Skip to main content
Log in

Geometrically Exact Finite Element Formulations for Slender Beams: Kirchhoff–Love Theory Versus Simo–Reissner Theory

  • Original Paper
  • Published:
Archives of Computational Methods in Engineering Aims and scope Submit manuscript

Abstract

The present work focuses on geometrically exact finite elements for highly slender beams. It aims at the proposal of novel formulations of Kirchhoff–Love type, a detailed review of existing formulations of Kirchhoff–Love and Simo–Reissner type as well as a careful evaluation and comparison of the proposed and existing formulations. Two different rotation interpolation schemes with strong or weak Kirchhoff constraint enforcement, respectively, as well as two different choices of nodal triad parametrizations in terms of rotation or tangent vectors are proposed. The combination of these schemes leads to four novel finite element variants, all of them based on a \(C^1\)-continuous Hermite interpolation of the beam centerline. Essential requirements such as representability of general 3D, large deformation, dynamic problems involving slender beams with arbitrary initial curvatures and anisotropic cross-section shapes, preservation of objectivity and path-independence, consistent convergence orders, avoidance of locking effects as well as conservation of energy and momentum by the employed spatial discretization schemes, but also a range of practically relevant secondary aspects will be investigated analytically and verified numerically for the different formulations. It will be shown that the geometrically exact Kirchhoff–Love beam elements proposed in this work are the first ones of this type that fulfill all these essential requirements. On the contrary, Simo–Reissner type formulations fulfilling these requirements can be found in the literature very well. However, it will be argued that the shear-free Kirchhoff–Love formulations can provide considerable numerical advantages such as lower spatial discretization error levels, improved performance of time integration schemes as well as linear and nonlinear solvers and smooth geometry representation as compared to shear-deformable Simo–Reissner formulations when applied to highly slender beams. Concretely, several representative numerical test cases confirm that the proposed Kirchhoff–Love formulations exhibit a lower discretization error level as well as a considerably improved nonlinear solver performance in the range of high beam slenderness ratios as compared to two representative Simo–Reissner element formulations from the literature.

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.

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
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26
Fig. 27
Fig. 28

Similar content being viewed by others

References

  1. Ademir LX (2014) Static Kirchhoff rods under the action of external forces: integration via Runge–Kutta method. J Comput Methods Phys 2014:650365

    MATH  Google Scholar 

  2. Antman SS (1974) Kirchhoff’s problem for nonlinearly elastic rods. Quar Appl Math 32(3):221–240

    MathSciNet  MATH  Google Scholar 

  3. Antman SS (1995) Nonlinear problems of elasticity. Springer, New York

    MATH  Google Scholar 

  4. Argyris JH, Balmer H, Doltsinis J St., Dunne PC, Haase M, Kleiber M, Malejannakis GA, Mlejnek H-P, Müller M, Scharpf DW (1979) Finite element method: the natural approach. Comput Methods Appl Mech Eng 17–18(Part 1):1–106

    MATH  Google Scholar 

  5. Armero F, Valverde J (2012) Invariant Hermitian finite elements for thin Kirchhoff rods. I. Comput Methods Appl Mech Eng 213–216:427–457

    MATH  Google Scholar 

  6. Armero F, Valverde J (2012) Invariant hermitian finite elements for thin Kirchhoff rods. II. Comput Methods Appl Mech Eng 213–216:458–485

    MATH  Google Scholar 

  7. Arnold M, Brüls O (2007) Convergence of the generalized-\(\alpha \) scheme for constrained mechanical systems. Multibody Syst Dynam 18(2):185–202

    MathSciNet  MATH  Google Scholar 

  8. Ashwell DG, Sabir AB (1971) Limitations of certain curved finite elements when applied to arches. Int J Mech Sci 13(2):133–139

    Google Scholar 

  9. Ashwell DG, Sabir AB, Roberts TM (1971) Further studies in the application of curved finite elements to circular arches. Int J Mech Sci 13(6):507–517

    Google Scholar 

  10. Avello A, de Jaln JG, Bayo E (1991) Dynamics of flexible multibody systems using cartesian co-ordinates and large displacement theory. Int J Numer Methods Eng 32(8):1543–1563

    MATH  Google Scholar 

  11. Bathe K-J, Bolourchi S (1979) Large displacement analysis of three-dimensional beam structures. Int J Numer Methods Eng 14(7):961–986

    MATH  Google Scholar 

  12. Bathe K-J, Iosilevich A, Chapelle D (2000) An inf-sup test for shell finite elements. Comput Struc 75(5):439–456

    MathSciNet  Google Scholar 

  13. Battini J-M, Pacoste C (2002) Co-rotational beam elements with warping effects in instability problems. Comput Methods Appl Mech Eng 191(17–18):1755–1789

    MATH  Google Scholar 

  14. Bauchau OA, Bottasso CL (1999) On the design of energy preserving and decaying schemes for flexible, nonlinear multi-body systems. Comput Methods Appl Mech Eng 169(12):61–79

    MathSciNet  MATH  Google Scholar 

  15. Bauchau OA, Han S, Mikkola A, Matikainen MK (2014) Comparison of the absolute nodal coordinate and geometrically exact formulations for beams. Multibody Syst Dynam 32(1):67–85

    MathSciNet  Google Scholar 

  16. Bauer AM, Breitenberger M, Philipp B, Wüchner R, Bletzinger K-U (2016) Nonlinear isogeometric spatial Bernoulli beam. Comput Methods Appl Mech Eng 303:101–127

    MathSciNet  Google Scholar 

  17. Belytschko T, Hseih BJ (1973) Nonlinear transient finite element analysis with convected coordinates. Int J Numer Methods Eng 7:255–271

    Google Scholar 

  18. Belytschko T, Lawrence WG (1979) Applications of higher order corotational stretch theories to nonlinear finite element analysis. Comput Struc 10(1):175–182

    MATH  Google Scholar 

  19. Bergou M, Wardetzky M, Robinson S, Audoly B, Grinspun E (2008) Discrete elastic rods. ACM Trans Graph 27(3):1–63

    Google Scholar 

  20. Bertails F, Audoly B, Cani M-P, Querleux B, Leroy F, Lévêque J-L (2006) Super-helices for predicting the dynamics of natural hair. ACM Trans Graph 25(3):1180–1187

    Google Scholar 

  21. Betsch P, Steinmann P (2001) Constrained integration of rigid body dynamics. Comput Methods Appl Mech Eng 191(35):467–488

    MathSciNet  MATH  Google Scholar 

  22. Betsch P, Steinmann P (2002) Frame-indifferent beam finite elements based upon the geometrically exact beam theory. Int J Numer Methods Eng 54(12):1775–1788

    MathSciNet  MATH  Google Scholar 

  23. Bishop RL (1975) There is more than one way to frame a curve. Am Math Month 82(3):246–251

    MathSciNet  MATH  Google Scholar 

  24. Borri M, Bottasso C (1994) An intrinsic beam model based on a helicoidal approximation—Part I: formulation. Int J Numer Methods Eng 37(13):2267–2289

    MATH  Google Scholar 

  25. Borri M, Bottasso C (1994) An intrinsic beam model based on a helicoidal approximation—Part II: linearization and finite element implementation. Int J Numer Methods Eng 37(13):2291–2309

    MATH  Google Scholar 

  26. Bottasso CL, Borri M (1998) Integrating finite rotations. Comput Methods Appl Mech Eng 164(34):307–331

    MathSciNet  MATH  Google Scholar 

  27. Boyer F, De Nayer G, Leroyer A, Visonneau M (2011) Geometrically exact Kirchhoff beam theory: application to cable dynamics. J Comput Nonlinear Dynam 6:1–14

    Google Scholar 

  28. Boyer F, Primault D (2004) Finite element of slender beams in finite transformations: a geometrically exact approach. Int J Numer Methods Eng 59(5):669–702

    MathSciNet  MATH  Google Scholar 

  29. Brezzi F (1974) On the existence, uniqueness and approximation of saddle-point problems arising from lagrangian multipliers. ESAIM 8(R2):129–151

    MathSciNet  MATH  Google Scholar 

  30. Brezzi F, Fortin M (1991) Mixed and hybrid finite elements. Springer, New York

    MATH  Google Scholar 

  31. Brüls O, Cardona A (2010) On the use of Lie group time integrators in multibody dynamics. J Comput Nonlinear Dynam 5:031002

    Google Scholar 

  32. Brüls O, Cardona A, Arnold M (2012) Lie group generalized-\(\alpha \) time integration of constrained flexible multibody systems. Mech Mach Theor 48:121–137

    Google Scholar 

  33. Cannarozzi M, Molari L (2008) A mixed stress model for linear elastodynamics of arbitrarily curved beams. Int J Numer Methods Eng 74(1):116–137

    MathSciNet  MATH  Google Scholar 

  34. Cardona A, Géradin M (1988) A beam finite element non-linear theory with finite rotations. Int J Numer Methods Eng 26(11):2403–2438

    MATH  Google Scholar 

  35. Cardona A, Géradin M, Doan DB (1991) Rigid and flexible joint modelling in multibody dynamics using finite elements. Comput Methods Appl Mech Eng 89(1):395–418

    Google Scholar 

  36. Choit JK, Lim JK (1995) General curved beam elements based on the assumed strain fields. Comput Struc 55(3):379–386

    MATH  Google Scholar 

  37. Chung J, Hulbert GM (1993) A time integration algorithm for structural dynamics with improved numerical dissipation: the generalized-\(\alpha \) method. J Appl Mech 60:371–375

    MathSciNet  MATH  Google Scholar 

  38. Cosserat E, Cosserat F (1909) Théorie des Corps Déformables, 2nd edn. Traité de Physique, Paris

  39. Crisfield MA (1990) A consistent co-rotational formulation for non-linear, three-dimensional, beam-elements. Comput Methods Appl Mech Eng 81(2):131–150

    MATH  Google Scholar 

  40. Crisfield MA (1997) Non-linear finite element analysis of solids and structures: advanced topics. Wiley, New York

    MATH  Google Scholar 

  41. Crisfield MA (2003) Non-linear finite element analysis of solids and structures: essentials. Wiley, New York

    Google Scholar 

  42. Crisfield MA, Galvanetto U, Jelenić G (1997) Dynamics of 3-D co-rotational beams. Comput Mech 20(6):507–519

    MATH  Google Scholar 

  43. Crisfield MA, Jeleni G (1999) Objectivity of strain measures in the geometrically exact three-dimensional beam theory and its finite-element implementation. Proc R Soc London 455(1983):1125–1147

    MathSciNet  MATH  Google Scholar 

  44. Crivelli LA, Felippa CA (1993) A three-dimensional non-linear Timoshenko beam based on the core-congruential formulation. Int J Numer Methods Eng 36(21):3647–3673

    MATH  Google Scholar 

  45. Cyron CJ, Wall WA (2012) Numerical method for the simulation of the Brownian dynamics of rod-like microstructures with three-dimensional nonlinear beam elements. Int J Numer Methods Eng 90(8):955–987

    MathSciNet  MATH  Google Scholar 

  46. Demoures F, Gay-Balmaz F, Leyendecker S, Ober-Blöbaum S, Ratiu TS, Weinand Y (2015) Discrete variational Lie group formulation of geometrically exact beam dynamics. Numer Math 130(1):73–123

    MathSciNet  MATH  Google Scholar 

  47. Dill EH (1992) Kirchhoff’s theory of rods. Arch Hist Exact Sci 44(1):1–23

    MathSciNet  MATH  Google Scholar 

  48. Dukić EP, Jelenić G, Gaćeša M (2014) Configuration-dependent interpolation in higher-order 2D beam finite elements. Finite Elem Anal Design 78:47–61

    MathSciNet  Google Scholar 

  49. Durville D (2010) Simulation of the mechanical behaviour of woven fabrics at the scale of fibers. Int J Mater Form 3(2):1241–1251

    Google Scholar 

  50. Dvorkin EN, Oñate E, Oliver J (1988) On a non-linear formulation for curved Timoshenko beam elements considering large displacement/rotation increments. Int J Numer Methods Eng 26(7):1597–1613

    MATH  Google Scholar 

  51. Eugster S (2015) Geometric continuum mechanics and induced beam theories. Springer, New York

    MATH  Google Scholar 

  52. Eugster SR, Hesch C, Betsch P, Glocker Ch (2014) Director-based beam finite elements relying on the geometrically exact beam theory formulated in skew coordinates. Int J Numer Methods Eng 97(2):111–129

    MathSciNet  MATH  Google Scholar 

  53. Fan WW, Zhu WD (2016) An accurate singularity-free formulation of a three-dimensional curved Euler–Bernoulli beam for flexible multibody dynamic analysis. J Vibrat Acoust 138(5):051001

    Google Scholar 

  54. Felippa CA, Haugen B (2005) A unified formulation of small-strain corotational finite elements: I. Theory. Comput Methods Appl Mech Eng 194(21–24):2285–2335

    MATH  Google Scholar 

  55. Fried I (1973) Shape functions and the accuracy of arch finite elements. Am Inst Aeronaut Astronaut J 11:287–291

    MATH  Google Scholar 

  56. Frischkorn J, Reese S (2013) A solid-beam finite element and non-linear constitutive modelling. Comput Methods Appl Mech Eng 265:195–212

    MathSciNet  MATH  Google Scholar 

  57. Gadot B, Martinez OR, du Roscoat SR, Bouvard D, Rodney D, Orgéas L (2015) Entangled single-wire NiTi material: a porous metal with tunable superelastic and shape memory properties. Acta Mater 96:311–323

    Google Scholar 

  58. Géradin M, Cardona A (1989) Kinematics and dynamics of rigid and flexible mechanisms using finite elements and quaternion algebra. Comput Mech 4(2):115–135

    MATH  Google Scholar 

  59. Géradin M, Cardona A (2001) Flexible multibody dynamics: a finite element approach. Wiley, New York

    Google Scholar 

  60. Gerstmayr J, Shabana AA (2006) Analysis of thin beams and cables using the absolute nodal co-ordinate formulation. Nonlinear Dynam 45(1):109–130

    MATH  Google Scholar 

  61. Gerstmayr J, Sugiyama H, Mikkola A (2013) Review on the absolute nodal coordinate formulation for large deformation analysis of multibody systems. J Comput Nonlinear Dynam 8(3):1–12

    Google Scholar 

  62. Ghosh S, Roy D (2008) Consistent quaternion interpolation for objective finite element approximation of geometrically exact beam. Comput Methods Appl Mech Eng 198(3–4):555–571

    MathSciNet  MATH  Google Scholar 

  63. Ghosh S, Roy D (2009) A frame-invariant scheme for the geometrically exact beam using rotation vector parametrization. Comput Mech 44(1):103–118

    MATH  Google Scholar 

  64. Gonzalez O (1996) Time integration and discrete Hamiltonian systems. J Nonlinear Sci 6(5):449–467

    MathSciNet  MATH  Google Scholar 

  65. Goyal S, Perkins NC, Lee CL (2005) Nonlinear dynamics and loop formation in Kirchhoff rods with implications to the mechanics of DNA and cables. J Comput Phys 209(1):371–389

    MathSciNet  MATH  Google Scholar 

  66. Greco L, Cuomo M (2013) B-Spline interpolation of Kirchhoff–Love space rods. Comput Methods Appl Mech Eng 256:251–269

    MathSciNet  MATH  Google Scholar 

  67. Greco L, Cuomo M (2016) An isogeometric implicit mixed finite element for Kirchhoff space rods. Comput Methods Appl Mech Eng 298:325–349

    MathSciNet  Google Scholar 

  68. Gruttmann F, Sauer R, Wagner W (1998) A geometrical nonlinear eccentric 3D-beam element with arbitrary cross-sections. Comput Methods Appl Mech Eng 160(34):383–400

    MATH  Google Scholar 

  69. Hsiao KM, Lin JY, Lin WY (1999) A consistent co-rotational finite element formulation for geometrically nonlinear dynamic analysis of 3-D beams. Comput Methods Appl Mech Eng 169(1–2):1–18

    MATH  Google Scholar 

  70. Hsiao KM, Yang RT (1995) A co-rotational formulation for nonlinear dynamic analysis of curved Euler beam. Comput Struc 54(6):1091–1097

    MATH  Google Scholar 

  71. Hsiao KM, Yang RT, Lee AC (1994) A consistent finite element formulation for non-linear dynamic analysis of planar beam. Int J Numer Methods Eng 37(1):75–89

    MATH  Google Scholar 

  72. Hughes TJR (2000) The finite element method: linear static and dynamic finite element analysis. Dover, New York

    MATH  Google Scholar 

  73. Hughes TJR, Taylor RL, Kanoknukulchai W (1977) A simple and efficient finite element for plate bending. Int J Numer Methods Eng 11(10):1529–1543

    MATH  Google Scholar 

  74. Ibrahimbegović A (1995) On finite element implementation of geometrically nonlinear Reissner’s beam theory: three-dimensional curved beam elements. Comput Methods Appl Mech Eng 122(1–2):11–26

    MATH  Google Scholar 

  75. Ibrahimbegović A (1997) On the choice of finite rotation parameters. Comput Methods Appl Mech Eng 149(1–4):49–71

    MathSciNet  MATH  Google Scholar 

  76. Ibrahimbegović A, Frey F (1993) Finite element analysis of linear and non-linear planar deformations of elastic initially curved beams. Int J Numer Methods Eng 36(19):3239–3258

    MATH  Google Scholar 

  77. Ibrahimbegović A, Frey F, Kozar I (1995) Computational aspects of vector-like parametrization of three-dimensional finite rotations. Int J Numer Methods Eng 38(21):3653–3673

    MathSciNet  MATH  Google Scholar 

  78. Ibrahimbegović A, Mamouri S (2002) Energy conserving/decaying implicit time-stepping scheme for nonlinear dynamics of three-dimensional beams undergoing finite rotations. Comput Methods Appl Mech Eng 191(3738):4241–4258

    MATH  Google Scholar 

  79. Ibrahimbegović A, Taylor RL (2002) On the role of frame-invariance in structural mechanics models at finite rotations. Comput Methods Appl Mech Eng 191(45):5159–5176

    MathSciNet  MATH  Google Scholar 

  80. Iosilevich A, Bathe K-J, Brezzi F (1997) On evaluating the infsup condition for plate bending elements. Int J Numer Methods Eng 40(19):3639–3663

    MATH  Google Scholar 

  81. Iura M, Atluri SN (1988) Dynamic analysis of finitely stretched and rotated three-dimensional space-curved beams. Comput Struc 29:875–889

    MATH  Google Scholar 

  82. Jelenić G, Crisfield MA (1998) Interpolation of rotational variables in nonlinear dynamics of 3D beams. Int J Numer Methods Eng 43(7):1193–1222

    MathSciNet  MATH  Google Scholar 

  83. Jelenić G, Crisfield MA (1999) Geometrically exact 3D beam theory: implementation of a strain-invariant finite element for statics and dynamics. Comput Methods Appl Mech Eng 171(1–2):141–171

    MathSciNet  MATH  Google Scholar 

  84. Jelenić G, Saje M (1995) A kinematically exact space finite strain beam model: finite element formulation by generalized virtual work principle. Comput Methods Appl Mech Eng 120(12):131–161

    MathSciNet  MATH  Google Scholar 

  85. Jung P, Leyendecker S, Linn J, Ortiz M (2011) A discrete mechanics approach to the Cosserat rod theory—Part 1: Static equilibria. Int J Numer Methods Eng 85(1):31–60

    MathSciNet  MATH  Google Scholar 

  86. Kane C, Marsden JE, Ortiz M, West M (2000) Variational integrators and the Newmark algorithm for conservative and dissipative mechanical systems. Int J Numer Methods Eng 49(10):1295–1325

    MathSciNet  MATH  Google Scholar 

  87. Kapania RK, Li J (2003) A formulation and implementation of geometrically exact curved beam elements incorporating finite strains and finite rotations. Comput Mech 30(5):444–459

    MATH  Google Scholar 

  88. Karamanlidis D, Jasti R (1987) Curved mixed beam elements for the analysis of thin-walled free-form arches. Ingenieur Arch 57(5):361–367

    MATH  Google Scholar 

  89. Kirchhoff G (1859) Ueber das Gleichgewicht und die Bewegung eines unendlich dünnen elastischen Stabes. J für die reine und angewandte Math 56:285–313

    MathSciNet  Google Scholar 

  90. Koiter WT (1966) On the nonlinear theory of thin elastic shells. In: Proceedings of the Koninklijke Nederlandse Akademie van Wetenschappen, vol 69.

  91. Kondoh K, Tanaka K, Atluri SN (1986) An explicit expression for the tangent-stiffness of a finitely deformed 3-D beam and its use in the analysis of space frames. Comput Struc 24(2):253–271

    MATH  Google Scholar 

  92. Koschnick F (2004) Geometrische locking-effekte bei Finiten Elementen und ein allgemeines Konzept zu ihrer Vermeidung. Ph.D thesis, Lehrstuhl für Statik, Technische Universität München

  93. Kulachenko A, Denoyelle T, Galland S, Lindström SB (2012) Elastic properties of cellulose nanopaper. Cellulose 19(3):793–807

    Google Scholar 

  94. Kulachenko A, Uesaka T (2012) Direct simulations of fiber network deformation and failure. Mech Mater 51:1–14

    Google Scholar 

  95. Lang H, Arnold M (2012) Numerical aspects in the dynamic simulation of geometrically exact rods. Appl Numer Math 62(10):1411–1427

    MathSciNet  MATH  Google Scholar 

  96. Lang H, Linn J, Arnold M (2010) Multi-body dynamics simulation of geometrically exact Cosserat rods. Multibody Syst Dynam 25(3):285–312

    MathSciNet  MATH  Google Scholar 

  97. Langer J, Singer DA (1996) Lagrangian aspects of the Kirchhoff elastic rod. SIAM Rev 38(4):605–618

    MathSciNet  MATH  Google Scholar 

  98. Lazarus A, Miller JJT, Reis PM (2013) Continuation of equilibria and stability of slender elastic rods using an asymptotic numerical method. J Mech Phys Solids 61(8):1712–1736

    MathSciNet  MATH  Google Scholar 

  99. Le T-N, Battini J-M, Hjiaj M (2014) A consistent 3D corotational beam element for nonlinear dynamic analysis of flexible structures. Comput Methods Appl Mech Eng 269:538–565

    MathSciNet  MATH  Google Scholar 

  100. Lee PG, Sin HC (1993) Locking-free straight beam element based on curvature. Commun Numer Methods Eng 9(12):1005–1011

    MATH  Google Scholar 

  101. Lens EV, Cardona A (2008) A nonlinear beam element formulation in the framework of an energy preserving time integration scheme for constrained multibody systems dynamics. Comput Struc 86(12):47–63

    Google Scholar 

  102. Linn J (2016) Discrete kinematics of Cosserat rods based on the difference geometry of framed curves. In: The 4th joint international conference on multibody system dynamics, Montréal, Canada

  103. Linn J, Lang H, Tuganov A (2013) Geometrically exact Cosserat rods with Kelvin–Voigt type viscous damping. Mech Sci 4(1):79–96

    Google Scholar 

  104. Love AEH (1944) A treatise on the mathematical theory of elasticity. Dover, New York

    MATH  Google Scholar 

  105. Lyly M, Stenberg R, Vihinen T (1993) A stable bilinear element for the Reissner–Mindlin plate model. Comput Methods Appl Mech Eng 110(3–4):343–357

    MathSciNet  MATH  Google Scholar 

  106. Maier M, Müller KW, Heussinger C, Köhler S, Wall WA, Bausch AR, Lieleg O (2015) A single charge in the actin binding domain of fascin can independently tune the linear and non-linear response of an actin bundle network. Eur Phys J E 38(5):50

    Google Scholar 

  107. Marsden JE, Hughes TJR (1994) Mathematical foundations of elasticity. Dover, New York

    MATH  Google Scholar 

  108. Meier C, Grill MJ, Wall WA, Popp A (2016) Geometrically exact beam elements and smooth contact schemes for the modeling of fiber-based materials and structures. arXiv preprint, arXiv:1611.06436

  109. Meier C, Popp A, Wall WA (2014) An objective 3D large deformation finite element formulation for geometrically exact curved Kirchhoff rods. Comput Methods Appl Mech Eng 278:445–478

    MathSciNet  MATH  Google Scholar 

  110. Meier C, Popp A, Wall WA (2015) A locking-free finite element formulation and reduced models for geometrically exact Kirchhoff rods. Comput Methods Appl Mech Eng 290:314–341

    MathSciNet  Google Scholar 

  111. Meier C, Popp A, Wall WA (2016) A finite element approach for the line-to-line contact interaction of thin beams with arbitrary orientation. Comput Methods Appl Mech Eng 308:377–413

    MathSciNet  Google Scholar 

  112. Meier C, Wall WA, Popp A (2017) A unified approach for beam-to-beam contact. Comput Methods Appl Mech Eng 315:972–1010

    MathSciNet  Google Scholar 

  113. Müller KW, Bruinsma RF, Lieleg O, Bausch AR, Wall WA, Levine AJ (2014) Rheology of semiflexible bundle networks with transient linkers. Phys Rev Lett 112:238102

    Google Scholar 

  114. Müller KW, Meier C, Wall WA (2015) Resolution of sub-element length scales in Brownian dynamics simulations of biopolymer networks with geometrically exact beam finite elements. J Comput Phys 303:185–202

    MathSciNet  MATH  Google Scholar 

  115. Noor AK, Peters JM (1981) Mixed models and reduced/selective integration displacement models for nonlinear analysis of curved beams. Int J Numer Methods Eng 17(4):615–631

    MATH  Google Scholar 

  116. Petrov E, Géradin M (1998) Finite element theory for curved and twisted beams based on exact solutions for three-dimensional solids Part 1: beam concept and geometrically exact nonlinear formulation. Comput Methods Appl Mech Eng 165(1–4):43–92

    MATH  Google Scholar 

  117. Prathap G (1985) The curved beam/deep arch/finite ring element revisited. Int J Numer Methods Eng 21(3):389–407

    MATH  Google Scholar 

  118. Prathap G, Naganarayana BP (1990) Analysis of locking and stress oscillations in a general curved beam element. Int J Numer Methods Eng 30(1):177–200

    Google Scholar 

  119. Quarteroni A, Sacco R, Fausto S (2000) Numerical mathematics. Springer, New York

    MATH  Google Scholar 

  120. Reissner E (1972) On one-dimensional finite-strain beam theory: the plane problem. Zeitschrift für Angewandte Mathematik und Physik (ZAMP) 23(5):795–804

    MATH  Google Scholar 

  121. Reissner E (1981) On finite deformations of space-curved beams. Zeitschrift für Angewandte Mathematik und Physik (ZAMP) 32(6):734–744

    MATH  Google Scholar 

  122. Rodney D, Gadot B, Martinez OR, Roscoat SR, Orgéas L (2016) Reversible dilatancy in entangled single-wire materials. Nat Mater 15:72–77

    Google Scholar 

  123. Romero I (2004) The interpolation of rotations and its application to finite element models of geometrically exact rods. Comput Mech 34:121–133

    MathSciNet  MATH  Google Scholar 

  124. Romero I (2008) A comparison of finite elements for nonlinear beams: the absolute nodal coordinate and geometrically exact formulations. Multibody Syst Dynam 20(1):51–68

    MathSciNet  MATH  Google Scholar 

  125. Romero I (2008) Formulation and performance of variational integrators for rotating bodies. Comput Mech 42(6):825–836

    MathSciNet  MATH  Google Scholar 

  126. Romero I, Armero F (2002) An objective finite element approximation of the kinematics of geometrically exact rods and its use in the formulation of an energy-momentum conserving scheme in dynamics. Int J Numer Methods Eng 54(12):1683–1716

    MathSciNet  MATH  Google Scholar 

  127. Romero I, Urrecha M, Cyron CJ (2014) A torsion-free non-linear beam model. Int J Non-Linear Mech 58:1–10

    Google Scholar 

  128. Sander O (2010) Geodesic finite elements for Cosserat rods. Int J Numer Methods Eng 82(13):1645–1670

    MathSciNet  MATH  Google Scholar 

  129. Sansour C, Nguyen TL, Hjiaj M (2015) An energy-momentum method for in-plane geometrically exact Euler–Bernoulli beam dynamics. Int J Numer Methods Eng 102(2):99–134

    MathSciNet  MATH  Google Scholar 

  130. Sansour C, Wagner W (2003) Multiplicative updating of the rotation tensor in the finite element analysis of rods and shells: a path independent approach. Comput Mech 31(1):153–162

    MathSciNet  MATH  Google Scholar 

  131. Santos HAFA, Pimenta PM, Almeida JPM (2011) A hybrid-mixed finite element formulation for the geometrically exact analysis of three-dimensional framed structures. Comput Mech 48(5):591–613

    MathSciNet  MATH  Google Scholar 

  132. Schmidt MG, Ismail AE, Sauer RA (2015) A continuum mechanical surrogate model for atomic beam structures. Int J Numer Methods Eng 13(5):413–442

    Google Scholar 

  133. Schulz M, Filippou FC (2001) Non-linear spatial Timoshenko beam element with curvature interpolation. Int J Numer Methods Eng 50(4):761–785

    MATH  Google Scholar 

  134. Shabana AA, Hussien HA, Escalona JL (1998) Application of the absolute nodal coordinate formulation to large rotation and large deformation problems. J Mech Design 120(2):188–195

    Google Scholar 

  135. Shabana AA, Yakoub RY (2001) Three dimensional absolute nodal coordinate formulation for beam elements: theory. J Mech Design 123(4):606–613

    Google Scholar 

  136. Shi Y, Hearst JE (1994) The Kirchhoff elastic rod, the nonlinear Schrödinger equation, and DNA supercoiling. J Chem Phys 101(6):5186–5200

    Google Scholar 

  137. Shoemake K (1985) Animating rotation with quaternion curves. ACM SIGGRAPH Comput Graph 19(3):245–254

    Google Scholar 

  138. Simo JC (1985) A finite strain beam formulation. The three-dimensional dynamic problem. Part I. Comput Methods Appl Mech Eng 49:55–70

    MATH  Google Scholar 

  139. Simo JC, Hughes TJR (1986) On the variational foundations of assumed strain methods. J Appl Mech 53:51–54

    MathSciNet  MATH  Google Scholar 

  140. Simo JC, Vu-Quoc L (1986) A three-dimensional finite strain rod model. Part II: computational aspects. Comput Methods Appl Mech Eng 58:79–116

    MATH  Google Scholar 

  141. Simo JC, Vu-Quoc L (1988) On the dynamics in space of rods undergoing large motions: a geometrically exact approach. Comput Methods Appl Mech Eng 66(2):125–161

    MathSciNet  MATH  Google Scholar 

  142. Simo JC, Wong KK (1991) Unconditionally stable algorithms for rigid body dynamics that exactly preserve energy and momentum. Int J Numer Methods Eng 31(1):19–52

    MathSciNet  MATH  Google Scholar 

  143. Smolenski WM (1998) Statically and kinematically exact nonlinear theory of rods and its numerical verification. Comput Methods Appl Mech Eng 178(1–2):89–113

    MathSciNet  MATH  Google Scholar 

  144. Sonneville V, Cardona A, Brüls O (2014) Geometric interpretation of a non-linear beam finite element on the Lie group SE(3). Arch Mech Eng 61:305–329

    MATH  Google Scholar 

  145. Sonneville V, Cardona A, Brüls O (2014) Geometrically exact beam finite element formulated on the special Euclidean group. Comput Methods Appl Mech Eng 268:451–474

    MathSciNet  MATH  Google Scholar 

  146. Spurrier RA (1978) Comment on “singularity-free extraction of a quaternion from a direction-cosine matrix”. J Spacecraft Rockets 15:255–255

    Google Scholar 

  147. Stolarski H, Belytschko T (1982) Membrane locking and reduced integration for curved elements. J Appl Mech 49:172–176

    MATH  Google Scholar 

  148. Strang G, Fix G (2008) An analysis of the finite elment method. Wellesley-Cambrigde Press, Cambrigde

    MATH  Google Scholar 

  149. Tessler A, Spiridigliozzi L (1986) Curved beam elements with penalty relaxation. Int J Numer Methods Eng 23(12):2245–2262

    Google Scholar 

  150. Timoshenko SP (1921) On the correction for shear of the differential equation for transverse vibrations of prismatic bars. Philos Magn Series 41(245):744–746

    Google Scholar 

  151. Češarek P, Saje M, Zupan D (2012) Kinematically exact curved and twisted strain-based beam. Int J Solids Struc 49(13):1802–1817

    Google Scholar 

  152. Vu TD, Durville D, Davies P (2015) Finite element simulation of the mechanical behavior of synthetic braided ropes and validation on a tensile test. Int J Solids Struc 58:106–116

    Google Scholar 

  153. Wall WA (2017) BACI: a multiphysics simulation environment. Technical report. Technical University of Munich, Munich

    Google Scholar 

  154. Wang Q, Wang CM (2007) The constitutive relation and small scale parameter of nonlocal continuum mechanics for modelling carbon nanotubes. Nanotechnology 18(7):075702

    Google Scholar 

  155. Weeger O, Yeung S-K, Dunn ML (2016) Isogeometric collocation methods for Cosserat rods and rod structures. Comput Methods Appl Mech Eng 316:100. doi:10.1016/j.cma.2016.05.009

    MathSciNet  Google Scholar 

  156. Weiss H (2002) Dynamics of geometrically nonlinear rods: I. Mechanical models and equations of motion. Nonlinear Dynam 30(4):357–381

    MathSciNet  MATH  Google Scholar 

  157. Weiss H (2002) Dynamics of geometrically nonlinear rods: II. Numerical methods and computational examples. Nonlinear Dynam 30(4):383–415

    MathSciNet  MATH  Google Scholar 

  158. Wempner G (1969) Finite elements, finite rotations and small strains of flexible shells. Int J Solids Struc 5(2):117–153

    MATH  Google Scholar 

  159. Yang Y, Tobias I, Olson WK (1993) Finite element analysis of DNA supercoiling. J Chem Phys 98(2):1673–1686

    Google Scholar 

  160. Zhang Z, Qi Z, Wu Z, Fang H (2015) A spatial Euler–Bernoulli beam element for rigid-flexible coupling dynamic analysis of flexible structures. Shock Vibrat 2015:208127

    Google Scholar 

  161. Zhao Z, Ren G (2012) A quaternion-based formulation of Euler–Bernoulli beam without singularity. Nonlinear Dynam 67(3):1825–1835

    MathSciNet  MATH  Google Scholar 

  162. Zupan D, Saje M (2003) Finite-element formulation of geometrically exact three-dimensional beam theories based on interpolation of strain measures. Comput Methods Appl Mech Eng 192(49–50):5209–5248

    MathSciNet  MATH  Google Scholar 

  163. Zupan D, Saje M (2006) The linearized three-dimensional beam theory of naturally curved and twisted beams: the strain vectors formulation. Comput Methods Appl Mech Eng 195(33–36):4557–4578

    MATH  Google Scholar 

  164. Zupan E, Saje M, Zupan D (2013) On a virtual work consistent three-dimensional Reissner–Simo beam formulation using the quaternion algebra. Acta Mech 224(8):1709–1729

    MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christoph Meier.

Ethics declarations

Conflict of interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.

Appendices

Appendix A: Rotational Shape Function Matrices

In this appendix, the shape functions \(\widetilde{\mathbf {I}}^i(\xi )\) required for the multiplicative rotation increments

$$\begin{aligned} \Delta \varvec{\theta }(\xi ) = \sum _{i=1}^{n_{\Lambda }} \widetilde{\mathbf {I}}^i(\xi ) \Delta \hat{\varvec{\theta }}^i, \quad \Delta \varvec{\theta }^{\prime }(\xi ) = \sum _{i=1}^{n_{\Lambda }} \widetilde{\mathbf {I}}^{i \prime }(\xi ) \Delta \hat{\varvec{\theta }}^i. \end{aligned}$$
(188)

associated with the triad interpolation (105) and originally derived in [83] shall be presented:

$$\begin{aligned} \begin{array}{ll} \widetilde{\mathbf {I}}^i(\xi ) &= L^i(\xi ) \varvec{\Lambda }_r \mathbf {T}^{-1}(\varvec{\Phi }_{lh}(\xi )) \mathbf {T}(\varvec{\Phi }^i_l)\varvec{\Lambda }_r^T\\ &\quad+\, \delta ^{iI} \varvec{\Lambda }_r \left[ \mathbf {I_3} - \mathbf {T}^{-1} (\varvec{\Phi }_{lh}(\xi )) \left\{ \sum \limits _{j=1}^{n_{\Lambda }} L^j(\xi ) \mathbf {T}(\varvec{\Phi }^j_l) \right\} \right] \mathbf {v}^I \varvec{\Lambda }_r^T \\ &\quad+\, \delta ^{iJ} \varvec{\Lambda }_r \left[ \mathbf {I_3} - \mathbf {T}^{-1} (\varvec{\Phi }_{lh}(\xi )) \left\{ \sum \limits _{j=1}^{n_{\Lambda }} L^j(\xi ) \mathbf {T}(\varvec{\Phi }^j_l) \right\} \right] \mathbf {v}^J \varvec{\Lambda }_r^T. \end{array} \end{aligned}$$
(189)

In (189), no summation over double indices is applied. The vectors \(\mathbf {v}^I\) and \(\mathbf {v}^J\) are defined as

$$\begin{aligned} \begin{array}{ll} \mathbf {v}^I &{}= \frac{1}{2}\left( \mathbf {I}_3 + \frac{1}{\Phi ^{IJ}}\tan {\left( \frac{\Phi ^{IJ}}{4}\right) }\mathbf {S}(\varvec{\Phi }^{IJ})\right) , \\ \mathbf {v}^J &{}= \frac{1}{2}\left( \mathbf {I}_3 - \frac{1}{\Phi ^{IJ}}\tan {\left( \frac{\Phi ^{IJ}}{4}\right) }\mathbf {S}(\varvec{\Phi }^{IJ})\right) ,\\ \end{array} \end{aligned}$$
(190)

with the common abbreviation \(\Phi ^{IJ} = ||\varvec{\Phi }^{IJ}||\). Moreover, the arc-length derivative \(\widetilde{\mathbf {I}}^{i \prime }(\xi )\) reads:

$$\begin{aligned} \begin{array}{ll} \widetilde{\mathbf {I}}^{i \prime }(\xi )&=L^{i \prime }(\xi ) \varvec{\Lambda }_r \mathbf {T}^{-1}(\varvec{\Phi }_{lh}(\xi )) \mathbf {T}(\varvec{\Phi }^i_l)\varvec{\Lambda }_r^T\\ &\quad+\,L^i(\xi ) \varvec{\Lambda }_r \mathbf {T}^{-1}_{,s} (\varvec{\Phi }_{lh}(\xi )) \mathbf {T}(\varvec{\Phi }^i_l)\varvec{\Lambda }_r^T \\ &\quad-\,\delta ^{iI} \varvec{\Lambda }_r \Bigg ( \mathbf {T}^{-1}_{,s}(\varvec{\Phi }_{lh}(\xi )) \Bigg \{ \sum \limits _{j=1}^{n_{\Lambda }} L^j(\xi ) \mathbf {T}(\varvec{\Phi }^j_l) \Bigg \}\\ & \quad+\, \mathbf {T}^{-1}(\varvec{\Phi }_{lh}(\xi )) \Bigg \{ \sum \limits _{j=1}^{n_{\Lambda }} L^{j \prime } (\xi ) \mathbf {T}(\varvec{\Phi }^j_l) \Bigg \} \Bigg ) \mathbf {v}^I\varvec{\Lambda }_r^T \\ &\quad-\,\delta ^{iJ} \varvec{\Lambda }_r \Bigg ( \mathbf {T}^{-1}_{,s}(\varvec{\Phi }_{lh}(\xi )) \Bigg \{ \sum \limits _{j=1}^{n_{\Lambda }} L^j(\xi ) \mathbf {T}(\varvec{\Phi }^j_l) \Bigg \}\\ &\quad+\, \mathbf {T}^{-1}(\varvec{\Phi }_{lh}(\xi )) \Bigg \{ \sum \limits _{j=1}^{n_{\Lambda }} L^{j \prime } (\xi ) \mathbf {T}(\varvec{\Phi }^j_l) \Bigg \} \Bigg ) \mathbf {v}^J\varvec{\Lambda }_r^T. \end{array} \end{aligned}$$
(191)

Finally, the required arc-length derivative \(\mathbf {T}^{-1}_{,s}(\varvec{\Phi }_{lh}(\xi ))\) is given by (see also [40, 82]):

$$ \begin{aligned} {\mathbf{T}}_{{,s}}^{{ - 1}} ({\mathbf{\Phi }}_{{lh}} (\xi )) & = {\mathbf{\Phi }}_{{lh}}^{T} {\mathbf{\Phi }}_{{lh}}^{\prime } \frac{{\Phi _{{lh}} \sin \Phi _{{lh}} - 2(1 - \cos \Phi _{{lh}} )}}{{\Phi _{{lh}}^{4} }}{\mathbf{S}}({\mathbf{\Phi }}_{{lh}} ) \\ & \quad + \frac{{1 - \cos \Phi _{{lh}} }}{{\Phi _{{lh}}^{2} }}{\mathbf{S}}({\mathbf{\Phi }}_{{lh}}^{\prime } ) \\ & \quad + \frac{1}{{\Phi _{{lh}}^{2} }}\left( {1 - \frac{{\sin \Phi _{{lh}} }}{{\Phi _{{lh}} }}} \right)\left( {{\mathbf{S}}({\mathbf{\Phi }}_{{lh}} ){\mathbf{S}}({\mathbf{\Phi }}_{{lh}}^{\prime } ) + {\mathbf{S}}({\mathbf{\Phi }}_{{lh}}^{\prime } ){\mathbf{S}}({\mathbf{\Phi }}_{{lh}} )} \right) \\ & \quad + {\mathbf{\Phi }}_{{lh}}^{T} {\mathbf{\Phi }}_{{lh}}^{\prime } \frac{{3\sin \Phi _{{lh}} - \Phi _{{lh}} (2 + \cos \Phi _{{lh}} )}}{{\Phi _{{lh}}^{5} }}{\mathbf{S}}({\mathbf{\Phi }}_{{lh}} ){\mathbf{S}}({\mathbf{\Phi }}_{{lh}} ). \\ \end{aligned} $$
(192)

Here, the abbreviations \(\varvec{\Phi }_{lh} = \varvec{\Phi }_{lh}(\xi )\) and \({\Phi }_{lh} = ||\varvec{\Phi }_{lh}(\xi )||\) have been applied. The limit \(\mathbf {T}^{-1}_{,s}(\varvec{\Phi }_{lh}(\xi )) \rightarrow 0.5 \mathbf {S}(\varvec{\Phi }_{lh}^{\prime }(\xi ))\) can be derived for small angles \(\varvec{\Phi }_{lh}(\xi ) \rightarrow \mathbf {0}\) (see [83]).

Appendix B: Dirichlet Boundary Conditions and Joints

For many applications, the formulation of proper Dirichlet boundary conditions or joints between the nodes of different beam elements are of a high practical relevance. This appendix represents a brief summary, where the possibility of formulating some basic constraint conditions will be investigated for the SK-ROT and the SK-TAN element.

1.1 Appendix B.1: SK-ROT Element

Since the SK-ROT element simplifies the formulation of Dirichlet boundary conditions and kinematic constraints in many practically relevant cases, it will be considered first.

Dirichlet boundary conditions A simple support at element node a can be realized via

$$\begin{aligned} \hat{\mathbf {d}}^a=\hat{\mathbf {d}}_u^a=\hat{\mathbf {d}}_0^a \rightarrow \Delta \hat{\mathbf {d}}^a=\mathbf {0}. \end{aligned}$$
(193)

If a clamped end should be modeled, also the cross-section orientation has to be fixed, i.e.

$$\begin{aligned} \varvec{\Lambda }^a=\varvec{\Lambda }_u^a=\varvec{\Lambda }_0^a, \quad \text {and} \quad \hat{\varvec{\psi }}^a=\hat{\varvec{\psi }}_0^a \rightarrow \Delta \hat{\varvec{\theta }} ^a=\mathbf {0}. \end{aligned}$$
(194)

Thus, the modeling of Dirichlet boundary conditions for the employed translational and rotational degrees of freedom is similar to standard finite elements that are purely based on translational degrees of freedom. This procedure can also be extended to inhomogeneous conditions. However, the determination of \(\Delta \hat{\varvec{\theta }} ^a\) requires special care in this case:

$$\begin{aligned} \begin{array}{ll} \hat{\mathbf {d}}^a&{}=\hat{\mathbf {d}}_u^a(t) \rightarrow \Delta \hat{\mathbf {d}}_{n+1}^a=\hat{\mathbf {d}}_{u,n+1}^a-\hat{\mathbf {d}}_{u,n}^a, \\ \varvec{\Lambda }^a&{}=\varvec{\Lambda }_u^a(t) \rightarrow \exp { ( \mathbf {S}( \Delta \hat{\varvec{\theta }}_{n+1}^a ) ) } =\varvec{\Lambda }_{u,n+1}^a \varvec{\Lambda }_{u,n}^{a T}. \end{array} \end{aligned}$$
(195)

The multiplicative procedure of the second line simplifies to the additive procedure according to the first line if the prescribed rotation is additive, which only holds for 2D rotations.

Connections and joints A simple (moment-free) joint between the two nodes a and b of two connected elements reads:

$$\begin{aligned} \hat{\mathbf {d}}^b=\hat{\mathbf {d}}^a, \quad \delta \hat{\mathbf {d}}^b=\delta \hat{\mathbf {d}}^a, \quad \Delta \hat{\mathbf {d}}^b=\Delta \hat{\mathbf {d}}^a. \end{aligned}$$
(196)

Thus, the degrees of freedom \(\hat{\mathbf {d}}^b\) can be eliminated from the global system of equations in a standard manner by simply assembling the corresponding lines and columns of the global residual vector and stiffness matrix properly. A rigid joint between two elements prescribed at the nodes a and b additionally requires to suppress any relative rotation between the associated nodal triads. It is assumed that these nodal triads differ by some fixed relative rotation \(\varvec{\Lambda }_{0}\):

$$ \begin{aligned} {\mathbf{\Lambda }}^{a} = {\mathbf{\Lambda }}^{b} {\mathbf{\Lambda }}_{0} {\text{ or }}\exp ({\mathbf{S}}(\Delta \widehat{\varvec{\theta }}^{a} )) & = \exp ({\mathbf{S}}(\Delta \widehat{\varvec{\theta }}^{b} )){\mathbf{\Lambda }}_{0} \\ \to {\mathbf{\Lambda }}_{0} & = {\mathbf{\Lambda }}^{{bT}} {\mathbf{\Lambda }}^{a} . \\ \end{aligned} $$
(197)

From (197), the following relations between the associated rotation increments can be derived:

$$\begin{aligned} \begin{array}{ll} \delta \varvec{\Lambda }^{a}=\delta \varvec{\Lambda }^{b}\varvec{\Lambda }_{0} &{}\rightarrow \mathbf {S}(\delta \hat{\varvec{\theta }}^a) \varvec{\Lambda }^{a}= \mathbf {S}(\delta \hat{\varvec{\theta }}^b) \varvec{\Lambda }^{b}\varvec{\Lambda }_{0} \\ &{}\rightarrow \delta \hat{\varvec{\theta }}^b = \delta \hat{\varvec{\theta }}^a \rightarrow \Delta \hat{\varvec{\theta }}^b = \Delta \hat{\varvec{\theta }}^a. \end{array} \end{aligned}$$
(198)

Consequently, also the rotational degrees of freedom \(\hat{\varvec{\psi }}^b\) can be eliminated in a standard manner by simply assembling the corresponding lines and columns of the global residual vector and of the global stiffness matrix properly.

Remark

It is emphasized that a rigid joint according to (197) is formulated via right-translation of the rotation tensor \(\varvec{\Lambda }_{0}\). This is mandatory since a rigid joint represents a fixed orientation difference between material quantities, i.e. a fixed relative rotation with respect to material axes. A left-translation via

$$\begin{aligned} \varvec{\Lambda }^{a} = \varvec{\Lambda }_{0}\varvec{\Lambda }^{b} \,\, \rightarrow \,\, \varvec{\Lambda }_{0} = \varvec{\Lambda }^{a}\varvec{\Lambda }^{bT} \,\, \rightarrow \,\, \delta \hat{\varvec{\theta }}^b = \varvec{\Lambda }_{0}^T \delta \hat{\varvec{\theta }}^a \ne \delta \hat{\varvec{\theta }}^a, \end{aligned}$$
(199)

i.e. a fixed relative rotation with respect to spatial axes, has a different physical meaning.

Remark

If additive increments \(\Delta \hat{\varvec{\psi }}^a\) and \(\Delta \hat{\varvec{\psi }}^b\) of the rotation vectors \(\hat{\varvec{\psi }}^a\) and \(\hat{\varvec{\psi }}^b\) instead of the multiplicative increments \(\Delta \hat{\varvec{\theta }}^a\) and \(\Delta \hat{\varvec{\theta }}^b\) were applied in the linearization process, Eq. (198) has to be replaced by:

$$\begin{aligned} \Delta \hat{\varvec{\theta }}^b = \Delta \hat{\varvec{\theta }}^a \,\,\,\rightarrow \,\,\, \Delta \hat{\varvec{\psi }}^b = \mathbf {T}(\hat{\varvec{\psi }}^b)\mathbf {T}^{-1}(\hat{\varvec{\psi }}^a)\Delta \hat{\varvec{\psi }}^a \ne \Delta \hat{\varvec{\psi }}^a. \end{aligned}$$
(200)

In this case, a direct elimination of the degrees of freedom \(\hat{\varvec{\psi }}^b\) via a proper assembly of the global stiffness matrix is not possible. Instead, the corresponding columns have to be scaled with the matrix \(\mathbf {T}(\hat{\varvec{\psi }}^b)\mathbf {T}^{-1}(\hat{\varvec{\psi }}^a)\).

Remark

Physically reasonable boundary conditions can be completely defined by the cross-section orientation and centroid position. For all considered types of boundary conditions, the degrees of freedom \(\hat{t}^{a}\) and \(\hat{t}^{b}\), which are a measure for the nodal axial force, are part of the FEM solution and must not be prescribed.

1.2 Appendix B.2: SK-TAN Element

The treatment of the translational degrees of freedom required for the subsequent boundary conditions is identical to the last section and will therefore be omitted here.

Dirichlet boundary conditions In order to model a clamped end with the SK-TAN element, the simplest case of a tangent vector that is parallel to a global base vector, e.g. \(\hat{\mathbf {t}}^a \parallel \mathbf {e}_1\), is considered. Then, (193) has to be supplemented by

$$ \begin{aligned} \widehat{{\mathbf{t}}}^{{aT}} {\mathbf{e}}_{2} = \widehat{{\mathbf{t}}}^{{aT}} {\mathbf{e}}_{3} = 0 & \to \Delta \widehat{t}_{2}^{a} = \Delta \widehat{t}_{3}^{a} = 0, \\ \hat{\varphi }^{a} = \hat{\varphi }_{0}^{a} & \to \Delta \hat{\varphi }^{a} = 0. \\ \end{aligned} $$
(201)

Here, the representation \(\hat{\mathbf {t}}^{a} = \hat{{t}}^{a}_i \mathbf {e}_i\) of the tangent in the global frame \(\mathbf {e}_i\) has been exploited. In order to prescribe boundary conditions with arbitrary triad orientation, the tangent has to be expressed in the basis of the prescribed triad:

$$ \begin{aligned} {\mathbf{\Lambda }}^{a} = {\mathbf{\Lambda }}_{u}^{a} = {\mathbf{\Lambda }}_{0}^{a} ,\,\widehat{{\mathbf{t}}}^{a} = \hat{T}_{i}^{a} {\mathbf{g}}_{i}^{a} & \to \hat{T}_{2}^{a} = \hat{T}_{3}^{a} = \Delta \hat{T}_{2}^{a} = \Delta \hat{T}_{3}^{a} = 0, \\ \hat{\varphi }^{a} = \hat{\varphi }_{0}^{a} & \to \Delta \hat{\varphi }^{a} = 0. \\ \end{aligned} $$
(202)

Consequently, in this case, the equations of the linearized residual vector that are associated with the degrees of freedom \(\hat{\mathbf {t}}^{a}\) have to be transformed by the rotation tensor \(\varvec{\Lambda }_0^a\) and the Dirichlet conditions have to be formulated in this rotated coordinate system. Again, the first component \(\hat{T}_1\) of the tangent vector, when expressed in the material frame, represents its magnitude and must not be prescribed. If the Dirichlet conditions are time-dependent, the prescribed evolution of the relative angle has to be adapted, since now the intermediate frame \(\mathbf {\Lambda }_{M_{\hat{\varphi }}}^a\) might change in time:

$$ \begin{aligned} \exp ({\mathbf{S}}[\hat{\varphi }_{{n + 1}}^{a} {\mathbf{g}}_{{1,n + 1}}^{a} ]) & = {\mathbf{\Lambda }}_{{n + 1}}^{a} {\mathbf{\Lambda }}_{{M_{{\hat{\varphi }}} ,n + 1}}^{{aT}} \\ {\text{with }}{\mathbf{\Lambda }}_{{M_{{\hat{\varphi }}} ,n + 1}}^{a} & = {\text{sr}}({\mathbf{\Lambda }}_{{M_{{\hat{\varphi }}} ,n}}^{a} ,{\mathbf{g}}_{{1,n + 1}}^{a} ). \\ \end{aligned} $$
(203)

Thus, the required value \(\hat{\varphi }^a_{n+1}\) has to be determined based on the prescribed current triad \(\mathbf {\Lambda }^a_{n+1}\) and the intermediate triad \(\mathbf {\Lambda }_{M_{\hat{\varphi }},n}^{a}\) of the last step (see Sect. 6.2.2). The remaining conditions remain unchanged as compared to (202).

Connections and joints Based on (197), (198) and (25), the relations between \((\hat{\mathbf {t}}^a,\hat{\varphi }^a)\) and \((\hat{\mathbf {t}}^b,\hat{\varphi }^b)\) can be stated:

$$\begin{aligned} \begin{array}{ll} \delta \hat{\mathbf {t}}^b&{}=-t^b \mathbf {S}(\mathbf {g}_1^b) \underbrace{\delta \varvec{\theta }^b}_{\dot{=}\delta \varvec{\theta }^a} + \mathbf {g}_{1}^b \delta t^b\\ &{}= -t^b \mathbf {S}(\mathbf {g}_1^b) \left( \frac{1}{t^a} \mathbf {S}(\mathbf {g}_1^a) \delta \hat{\mathbf {t}}^a + \mathbf {g}_{1}^a \delta \hat{\Theta }_1^a \right) + \mathbf {g}_{1}^b \delta \hat{t}^b, \\ \delta \hat{\Theta }_1^b&{}=\mathbf {g}_{1}^{bT} \underbrace{\delta \varvec{\theta }^b}_{\dot{=}\delta \varvec{\theta }^a}= \mathbf {g}_{1}^{bT} \left( \frac{1}{t^a} \mathbf {S}(\mathbf {g}_1^a) \delta \hat{\mathbf {t}}^a + \mathbf {g}_{1}^a \delta \hat{\Theta }_1^a \right) . \end{array} \end{aligned}$$
(204)

Combining these two relations eventually yields the following total transformation matrix:

$$ \begin{aligned} \left( {\begin{array}{*{20}c} {\delta \widehat{{\mathbf{t}}}^{b} } \\ {\delta \hat{\Theta }_{1}^{b} } \\ \end{array} } \right) & = {\mathbf{T}}_{{RC}} \left( {\begin{array}{*{20}c} {\delta \widehat{{\mathbf{t}}}^{a} } \\ {\delta \hat{\Theta }_{1}^{a} } \\ {\delta \hat{t}^{b} } \\ \end{array} } \right), \\ {\mathbf{T}}_{{RC}} : & = \left( {\begin{array}{*{20}c} { - t^{b} {\mathbf{S}}({\mathbf{g}}_{1}^{b} )\frac{1}{{t^{a} }}{\mathbf{S}}({\mathbf{g}}_{1}^{a} )} & { - t^{b} {\mathbf{S}}({\mathbf{g}}_{1}^{b} ){\mathbf{g}}_{1}^{a} } & {{\mathbf{g}}_{1}^{b} } \\ {{\mathbf{g}}_{1}^{{bT}} \frac{1}{{t^{a} }}{\mathbf{S}}({\mathbf{g}}_{1}^{a} )} & {{\mathbf{g}}_{1}^{{bT}} {\mathbf{g}}_{1}^{a} } & 0 \\ \end{array} } \right). \\ \end{aligned} $$
(205)

A similar relation can also be formulated for the iterative increments. Since the multiplicative rotation increment components \(\Delta \Theta _1 = \mathbf {T}_{\Theta _{M1} \mathbf {t}} \Delta \mathbf {t} + \Delta \varphi \) [see (20)] have to be expressed by additive increments \(\Delta \mathbf {t}\) and \(\Delta \varphi \) for the chosen linearization, an additional transformation is required compared to (205):

$$ \begin{aligned} \left( {\begin{array}{*{20}c} {\Delta \widehat{{\mathbf{t}}}^{b} } \\ {\Delta \hat{\varphi }^{b} } \\ \end{array} } \right) & = \widetilde{{\mathbf{T}}}_{{RC1}} {\mathbf{T}}_{{RC}} \widetilde{{\mathbf{T}}}_{{RC2}} \left( {\begin{array}{*{20}c} {\Delta \widehat{{\mathbf{t}}}^{a} } \\ {\Delta \hat{\varphi }_{1}^{a} } \\ {\Delta \hat{t}^{b} } \\ \end{array} } \right), \\ \widetilde{{\mathbf{T}}}_{{RC1}} : & = \left( {\begin{array}{*{20}c} {{\mathbf{I}}_{3} } & {\mathbf{0}} \\ { - {\mathbf{T}}_{{\Theta _{{M1}} {\mathbf{t}}}} } & 1 \\ \end{array} } \right),\,\widetilde{{\mathbf{T}}}_{{RC2}} : = \left( {\begin{array}{*{20}c} {{\mathbf{I}}_{3} } & {\mathbf{0}} & {\mathbf{0}} \\ {{\mathbf{T}}_{{\Theta _{{M1}} {\mathbf{t}}}} } & 1 & 0 \\ {{\mathbf{0}}^{T} } & 0 & 1 \\ \end{array} } \right). \\ \end{aligned} $$
(206)

Equations (205) and (206) allow to transform the corresponding lines and columns of the global residual vector and of the global tangent stiffness matrix properly and to eliminate the degrees of freedom \((\hat{\mathbf {t}}^b,\hat{\varphi }^b)\) from the global system of equations. Again, the magnitude of the tangent vector \(\hat{t}^b\) is not influenced by the rigid joint and enters the system of equations as new degree of freedom. While in the last section, the motion of the rigid joint was completely determined by the set \((\hat{\mathbf {d}}^a, \hat{\varvec{\psi }}^a,\hat{t}^a,\hat{t}^b)\), in this section the alternative set \((\hat{\mathbf {d}}^a,\hat{\mathbf {t}}^a,\hat{\varphi }^a,\hat{t}^b)\) is employed.

All in all, it can be concluded that the realization of clamped ends with arbitrary orientation or of rigid joints between beams is simpler for the SK-ROT formulation based on nodal rotation vectors. While for these elements such conditions can directly be formulated in the global coordinate system, the tangent vector-based SK-TAN formulation requires an additional transformation of the corresponding lines and columns of the global residual vector and stiffness matrix. In Sect. 11, some properties of the tangent vector-based variant will become apparent which make this type of formulation favorable for many applications. If certain element nodes require Dirichlet conditions of the type considered here, it is still possible to apply a hybrid approach, and to replace the nodal tangents by nodal rotation vectors as primary variables only at the specific nodes where such conditions are required. All the results derived so far apply in a similar manner to the WK-TAN and WK-ROT elements.

Appendix C: Linearization of SK-TAN Element

Before deriving the linearization of the SK-TAN element, some former definitions are repeated:

$$\begin{aligned} \begin{array}{ll} \mathbf {t} :=&{}\mathbf {r}^{\prime }, \quad \mathbf {g}_1 := \frac{\mathbf {r}^{\prime }}{||\mathbf {r}^{\prime }||}, \quad \,\,\,\, \widetilde{\mathbf {t}} := \frac{\mathbf {r}^{\prime }}{||\mathbf {r}^{\prime }||^2}, \\ \mathbf {g}_1^{\prime } =&{} \frac{\mathbf {\mathbf {r}^{\prime \prime }}}{||\mathbf {r}^{\prime }||} - \frac{(\mathbf {r}^{\prime T}\mathbf {r}^{\prime \prime })\mathbf {r}^{\prime }}{||\mathbf {r}^{\prime }||^3}, \quad \widetilde{\mathbf {t}}^{\prime } = \frac{\mathbf {r}^{\prime \prime }}{||\mathbf {r}^{\prime }||^2} - \frac{2(\mathbf {r}^{\prime T}\mathbf {r}^{\prime \prime })\mathbf {r}^{\prime }}{||\mathbf {r}^{\prime }||^4}. \end{array} \end{aligned}$$
(207)

These quantities will be required for later derivations. Linearization of (207) yields:

$$ \begin{aligned} \Delta {\mathbf{g}}_{1} & = \frac{1}{{||{\mathbf{r}}^{\prime } ||}}\left( {{\mathbf{I}}_{3} - {\mathbf{g}}_{1} \otimes {\mathbf{g}}_{1}^{T} } \right)\mathsf{H}^{\prime } \Delta \widehat{\mathsf{d}}, \\ \Delta \widetilde{{\mathbf{t}}} & = \frac{1}{{||{\mathbf{r}}^{\prime } ||^{2} }}\left( {{\mathbf{I}}_{3} - 2{\mathbf{g}}_{1} \otimes {\mathbf{g}}_{1}^{T} } \right)\mathsf{H}^{\prime } \Delta \widehat{\mathsf{d}}, \\ \Delta {\mathbf{g}}_{1}^{\prime } & = - \frac{{({\mathbf{r}}^{{\prime T}} {\mathbf{r}}^{{\prime \prime }} )}}{{||{\mathbf{r}}^{\prime } ||^{3} }}\left( {{\mathbf{I}}_{3} - {\mathbf{g}}_{1} \otimes {\mathbf{g}}_{1}^{T} } \right)\mathsf{H}^{\prime } \Delta \widehat{\mathsf{d}} \\ & \quad - \frac{1}{{||{\mathbf{r}}^{\prime } ||}}\left( {{\mathbf{g}}_{1}^{\prime } \otimes {\mathbf{g}}_{1}^{T} + {\mathbf{g}}_{1} \otimes {\mathbf{g}}_{1}^{{\prime T}} } \right)\mathsf{H}^{\prime } \Delta \widehat{\mathsf{d}} \\ & \quad + \frac{1}{{||{\mathbf{r}}^{\prime } ||}}\left( {{\mathbf{I}}_{3} - {\mathbf{g}}_{1} \otimes {\mathbf{g}}_{1}^{T} } \right)\mathsf{H}^{{\prime \prime }} \Delta \widehat{\mathsf{d}}, \\ \Delta \widetilde{{\mathbf{t}}}^{\prime } & = - \frac{{2({\mathbf{r}}^{{\prime T}} {\mathbf{r}}^{{\prime \prime }} )}}{{||{\mathbf{r}}^{\prime } ||^{4} }}\left( {{\mathbf{I}}_{3} - 2{\mathbf{g}}_{1} \otimes {\mathbf{g}}_{1}^{T} } \right)\mathsf{H}^{\prime } \Delta \widehat{\mathsf{d}} \\ & \quad - \frac{2}{{||{\mathbf{r}}^{\prime } ||^{2} }}\left( {{\mathbf{g}}_{1}^{\prime } \otimes {\mathbf{g}}_{1}^{T} + {\mathbf{g}}_{1} \otimes {\mathbf{g}}_{1}^{{\prime T}} } \right)\mathsf{H}^{\prime } \Delta \widehat{\mathsf{d}} \\ & \quad + \frac{1}{{||{\mathbf{r}}^{\prime } ||^{2} }}\left( {{\mathbf{I}}_{3} - 2{\mathbf{g}}_{1} \otimes {\mathbf{g}}_{1}^{T} } \right)\mathsf{H}^{{\prime \prime }} \Delta \widehat{\mathsf{d}}. \\ \end{aligned} $$
(208)

In the following, the linearization of the SK-TAN element based on the residual vector (165) will be derived. The linearization of (165) obeys the following general form:

$$ \begin{aligned} \Delta \mathsf{r}_{{\widehat{{\mathbf{d}}}}} & = \int\limits_{{ - 1}}^{1} {\left( {\Delta \mathsf{v}_{{\theta _{ \bot } }}^{\prime } {\mathbf{m}} + \mathsf{v}_{{\theta _{ \bot } }}^{\prime } \Delta {\mathbf{m}} + \Delta \overline{\mathsf{v}} _{\epsilon{}} \bar{F}_{1} + \overline{\mathsf{v}} _{\epsilon{}} \Delta \bar{F}_{1} } \right)Jd\xi } \\ & \quad - \int\limits_{{ - 1}}^{1} {\left( {\mathsf{H}^{T} \Delta {\mathbf{f}}_{\rho } + \Delta \mathsf{v}_{{\theta _{ \bot } }} \widetilde{{\mathbf{m}}}_{\rho } + \mathsf{v}_{{\theta _{ \bot } }} \Delta {\mathbf{m}}_{\rho } } \right)Jd\xi - [\Delta \mathsf{v}_{{\theta _{ \bot } }} {\mathbf{m}}_{\sigma } ]_{{\varGamma_{\sigma } }} } \\ \Delta \mathsf{r}_{{{\mathbf{\hat{\Theta }}}_{1} }} & = \int\limits_{{ - 1}}^{1} {(\Delta \mathsf{v}_{{\theta _{{\parallel \Theta }} }}^{\prime } {\mathbf{m}} + \mathsf{v}_{{\theta _{{\parallel \Theta }} }}^{\prime } \Delta {\mathbf{m}} - \Delta \mathsf{v}_{{\theta _{{\parallel \Theta }} }} \widetilde{{\mathbf{m}}}_{\rho } } \\ & \quad - \mathsf{v}_{{\theta _{{\parallel \Theta }} }} \Delta {\mathbf{m}}_{\rho } )Jd\xi - [\Delta \mathsf{v}_{{\theta _{{\parallel \Theta }} }} {\mathbf{m}}_{\sigma } ]_{{\varGamma_{\sigma } }} . \\ \end{aligned} $$
(209)

In order to identify the element stiffness matrix \({{\mathbf {\mathsf{{k}}}}}_{SK-TAN},\) (209) has to be brought in the form

$$\begin{aligned} \begin{array}{ll} \Delta {{\mathbf {\mathsf{{r}}}}}_{SK-TAN}=:{{\mathbf {\mathsf{{k}}}}}_{SK-TAN} \Delta \hat{{{\mathbf {\mathsf{{x}}}}}}_{TAN}. \end{array} \end{aligned}$$
(210)

The vector \(\Delta \hat{{{\mathbf {\mathsf{{x}}}}}}_{TAN}\) has already been defined in Sect. 9.1. The moment-related terms yield:

$$ \begin{aligned} \Delta \mathsf{v}_{{\theta _{ \bot } }}^{\prime } {\mathbf{m}} & = \mathsf{H}^{{\prime \prime T}} {\mathbf{S}}({\mathbf{m}})\Delta \widetilde{{\mathbf{t}}} + \mathsf{H}^{{\prime T}} {\mathbf{S}}({\mathbf{m}})\Delta \widetilde{{\mathbf{t}}}^{\prime } , \\ \Delta {\mathbf{m}} & = - {\mathbf{S}}({\mathbf{m}})\Delta \varvec{\theta } + {\mathbf{c}}_{m} \Delta \varvec{\theta }^{\prime } , \\ \Delta \mathsf{v}_{{\theta _{ \bot } }} \widetilde{{\mathbf{m}}}_{\rho } & = \mathsf{H}^{{\prime T}} {\mathbf{S}}(\widetilde{{\mathbf{m}}}_{\rho } )\Delta \widetilde{{\mathbf{t}}}, \\ \Delta \mathsf{v}_{{\theta _{ \bot } }} {\mathbf{m}}_{\sigma } & = \mathsf{H}^{{\prime T}} {\mathbf{S}}({\mathbf{m}}_{\sigma } )\Delta \widetilde{{\mathbf{t}}}, \\ \Delta \mathsf{v}_{{\theta _{{\parallel \Theta }} }} \widetilde{{\mathbf{m}}}_{\rho } & = (\mathsf{L}_{\parallel }^{T} \otimes \widetilde{{\mathbf{m}}}_{\rho }^{T} )\Delta {\mathbf{g}}_{1} , \\ \Delta \mathsf{v}_{{\theta _{{\parallel \Theta }} }} {\mathbf{m}}_{\sigma } & = (\mathsf{L}_{\parallel }^{T} \otimes {\mathbf{m}}_{\sigma }^{T} )\Delta {\mathbf{g}}_{1} , \\ \Delta \mathsf{v}_{{\theta _{{\parallel \Theta }} }}^{\prime } {\mathbf{m}} & = (\mathsf{L}_{\parallel }^{{\prime T}} \otimes {\mathbf{m}}^{T} )\Delta {\mathbf{g}}_{1} + (\mathsf{L}_{\parallel }^{T} \otimes {\mathbf{m}}^{T} )\Delta {\mathbf{g}}_{1}^{\prime } . \\ \end{aligned} $$
(211)

Here, many of the relations already derived in Sect. 6.2.4 could be re-used. The field of multiplicative rotation vector increments \(\Delta \varvec{\theta }\) follows directly from Eq. (122):

$$ \begin{aligned} \Delta \varvec{\theta } & = \mathsf{v}_{{\theta _{{\parallel \Theta }} }}^{T} \Delta \widehat{\varvec{\Theta }}_{1} + (\mathsf{v}_{{\theta _{ \bot } }}^{T} + \mathsf{v}_{{\theta _{{\parallel d}} }}^{T} )\Delta \widehat{\mathsf{d}}, \\ \mathsf{v}_{{\theta _{{\parallel \Theta }} }} & = \mathsf{L}_{\parallel }^{T} \otimes {\mathbf{g}}_{1}^{T} ,{\mkern 1mu} {\mkern 1mu} {\mkern 1mu} \mathsf{v}_{{\theta _{ \bot } }} = - \mathsf{H}^{{\prime T}} {\mathbf{S}}(\widetilde{{\mathbf{t}}}), \\ \mathsf{v}_{{\theta _{{\parallel d}} }} & = \left( {\sum\limits_{{i = 1}}^{{n_{\Lambda } }} {L^{i} } {\mkern 1mu} \mathsf{v}_{1} (\xi _{i} ) - \mathsf{v}_{1} (\xi )} \right) \otimes {\mathbf{g}}_{1}^{T} , \\ \mathsf{v}_{1} (\xi ) & = \frac{{\mathsf{H}^{{\prime T}} (\xi )({\mathbf{g}}_{1}^{I} \times \widetilde{{\mathbf{t}}}(\xi )) - \mathsf{H}^{{\prime T}} (\xi _{I} )({\mathbf{g}}_{1} (\xi ) \times \widetilde{{\mathbf{t}}}^{I} )}}{{1 + {\mathbf{g}}_{1}^{{{\kern 1pt} T}} (\xi ){\mathbf{g}}_{1}^{I} }}. \\ \end{aligned} $$
(212)

In a similar manner, the associated arc-length derivative \(\Delta \varvec{\theta }^{\prime }\) follows from Eq. (123):

$$ \begin{aligned} \Delta \varvec{\theta}^{\prime} & = \mathsf{v}_{{\theta _{{\parallel \Theta }} }}^{{\prime T}} \Delta \widehat{{\mathbf{\Theta }}}_{1} + (\mathsf{v}_{{\theta _{ \bot } }}^{{\prime T}} + \mathsf{v}_{{\theta _{{\parallel d}} }}^{{\prime T}} )\Delta \widehat{\mathsf{d}}, \\ \mathsf{v}_{{\theta _{{\parallel \Theta }} }}^{\prime } & = \mathsf{L}_{\parallel }^{{\prime T}} \otimes {\mathbf{g}}_{1}^{T} + \mathsf{L}_{\parallel }^{T} \otimes {\mathbf{g}}_{1}^{{\prime T}} , \\ \mathsf{v}_{{\theta _{ \bot } }}^{\prime } & = - \mathsf{H}^{{\prime \prime T}} {\mathbf{S}}(\widetilde{{\mathbf{t}}}) - \mathsf{H}^{{\prime T}} {\mathbf{S}}(\widetilde{{{\mathbf{t^{\prime}}}}}), \\ \mathsf{v}_{{\theta _{{\parallel d}} }}^{\prime } & = \left( {\sum\limits_{{i = 1}}^{{n_{\Lambda } }} {L^{{i\prime }} } \mathsf{v}_{1} (\xi _{i} ) - \mathsf{v}_{1}^{\prime } (\xi )} \right) \otimes {\mathbf{g}}_{1}^{T} \\ & \quad + \left( {\sum\limits_{{i = 1}}^{{n_{\Lambda } }} {L^{i} } \mathsf{v}_{1} (\xi _{i} ) - \mathsf{v}_{1} (\xi )} \right) \otimes {\mathbf{g}}_{1}^{{\prime T}} , \\ \mathsf{v}_{1}^{\prime } (\xi ) & = \frac{{\mathsf{H}^{{\prime T}} (\xi )\left( {{\mathbf{g}}_{1}^{I} \times \widetilde{{{\mathbf{t^{\prime}}}}}(\xi )} \right) + \mathsf{H}^{{\prime \prime T}} (\xi )\left( {{\mathbf{g}}_{1}^{I} \times \widetilde{{\mathbf{t}}}(\xi )} \right)}}{{1 + {\mathbf{g}}_{1}^{T} (\xi ){\mathbf{g}}_{1}^{I} }} \\ & \quad - \frac{{\mathsf{H}^{{\prime T}} (\xi _{I} )\left( {{\mathbf{g}}_{1}^{\prime } (\xi ) \times \widetilde{{\mathbf{t}}}^{I} } \right) + \left( {{\mathbf{g}}_{1}^{{\prime T}} (\xi ){\mathbf{g}}_{1}^{I} } \right){\mathbf{v}}_{1} (\xi )}}{{1 + {\mathbf{g}}_{1}^{T} (\xi ){\mathbf{g}}_{1}^{I} }}, \\ \end{aligned} $$
(213)

with \(\widetilde{\mathbf {t}}^{\prime } = \mathbf {r}^{\prime \prime }/||\mathbf {r}^{\prime }||^2 - 2(\mathbf {r}^{\prime T}\mathbf {r}^{\prime \prime })\mathbf {r}^{\prime }/||\mathbf {r}^{\prime }||^4\). The remaining linearizations in (211) have already been derived in (208). In contrast to the spin vector field \(\delta \varvec{\theta }\), the increment field \(\Delta \varvec{\theta }\) has to be expressed via additive increments \(\Delta \hat{\varphi }^i\). The required relation is given by (129) and repeated here:

$$ \begin{aligned} \Delta \widehat{\varvec{\Theta }}_{1} & = (\hat{\Theta }_{1}^{1} ,\hat{\Theta }_{1}^{2} ,\hat{\Theta }_{1}^{3} )^{T} , \\ \Delta \hat{\Theta }_{1}^{i} & = - \frac{{\overline{{\mathbf{g}}} _{1}^{{iT}} {\mathbf{S}}({\mathbf{g}}_{1}^{i} )}}{{1 + {\mathbf{g}}_{1}^{{iT}} \overline{{\mathbf{g}}} _{1}^{i} }}\frac{{\mathsf{H}^{\prime } (\xi ^{i} )\Delta \widehat{\mathsf{d}}}}{{||{\mathbf{t}}^{i} ||}} + \Delta \hat{\varphi }^{i} . \\ \end{aligned} $$
(214)

The linearization of the element residual terms associated with axial tension results in:

$$ \begin{aligned} \Delta \bar{F}_{1} & = EA\Delta {\bar{\epsilon}} = EA\overline{\mathsf{v}} _{\epsilon{}}^{T} \Delta \widehat{\mathsf{d}}, \\ \Delta \overline{\mathsf{v}} _{\epsilon{}} & = \sum\limits_{{i = 1}}^{3} {L^{i} } (\xi )\Delta \mathsf{v}_{\epsilon{}} (\xi ^{i} ), \\ \Delta \mathsf{v}_{\epsilon{}} & = \frac{{\mathsf{H}^{{\prime T}} }}{{||{\mathbf{r}}^{\prime } ||}}({\mathbf{I}}_{3} - {\mathbf{g}}_{1} \otimes {\mathbf{g}}_{1}^{T} )\mathsf{H}^{\prime } \Delta \widehat{\mathsf{d}}. \\ \end{aligned} $$
(215)

Based (93), the linearization of the inertia forces reads:

$$\begin{aligned} \begin{array}{ll} -{{\mathbf {\mathsf{{H}}}}}^T \Delta \mathbf {f}_{\rho } = \rho A c_{\ddot{\mathbf {r}}1} {{\mathbf {\mathsf{{H}}}}}^T {{\mathbf {\mathsf{{H}}}}} \Delta \hat{{{\mathbf {\mathsf{{d}}}}}}, \quad c_{\ddot{\mathbf {r}}1} = \frac{1 - \alpha _m}{(1 - \alpha _f)\beta \Delta t^2}. \end{array} \end{aligned}$$
(216)

The time integration factor \(c_{\ddot{\mathbf {r}}1}\) of the modified generalized-\(\alpha \) scheme according to Sect. 5 slightly differs from the corresponding factor of the standard generalized-\(\alpha \) scheme. The linearization of the inertia moments yields:

$$ \begin{aligned} - \Delta {\mathbf{m}}_{\rho } & = {\mathbf{S}}({\mathbf{m}}_{\rho } )\Delta \varvec{\theta } \\ & \quad + {\mathbf{\Lambda }}[c_{{\mathbf{W}}} \{ {\mathbf{S}}({\mathbf{W}}){\mathbf{C}}_{\rho } - {\mathbf{S}}({\mathbf{C}}_{\rho } {\mathbf{W}})\} + c_{{\mathbf{A}}} {\mathbf{C}}_{\rho } ]\Delta \widetilde{{\mathbf{\Theta }}}_{{n + 1}} , \\ \Delta \widetilde{{\mathbf{\Theta }}}_{{n + 1}} & = \Lambda _{n}^{T} \Delta \widetilde{\varvec{\theta }}_{{n + 1}} = \Lambda _{n}^{T} {\mathbf{T}}(\widetilde{\varvec{\theta }}_{{n + 1}} )\Delta \varvec{\theta }, \\ c_{{\mathbf{A}}} & = \frac{{1 - \alpha _{m} }}{{(1 - \alpha _{f} )\beta \Delta t^{2} }},\quad c_{{\mathbf{W}}} = \frac{\gamma }{{\beta \Delta t}}. \\ \end{aligned} $$
(217)

For clarity, the indices \(n + 1\) and n of the current and previous time step have explicitly been noted for some of the quantities occurring in (217). All the other quantities are evaluated at \(t_{n + 1}.\) As already introduced in Sect. 5, the fields \(\widetilde{\varvec{\Theta }}_{n+1}\) and \(\widetilde{\varvec{\theta }}_{n+1}\) are the material and spatial multiplicative rotation increments relating the current configuration and the converged configuration of the previous time step \(t_n.\) The two vectors are related by the transformation

$$ \begin{aligned} \widetilde{{\mathbf{\Theta }}}_{{n + 1}} & = {\mathbf{\Lambda }}_{{n + 1}}^{T} \widetilde{\varvec{\theta }}_{{n + 1}} {\mathbf{ = \Lambda }}_{n}^{T} \widetilde{\varvec{\theta }}_{{n + 1}} \\ \to \Delta \widetilde{{\mathbf{\Theta }}}_{{n + 1}} & = {\mathbf{\Lambda }}_{n}^{T} \Delta \widetilde{\varvec{\theta }}_{{n + 1}} . \\ \end{aligned} $$
(218)

The second step in (218) is valid since \(\widetilde{\varvec{\theta }}_{n+1}\) is an eigenvector with eigenvalue one of the rotation tensor \(\varvec{\Lambda }_{n+1}\varvec{\Lambda }^T_{n}\) between the configurations n and \(n + 1,\) thus \(\varvec{\Lambda }_{n+1}\varvec{\Lambda }^T_{n}\widetilde{\varvec{\theta }}_{n+1} = \widetilde{\varvec{\theta }}_{n+1}.\) Furthermore, \(\Delta \widetilde{\varvec{\Theta }}_{n+1}\) and \(\Delta \widetilde{\varvec{\theta }}_{n+1}\) represent the fields of additive increments of \(\widetilde{\varvec{\Theta }}_{n+1}\) and \(\widetilde{\varvec{\theta }}_{n+1}\) between two successive Newton iterations, whereas \(\Delta \varvec{\theta }\) as given by (212) represents the field of multiplicative rotation increments between two successive Newton iterations.

Appendix D: Linearization of WK-TAN Element

The residual vector of the WK-TAN element is given in Eq. (184). The linearization of (184) has the general form:

$$ \begin{aligned} \Delta \mathsf{r}_{{\widehat{{\mathbf{d}}}}} & = \int\limits_{{ - 1}}^{1} {\left( {\Delta \overline{\mathsf{v}} _{{\theta _{ \bot } }}^{\prime } {\mathbf{m}} + \overline{\mathsf{v}} _{{\theta _{ \bot } }}^{\prime } \Delta {\mathbf{m}} + \Delta \overline{\mathsf{v}} _{\epsilon{}} \bar{F}_{1} + \overline{\mathsf{v}} _{\epsilon{}} \Delta \bar{F}_{1} } \right)Jd\xi } \\ & \quad - \int\limits_{{ - 1}}^{1} {\left( {\mathsf{H}^{T} \Delta {\mathbf{f}}_{\rho } + \Delta \overline{\mathsf{v}} _{{\theta _{ \bot } }} \widetilde{{\mathbf{m}}}_{\rho } + \overline{\mathsf{v}} _{{\theta _{ \bot } }} \Delta {\mathbf{m}}_{\rho } } \right)Jd\xi } - [\Delta \overline{\mathsf{v}} _{{\theta _{ \bot } }} {\mathbf{m}}_{\sigma } ]_{{\varGamma_{\sigma } }} \\ \Delta \mathsf{r}_{{{\mathbf{\hat{\Theta }}}_{1} }} & = \int\limits_{{ - 1}}^{1} {(\Delta \overline{\mathsf{v}} _{{\theta _{{\parallel \Theta }} }}^{\prime } {\mathbf{m}} + \overline{\mathsf{v}} _{{\theta _{{\parallel \Theta }} }}^{\prime } \Delta {\mathbf{m}} - \Delta \overline{\mathsf{v}} _{{\theta _{{\parallel \Theta }} }} \widetilde{{\mathbf{m}}}_{\rho } } \\ & \quad - \overline{\mathsf{v}} _{{\theta _{{\parallel \Theta }} }} \Delta {\mathbf{m}}_{\rho } )Jd\xi - [\Delta \overline{\mathsf{v}} _{{\theta _{{\parallel \Theta }} }} {\mathbf{m}}_{\sigma } ]_{{\varGamma_{\sigma } }} . \\ \end{aligned} $$
(219)

In order to identify the element stiffness matrix \({{\mathbf {\mathsf{{k}}}}}_{WK-TAN},\) (219) has to be brought in the form

$$\begin{aligned} \begin{array}{ll} \Delta {{\mathbf {\mathsf{{r}}}}}_{WK-TAN}=:{{\mathbf {\mathsf{{k}}}}}_{WK-TAN} \Delta \hat{{{\mathbf {\mathsf{{x}}}}}}_{TAN}. \end{array} \end{aligned}$$
(220)

The linearization of vectors of the form \(\bar{{{\mathbf {\mathsf{{v}}}}}}_{...}\) and \(\bar{{{\mathbf {\mathsf{{v}}}}}}_{...}^{\prime }\) as originally defined in (184) yields:

$$ \begin{aligned} \Delta \overline{\mathsf{v}} _{{\theta _{ \bot } }} & = - \sum\limits_{{i = 1}}^{3} {L^{i} } (\xi )\Delta \mathsf{v}_{{\theta _{ \bot } }} (\xi ^{i} ), \\ \Delta \overline{\mathsf{v}} _{\epsilon} & = \sum\limits_{{i = 1}}^{3} {L^{i} } (\xi )\Delta \mathsf{v}_{\epsilon} (\xi ^{i} ), \\ \Delta \overline{\mathsf{v}} _{{\theta _{{\parallel \Theta }} }} & = \sum\limits_{{i = 1}}^{3} {L^{i} } (\xi )\Delta \mathsf{v}_{{\theta _{{\parallel \Theta }} }} (\xi ^{i} ), \\ \end{aligned} $$
(221)

with \(\Delta \bar{{{\mathbf {\mathsf{{v}}}}}}^{\prime }_{...} = \sum _{i=1}^{3} L^i{,\xi }(\xi )/J(\xi ) \Delta {{\mathbf {\mathsf{{v}}}}}_{...} (\xi ^i)\). The linearization of the vectors \({{\mathbf {\mathsf{{v}}}}}_{...}\) and \({{\mathbf {\mathsf{{v}}}}}_{...}^{\prime }\) has already been stated in the last section. Also the linearization of the moment stress resultant has the same form as in the last section:

$$\begin{aligned} \begin{array}{ll} \Delta \mathbf {m}&= -\mathbf {S}(\mathbf {m}) \Delta \varvec{\theta } + \mathbf {c}_m \Delta \varvec{\theta }^{\prime }. \end{array} \end{aligned}$$
(222)

However, the fields \(\Delta \varvec{\theta }\) and \(\Delta \varvec{\theta }^{\prime }\) originally defined in Sect. 6.2.3 are this time given by

$$\begin{aligned} \begin{array}{ll} \Delta \varvec{\theta } = \sum\limits _{i=1}^{3} \widetilde{\mathbf {I}}^i(\xi ) \Delta {\varvec{\theta }}(\xi ^i), \,\,\,\, \Delta \varvec{\theta }^{\prime } = \sum\limits _{i=1}^{3} \frac{1}{J(\xi )}\widetilde{\mathbf {I}}_{,\xi }^i(\xi ) \Delta {\varvec{\theta }}(\xi ^i). \end{array} \end{aligned}$$
(223)

Due to the Kirchhoff constraint, the nodal increments \(\Delta {\varvec{\theta }}(\xi ^i)\) can be expressed according to:

$$ \begin{aligned} \Delta \varvec{\theta }(\xi ^{i} ) & = \Delta \hat{\Theta }_{1}^{i} {\mathbf{g}}_{1} (\xi ^{i} ) + \mathsf{v}_{{\theta _{ \bot } }}^{T} (\xi ^{i} )\Delta \widehat{\mathsf{d}}, \\ \Delta \hat{\Theta }_{1}^{i} & = - \frac{{\overline{{\mathbf{g}}} _{1}^{{iT}} {\mathbf{S}}({\mathbf{g}}_{1}^{i} )}}{{1 + {\mathbf{g}}_{1}^{{iT}} \overline{{\mathbf{g}}} _{1}^{i} }}\frac{{\mathsf{H}^{\prime } (\xi ^{i} )\Delta \widehat{\mathsf{d}}}}{{||{\mathbf{t}}^{i} ||}} + \Delta \hat{\varphi }^{i} . \\ \end{aligned} $$
(224)

The linearization of the inertia forces and moments is identical to the corresponding results of the last section given in (216) and (217). However, for the WK-TAN element, the rotation increment field \(\Delta \varvec{\theta }\) is given by Eq. (223).

Appendix E: Linearization of SK/WK-ROT Elements

The nodal primary variable variations of the SK/WK-TAN and the SK/WK-ROT elements read:

$$ \begin{aligned} \delta \widehat{\mathsf{x}}_{{TAN}} : & = (\delta \widehat{{\mathbf{d}}}^{{1T}} ,\delta \widehat{{\mathbf{t}}}^{{1T}} ,\delta \hat{\Theta }_{1}^{1} ,\delta \widehat{{\mathbf{d}}}^{{2T}} ,\delta \widehat{{\mathbf{t}}}^{{2T}} ,\delta \hat{\Theta }_{1}^{2} ,\delta \hat{\Theta }_{1}^{3} )^{T} , \\ \delta \widehat{\mathsf{x}}_{{ROT}} : & = (\delta \widehat{{\mathbf{d}}}^{{1T}} ,\delta \widehat{\varvec{\theta }}^{{1T}} ,\delta \hat{t}^{1} ,\delta \widehat{{\mathbf{d}}}^{{2T}} ,\delta \widehat{\varvec{\theta }}^{{2T}} ,\delta \hat{t}_{1} ,\delta \hat{\Theta }_{1}^{3} )^{T} . \\ \end{aligned} $$
(225)

In a similar manner, the set of iterative nodal primary variable increments have been defined as:

$$ \begin{aligned} \Delta \widehat{\mathsf{x}}_{{TAN}}: & = (\Delta \widehat{{\mathbf{d}}}^{{1T}} ,\Delta \widehat{{\mathbf{t}}}^{{1T}} ,\Delta \hat{\varphi }^{1} ,\Delta \widehat{{\mathbf{d}}}^{{2T}} ,\Delta \widehat{{\mathbf{t}}}^{{2T}} ,\Delta \hat{\varphi }^{2} ,\Delta \hat{\varphi }^{3} )^{T} , \\ \Delta \widehat{\mathsf{x}}_{{ROT}}: & = (\Delta \widehat{{\mathbf{d}}}^{{1T}} ,\Delta \widehat{\varvec{\theta }}^{{1T}} ,\Delta \hat{t}^{1} ,\Delta \widehat{{\mathbf{d}}}^{{2T}} ,\Delta \widehat{\varvec{\theta }}^{{2T}} ,\Delta \hat{t}^{2} ,\Delta \hat{\varphi }^{3} )^{T} . \\ \end{aligned} $$
(226)

The transformations between these primary variable variations and increments is given by:

$$\begin{aligned} \begin{array}{ll} \delta \hat{{{\mathbf {\mathsf{{x}}}}}}_{TAN} = \widetilde{{{\mathbf {\mathsf{{T}}}}}}_{\hat{{{\mathbf {\mathsf{{x}}}}}}} \delta \hat{\mathbf {x}}_{ROT} \quad \text {and} \quad \Delta \hat{{{\mathbf {\mathsf{{x}}}}}}_{TAN} = {{\mathbf {\mathsf{{T}}}}}_{M\hat{{{\mathbf {\mathsf{{x}}}}}}} \Delta \hat{\mathbf {x}}_{ROT}. \end{array} \end{aligned}$$
(227)

The transformation matrices \(\widetilde{{{\mathbf {\mathsf{{T}}}}}}_{\hat{{{\mathbf {\mathsf{{x}}}}}}}\), originally defined in (169), and \({{\mathbf {\mathsf{{T}}}}}_{M\hat{{{\mathbf {\mathsf{{x}}}}}}}\) have the following form:

$$\begin{aligned} \widetilde{{{\mathbf {\mathsf{{T}}}}}}_{\hat{\mathbf {x}}} = \left( \begin{array}{ccccc} \mathbf {I}_3 &{}&{}&{}&{}\\ &{}\widetilde{\mathbf {T}}^1&{}&{}&{}\\ &{}&{}\mathbf {I}_3&{}&{}\\ &{}&{}&{}\widetilde{\mathbf {T}}^2&{}\\ &{}&{}&{}&{}1 \end{array} \right) , \,\, {{\mathbf {\mathsf{{T}}}}}_{M\hat{\mathbf {x}}} = \left( \begin{array}{ccccc} \mathbf {I}_3 &{}&{}&{}&{}\\ &{}\mathbf {T}_M^1&{}&{}&{}\\ &{}&{}\mathbf {I}_3&{}&{}\\ &{}&{}&{}\mathbf {T}_M^2&{}\\ &{}&{}&{}&{}1 \end{array} \right) . \end{aligned}$$
(228)

These two different matrices are required, since the primary variable variations of the SK/WK-TAN elements are based on the multiplicative quantities \(\delta \hat{\Theta }_{1}^i\), whereas the corresponding iterative primary variable increments are based on the additive quantities \(\Delta \hat{\varphi }^i\). The matrices \(\widetilde{\mathbf {T}}^i\) and \(\mathbf {T}_M^i\) (25, 24) are evaluated at the element boundary nodes:

$$\begin{aligned} \widetilde{\mathbf {T}}^i := \widetilde{\mathbf {T}}(\xi ^i) \quad \text {and} \quad \mathbf {T}_M^i := \mathbf {T}_M(\xi ^i) \quad \text {for} \quad i=1,2. \end{aligned}$$
(229)

In Sect. 9.2, it has already been shown that the following residual transformation is valid:

$$\begin{aligned} {{\mathbf {\mathsf{{r}}}}}_{ROT} = \widetilde{{{\mathbf {\mathsf{{T}}}}}}_{\hat{\mathbf {x}}}^T {{\mathbf {\mathsf{{r}}}}}_{TAN}. \end{aligned}$$
(230)

In a similar manner, also the linearized element residual vector can be transformed:

$$\begin{aligned} \begin{array}{ll} \Delta {{\mathbf {\mathsf{{r}}}}}_{ROT} =&{} \Delta \widetilde{{{\mathbf {\mathsf{{T}}}}}}_{\hat{\mathbf {x}}}^T {{\mathbf {\mathsf{{r}}}}}_{TAN} + \widetilde{{{\mathbf {\mathsf{{T}}}}}}_{\hat{\mathbf {x}}}^T \underbrace{\Delta {{\mathbf {\mathsf{{r}}}}}_{TAN}}_{={{\mathbf {\mathsf{{k}}}}}_{TAN}\Delta \hat{{{\mathbf {\mathsf{{x}}}}}}_{TAN}} =: \ {{\mathbf {\mathsf{{k}}}}}_{ROT} \Delta \hat{{{\mathbf {\mathsf{{x}}}}}}_{ROT}, \\ {{\mathbf {\mathsf{{k}}}}}_{ROT} :=&{} \left( \widetilde{{{\mathbf {\mathsf{{H}}}}}}_{\hat{\mathbf {x}}} ({{\mathbf {\mathsf{{r}}}}}_{TAN}) + \widetilde{{{\mathbf {\mathsf{{T}}}}}}_{\hat{\mathbf {x}}}^T {{\mathbf {\mathsf{{k}}}}}_{TAN} {{\mathbf {\mathsf{{T}}}}}_{M\hat{\mathbf {x}}}\right) . \end{array} \end{aligned}$$
(231)

Here, the matrix \( \widetilde{{{\mathbf {\mathsf{{H}}}}}}_{\hat{\mathbf {x}}} ({{\mathbf {\mathsf{{r}}}}}_{TAN})\) has been introduced, given by

$$\begin{aligned} \widetilde{{{\mathbf {\mathsf{{H}}}}}}_{\hat{\mathbf {x}}} ({{\mathbf {\mathsf{{r}}}}}_{TAN}) = \left( \begin{array}{ccccc} \mathbf {0} &{}&{}&{}&{}\\ &{}\widetilde{{{\mathbf {\mathsf{{H}}}}}}^1&{}&{}&{}\\ &{}&{}\mathbf {0}&{}&{}\\ &{}&{}&{}\widetilde{{{\mathbf {\mathsf{{H}}}}}}^2&{}\\ &{}&{}&{}&{}0 \end{array} \right) . \end{aligned}$$
(232)

This matrix is used for representing the linearization of \(\widetilde{{{\mathbf {\mathsf{{T}}}}}}_{\hat{\mathbf {x}}}\):

$$\begin{aligned} \widetilde{{{\mathbf {\mathsf{{H}}}}}}_{\hat{\mathbf {x}}} ({{\mathbf {\mathsf{{r}}}}}_{TAN}) \Delta \hat{{{\mathbf {\mathsf{{x}}}}}}_{ROT} := \Delta \widetilde{{{\mathbf {\mathsf{{T}}}}}}_{\hat{\mathbf {x}}}^T {{\mathbf {\mathsf{{r}}}}}_{TAN}. \end{aligned}$$
(233)

After calculating the derivative of \(\widetilde{{{\mathbf {\mathsf{{T}}}}}}_{\hat{\mathbf {x}}}\) and re-ordering the result, the submatrices \(\widetilde{{{\mathbf {\mathsf{{H}}}}}}^i\) can be stated:

$$\begin{aligned} \widetilde{{{\mathbf {\mathsf{{H}}}}}}^i = \left( \begin{array}{cc} \mathbf {S}(\mathbf {r}_{ T A N ,\hat{\mathbf {t}}^{i}} ) \mathbf {S}(\mathbf {g}_1^{i}) - r_{ T A N ,\hat{\Theta }_1^{i}} \mathbf {S}(\mathbf {g}_1^{i}) \,\,\,&{}\,\,\, \mathbf {S}(\mathbf {g}_1^{i})\mathbf {r}_{ T A N ,\hat{\mathbf {t}}^{i}} \\ -\mathbf {r}_{ T A N ,\hat{\mathbf {t}}^{i}}^T\mathbf {S}(\mathbf {g}_1^{i}) &{} 0 \end{array} \right) \end{aligned}$$
(234)

for \(i=1,2.\) From (231), the following transformation rule for the the element stiffness matrix can be stated:

$$ \begin{aligned} \Delta {\mathbf{r}}_{{ROT}} & = {\mathbf{k}}_{{ROT}} \Delta \widehat{{\mathbf{x}}}_{{ROT}} \\ {\text{with }}{\mathbf{k}}_{{ROT}} & = {\mathbf{\widetilde{H}}}_{{\widehat{{\mathbf{x}}}}} ({\mathbf{r}}_{{TAN}} ) + {\mathbf{\widetilde{T}}}_{{\widehat{{\mathbf{x}}}}}^{T} {\mathbf{k}}_{{TAN}} {\mathbf{T}}_{{M\widehat{{\mathbf{x}}}}} . \\ \end{aligned} $$
(235)

In order to apply this transformation, the components of the element stiffness matrices \({{\mathbf {\mathsf{{k}}}}}_{TAN}\) and \({{\mathbf {\mathsf{{k}}}}}_{ROT}\) have to be arranged in the same order as the element residual vectors:

$$ \begin{aligned} \mathsf{r}_{{TAN}}: & = ({\mathbf{r}}_{{TAN,\widehat{{\mathbf{d}}}^{1} }}^{T} ,{\mathbf{r}}_{{TAN,\widehat{{\mathbf{t}}}^{1} }}^{T} ,r_{{TAN,\hat{\Theta }_{1}^{1} }} , \\ & \quad {\mathbf{r}}_{{TAN,\widehat{{\mathbf{d}}}^{2} }}^{T} ,{\mathbf{r}}_{{TAN,\widehat{{\mathbf{t}}}^{2} }}^{T} ,r_{{TAN,\hat{\Theta }_{1}^{2} }} ,r_{{TAN,\hat{\Theta }_{1}^{3} }} )^{T} , \\ \mathsf{r}_{{ROT}}: & = (\mathsf{r}_{{ROT,\widehat{{\mathbf{d}}}^{1} }}^{T} ,\mathsf{r}_{{ROT,\widehat{\theta }^{1} }}^{T} ,r_{{ROT,\hat{t}^{1} }} , \\ & \quad \mathsf{r}_{{ROT,\widehat{{\mathbf{d}}}^{2} }}^{T} ,\mathsf{r}_{{ROT,\widehat{{\theta ^{2} }}}}^{T} ,r_{{ROT,\hat{t}^{2} }} ,r_{{ROT,\hat{\Theta }_{1}^{3} }} )^{T} . \\ \end{aligned} $$
(236)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Meier, C., Popp, A. & Wall, W.A. Geometrically Exact Finite Element Formulations for Slender Beams: Kirchhoff–Love Theory Versus Simo–Reissner Theory. Arch Computat Methods Eng 26, 163–243 (2019). https://doi.org/10.1007/s11831-017-9232-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11831-017-9232-5

Navigation