Skip to main content
Log in

Mechanism creation in tensegrity structures by cellular morphogenesis

  • Original Paper
  • Published:
Acta Mechanica Aims and scope Submit manuscript

Abstract

Tensegrity structures are self-equilibrated statically and kinematically indeterminate structures. Cellular morphogenesis represents a generative process for the composition of complex tensegrity structures based on elementary cells. This article discusses the mechanism creation by the integration of mobility conditions in the generation of tensegrity structures using cellular morphogenesis. The creation of finite and infinitesimal mechanisms in tensegrity structures is described by the adhesion of cells sharing less than d nodes (d being the dimension of the workspace) and by the fusion of cells with the removal of two edges. Parametric descriptions of the infinitesimal displacements in the case of trivial and finite mechanisms are derived by the analysis of rigid assemblies corresponding to the rigid parts of the structure. In addition, an interpretation of the degeneracy of tensegrity structures in the case of self-stressable mechanisms is also presented. Analytical solutions of the degeneracy space for configurations resulting from adhesion and the fusion of two cells with the removal of two edges are described using symbolic calculations on the rigidity matrices of tensegrity structures. Although the study focuses on selected configurations and arrangements, the generalization of the ideas and findings included in this paper can lead to the generation of tensegrity structures with predefined static as well as kinematic properties, thus further enabling the application of tensegrity systems in science and engineering.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17

Similar content being viewed by others

References

  1. Karnovsky, I.A., Lebed, O.: Advanced Methods of Structural Analysis. Springer, Berlin (2010)

    Book  Google Scholar 

  2. Maxwell, J.C.: On the calculation of the equilibrium and stiffness of frames. Lond. Edinb. Dublin Phil. Mag. J. Sci. 27(182), 294–299 (1864)

    Article  Google Scholar 

  3. Mohr, O.: Beitrag zur Theorie des Fachwerks. Der Civilingenieur 31, 289–310 (1885)

    MATH  Google Scholar 

  4. Levi-Civita, T., Amaldi, U.: Lezioni di meccanica razionale. Bull. Am. Math. Soc. 36, 781–782 (1930)

    Article  MathSciNet  Google Scholar 

  5. Laman, G.: On graphs and rigidity of plane skeletal structures. J. Eng. Math. 4(4), 331–340 (1970)

    Article  MathSciNet  Google Scholar 

  6. Asimow, L., Roth, B.: The rigidity of graphs. Trans. Am. Math. Soc. 245, 279–289 (1978)

    Article  MathSciNet  Google Scholar 

  7. Asimow, L., Roth, B.: The rigidity of graphs, II. J. Math. Anal. Appl. 68(1), 171–190 (1979)

    Article  MathSciNet  Google Scholar 

  8. Crapo, H.: Structural rigidity. Struct. Topol. 1979, núm. 1 (1979)

  9. Connelly, R.: Rigidity. In: Handbook of Convex Geometry. North-Holland, pp. 223–271 (1993)

  10. Connelly, R.: Generic global rigidity. Discrete Comput. Geom. 33(4), 549–563 (2005)

    Article  MathSciNet  Google Scholar 

  11. Whiteley, W.: The union of matroids and the rigidity of frameworks. SIAM J. Discrete Math. 1(2), 237–255 (1988)

    Article  MathSciNet  Google Scholar 

  12. Phillips Jr., C.: Topology of covalent non-crystalline solids I: short-range order in chalcogenide alloys. J. Non Cryst. Solids 34(2), 153–181 (1979)

    Article  Google Scholar 

  13. Phillips, J.C.: Topology of covalent non-crystalline solids II: medium-range order in chalcogenide alloys and A-Si (Ge). J. Non Cryst. Solids 43(1), 37–77 (1981)

    Article  Google Scholar 

  14. Thorpe, M.F., Duxbury, P.M. (eds.): Rigidity Theory and Applications. Springer, Berlin (1999)

    Google Scholar 

  15. Thomas, S., Tang, X., Tapia, L., Amato, N.M.: Simulating protein motions with rigidity analysis. J. Comput. Biol. 14(6), 839–855 (2007)

    Article  Google Scholar 

  16. Hermans, S.M.A., Pfleger, C., Nutschel, C., Hanke, C.A., Gohlke, H.: Rigidity theory for biomolecules: concepts, software, and applications. Wiley Interdisc. Rev. Comput. Mol. Sci. 7(4), e1311 (2017)

    Article  Google Scholar 

  17. Eren, T., Goldenberg, O.K., Whiteley, W., Yang, Y.R., Morse, A.S., Anderson, B.D. and Belhumeur, P.N.: Rigidity, computation, and randomization in network localization. In: IEEE INFOCOM 2004, vol. 4. IEEE, pp. 2673–2684 (2004)

  18. Zhu, Z., So, A.M.-C., Ye, Y.: Universal rigidity and edge sparsification for sensor network localization. SIAM J. Optim. 20(6), 3059–3081 (2010)

    Article  MathSciNet  Google Scholar 

  19. Kitson, D., Levene, R.H.: Graph rigidity for unitarily invariant matrix norms. J. Math. Anal. Appl. 491, 124353 (2020)

    Article  MathSciNet  Google Scholar 

  20. Schulze, B., Serocold, H., Theran, L.: Frameworks with coordinated edge motions. arXiv preprint arXiv:1807.05376 (2018)

  21. Karpenkov, O.: Tensegrities on the space of generic functions. arXiv preprint arXiv:1905.11262 (2019)

  22. Karpenkov, O., Müller, C.: Geometric criteria for realizability of tensegrities in higher dimensions. arXiv preprint arXiv:1907.02830 (2019)

  23. Sitharam, M., St John, A., Sidman, J.: Handbook of Geometric Constraint Systems Principles. CRC, Boca Raton (2018)

    Book  Google Scholar 

  24. Calladine, C.R., Pellegrino, S.: First-order infinitesimal mechanisms. Int. J. Solids Struct. 27(4), 505–515 (1991)

    Article  MathSciNet  Google Scholar 

  25. Calladine, C.R., Pellegrino, S.: Further remarks on first-order infinitesimal mechanisms. Int. J. Solids Struct. 29(17), 2119–2122 (1992)

    Article  Google Scholar 

  26. Pellegrino, S., Calladine, C.R.: Matrix analysis of statically and kinematically indeterminate frameworks. Int. J. Solids Struct. 22(4), 409–428 (1986)

    Article  Google Scholar 

  27. Pellegrino, S.: Analysis of prestressed mechanisms. Int. J. Solids Struct. 26(12), 1329–1350 (1990)

    Article  Google Scholar 

  28. Tarnai, T., Szabó, J.: Rigidity and stability of prestressed infinitesimal mechanisms. In New Approaches to Structural Mechanics, Shells and Biological Structures. Springer, Dordrecht, pp. 245–256 (2002)

  29. Tarnai, T., Lengyel, A.: A remarkable structure of Leonardo and a higher-order infinitesimal mechanism. J. Mech. Mater. Struct. 6(1), 591–604 (2011)

    Article  Google Scholar 

  30. Kuznetsov, E.N.: Singular configurations of structural systems. Int. J. Solids Struct. 36(6), 885–897 (1999)

    Article  Google Scholar 

  31. Kuznetsov, E.N.: Systems with infinitesimal mobility: part I—matrix analysis and first-order infinitesimal mobility. pp. 513–519 (1991)

  32. Kuznetsov, E. N.: Systems with infinitesimal mobility: part II—compound and higher-order infinitesimal mechanisms. pp. 520–526 (1991)

  33. Guest, S., Fowler, P.: Symmetry conditions and finite mechanisms. J. Mech. Mater. Struct. 2(2), 293–301 (2007)

    Article  Google Scholar 

  34. Schenk, M., Herder, J.L., Guest, S.D.: Design of a statically balanced tensegrity mechanism. In: ASME 2006 International Design Engineering Technical Conferences and Computers and Information in Engineering Conference. American Society of Mechanical Engineers Digital Collection, pp. 501–511 (2006)

  35. Salerno, G.: How to recognize the order of infinitesimal mechanisms: a numerical approach. Int. J. Numer. Methods Eng. 35(7), 1351–1395 (1992)

    Article  Google Scholar 

  36. Yuan, X.F., Dong, S.L.: Integral feasible prestress of cable domes. Comput. Struct. 81(21), 2111–2119 (2003)

    Article  Google Scholar 

  37. Geiger, D.H., Stefaniuk, A., Chen, D.: The design and construction of two cable domes for the Korean Olympics. In: Proceedings of the IASS Symposium on Shells, Membranes and Space Frames, vol. 2. Elsevier Science Publishers BV, pp. 265–272 (1986)

  38. Dhakal, S., Bhandary, N.P., Yatabe, R., Kinoshita, N., Glade, T.: Experimental, numerical and analytical modelling of a newly developed rockfall protective cable-net structure. Nat. Hazards Earth Syst. Sci. 11(12), 3197–3212 (2011)

    Article  Google Scholar 

  39. Jiang, P., Wang, Q.M., Zhao, Q.: Optimization and analysis on cable net structure supporting the reflector of the large radio telescope FAST. Appl. Mech. Mater. 94, 979–982 (2011)

    Article  Google Scholar 

  40. Tibert, A.G., Pellegrino, S.: Deployable tensegrity reflectors for small satellites. J. Spacecr. Rock. 39(5), 701–709 (2002)

    Article  Google Scholar 

  41. Schenk, M.: Statically Balanced Tensegrity Mechanisms. Delft University of Technology, A literature review. Department of BioMechanical Engineering (2005)

  42. Rhode-Barbarigos, L., Hadj Ali, N.B., Motro, R., Smith, I.F.C.: Designing tensegrity modules for pedestrian bridges. Eng. Struct. 32(4), 1158–1167 (2010)

    Article  Google Scholar 

  43. Arsenault, M., Gosselin, C.M.: Kinematic, static, and dynamic analysis of a spatial three-degree-of-freedom tensegrity mechanism, pp. 1061–1069 (2006)

  44. Arsenault, M., Gosselin, C.M.: Kinematic, static, and dynamic analysis of a planar one-degree-of-freedom tensegrity mechanism, pp. 1152–1160 (2005)

  45. Arsenault, M., Gosselin, C.M.: Static balancing of tensegrity mechanisms. ASME J. Mech. Des. 129(3), 295–300 (2007)

    Article  Google Scholar 

  46. Aloui, O., Orden, D., Rhode-Barbarigos, L.: Generation of planar tensegrity structures through cellular multiplication. Appl. Math. Model. 64, 71–92 (2018)

    Article  MathSciNet  Google Scholar 

  47. Aloui, O., Flores, J., Orden, D., Rhode-Barbarigos, L.: Cellular morphogenesis of three-dimensional tensegrity structures. Comput. Methods Appl. Mech. Eng. 346, 85–108 (2019)

    Article  MathSciNet  Google Scholar 

  48. De Guzman, M., Orden, D.: From graphs to tensegrity structures: geometric and symbolic approaches. Publicacions Matemàtiques, pp. 279–299 (2006)

  49. Aloui, O., Lin, J., Rhode-Barbarigos, L.: A theoretical framework for sensor placement, structural identification and damage detection in tensegrity structures. Smart Mater. Struct. 28(12), 125004 (2019)

    Article  Google Scholar 

  50. Koiter, W.T.: On Tarnai’s conjecture with reference to both statically and kinematically indeterminate structures. Report no. 788 of Laboratory for Engineering Mechanics (1984)

  51. Khosravi, M., Taylor, M.D.: The wedge product and analytic geometry. Am. Math. Mon. 115(7), 623–644 (2008)

    Article  MathSciNet  Google Scholar 

  52. Stein, P.: A note on the volume of a simplex. Am. Math. Mon. 73(3), 299–301 (1966)

    Article  MathSciNet  Google Scholar 

  53. Saxe, J.B.: Embeddability of weighted graphs in k-space is strongly NP-hard. In: Proceedings of 17th Allerton Conference in Communications, Control and Computing, Monticello, IL, pp. 480–489 (1979)

  54. Kim, J.W., Park, F.C.: Manipulability of closed kinematic chain. ASME J. Mech. Des. 120, 542–548 (1998)

    Article  Google Scholar 

  55. Purwar, A., Jin, Z., Ge, Q.J.: Rational motion interpolation under kinematic constraints of spherical 6r closed chains. J. Mech. Des. 130(6), 1223–1234 (2008)

    Article  Google Scholar 

  56. Park, F.C., Kim, J.W.: Singularity analysis of closed kinematic chains. ASME J. Mech. Des. 121(1), 32–38 (1999)

    Article  Google Scholar 

  57. Yuan, X., Li, A., Shen, Y., Qian, R.: Kinematic path analysis of kinematically indeterminate systems. KSCE J. Civ. Eng. 20(2), 813–819 (2016)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Landolf Rhode-Barbarigos.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A

  • Proving that \(\mathbf{t }_{{\mathrm{e}}_{\mathrm{i}}}\) is a trivial infinitesimal motion

Let \(\mathbf{t }_{{\mathrm{e}}_{\mathrm{i}}}\) be the infinitesimal displacement vector defined by:

$$\begin{aligned} \mathbf{t }_{{\mathbf{e }}_{\mathrm{i}}} = \left( {\begin{array}{c} {\mathbb {O}}_{{n_{v}}}\}x_{1} \\ \vdots \\ {\mathbb {l}}_{{n_{v}}}\}x_{i} \\ \vdots \\ {\mathbb {O}}_{{n_{v}}}\}x_{d} \\ \end{array}} \right) = \mathbf{e }_{\mathbf{i }}\otimes {\mathbb {l}}_{{n_{v}}}, \end{aligned}$$
(A.1)

and \(\mathrm {R}\) is the rigidity matrix of the structure defined by:

$$\begin{aligned} \mathrm {R = }\left( {\begin{array}{*{20}c} \mathrm {R}_{\mathrm{1}} &{} \mathrm {R}_{\mathrm{2}} &{} {\cdots } &{} \mathrm {R}_{\mathrm{3}}\\ \end{array} } \right) = \left( {\begin{array}{*{20}c} \mathrm {diag}\left( \mathrm {C}\mathrm {X}_{\mathrm{1}} \right) \mathrm {C} &{} \mathrm {diag}\left( \mathrm {C}\mathrm {X}_{\mathrm{2}} \right) \mathrm {C} &{} {\cdots } &{} \mathrm {diag}\left( \mathrm {C}\mathrm {X}_{\mathrm{d}} \right) \mathrm {C}\\ \end{array} } \right) . \end{aligned}$$
(A.2)

The vector \(\mathrm {R}\mathbf{t }_{{\mathrm{e}}_{\mathrm{i}}}\) can be written as:

$$\begin{aligned} \sum \limits _{\mathrm{k=1}}^\mathrm {d} {\mathrm {R}_{\mathrm{k}}\mathbf{t }}_{{\mathbf{e }}_{\mathrm{i}_{\mathrm{k}}}} \end{aligned}$$
(A.3)

where \(\mathrm {R}_{\mathrm{k}}\mathrm {=diag}\left( \mathrm {C}\mathrm {X}_{\mathrm{k}} \right) \mathrm {C}\) and \(\mathbf{t }_{{\mathrm{e}}_{{\mathrm{i}}_{\mathrm{k}}}}=\delta _{ik}{\mathbb {l}}_{{n_{v}}}\). Equation (A.2) reduces to:

$$\begin{aligned} \mathrm {diag}\left( \mathrm {C}\mathrm {X}_{\mathrm{i}} \right) \mathrm {C} {\mathbb {l}}_{{n_{v}}}={\mathbb {O}}_{{n_{v}}}. \end{aligned}$$
(A.4)

C is the connectivity matrix of the structure also known as the node-branch incidence matrix of the underlying graph \(G\left( V,E \right) \). In the graph \(G\left( V,E \right) \), directions to the edges \(\left( v_{i},v_{j} \right) \) can be assigned arbitrarily such that \(\left( v_{i},v_{j} \right) \ne \left( v_{j},v_{i} \right) \). The \(n_{e}\times n_{v}\) corresponding connectivity matrix C is indexed by the list of nodes V and the list of oriented edges E, and it is described by:

$$\begin{aligned} c_{\left( v_{i},v_{j} \right) ,v_{k}}=\left\{ {\begin{array}{cc} 1 &{} \quad \mathrm{if}\; v_{i} = v_{k} \\ -1 &{}\quad \mathrm{if}\; v_{j} = v_{k} \\ 0 &{} \quad \mathrm{otherwise} \end{array} }. \right. \end{aligned}$$
(A.5)

Clearly, the sum of the row elements of C is always 0 regardless of the topology of the graph. Consequently, vector \({\mathbb {l}}_{{n_{v}}}\) is always in the right nullspace of C. Thus,\( \mathbf{t }_{{\mathrm{e}}_{\mathrm{i}}}\)is in the right nullspace of the rigidity matrix \(\mathrm {R}\) even when the structure is rigid, proving that it is a trivial infinitesimal translation. In addition, relation (A.6)

$$\begin{aligned} \mathbf{t }_{\mathbf{e }_{\mathrm{i}}}^{\mathbf{T }} \mathbf{t }_{\mathbf{e }_{\mathrm{j}}} = \sum \limits _{k=1}^{n_{v}} {\delta _{ik}\delta _{kj}} {\mathbb {l}}_{{n_{v}}}^{\mathrm{T}}{\mathbb {l}}_{{n_{v}}}=\delta _{ij}n_{v}, \end{aligned}$$
(A.6)

proves that the basis is orthogonal. The number of these vectors is equal to the dimension of the infinitesimal translation space T. Therefore, \(\mathcal {B}\left( T \right) =\left( \mathbf{t }_{\mathbf{e }_{\mathbf{i }}} \right) _{\mathbf{e }_{\mathbf{i }}\varvec{\in }\mathcal {B}\left( R^{d} \right) }\) is a basis for T

  • Proving that \(\mathbf{r }_{x_{i}x_{j}}\)is a trivial infinitesimal motion

Let \(\mathbf{r }_{x_{i}x_{j}}\) be the infinitesimal displacement defined by:

$$\begin{aligned} \mathbf{r }_{x_{i}x_{j}} = \left( {\begin{array}{c} {\mathbb {O}}_{{n_{v}}}\}x_{1} \\ \vdots \\ - \mathbf{x} _{\mathbf{j }}\}x_{i} \\ \vdots \\ \mathbf{x} _{\mathbf{i }}\}x_{j} \\ \vdots \\ {\mathbb {O}}_{{n_{v}}}\}x_{d} \\ \end{array}} \right) = \left( \mathbf{e }_{\mathbf{j }}\otimes \mathbf{x }_{\mathbf{i }} \right) -\left( \mathbf{e }_{\mathbf{i }}\otimes \mathbf{x }_{\mathbf{j }} \right) . \end{aligned}$$
(A.7)

The vector \(\mathrm {R}\mathbf{r }_{x_{i}x_{j}}\) is equal to:

$$\begin{aligned} \mathrm {R}\mathbf{r }_{x_{i}x_{j}} = \sum \limits _{\mathrm{k=1}}^\mathrm {d} {\mathrm {R}_{\mathrm{k}}}\mathbf{r }_{{\mathbf{x _\mathbf{i} }\mathbf{x }}_{{\mathrm{j}}_{\mathrm{k}}}} \end{aligned}$$
(A.8)

where \(\mathrm {R}_{\mathrm{k}}\mathrm {=diag}\left( \mathrm {C}\mathbf{x }_{\mathbf{k }} \right) \mathrm {C}\) and \(\mathbf{r }_{{\mathbf{x _\mathbf{i} }\mathbf{x }}_{{\mathrm{j}}_{\mathrm{k}}}}={-\delta }_{ik}\mathbf{x }_{\mathbf{j }}+\delta _{jk}\mathbf{x }_{\mathbf{i }}\). Equation (A.7) reduces to:

$$\begin{aligned} diag\left( C\mathbf{x }_{\mathbf{j }} \right) C\mathbf{x }_{\mathbf{i }}-diag\left( C\mathbf{x }_{\mathbf{i }} \right) C\mathbf{x }_{{\varvec{j}}}. \end{aligned}$$
(A.9)

Noticing that \(diag\left( \mathbf{u } \right) \mathbf{v }=diag\left( \mathbf{v } \right) \mathbf{u }\), it follows that Eq. (A.7) evaluates to exactly \({\mathbb {O}}_{{n_{v}}}\) proving that \(\mathbf{r }_{x_{i}x_{j}}\) is in the right nullspace of the rigidity matrix.

  • Proving that \(\mathcal {B}\left( M^{t} \right) =\left( \left( \mathbf{t }_{\mathrm{e}_{\mathrm{i}}} \right) _{\mathrm{e}_{i}\in \mathcal {B}\left( R^{d} \right) },\left( \mathbf{r }_{x_{i}x_{j}} \right) _{\left( i,j \right) \in \left( {\begin{array}{l} \llbracket 1,d \rrbracket \\ 2 \\ \end{array}} \right) } \right) \) is a basis of the infinitesimal trivial motions

In the planar case, \(\mathcal {B}\left( M^{t} \right) =\left( \mathbf{t }_{\mathbf{e }_{1}},\mathbf{t }_{\mathbf{e }_{2}},\mathbf{r }_{xy} \right) \). Assume now that the structure has at least two distinct nodes. Let’s evaluate:

$$\begin{aligned} {\alpha }_{\mathrm{1}}\mathbf{t }_{\mathbf{e }_{1}}+\alpha _{2}\mathbf{t }_{\mathbf{e }_{2}}+\alpha _{3}\mathbf{r }_{xy} = {\mathbb {O}}_{{n_{v}}}. \end{aligned}$$
(A.10)

Then, consider the components of the summation vector that correspond to two distinct nodes in the structure \(v_{1}\left( x_{1},y_{1} \right) \) and \(v_{2}\left( x_{2},y_{2} \right) \). The corresponding equations are described by:

$$\begin{aligned} \alpha _{1}\left( {\begin{array}{l} 1 \\ 1 \\ 0 \\ 0 \\ \end{array}} \right) +\alpha _{2}\left( {\begin{array}{l} 0 \\ 0 \\ 1 \\ 1 \\ \end{array}} \right) +\alpha _{3}\left( {\begin{array}{l} -y_{1} \\ -y_{2} \\ x_{1} \\ x_{2} \\ \end{array}} \right) = \left( {\begin{array}{*{20}c} {\begin{array}{l} 1 \\ 1 \\ 0 \\ 0 \\ \end{array}} &{} {\begin{array}{l} 0 \\ 0 \\ 1 \\ 1 \\ \end{array}} &{} {\begin{array}{l} -y_{1} \\ -y_{2} \\ x_{1} \\ x_{2} \\ \end{array}}\\ \end{array} } \right) \left( {\begin{array}{l} \alpha _{1} \\ \alpha _{2} \\ \alpha _{3} \\ \end{array}} \right) = {\mathbb {O}}_{4}. \end{aligned}$$
(A.11)

A row echelon transformation of the matrix above can be described by:

$$\begin{aligned} \left( {\begin{array}{ccc} 1 &{} 0 &{} -y_{1} \\ 0 &{} 1 &{} x_{1} \\ 0 &{} 0 &{} y_{1}-y_{2} \\ 0 &{} 0 &{} x_{2}-x_{1} \\ \end{array}} \right) . \end{aligned}$$
(A.12)

This matrix is full rank when \(x_{1}\ne x_{2}\) or \(y_{1}\ne y_{2}\), which is always valid when two distinct nodes exist in the structure. Hence, the solution to Eq. (A.11) is \(\alpha _{1}=0, \alpha _{2}=\)0, and \(\alpha _{3}=0\). Thus, the vectors of the set \(\mathcal {B}\left( M^{t} \right) \) are linearly independent, and their number is equal to the dimension of \(M^{t}\). Consequently, \(\mathcal {B}\left( M^{t} \right) \) represents a basis for \(M^{t}\) when the non-degeneracy condition is fulfilled.

In three-dimensional space, \(\mathcal {B}\left( M^{t} \right) =\left( \mathbf{t }_{\mathbf{e }_{1}},\mathbf{t }_{\mathbf{e }_{2}},\mathbf{t }_{\mathbf{e }_{3}},\mathbf{r }_{xy},\mathbf{r }_{yz},\mathbf{r }_{zx} \right) \). Assume that the structure has at least three distinct nodes \(v_{1}\), \(v_{2}\), and \(v_{3}\) that do not lie on a line. Let’s evaluate:

$$\begin{aligned} {\alpha }_{\mathrm{1}}\mathbf{t }_{\mathbf{e }_{1}}+\alpha _{2}\mathbf{t }_{\mathbf{e }_{2}}+\alpha _{3}\mathbf{t }_{\mathbf{e }_{{\varvec{3}}}}{} \mathbf + \alpha _{4}\mathbf{r }_{xy}{} \mathbf + \alpha _{5}\mathbf{r }_{yz}{} \mathbf + \alpha _{6}\mathbf{r }_{zx} = {\mathbb {O}}_{{n_{v}}}. \end{aligned}$$
(A.13)

Then, consider the components of the summation vector that correspond to \(v_{1}\left( x_{1},y_{1},z_{1} \right) \), \(v_{2}\left( x_{2},y_{2},z_{2} \right) \) and \(v_{3}\left( x_{3},y_{3},z_{3} \right) \). The corresponding equations are described by:

$$\begin{aligned}&\alpha _{1}\left( {\begin{array}{l} 1 \\ 1 \\ 1 \\ 0 \\ 0 \\ 0 \\ 0 \\ 0 \\ 0 \\ \end{array}} \right) +\alpha _{2}\left( {\begin{array}{l} 0 \\ 0 \\ 0 \\ 1 \\ 1 \\ 1 \\ 0 \\ 0 \\ 0 \\ \end{array}} \right) +\alpha _{3}\left( {\begin{array}{l} 0 \\ 0 \\ 0 \\ 0 \\ 0 \\ 0 \\ 1 \\ 1 \\ 1 \\ \end{array}} \right) +\alpha _{4}\left( {\begin{array}{l} -y_{1} \\ -y_{2} \\ -y_{3} \\ x_{1} \\ x_{2} \\ x_{3} \\ 0 \\ 0 \\ 0 \\ \end{array}} \right) +\alpha _{5}\left( {\begin{array}{l} 0 \\ 0 \\ 0 \\ -z_{1} \\ -z_{2} \\ -z_{3} \\ y_{1} \\ y_{2} \\ y_{3} \\ \end{array}} \right) +\alpha _{6}\left( {\begin{array}{l} z_{1} \\ z_{2} \\ z_{3} \\ 0 \\ 0 \\ 0 \\ {-x}_{1} \\ -x_{2} \\ -x_{3} \\ \end{array}} \right) \nonumber \\&\quad = \left( {\begin{array}{*{20}c} {\begin{array}{l} 1 \\ 1 \\ 1 \\ 0 \\ 0 \\ 0 \\ 0 \\ 0 \\ 0 \\ \end{array}} &{} {\begin{array}{l} 0 \\ 0 \\ 0 \\ 1 \\ 1 \\ 1 \\ 0 \\ 0 \\ 0 \\ \end{array}} &{} {\begin{array}{l} 0 \\ 0 \\ 0 \\ 0 \\ 0 \\ 0 \\ 1 \\ 1 \\ 1 \\ \end{array}} &{} {\begin{array}{l} -y_{1} \\ -y_{2} \\ -y_{3} \\ x_{1} \\ x_{2} \\ x_{3} \\ 0 \\ 0 \\ 0 \\ \end{array}} &{} {\begin{array}{l} 0 \\ 0 \\ 0 \\ -z_{1} \\ -z_{2} \\ -z_{3} \\ y_{1} \\ y_{2} \\ y_{3} \\ \end{array}} &{} {\begin{array}{l} z_{1} \\ z_{2} \\ z_{3} \\ 0 \\ 0 \\ 0 \\ {-x}_{1} \\ -x_{2} \\ -x_{3} \\ \end{array}}\\ \end{array} } \right) \left( {\begin{array}{l} \alpha _{1} \\ \alpha _{2} \\ \alpha _{3} \\ \alpha _{4} \\ \alpha _{5} \\ \alpha _{6} \\ \end{array}} \right) = {\mathbb {O}}_{9}. \end{aligned}$$
(A.14)

The row echelon form of the matrix above can be calculated by symbolic calculations in MATLAB which show that the matrix is full rank when the non-degeneracy condition is fulfilled. This proves that the vectors of \(B\left( M^{t} \right) \) are linearly independent and their number is equal to the dimension of \(M^{t}\), which shows that this is a basis for the trivial infinitesimal motions in the space.

Appendix B

  • Proving that \(\mathbf{r }_{x_{i}x_{j}}^{{\bot }}\)is a trivial infinitesimal motion

Let \(\mathbf{r }_{x_{i}x_{j}}^{{\bot }}\) be the infinitesimal displacement defined by:

$$\begin{aligned} \mathbf{r }_{x_{i}x_{j}}^{{\bot }} = \left( {\begin{array}{c} {\mathbb {O}}_{{n_{v}}}\}x_{1} \\ \vdots \\ - \mathbf{(x }_{\mathbf{j }}-{\varvec{\omega }}_{j}\}x_{i} \\ \vdots \\ \mathbf ( {} \mathbf{x} _{\mathbf{i }}-{\varvec{\omega }}_{i}\}x_{j} \\ \vdots \\ {\mathbb {O}}_{{n_{v}}}\}x_{d} \\ \end{array}} \right) = \left( \mathbf{e }_{\mathbf{j }}\otimes {\mathbf{(x }}_{\mathbf{i }}-{\varvec{\omega }}_{\mathbf{i }} \right) -\left( \mathbf{e }_{\mathbf{i }}\otimes {\mathbf{(x }}_{\mathbf{j }}-{\varvec{\omega }}_{{\varvec{j}}} \right) = \mathbf{r }_{x_{i}x_{j}}{} \mathbf + \omega _{j}\mathbf{t }_{\mathbf{e }_{\mathbf{i }}}-\omega _{i}\mathbf{t }_{\mathbf{e }_{\mathbf{j }}}. \end{aligned}$$
(B.1)

It as shown in “Appendix A” that \(\mathbf{r }_{x_{i}x_{j}}\) , \(\mathbf{t }_{\mathbf{i }}\) and \(\mathbf{t }_{\mathbf{j }}\) are in the right nullspace of the rigidity matrix \(\mathrm {R}\). Consequently, \(\mathbf{r }_{x_{i}x_{j}}^{{\bot }}\) is in the right nullspace of \(\mathrm {R}\).

  • Proving that \(\mathbf{r }_{x_{i}x_{j}}^{{\bot }} {\bot }\mathbf{t }_{\mathbf{e }_{\mathbf{k }}}\) \(\varvec{\forall }\left( i\mathbf , j \right) \varvec{\in }\left( {\begin{array}{c} \llbracket 1,d \rrbracket \\ 2 \\ \end{array}} \right) \) and \(\forall \mathbf{e }_{\mathbf{k }}\in B\left( R^{d} \right) \)

Let’s evaluate the product \(\mathbf{t }_{\mathbf{e }_{\mathbf{k }}}^{\mathbf{T }}\mathbf{r }_{x_{i}x_{j}}^{{\bot }}\):

$$\begin{aligned} \begin{array}{l} \mathbf{t }_{\mathbf{e }_{\mathbf{k }}}^{\mathbf{T }}\mathbf{r }_{x_{i}x_{j}}^{{\bot }} = \mathbf{t }_{\mathbf{e }_{\mathbf{k }}}^{\mathbf{T }}\left( \mathbf{r }_{x_{i}x_{j}}{} \mathbf + \omega _{j}\mathbf{t }_{\mathbf{e }_{\mathbf{i }}}-\omega _{i}\mathbf{t }_{\mathbf{e }_{\mathbf{j }}} \right) \\ =\mathbf{t }_{\mathbf{e }_{\mathbf{k }}}^{\mathbf{T }}\mathbf{r }_{x_{i}x_{j}}{} \mathbf + \omega _{j}\mathbf{t }_{\mathbf{e }_{\mathbf{k }}}^{\mathbf{T }}\mathbf{t }_{\mathbf{e }_{\mathbf{i }}}-\omega _{i}\mathbf{t }_{\mathbf{e }_{\mathbf{k }}}^{\mathbf{T }}\mathbf{t }_{\mathbf{e }_{\mathbf{j }}} \\ =\left( \mathbf{e }_{{\varvec{k}}}\otimes {\mathbb {l}}_{{n_{v}}} \right) ^\mathbf{T }\left( \left( \mathbf{e }_{\mathbf{j }}\otimes \mathbf{x }_{\mathbf{i }} \right) -\left( \mathbf{e }_{\mathbf{i }}\otimes \mathbf{x }_{\mathbf{j }} \right) \right) +\omega _{j}\left( \mathbf{e }_{{\varvec{k}}}\otimes {\mathbb {l}}_{{n_{v}}} \right) ^\mathbf{T }\left( \mathbf{e }_{{\varvec{i}}}\otimes {\mathbb {l}}_{{n_{v}}} \right) -\omega _{i}\left( \mathbf{e }_{{\varvec{k}}}\otimes {\mathbb {l}}_{{n_{v}}} \right) ^\mathbf{T }\left( \mathbf{e }_{\mathbf{j }}\otimes {\mathbb {l}}_{{n_{v}}} \right) . \\ \end{array} \end{aligned}$$
(B.2)

Let ABC,  and D be 4 matrices such that the products AC and BD are defined. The mixed product property of the Kronecker product states that \((A\otimes B)(C\otimes D)=(AC\otimes BD)\). This property can be used to develop the expression in (B.2) which results in:

$$\begin{aligned} \begin{array}{l} \mathbf{t }_{\mathbf{e }_{\mathbf{k }}}^{\mathbf{T }}\mathbf{r }_{x_{i}x_{j}}^{{\bot }} = \left( \left( \mathbf{e }_{{\varvec{k}}}^{\mathbf{T }}\mathbf{e }_{\mathbf{j }}\otimes {{\mathbb {l}}_{{n_{v}}}^{\mathrm{T}}\mathbf{x }}_{\mathbf{i }} \right) -\left( \mathbf{e }_{{\varvec{k}}}^{\mathbf{T }}\mathbf{e }_{\mathbf{i }}\otimes {{\mathbb {l}}_{{n_{v}}}^{\mathrm{T}}\mathbf{x }}_{\mathbf{j }} \right) \right) +\omega _{j}\left( \mathbf{e }_{{\varvec{k}}}^{\mathbf{T }}\mathbf{e }_{{\varvec{i}}}\otimes {\mathbb {l}}_{{n_{v}}}^{\mathrm{T}}{\mathbb {l}}_{{n_{v}}} \right) -\omega _{i}\left( {\mathbf{e }_{{\varvec{k}}}^{\mathbf{T }}\mathbf{e }}_{\mathbf{j }}\otimes {{\mathbb {l}}_{{n_{v}}}^{\mathrm{T}}{\mathbb {l}}}_{{n_{v}}} \right) \\ =\delta _{jk}\left( \sum \limits _{\lambda =1}^{n_{v}} x_{i\lambda } \right) -\delta _{ik}\left( \sum \limits _{\lambda =1}^{n_{v}} x_{j\lambda } \right) +\omega _{j}\delta _{ik}n_{v}-\omega _{i}\delta _{jk}n_{v}. \\ \end{array} \end{aligned}$$
(B.3)

Obviously, when \(k\ne i\ne j\), this product evaluates to 0 since \(\delta _{ik}=\delta _{jk}=0\). Notice that \(\omega _{i}=\frac{1}{n_{v}}\left( \sum \limits _{\lambda =1}^{n_{v}} x_{i\lambda } \right) , \quad \omega _{j}=\frac{1}{n_{v}}\left( \sum \limits _{\lambda =1}^{n_{v}} x_{j\lambda } \right) \) by the definition of the geometric centroid. Consequently, this product is 0 also when \(k=i \)or\( k=j\). Hence, \(\mathbf{r }_{x_{i}x_{j}}^{{\bot }} { \bot }\mathbf{t }_{\mathbf{e }_{\mathbf{k }}}\) \(\varvec{\forall }\left( i\mathbf , j \right) \varvec{\in }\left( {\begin{array}{c} \llbracket 1,d \rrbracket \\ 2 \\ \end{array}} \right) \) and \(\forall \mathbf{e }_{\mathbf{k }}\in B\left( R^{d} \right) \).

  • Proving that \(B\left( M^{t} \right) =\left( \left( \mathbf{t }_{\mathrm{e}_{\mathrm{i}}}^{{\bot }} \right) _{\mathrm{e}_{i}\in B\left( R^{d} \right) },\left( \mathbf{r }_{x_{i}x_{j}}^{{\bot }} \right) _{\left( i,j \right) \in \left( {\begin{array}{c} \llbracket 1,d \rrbracket \\ 2 \\ \end{array}} \right) } \right) \) is a basis of the infinitesimal trivial motions

Let’s evaluate the linear combination:

$$\begin{aligned} {\alpha }_{\mathrm{1}}\mathbf{t }_{\mathbf{e }_{1}}^{{\bot }}+\alpha _{2}\mathbf{t }_{\mathbf{e }_{2}}^{{\bot }}+\alpha _{3}\mathbf{t }_{\mathbf{e }_{{\varvec{3}}}}^{{\bot }}{} \mathbf + \alpha _{4}\mathbf{r }_{xy}^{{\bot }}{} \mathbf + \alpha _{5}\mathbf{r }_{yz}^{{\bot }}{} \mathbf + \alpha _{6}\mathbf{r }_{zx}^{{\bot }} = {\mathbb {O}}_{{n_{v}}}. \end{aligned}$$
(B.4)

Equation (B.4) can be rewritten as:

$$\begin{aligned} {\alpha }_{\mathrm{1}}\mathbf{t }_{\mathbf{e }_{{\varvec{1}}}}+\alpha _{2}\mathbf{t }_{\mathbf{e }_{{\varvec{2}}}}+\alpha _{3}\mathbf{t }_{\mathbf{e }_{{\varvec{3}}}}{} \mathbf + \alpha _{4}\left( \mathbf{r }_{xy}{} \mathbf + \omega _{{\varvec{y}}}\mathbf{t }_{\mathbf{e }_{{\varvec{1}}}}-\omega _{{\varvec{x}}}\mathbf{t }_{\mathbf{e }_{{\varvec{2}}}} \right) \mathbf + \alpha _{5}\left( \mathbf{r }_{yz}{} \mathbf + \omega _{{\varvec{z}}}\mathbf{t }_{\mathbf{e }_{{\varvec{2}}}}-\omega _{{\varvec{y}}}\mathbf{t }_{\mathbf{e }_{{\varvec{3}}}} \right) \mathbf + \alpha _{6}\left( \mathbf{r }_{zx}{} \mathbf + \omega _{{\varvec{x}}}\mathbf{t }_{\mathbf{e }_{{\varvec{3}}}}-\omega _{{\varvec{z}}}\mathbf{t }_{\mathbf{e }_{{\varvec{1}}}} \right) = {\mathbb {O}}_{{n_{v}}}\nonumber \\ \end{aligned}$$
(B.5)

When the coefficients of each vector in Eq. (B.5) are grouped together, it becomes:

$$\begin{aligned} \left( {\alpha }_{\mathrm{1}}+\alpha _{4}\omega _{y}-\alpha _{6}\omega _{z} \right) \mathbf{t }_{\mathbf{e }_{1}}+\left( \alpha _{2}+\alpha _{5}\omega _{z}-\alpha _{4}\omega _{x} \right) \mathbf{t }_{\mathbf{e }_{2}}+\left( \alpha _{3}+\alpha _{6}\omega _{x}-\alpha _{5}\omega _{y} \right) \mathbf{t }_{\mathbf{e }_{{\varvec{3}}}}{} \mathbf + \alpha _{4}\mathbf{r }_{xy}{} \mathbf + \alpha _{5}\mathbf{r }_{yz}{} \mathbf + \alpha _{6}\mathbf{r }_{zx} = {\mathbb {O}}_{{n_{v}}}\nonumber \\ \end{aligned}$$
(B.6)

The vectors \(\left( \mathbf{t }_{\mathbf{e }_{1}},\mathbf{t }_{\mathbf{e }_{2}},\mathbf{t }_{\mathbf{e }_{3}},\mathbf{r }_{xy},\mathbf{r }_{yz},\mathbf{r }_{zx} \right) \) are already shown to be linearly independent in “Appendix A”. It follows that \(\alpha _{4}=\alpha _{5}=\alpha _{6}=0\), and consequently, \(\alpha _{1}=\alpha _{2}=\alpha _{3}=0\). Since the number of vectors is equal to the dimension of \(M^{T}\), \(B\left( M^{t} \right) =\left( \mathbf{t }_{\mathbf{e }_{1}}^{{\bot }},\mathbf{t }_{\mathbf{e }_{2}}^{{\bot }},\mathbf{t }_{\mathbf{e }_{3}}^{{\bot }},\mathbf{r }_{xy}^{{\bot }},\mathbf{r }_{yz}^{{\bot }},\mathbf{r }_{zx}^{{\bot }} \right) \) is a basis for \(M^{t}\) which concludes the proof for dimension \(d=3\). For the planar case, the proof is similar.

  • Geometric interpretation of \(\mathbf{r }_{x_{i}x_{j}}^{{\bot }}\)

Let \({\Omega }\left( \omega _{1},\cdots ,\omega _{d} \right) \) be the geometric centroid of \(n_{v}\) points \(\mathbf{p }_{\mathbf{i }}\left( x_{i1},\cdots ,x_{id} \right) \) in a d-dimensional space (\(d=2\) or \(d=3)\) . \(\omega _{k}\) is defined as \(\frac{1}{n_{v}}\sum \limits _{\lambda =1}^{n_{v}} x_{\lambda k} \). Consider the finite rotation in the plane around the centroid \({\Omega }\) or in space around an axis \(({\Omega , }\mathbf{e }_{\mathbf{1 }})\) with \(\mathbf{e }_{\mathbf{1 }}\in B\left( R^{d} \right) \). In both cases, the finite rotation matrix \({\mathrm {Rot}}_{d}\) is described by:

$$\begin{aligned} {\mathrm {Rot}}_{2}=\left( {\begin{array}{cc} \cos \left( \theta \right) &{} -\sin \left( \theta \right) \\ \sin \left( \theta \right) &{} cos\left( \theta \right) \\ \end{array} } \right) , \quad {\mathrm {Rot}}_{\mathrm{3}} = \left( {\begin{array}{ccc} \cos \left( \theta \right) &{} -\sin \left( \theta \right) &{} 0\\ \sin \left( \theta \right) &{} \cos \left( \theta \right) &{} 0\\ 0 &{} 0 &{} 1\\ \end{array} } \right) . \end{aligned}$$
(B.7)

However, the rotational transformation is described by:

$$\begin{aligned} {\mathbf{p' }_{\mathbf{i }}\mathrm {=Rot}}_{d}\left( \mathbf{p} _{\mathbf{i }}-{\varvec{\Omega }} \right) +{\varvec{\Omega }}. \end{aligned}$$
(B.8)

When \(\theta \) is infinitesimal, the transformations in Eq. (B.8) can be rewritten as:

$$\begin{aligned}&\mathbf{p' }_{\mathbf{i }} = \left( {\begin{array}{*{20}c} 1 &{} -\delta \theta \\ \delta \theta &{} 1\\ \end{array} } \right) \left( \mathbf{p} _{\mathbf{i }}-{\varvec{\Omega }} \right) +{\varvec{\Omega }}, \nonumber \\&\mathbf{p' }_{\mathbf{i }} = \left( {\begin{array}{*{20}c} 1 &{} -\delta \theta &{} 0\\ \delta \theta &{} 1 &{} 0\\ 0 &{} 0 &{} 1\\ \end{array} } \right) \left( \mathbf{p} _{\mathbf{i }}-{\varvec{\Omega }} \right) +{\varvec{\Omega }} \end{aligned}$$
(B.9)

which gives the infinitesimal displacements:

$$\begin{aligned}&\mathbf{r }_{xy}^{{\bot }}=\delta \theta \left( {\begin{array}{l} {-(y}_{i}-\omega _{y} \\ {(x}_{i}-\omega _{x}) \\ \end{array}} \right) \hbox {in } 2 d, \nonumber \\&\mathbf{r }_{xy}^{{\bot }}=\delta \theta \left( {\begin{array}{c} -{(y}_{i}-\omega _{y}) \\ (x_{i}-\omega _{x}) \\ 0 \\ \end{array}} \right) \hbox {in } 3 d. \end{aligned}$$
(B.10)

This proves that the infinitesimal rotations \(\mathbf{r }_{xy}^{{\bot }}\) when integrated describe finite rotations around the centroid of the structure or an axis passing by \({\Omega }\) and in the same direction as \(\mathbf{e }_{\mathbf{3 }}\). For the rotations, the interpretations are similar.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Aloui, O., Rhode-Barbarigos, L. Mechanism creation in tensegrity structures by cellular morphogenesis. Acta Mech 231, 4891–4917 (2020). https://doi.org/10.1007/s00707-020-02803-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00707-020-02803-7

Navigation