. The load path approach using GA, global search, and local search, is shown in Figure 10.

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

Figure 10. The GA based load path approach followed by additional global/local search to improve the convergence.

EXAMPLES AND DISCUSSIONS

Shape morphing can be seen useful in many areas, such as changing the aircraft wing shape to reduce drag, or changing the lumbar support shape in chairs to enhance comfort. To study the performance of the synthesis approach for compliant mechanisms, the binary ground structure approach and the improved load path approach are applied to several examples to understand their capabilities and limitations. All of the examples use the Least Square Error (LSE) deviation shown in Equation (1) as the objective function in GA to evaluate the difference between the achieved shape (deformed curve) and the desired target curve shape, subject to size, node locations, stress, stiffness, and connectivity constraints. Interested readers can refer to our previous paper (Lu and Kota, 2003) for more details regarding the problem formulation. LSE dev =

1 n ∑ ( xDEF ,i − xTAR ,i ) 2 + ( y DEF ,i − yTAR ,i ) 2 n i =1

(1)

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

where n is the total number of data points along the morphing boundary; (xDEF,i ,yDEF,i) and (xTAR,i ,yTAR,i) are the ith data point on the deformed and target morphing boundary.

Morphing Aircraft Leading Edge Most aircraft wings are optimized to produce minimum drag under a particular flying speed, at which the largest proportion of fuel is expended. However, in reality, flying speed varies continuously throughout flight. Hence, to obtain optimal fuel efficiency, the wing shape should be able to change in response to the change in flying speed (Lu and Kota, 2002). The shape morphing of a hypothetical airfoil leading edge is investigated here to compare the performance of the binary ground structure and load path approaches.

Figure 11 and Figure 12 are the results obtained from the binary ground structure and load path approaches respectively. The structural topologies are shown in solid lines, while the dark dash lines show the target shape, and the light dash lines represent the actual shape (deformed curve) achieved due to input actuation. Both solutions are obtained using the same number of population (150), number of generation (50), crossover probability (0.8), and mutation probability (0.5). In addition, the output points along the morphing boundary are determined in the preprocessor (Lu and Kota, 2003). The overall dimension is 260mm (10.24inch) by 230mm (9.06inch) by 20mm (0.79inch) (out-of-plane), and the material is aluminum.

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

Figure 11. The morphing leading edge design obtained from ten trials of the binary ground structure approach.

Figure 12. The morphing leading edge design obtained from ten trial runs of the load path approach.

Due to the heuristic nature, GA can provide a different result for the same problem in each run. The designs shown in Figure 11 and Figure 12 are the best solutions from ten trial runs of each approach, with LSE deviations of 8.92mm (0.35inch) and 3.72mm (0.15inch) respectively. The average computation time and LSE deviation for the ten trials are shown in Table 5. As can be seen, the binary ground structure approach requires almost twice the computation time of the load path approach. This may result from the

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

larger number of design variables used in the binary ground structure approach. Moreover, the verifying algorithm for structural connectivity may also lead to excessive computation time. Since the mesh is fixed in this approach, the optimal solution is always a subset of all possible designs embedded in the initial discretization mesh. However, the true optimal solution might not be included in the initial mesh. Therefore, the selection of the initial mesh is critical to the quality of the final solution. Note that there are several ‘trivial’ elements that have one ‘free end’ as shown in Figure 11. These elements have no strain/stress in them (only subject to rigid body motion), so they can be removed without affecting the compliant mechanism performance.

Table 5. The LSE deviation value and computation time from ten trials of both approaches.

Aircraft Leading Edge Example LSE dev. of best design Average LSE dev. Average CPU time

Binary Ground Structure Approach 8.9193mm 9.4540mm 533sec (8.88min)

Load Path Approach 3.7189mm 6.7385mm 274sec (4.57min)

The load path approach, on the other hand, can generate various structural topologies, because the locations of the connection ports are part of the design variables. The use of load path representation also eliminates the need of an additional verifying algorithm for connectivity. Therefore, the computation time is reduced and the desired shape morphing can be achieved with smaller deviation. As seen in Figure 12, all elements are connected at both ends because the topology is now represented in terms of the load path, thus there are no more trivial elements.

Flexible Antenna Reflector

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

Recent studies (Angelino and Washington, 2001; Martin et al., 2000; Washington, 1996; Yoon and Washington, 1998) have shown that antenna reflector adaptation can potentially enhance system performance and increase flexibility, such as changing the signal pattern or coverage area. In this example, a flexible antenna reflector changes its shape to direct the radiation signal to a different direction. As shown in Figure 13 and Figure 14, the two tips of the cylindrical reflector move in opposite directions to redirect the signal to the right. Figure 13 and Figure 14 are the optimal solutions from the binary ground structure approach and load path approach respectively with the corresponding LSE deviation values of 0.53mm (0.02inch) and 0.51mm (0.02inch). The overall dimension of the reflector is approximately 200mm (7.87inch) by 40mm (1.57inch) by 4mm (0.16inch) (out-of-plane) and the material is ABS (Acrylonitrile-Butadiene-Styrene) plastic. As can be seen, both designs are able to achieve the desired shape morphing of less than 2% of the shorter overall dimension. Note that both designs were obtained from only one trial run of each algorithm, while multiple trials are sometimes necessary for more complicated shape morphing. This suggests that the binary ground structure approach and load path approach are equally effective in finding a design that can achieve the desired shape morphing when the problem is less complex. As we will describe later, the complexity is typically related to the number of inflection points in the problem. Since the shape change required in the reflector is similar to cantilever beam bending without generating any inflection point, the problem is thus considered ‘simple’ shape change.

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

Figure 13. The optimal solution for antenna reflector beam steering obtained from the binary ground structure approach.

Figure 14. The optimal solution for antenna reflector beam steering obtained from the load path approach.

Shape Morphing Lumbar Support Lower back pain occurs frequently and is one of the most costly health problems affecting industry and society. Lifetime prevalences of 60% to 90% have been reported (Andersson, 1991). Lumbar support is one of the commonly used preventive strategies (Lahad et al., 1994). This example is inspired by the lumbar support system that is commonly used in car seats and office chairs to prevent lower back pain. The downward

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

‘seating’ motion from a person is used as an input to actuate the lumbar support that changes the initially straight back support shape into a curved profile. The curved profile should match the natural profile of human spine, which typically includes an inflection point as shown in Figure 15.

Figure 15. Natural sitting spinal model in an ideal driver’s seat (Harrison et al., 2000).

To study the performance of the binary ground structure and load path approaches, we created a curve to roughly approximate the spinal shape. The initial (straight) and target (spine) curve are shown in Figure 16 and Figure 17, utilizing the downward ‘seating’ motion as the input. Figure 16 and Figure 17 are the best designs obtained from 10 trial runs of the binary ground structure approach and load path approach each. Asterisk symbols are placed beside physical element connections to distinguish them from the visual intersections created by overlapping elements. The corresponding LSE deviation values are 11.24mm (0.44inch) and 10.55mm (0.42inch) respectively. The overall dimension is 200mm (7.87inch) by 500mm (19.69inch) by 5mm (0.2inch) (out-of-plane)

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

and the material is ABS plastic. The average LSE deviation and required computation time are listed in Table 6.

Figure 16. The optimal lumbar support design obtained from the binary ground structure approach.

Figure 17. The optimal lumbar support design obtained from the load path approach. Note that relative motions are allowed between overlapping paths (physical element connections are denoted by *).

As can be seen in Table 6, the binary ground structure approach requires almost twice as much computation time as the load path approach. Although the LSE deviation values are close in Figure 16 and Figure 17, the shape morphing does look different (visually). This

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

may result from the insufficient numbers of data points along the initial and target curves. There are currently 11 data points along the curve, and the Euclidian distance between the deformed and target location of each point are used to measure the LSE deviation in Equation (1). Additional data points along the morphing boundary can potentially lead to more significant difference between the LSE deviations obtained from the two approaches.

Table 6. The LSE deviation value and computation time from ten trials of both approaches.

Lumbar Support Example LSE dev. of best design Average LSE dev. Average CPU time

Binary Ground Structure Approach 11.2738mm 12.3716mm 460sec (7.67min)

Load Path Approach 10.5463mm 10.6284mm 243sec (4.05min)

Simple cantilever beam bending can be seen as the whole beam ‘pivoting’ about the fixed end (Howell and Midha, 1995). However, creating an inflection point requires changing the center of curvature from one side to the opposite side of the beam. Since it requires a moment or some opposite (push/pull) motions to generate a couple at the inflection point, shape morphing involving creating an inflection point, such as this lumbar support example, is considered more complicated than simple cantilever beam bending (antenna reflector example). It is observed that, from the 10 trial runs using binary ground structure approach, the resulting designs can typically achieve the shape morphing on the right hand side (flatter side) in Figure 16, but most of them fail to match the portion on the left of the inflection point. In other words, it is relatively straightforward for the algorithm to find a solution without an inflection point, but generating an inflection point can be quite difficult. On the other hand, the load path approach appears to capture the inflection point better. As can be seen in Figure 17, the overlapping paths from the input

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

and fixed point to the opposite sides of the inflection point seem to provide a push/pull motion, leading to a moment about the inflection point. This overlapping ‘X’ configuration can also be vaguely seen in the result in Figure 16. However, if this X configuration is not a subset of the initial discretization mesh, the binary ground structure approach may be difficult to find a topology that can produce an inflection point on the morphing boundary. When selecting an initial mesh for Figure 16, we intentionally included the X configuration in the initial mesh. The decision was based on our understanding of the problem and some results obtained from the load path approach, but this kind of information is generally unavailable beforehand. Therefore, we believe the load path approach is a better means to systematically design shape morphing compliant mechanisms without the need of intuition or prior experience to select an initial discretization mesh.

FINAL REMARKS

In this paper, we introduced the load path representation method and its incorporation into the GA based synthesis approach to design shape morphing compliant mechanisms. Several examples are included to study the performance of the load path approach and the binary ground structure approach. The results showed that the load path approach requires almost half of the computation time required by the binary ground structure approach, due to the absence of additional verifying algorithm for structural connectivity. Generally, both approaches are able to achieve simple shape morphing when no inflection points are involved. However, even for more complicated problems with

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

inflection points, load path approach consistently yields solutions that successfully achieve the desired shape change. More importantly, unlike the binary ground structure approach, the load path approach does not have an initial discretization network which typically requires intuition or prior experience when determining the complexity and configuration of the mesh. Therefore, the load path approach can potentially lead to a fully systematic synthesis approach for shape morphing. Although the load path representation was developed for synthesis of shape morphing compliant mechanisms in this context, it is not limited to multiple output problems. For structural optimization or single output compliant mechanism synthesis, the load path representation can be used by simply replacing the curve comparison objective function with other objective functions, such as minimizing strain energy or maximizing geometric advantages. In other words, the load path representation is a general method to parameterize the design domain into appropriate design variables to simultaneously optimize the compliant mechanism topology and dimensions. Future and on-going research includes the study of boundary conditions since the locations of the input actuator and ground supports also play significant roles in the structural topology. The load path representation is also being applied to structural optimization and other compliant mechanism design problems to understand the generality of this approach.

ACKNOWLEDGEMENTS

Authors gratefully acknowledge the funding support of U.S. Air Force Office of Scientific Research for this work under the research contract number F49620-96-1-0205.

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

REFERENCES Ananthasuresh, G. K., Kota, S. and Kikuchi, K. 1994. "Strategies for Systematic Synthesis of Compliant MEMS," 1994 International Mechanical Engineering Congress and Exposition, 55:677-686. Andersson, G. 1991. "The Epidemiology of Spinal Disorders," Raven Press, 107-146. Angelino, M. and Washington, G. 2001. "Point Actuated Aperture Antenna Development," Proceedings of SPIE, 4334:147-155. Frecker, M. I., Ananthasuresh, G. K., Nishiwaki, S., Kikuchi, N. and Kota, S. 1997. "Topological Synthesis of Compliant Mechanisms Using Multi-Criteria Optimization," ASME Journal of Mechanical Design, 119(2):238-245. Goldberg, D. 1989. "Genetic Algorithms in Search, Optimization, and Machine Learning," Addison-Wesley. Harrison, D. D., Harrison, S. O., Croft, A. C., Harrison, D. E. and Troyanovich, S. J. 2000. "Sitting Biomechanics, Part Ii: Optimal Car Driver's Seat and Optimal Driver's Spinal Model," Journal of Manipulative and Physiological Therapeutics, 23(1):37-47. Hetrick, J. and Kota, S. 1999. "An Energy Formulation for Parametric Size and Shape Optimization of Compliant Mechanisms," ASME Journal of Mechanical Design, 121:229-234. Holland, J. H. 1975. "Adaptation in Natural and Artificial Systems," The University of Michigan Press. Howell, L. L. and Midha, A. 1995. "Parametric Deflection Approximations for EndLoaded, Large-Deflection Beams in Compliant Mechanisms," ASME Journal of Mechanical Design, 117(1):156-165. Jones, D. R., Perttunen, C. D. and Stuckman, B. E. 1993. "Lipschitzian Optimization without the Lipschitz Constant," Journal of Optimization Theory and Applications, 79(1):157-181. Joo, J., Kota, S. and Kikuchi, N. 2000. "Topological Synthesis of Compliant Mechanisms Using Linear Beam Elements," Mechanics Based Design of Structures and Machines, 28(4):245-280. Lahad, A., Malter, A. D., Berg, A. O. and Deyo, R. A. 1994. "The Effectiveness of Four Interventions for the Prevention of Low Back Pain," Journal of the American Medical Association, 272(16):1286-1291.

Accepted by the Journal of Intelligent Material Systems and Structures (in press 2004)

Larsen, U. D., Sigmund, O. and Bouwstra, S. 1997. "Design and Fabrication of Compliant Micromechanisms and Structures with Negative Poisson's Ratio," Journal of Microelectromechanical systems, 6(2):99-106. Lu, K. J. and Kota, S. 2002. "Compliant Mechanism Synthesis for Shape-Change Applications: Preliminary Results," SPIE Modeling, Signal Processing, and Control Conference, 4693:161-172. Lu, K. J. and Kota, S. 2003. "Design of Compliant Mechanisms for Morphing Structural Shapes," Journal of Intelligent Material Systems and Structures, 14(6):379-391. Martin, J. W., Redmond, J. M., Barney, P. S., Henson, T. D., Wehlburg, J. C. and Main, J. A. 2000. "Distributed Sensing and Shape Control of Piezoelectric Bimorph Mirrors," Journal of Intelligent Material Systems and Structures, 11:744-757. Saggere, L. and Kota, S. 1999. "Static Shape Control of Smart Structures Using Compliant Mechanisms," AIAA Journal, 37(5):572-578. Tai, K. and Chee, T. H. 2000. "Design of Structures and Compliant Mechanisms by Evolutionary Optimization of Morphological Representations of Topology," ASME Journal of Mechanical Design, 122:560-566. Washington, G. 1996. "Smart Aperture Antennas," Journal of Smart Materials and Structures, 5(6):801-805. Yoon, H. S. and Washington, G. 1998. "Piezoceramic Actuated Aperture Antennae," Journal of Smart Materials and Structures, 7(4):537-542.