AIAA 2009-3996

19th AIAA Computational Fluid Dynamics 22 - 25 June 2009, San Antonio, Texas

Explicit Mesh Deformation Using Inverse Distance Weighting Interpolation Jeroen A.S. Witteveen∗, Hester Bijl† Faculty of Aerospace Engineering, Delft University of Technology, Kluyverweg 1, 2629HS Delft, The Netherlands Mesh deformation algorithms usually require solving a system of equations which can be computationally intensive for complex three-dimensional applications. Here an explicit mesh deformation method is proposed based on Inverse Distance Weighting (IDW) interpolation of the boundary node displacements to the interior of the flow domain. The point-by-point mesh deformation algorithm results in a straightforward implementation and parallelization. Fluid-structure interaction applications to a two-dimensional airfoil and the three-dimensional AGARD 445.6 wing benchmark show a reduction of the computational costs up to a factor 100 with respect to Radial Basis Function (RBF) mesh deformation for a comparable simulation accuracy.

I.

Introduction

In fluid-structure interaction simulations the dynamics of the structure and the flow are coupled by forces and displacements on their interface. Flow forces result through deformations of the structure in moving boundaries for the flow domain. There is, therefore, a need for accurate and efficient mesh motion algorithms to propagate these boundary displacements to the mesh in the interior of the flow domain. For structured meshes the efficient Transfinite Interpolation method is available. Methods for the deformation of unstructured meshes often use the connectivity of the flow mesh in algorithms based on spring analogy, body elasticity, or Laplacian and Biharmonic operators. These methods can be computationally intensive, since they result in solving a system of equations of the size of the number of flow points. Point-by-point methods do not rely on connectivity information as they determine the displacement of each point in the flow mesh based on its relative position with respect to the domain boundary. Recently, a point-by-point mesh deformation method was developed based on Radial Basis Function (RBF) interpolation.1, 6 The resulting flexible and robust mesh motion algorithm can deal with large deformations, hanging nodes, and is easily implemented in parallel. However, radial basis function mesh motion requires the solution of a system of equations of the size of the number of boundary nodes. Solving this system can still be expensive as it accounts for a significant part of the computational time in large-scale three-dimensional simulations. In this paper, we present an explicit mesh motion algorithm based on Inverse Distance Weighting (IDW) interpolation,7 which does not lead to solving a system of equations. The proposed point-by-point method maintains the robustness and flexibility for dealing with large deformations, hanging nodes, and parallelization of the previous method. In addition, the explicit formulation results in a faster mesh motion algorithm and an easier implementation. The mesh deformation method is studied for unstructured triangular and hexahedral meshes in a NACA0012 airfoil fluid-structure interaction to optimize the shape parameter value c. Subsequently IDW mesh deformation is applied to fluid-structure interaction simulation for a twodimensional airfoil and the three-dimensional AGARD 445.6 wing benchmark. The results show an decrease of computational costs up to a factor 100 with respect to the RBF method, while maintaining comparable simulation result accuracy. ∗ Postdoctoral † Full

researcher, Member AIAA, +31(0)15 2782046, [email protected], http://www.jeroenwitteveen.com. Professor, Member AIAA, +31(0)15 2785373.

1 of 10 American Aeronautics and Astronautics Copyright © 2009 by J.A.S. Witteveen, H. Bijl. Published by the AmericanInstitute Institute ofof Aeronautics and Astronautics, Inc., with permission.

II.

Inverse distance weighting interpolation

Inverse distance weighting interpolation7 is an explicit method for multivariate interpolation of scattered data points. The interpolation surface w(x) through n data samples v = {v1 , .., vn } of the exact function u(x) with vi ≡ u(xi ) is given in inverse distance weighting by Pn vi φ(ri ) w(x) = Pi=1 , (1) n i=1 φ(ri )

with weighting function

φ(r) = r−c ,

(2)

where ri = kx−xi k ≥ 0 is the Euclidian distance between x and data point xi , and c is a power parameter. In the mesh deformation algorithm separate shape parameters are introduced for dynamic and static boundary nodes, and for translations and rotations: cd,T , cd,R , cs,T , and cs,R .

III.

Triangular grid around a NACA0012 airfoil

First the mesh deformation of an unstructured triangular mesh around a NACA0012 airfoil in a square domain of size 10c × 10c is considered with c the airfoil chord, as shown in Figure 1. The mesh consists of 1524 cells with 112 nodes on the airfoil, 24 nodes on the outer boundary, and 694 internal nodes. The mesh is subject to a given displacement of the airfoil consisting of a translation in both directions of ∆x = ∆y = −2.5c and a rotation over ∆α = −60deg in one step from the equilibrium position of the initial mesh. The mesh quality of the resulting mesh is shown in Figure 2. A combination of a size and skew measure between 0 and 1 is used as measure of the mesh quality.4 The proposed explicit mesh deformation method results in a mesh of good quality without degenerate cells even for this extreme displacement of the airfoil in one step. Especially near the airfoil the grid is practically undistorted with a mesh quality close to 1, see Figure 2b. This is important for a robust deformation of the boundary layer cells and an accurate resolution of the aerodynamic forces on the airfoil. The overall quality of the mesh is in this case 0.426. This mesh quality is obtained by optimizing the shape parameter values, which results in this case in cd,T = 2.2, cd,R = 2.2, cs,T = 2.0, and cs,R = 0.1. Similar optimizations for several other displacements point to general optimal integer values of the shape parameters of cd,T = cd,R = 2, cs,T = cs,R = 1. These values are used in the fluid-structure applications below.

(a) Initial mesh

(b) Initial mesh (zoom)

Figure 1. Unstructured triangular mesh around a NACA0012 airfoil with 1524 cells.

A.

Two-dimensional airfoil fluid-structure interaction example

The two-dimensional example is the post-flutter simulation of an elastically mounted rigid airfoil with nonlinear structural stiffness in an inviscid flow. The nonlinear structural stiffness is modeled by a cubic spring stiffness term in the following two-degree-of-freedom model for coupled pitch and plunge motion of the airfoil:3  ω 1 ¯ 2 (ξ + βξ ξ 3 ) = − Cl (τ ), (3) ξ ′′ + xα α′′ + ∗ U πµ 2 of 10 American Institute of Aeronautics and Astronautics

(a) Mesh quality

(b) Mesh quality (zoom)

Figure 2. Mesh quality of the unstructured triangular mesh around a NACA0012 airfoil for a deflection of ∆x = ∆y = −2.5c and ∆α = −60deg.

1 2 xα ′′ ξ + α′′ + ∗2 (α + βα α3 ) = Cm (τ ), rα2 U πµrα2

(4)

where βξ = 0m−2 and βα = 3rad−2 are nonlinear spring constants, ξ(τ ) = h/b is the nondimensional plunge displacement of the elastic axis, α(τ ) is the pitch angle, and (′ ) denotes differentiation with respect to nondimensional time τ = U t/b, with half-chord length b = c/2 = 0.5m and free stream velocity U = 103.6m/s, which corresponds to a Mach number of M∞ = 0.3 for free stream density ρ∞ = 0.12kg/m3 and pressure p∞ = 1.0 · 105 Pa. The radius of gyration around the elastic axis is rα b = 0.25m, bifurcation parameter U ∗ is defined as U ∗ = U/(bωα ), and the airfoil-air mass ratio is µ = m/πρ∞ b2 = 100, with m the airfoil mass. The elastic axis is located at a distance ah b = −0.25m from the mid-chord position and the mass center is located at a distance xα b = 0.125m from the elastic axis. The ratio of natural frequencies ω(ω) = ωξ /ωα = 0.2, with ωξ and ωα the natural frequencies of the airfoil in pitch and plunge, respectively.5 The nondimensional aerodynamic lift and moment coefficients, Cl (τ ) and Cm (τ ), for the NACA0012 airfoil are determined by solving the nonlinear Euler equations for inviscid flow2 using a second-order finite volume scheme on an unstructured hexahedral mesh in spatial domain D with dimensions 30c × 20c. An Arbitrary Lagrangian-Eulerian formulation is employed to couple the fluid mesh with the movement of the structure. Time integration is performed using the BDF-2 method with stepsize ∆τ = 0.4 until τmax = 1000. Initially the airfoil is at rest at a deflection of α(0) = 0.1deg and ξ(0) = 0 from its equilibrium position. The initial condition of the flow field is given by the steady state solution for the initial deflection. In the numerical results we compare the results of Inverse Distance Weighting (IDW) and Radial Basis Function (RBF) mesh deformation on three different meshes with increasing number of cells. The meshes with 4198, 11112, and 37574 cells are shown in Figure 3. Euler flow simulations are used here in order to be able to use this large range of mesh sizes including the coarse mesh of 4198 cells. For the Inverse Distance Weighting mesh deformation two different implementations are used. The first formulation, IDW, is the straightforward implementation based on the boundary node displacements with respect to the previous time step, which is in general robust and accurate. In the second approach the boundary node displacements are written as function of the generalized structural eigenmode displacements. For this eigenmode formulation, IDWeig, the displacements are computed with respect to the equilibrium position by storing the significantly smaller interpolation matrix of size ndof × nin , with ndof = 2 the number of eigenmodes and nin the number of internal mesh points. The results are compared with a Radial Basis Function mesh deformation reference based on the Thin Plate Splines (TPS) function deformation with respect to the previous time step, where the system of equations is solved directly. Other matrix solution methods are possible here, such that the CPU results should be interpret as a first comparison. The simulation results are compared in terms of the pitch angle α(τ ), the computational time, and the mesh quality fopt (τ ). The pitch angle α(τ ) as function of nondimensional time is given in Figure 4 for the IDW, IDWeig, and RBF mesh deformation algorithms on the finest mesh with 37574 cells. The three time histories show a practically identical oscillatory increase of α(τ ) until the system reaches a periodic limit cycle oscillation with constant amplitude. The quantitative differences in the prediction of the limit cycle oscillation amplitude with respect to the IDW results are given in Table 1 for the three mesh sizes. These

3 of 10 American Institute of Aeronautics and Astronautics

(a) 4198 cells (zoom)

(b) 11112 cells (zoom)

(c) 37574 cells (zoom)

(d) 37574 cells

Figure 3. Three meshes for the two-dimensional airfoil fluid-structure interaction example.

4 of 10 American Institute of Aeronautics and Astronautics

results confirm that the relative errors in the amplitude prediction are small with a maximum of 1.81 · 10−3 .

10 8

RBF IDW IDWeig

angle of attack α [deg]

6 4 2 0 −2 −4 −6 −8 −10 0

1

2 3 nondimensional time τ

4

5

Figure 4. Pitch angle α(τ ) of the mesh with 37574 cells for the two-dimensional airfoil fluid-structure interaction example.

Table 1. Relative error in pitch angle amplitude Aα with respect to the IDW solution as function of the number of cells for the two-dimensional airfoil fluid-structure interaction example.

RBF IDWeig

4198 cells 1.39 · 10−3 2.27 · 10−4

11112 cells 1.74 · 10−3 3.12 · 10−4

37574 cells 1.81 · 10−3 2.38 · 10−4

In contrast, the CPU time results for the average computational costs for the mesh deformation per time step shows significant differences. The computational times for the three mesh deformation methods increase monotonically with the mesh size in Figure 5. The explicit IDW algorithm results, however, in a substantial slower increase of the computational costs with increasing number of mesh cells than RBF mesh deformation, since it does not involve solving a system of equations. Table 2 indicates that IDW reduces the computational work for mesh deformation up to a factor 10 compared to RBF for the mesh with 37574 cells. The IDWeig formulation based on the structural eigenmodes further reduces the costs to a factor 100 compared to RBF. A linear increase of the CPU time is actually achieved here for IDWeig, where the other algorithms show a quadratic growth rate. Table 2. Average computational time per time step in seconds as function of the number of cells for the two-dimensional airfoil fluid-structure interaction example.

RBF IDW IDWeig

4198 cells 0.130 0.0210 0.0257

11112 cells 0.975 0.160 0.0729

37574 cells 20.6 1.92 0.225

The differences in computational costs are reflected to some extend in the mesh qualities given in Figure 6. The global mesh quality metric fopt shows for the three methods an oscillatory behavior due to the oscillation of the airfoil around the equilibrium position. The mesh quality decreases up to τ = 2 due to the increasing amplitude in the transient part of the airfoil oscillation. For IDW and RBF the metric fopt further decreases for τ > 2 caused by the build up of errors due to the deformation of the mesh with respect to the previous time step. Since IDWeig is based on the deformation with respect to the equilibrium position, the mesh quality does not detoriate after the transient part with maxima close to fopt = 100%. Overall the RBF mesh deformation reaches a factor 2 higher mesh quality than IDW and IDWeig. This leads, however, not 5 of 10 American Institute of Aeronautics and Astronautics

25

RBF IDW IDWeig

average CPU time [s]

20

15

10

5

0 0

0.5

1

1.5 2 2.5 number of cells

3

3.5

4 4

x 10

Figure 5. Average computational time per time step as function of the number of cells for the two-dimensional airfoil fluid-structure interaction example.

1

1

0.995

0.995

0.99 0.985 0.98 0.975 0.97 0.965 0

RBF global mesh quality fopt

1 0.995

IDWeig global mesh quality fopt

IDW global mesh quality fopt

in significantly more accurate fluid-structure interaction response predictions as discussed in relation with Figure 4 and Table 1. Also IDW and IDWeig mesh deformation result in high mesh quality of fopt” > 96.5%. The IDW mesh quality can be improved by using optimized shape parameter values c for this specific problem instead of the used integer values to enable fast evaluation.

0.99 0.985 0.98 0.975 0.97

1

2 3 nondimensional time τ

(a) IDW

4

5

0.965 0

0.99 0.985 0.98 0.975 0.97

1

2 3 nondimensional time τ

4

5

0.965 0

(b) IDWeig

1

2 3 nondimensional time τ

4

5

(c) RBF

Figure 6. Mesh quality fopt (τ ) of the mesh with 37574 cells for the two-dimensional airfoil fluid-structure interaction example.

B.

Three-dimensional AGARD 445.6 wing aeroelastic benchmark

The AGARD aeroelastic wing 445.6 configuration number 3 known as the weakened model8 is considered here with a NACA 65A004 symmetric airfoil, taper ratio of 0.66, 45o quarter-chord sweep angle, and a 2.5-foot semi-span subject to an inviscid flow. The structure is described by a nodal discretization using an undamped linear finite element model in the Matlab finite element toolbox OpenFEM. The discretization contains in the chordal and spanwise direction 6 × 6 brick-elements with 20 nodes and 60 degrees-of-freedom, and at the leading and trailing edge 2 × 6 pentahedral elements with 15 nodes and 45 degrees-of-freedom.9 Orthotropic material properties are used and the fiber orientation is taken parallel to the quarter-chord line. The Euler equations for inviscid flow2 are solved using a second-order central finite volume discretization on a 60 × 15 × 30m domain using an unstructured hexahedral mesh. The free stream conditions are for the density ρ∞ = 0.099468kg/m3 and the pressure p∞ = 7704.05Pa.8 Time integration is performed using a third-order implicit multi-stage Runge-Kutta scheme with step size ∆t = 2.5 · 10−3 s until t = 1.25s to determine the stochastic solution until t = 1s. The first bending mode with a vertical tip displacement of ytip = 0.01m is used as initial condition for the structure. The coupled fluid-structure interaction system is solved using a partitioned IMEX scheme with explicit

6 of 10 American Institute of Aeronautics and Astronautics

treatment of the coupling terms without sub-iterations. An Arbitrary Lagrangian-Eulerian formulation is employed to couple the fluid mesh with the movement of the structure. The flow forces and the structural displacements are imposed on the structure and the flow using nearest neighbor and radial basis function interpolation,9 respectively. This three-dimensional test problem is studied in analogy to the analysis for the two-dimensional airfoil fluid-structure interaction application. The results for IDW and RBF mesh deformation are compared from three different meshes with increasing number of cells shown in Figure 7. In addition a formulation IDWnorot is considered which does not take into account the effect of rotations of the mesh boundary normal vector rotation, since rotations are expected to be small for this problem. This is a possible approach to reduce the computational costs of IDW mesh deformation further, since IDW treats boundary node displacements and rotations separately. The RBF algorithm handles them combined by implicitly interpreting relative translations as rotations.

(a) 15122 cells

(b) 31274 cells

(c) 75656 cells

Figure 7. Three surface meshes for the three-dimensional AGARD 445.6 wing aeroelastic benchmark.

The response output functional that is considered in this case is the lift force L(t) as function of time. The time series for L(t) of the three mesh deformation algorithms closely agree as shown in Figure 8. The maximum relative error in the lift amplitude AL with respect to the IDW solution is smaller than 3 · 10−2 , see Table 3. 300

RBF IDW IDWnorot

200

lift force [N]

100

0

−100

−200

−300 0

0.2

0.4

time t

0.6

0.8

1

Figure 8. Lift force L(t) of the mesh with 75656 cells for the three-dimensional AGARD 445.6 wing aeroelastic benchmark.

The computational costs per time step are in this case the sum of the CPU times for the three stages per time step of the multi-stage Runge-Kutta time integration scheme employed here. The mesh deformation algorithms result in three dimensions in higher computational costs due to the additional translation and rotation dimensions as can be concluded from Figure 9 and Table 4. The RBF method results again in a fast increase of computational costs with increasing mesh size up to 2544 seconds per time step for the finest mesh with 75656 cells, which constitutes an impractical contribution to the total CPU time for the fluid-structure interaction simulation of approximately 90%. The reduction of the CPU time by IDW with

7 of 10 American Institute of Aeronautics and Astronautics

Table 3. Relative error in lift force amplitude AL with respect to the IDW solution as function of the number of cells for the three-dimensional AGARD 445.6 wing aeroelastic benchmark.

15122 cells 5.90 · 10−3 2.95 · 10−2

RBF IDWnorot

31274 cells 2.08 · 10−3 1.21 · 10−2

75656 cells 9.57 · 10−5 1.24 · 10−2

a factor 20 demonstrates that the efficiency gain of the explicit IDW algorithm increases with dimension compared to RBF mesh deformation. Neglecting the constributions of rotations reduces the computational costs for IDWnorot further to 39.7 seconds which corresponds to a factor 50 decrease with respect to the RBF method. 3000

average CPU time [s]

2500

RBF IDW IDWnorot

2000

1500

1000

500

0 0

1

2

3 4 5 number of cells

6

7

8 4

x 10

Figure 9. Average computational time per time step as function of the number of cells for the three-dimensional AGARD 445.6 wing aeroelastic benchmark.

Table 4. Average computational time per time step in seconds as function of the number of cells for the three-dimensional AGARD 445.6 wing aeroelastic benchmark.

RBF IDW IDWnorot

15122 cells 113 3.99 2.46

31274 cells 264 15.6 7.68

75656 cells 2544 108 39.7

However, not taking into account rotations results in a mean mesh quality fmean of 95% to 96% as shown in Figure 10. The mesh quality metric increases with time for this case due to the decaying oscillation amplitude illustrated by the lift force L(t) in Figure 8. RBF mesh deformation gives here a mean mesh quality of more than 99%, which is a factor 3 larger than the quality for IDW of 97%. This difference in mesh quality would in general not justify the significant additional computational costs for the RBF algorithm presented in Table 4. Moreover the mesh qualities of the three methods are not notably reflected in the aeroelastic simulation results in terms of the lift force amplitude AL as illustrated in Table 3. An implementation of the IDWeig formulation based on the structural eigenmodes of the AGARD 445.6 wing is a direction to further reduced the computational costs possibly in combination with neglecting the effect of boundary normal vector rotations in IDWnorot.

8 of 10 American Institute of Aeronautics and Astronautics

mean global mesh quality f

mean

1

0.99

0.98

0.97

0.96

0.95

0.94 0

0.2

0.4

time t

0.6

RBF IDW IDWnorot 0.8 1

(a) IDW Figure 10. Mesh quality fmean (t) of the mesh with 75656 cells for the three-dimensional AGARD 445.6 wing aeroelastic benchmark.

IV.

Conclusions

An explicit mesh deformation method is presented based on Inverse Distance Weighting (IDW) interpolation of the boundary node displacements. The point-by-point approach results in a significant reduction of computational costs, since the proposed mesh deformation algorithm does not involve the solution of a matrix system of equations. This enables an easy implementation and parallelization of the IDW mesh deformation routine. For an unstructured triangular mesh around a NACA0012 airfoil the optimal integer values of the shape parameter values cd,T = cd,R = 2, cs,T = cs,R = 1 are established. These shape parameter values are subsequently implemented in a fluid-structure interaction solver applied to a two-dimensional airfoil and the three-dimensional AGARD 445.6 wing benchmark. The results demonstrate a reduction of computational costs for IDW mesh deformation with respect to the RBF method of a factor 10 and 20, respectively. To addional formulations of IDWeig based on the structural eigenmodes and neglecting boundary normal vector rotations in IDWnorot further reduce the CPU time to a factor 100 and 50 with respect to RBF mesh deformation. The factor 2–5 higher mesh quality of the RBF algorithm are not reflected in the simulation results for the amplitudes of the pitch angle α and lift force L, for which the mesh deformation methods agree up to a relative error of 3 · 10−2 . In future work the mesh quality of IDW mesh deformation can be improved by optimizing the shape parameter values c for each case separately. An implementation of the IDWeig formulation for the AGARD 445.6 wing can further reduce the computational costs in the three-dimensional problem.

References 1 A. de Boer, M.S. van der Schoot, H. Bijl. Mesh deformation based on radial basis function interpolation, Comput. Struct. 85 (2007) 784–795. 2 A.J. Chorin, J.E. Marsden, A mathematical introduction to fluid mechanics, Springer-Verlag, New York (1979). 3 Y. Fung, An introduction to aeroelasticity, Dover Publications, New York (1969). 4 P.M. Knupp, Algebraic mesh quality metrics for unstructured initial meshes, Finite Elem. Anal. Des. 39 (2003) 217–241. 5 B.H.K. Lee, L.Y. Jiang, Y.S. Wong, Flutter of an airfoil with a cubic nonlinear restoring force, AIAA-1998-1725, 39th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, Long Beach, California (1998). 6 T.C.S. Rendall, C. Allen, Unified CFD-CSD interpolation and mesh motion using radial basis functions, Int. J. Numer. Meth. Eng. 74 (2008) 1519–1559. 7 D. Shepard. A two-dimensional interpolation function for irregularly-spaced data, Proceedings of the 1968 ACM National Conference (1968) 517–524. 8 E. Yates Jr., AGARD standard aeroelastic configurations for dynamic response. Candidate configuration I.-Wing 445.6, Technical Memorandum 100492, NASA, 1987.

9 of 10 American Institute of Aeronautics and Astronautics

9 A.H. van Zuijlen, A. de Boer, H. Bijl, Higher-order time integration through smooth mesh deformation for 3D fluidstructure interaction simulations, J. Comput. Phys. 224 (2007) 414430.

10 of 10 American Institute of Aeronautics and Astronautics

Explicit Mesh Deformation Using Inverse Distance ...

putational costs up to a factor 100 with respect to Radial Basis Function (RBF) mesh deformation for a comparable simulation accuracy. I. Introduction.

2MB Sizes 4 Downloads 169 Views

Recommend Documents

Explicit and Robust Inverse Distance Weighting ... - Jeroen Witteveen
Center for Turbulence Research, Stanford University, ... equations of the size of the number of internal flow points ni to determine the deformed state of the mesh. ..... 36th AIAA Fluid Dynamics Conference and Exhibit, San Francisco, California ...

Nonrigid Image Deformation Using Moving ... - Semantic Scholar
500×500). We compare our method to a state-of-the-art method which is modeled by rigid ... Schematic illustration of image deformation. Left: the original image.

Nonrigid Image Deformation Using Moving ... - Semantic Scholar
To illustrate, consider Fig. 1 where we are given an image of Burning. Candle and we aim to deform its flame. To this end, we first choose a set of control points, ...

A Mesh Meaningful Segmentation Algorithm Using Skeleton ... - Kai Xu
turning space region, as real critical points, since the straight absorption would lead to .... Latin American Conference on Robotics and Automation, pp. 91–97 ...

Operating a UAV Mesh & Internet Backhaul Network using ...
Figure 1. A high-level overview of the software-defined networking architecture. UAVs for this ... implementation of services and applications that control, monitor ...

Interior Distance Using Barycentric Coordinates
using visibility graphs [CSY94,BCKO08]), they are still ex- pensive. .... the tools from Multi-Dimensional Scaling (MDS) [BG05, ..... Figure 8: 3D Distance visualization within the human, dino-pet, hand, cheetah, octopus, and fertility models.

A Mesh Meaningful Segmentation Algorithm Using Skeleton ... - Kai Xu
would not be taken account into subsequent computation of critical points. Basically, .... treat all selected points found by before checking procedure (Fig. 4.b), lying in the ..... International Journal of Robotics Research 5(4),. 38–55 (1986).

A Mesh Meaningful Segmentation Algorithm Using Skeleton ... - Kai Xu
sion, mesh editing, mesh deformation [7], collision detection [8], shape analysis and matching [9][10]. Especially, the process that decomposes a model into visually meaningful components is called part-type segmentation [1][2] (shape decomposition o

Inverse Functions and Inverse Trigonometric Functions.pdf ...
Sign in. Loading… Whoops! There was a problem loading more pages. Retrying... Whoops! There was a problem previewing this document. Retrying.

Robust Interactive Mesh Cutting Using Fast Geodesic ...
on Dijkstra's algorithm and the graph cut algorithm can be applied here for cutting ... a geodesic curvature flow based interactive mesh cutting framework named ...

A Mesh Meaningful Segmentation Algorithm Using Skeleton ... - Kai Xu
cluding texture mapping [4], shape manipulation[5][6], simplification and compres- sion, mesh ... 1.e) of a mesh is computed by using a repulsive force field. (Fig.

3D Mesh Segmentation Using Mean-Shifted Curvature
meaningful and pleasing results, several ingredients are introduced into the .... function definition which uses the product of kernels to filter two feature components ..... Pan, X., Ye, X. and Zhang, S.: 3D Mesh segmentation using a two-stage ...

Online Determination of Track Loss Using Template Inverse Matching
Sep 29, 2008 - match between the real target and one of the virtual targets. Under this .... The experiments are implemented on a standard PC. (Pentium IV at ...

Inverse Aerodynamic Design using Double Adjoint ...
Dec 5, 2013 - communication with an optimization software, calculation of gradients, and verification and validation of the utilization of sensitivity analysis with ...

Using Inverse Optimal Control To Predict Human ...
σ min max. Joint center distances ... (red) and predicted motions with (blue) and without (green) replanning. We applied our framework to data gathered from two.

Procedural Mesh Splitting - GitHub
Jun 1, 2012 - Email: [email protected]. Website: http://danni.foxesgames.com ...... part of the object hierarchy making it obligatory. Components.

Simultaneous multidimensional deformation ...
Jul 20, 2011 - whose real part constitutes a moiré interference fringe pattern. Moiré fringes encode information about multiple phases which are extracted by introducing a spatial carrier in one of the object beams and subsequently using a Fourier

Explicit Meaning Transmission
Agents develop individual, distinct meaning structures, ..... In Proceed- ings of the AISB Symposium: Starting from Society – the application of social analogies to ...

deformation and quantization
Derivations of (1) have been given using brane quantization ... CP1 of complex structures aI + bJ + cK, a2 + b2 + c2 = 1, .... Use V to introduce the Ω-deformation:.

Mixtures of Inverse Covariances
class. Semi-tied covariances [10] express each inverse covariance matrix 1! ... This subspace decomposition method is known in coding ...... of cepstral parameter correlation in speech recognition,” Computer Speech and Language, vol. 8, pp.

Skeleton Extraction Using SSM of the Distance Transform
Sep 17, 2008 - Abstract. This paper proposes a novel approach of skeletonization for both binary and gray-scale images based on the Skeleton Strength Map (SSM) calculated from. Euclidean distance transform. The distance transform is firstly computed

Learning from weak representations using distance ...
1.3.4 Similar object classes and knowledge transfer . .... growing attention in the machine learning and computer vision communities .... “Learning distance function by Coding similarity” - Here we define 'similarity' and its goals using informat