Skip to main content
Log in

Design of a large stroke compliant gripping mechanism for constant-force applications

  • Technical Paper
  • Published:
Journal of the Brazilian Society of Mechanical Sciences and Engineering Aims and scope Submit manuscript

Abstract

A constant force compliant mechanism (CFCM) prevents a mechanical gripper from exerting excessive contact force through passive force control. In response to changes in mechanical stiffness induced by structural deformation, it can generate nearly constant extrinsic loads throughout the effective range of input displacement. This research aims to design an embedded CFCM end effector for manipulators capable of grasping geometrically and mechanically unknown objects without the need of sensory equipment. The proposed topology optimization method synthesizes CFCMs through discrete parameterization, loop detection, deformation analysis, objective domain enumeration, and function error evaluation, utilizing a genetic algorithm. The algorithm minimizes errors and improves solution accuracy by deducing solution feasibility through several penalty functions. Finite element analysis verifies the performance, deformation, and stress conditions of the optimal CFCM. The study includes manufacturing a CFCM prototype and assembling for end-effector, followed by experiments to validate analytical results and assessing actual grasping ability. The proposed design provides an extended range of motion and improved force stability, laying a groundwork for future research in this area.

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

Similar content being viewed by others

References

  1. Lins RG, Givigi SN (2021) Cooperative robotics and machine learning for smart manufacturing: platform design and trends within the context of industrial internet of things. IEEE Access 9:95444–95455. https://doi.org/10.1109/ACCESS.2021.3094374

    Article  Google Scholar 

  2. Kiguchi K, Fukuda T (2000) Position/force control of robot manipulators for geometrically unknown objects using fuzzy neural networks. IEEE Trans Ind Electron 47(3):641–649. https://doi.org/10.1109/41.847905

    Article  Google Scholar 

  3. Kumar N, Panwar V, Sukavanam N, Sharma SP, Borm JH (2011) Neural network based hybrid force/position control for robot manipulators. Int J Precis Eng Manuf 12(3):419–426. https://doi.org/10.1007/s12541-011-0054-3

    Article  Google Scholar 

  4. Chen YH, Lan CC (2012) An adjustable constant-force mechanism for adaptive end-effector operations. ASME J Mech Des 134(3):031005. https://doi.org/10.1115/1.4005865

    Article  Google Scholar 

  5. Liu CH, Hsu MC, Chen TL, Chen Y (2020) Optimal design of a compliant constant-force mechanism to deliver a nearly constant output force over a range of input displacements. Soft Robot 7(6):758–769. https://doi.org/10.1089/soro.2019.0122

    Article  Google Scholar 

  6. Jutte CV, Kota S (2008) Design of nonlinear springs for prescribed load-displacement functions. ASME J Mech Des 130(8):081403. https://doi.org/10.1115/1.2936928

    Article  Google Scholar 

  7. Wang Y (1997) Dynamics of an elastic four bar linkage mechanism with geometric nonlinearities. Nonlinear Dyn 14:357–375. https://doi.org/10.1023/A:1008269731024

    Article  Google Scholar 

  8. Pham H, Wang D (2011) A constant-force bistable mechanism for force regulation and overload protection. Mech Mach Theory 46(7):899–909. https://doi.org/10.1016/j.mechmachtheory.2011.02.008

    Article  Google Scholar 

  9. Wei Y, Xu Q (2022) Design of a new passive end-effector based on constant-force mechanism for robotic polishing. Robot Comput Integr Manuf. https://doi.org/10.1016/j.rcim.2021.102278

    Article  Google Scholar 

  10. Tolman KA, Merriam EG, Howell LL (2016) Compliant constant-force linear-motion mechanism. Mech Mach Theory 106:68–79. https://doi.org/10.1016/j.mechmachtheory.2016.08.009

    Article  Google Scholar 

  11. Boyle C, Howell LL, Magleby SP, Evans MS (2003) Dynamic modeling of compliant constant-force compression mechanisms. Mech Mach Theory 38(12):1469–1487. https://doi.org/10.1016/S0094-114X(03)00098-3

    Article  Google Scholar 

  12. Zhang X, Xu Q (2019) Design and testing of a novel 2-DOF compound constant-force parallel gripper. Precis Eng 56:53–61. https://doi.org/10.1016/j.precisioneng.2018.09.004

    Article  Google Scholar 

  13. Pedersen CBW, Fleck NA, Ananthasuresh GK (2006) Design of a compliant mechanism to modify an actuator characteristic to deliver a constant output force. ASME J Mech Des 128(5):1101–1112. https://doi.org/10.1115/1.2218883

    Article  Google Scholar 

  14. Liu CH, Chung FM, Ho YP (2021) Topology optimization for design of a 3D-printed constant-force compliant finger. IEEE/ASME Trans Mechatron 26(4):1828–1836. https://doi.org/10.1109/TMECH.2021.3077947

    Article  Google Scholar 

  15. Howell LL, Midha A (1994) A method for the design of compliant mechanisms with small-length flexural pivots. ASME J Mech Des 116(1):280–290. https://doi.org/10.1115/1.2919359

    Article  Google Scholar 

  16. Howell LL, Midha A (1995) Parametric deflection approximations for end-loaded, large-deflection beams in compliant mechanisms. ASME J Mech Des 117(1):156–165. https://doi.org/10.1115/1.2826101

    Article  Google Scholar 

  17. Saxena A, Kramer SN (1999) Erratum: a simple and accurate method for determining large deflections in compliant mechanisms subjected to end forces and moments. ASME J Mech Des 121(2):1. https://doi.org/10.1115/1.2829442

    Article  Google Scholar 

  18. Su HJ (2009) A pseudorigid-body 3R model for determining large deflection of cantilever beams subject to tip loads. ASME J Mech Robot 1(2):021008. https://doi.org/10.1115/1.3046148

    Article  MathSciNet  Google Scholar 

  19. Leishman LC, Colton MB (2011) A pseudo-rigid-body model approach for the design of compliant mechanism springs for prescribed force-deflections. In: Proceedings of ASME international design engineering technical conferences and computers and information in engineering conference, Washington, DC, USA. August 28–31. pp 93–102. https://doi.org/10.1115/DETC2011-47590

  20. Bilancia P, Berselli G (2020) Design and testing of a monolithic compliant constant force mechanism. Smart Mater Struct. https://doi.org/10.1088/1361-665X/ab6884

    Article  Google Scholar 

  21. Awtar S, Quint J, Ustick J (2021) Experimental characterization of a large-range parallel kinematic XYZ flexure mechanism. ASME J Mech Robot 10(1115/1):4048259

    Google Scholar 

  22. Ling M, Zhang X (2023) Nonlinear analysis of moderately large-stroke flexure beams in compliant mechanisms using a dynamic beam constraint model (DBCM). In: Liu X (eds) Advances in mechanism, machine science and engineering in China. CCMMS 2022. Lecture notes in mechanical engineering. Springer, Singapore. https://doi.org/10.1007/978-981-19-9398-5_69

  23. Milojević A, Linß S, Ćojbašić Ž, Handroos H (2021) A novel simple, adaptive, and versatile soft-robotic compliant two-finger gripper with an inherently gentle touch. ASME J Mech Robot 13(1):011015. https://doi.org/10.1115/1.4048752

    Article  Google Scholar 

  24. Luo Z, Chen L, Yang J, Zhang Y, Abdel-Malek K (2005) Compliant mechanism design using multi-objective topology optimization scheme of continuum structures. Struct Multidiscip Optim 30(2):142–154. https://doi.org/10.1007/s00158-004-0512-y

    Article  Google Scholar 

  25. Da Silva GA, Beck AT, Sigmund O (2020) Topology optimization of compliant mechanisms considering stress constraints, manufacturing uncertainty and geometric nonlinearity. Comput Methods Appl Mech Eng 365:112972. https://doi.org/10.1016/j.cma.2020.112972

    Article  MathSciNet  Google Scholar 

  26. Lu KJ, Kota S (2006) Topology and dimensional synthesis of compliant mechanisms using discrete optimization. ASME J Mech Des 128(5):1080–1091. https://doi.org/10.1115/1.2216729

    Article  Google Scholar 

  27. Lee J, Detroux T, Kerschen G (2020) Enforcing a force-displacement curve of a nonlinear structure using topology optimization with slope constraints. Appl Sci 10(8):2676. https://doi.org/10.3390/app10082676

    Article  Google Scholar 

  28. Pedersen CB, Buhl T, Sigmund O (2001) Topology synthesis of large-displacement compliant mechanisms. Int J Numer Methods Eng 50(12):2683–2705. https://doi.org/10.1002/nme.148

    Article  Google Scholar 

  29. Miao YB, Zheng J (2020) Optimization design of compliant constant-force mechanism for apple picking actuator. Comput Electron Agric 170:105232. https://doi.org/10.1016/j.compag.2020.105232

    Article  Google Scholar 

  30. Tessema B, Yen GG (2006) A self adaptive penalty function based algorithm for constrained optimization. In: 2006 IEEE international conference on evolutionary computation, pp 246–253

  31. Naves M, Brouwer DM, Aarts RG (2017) Building block-based spatial topology synthesis method for large-stroke flexure hinges. ASME J Mech Robot 9(4):041006. https://doi.org/10.1115/1.4036223

    Article  Google Scholar 

  32. Salajegheh E, Gholizadeh S (2005) Optimum design of structures by an improved genetic algorithm using neural networks. Adv Eng Softw 36(11–12):757–767. https://doi.org/10.1016/j.advengsoft.2005.03.022

    Article  Google Scholar 

  33. Ultimaker Support (2020) Ultimaker Nylon TDS. https://support.ultimaker.com/hc/en-us/articles/360011962600-Ultimaker-Nylon-TDS

  34. Kumar GS, Kalra PK, Dhande SG (2003) Parameter optimization for B-spline curve fitting using genetic algorithms. In: The 2003 congress on evolutionary computation, 2003. CEC'03, vol 3. pp 1871–1878

  35. Lo CW, Chang Y, Wang ML, Lin CR, Lee JJ (2021) Design of a monolithic constant-force compliant mechanism for extended range of motion and minimal force variation. In: Proceedings of ASME international design engineering technical conferences and computers and information in engineering conference, Virtual Online. August 17–19. V08AT08A015. https://doi.org/10.1115/DETC2021-69726

  36. Ting EC, Shih C, Wang YK (2004) Fundamentals of a vector form intrinsic finite element: part I. Basic procedure and a plane frame element. J Mech 20(2):113–122. https://doi.org/10.1017/S1727719100003336

    Article  Google Scholar 

  37. Ting EC, Shi C, Wang YK (2004) Fundamentals of a vector form intrinsic finite element. Part II. Plane solid element. J Mech 20(2):123–132

    Article  Google Scholar 

  38. Day AS (1965) An introduction to dynamic relaxation (dynamic relaxation method for structural analysis, using computer to calculate internal forces following development from initially unloaded state). The Engineer 219:218–221

    Google Scholar 

  39. Wang ML, Chuang CC, Lee JJ (2021) Development of an adaptive explicit algorithm for static simulation using the vector form intrinsic finite element method. J Mech 37:566–583. https://doi.org/10.1093/jom/ufab022

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jyh-Jone Lee.

Ethics declarations

Conflict of interest

C.W. Lo, Y. Chang, M.L. Wang, and J.J. Lee declare that they have no competing interests discussed in this study.

Additional information

Technical Editor: Pedro Manuel Calas Lopes Pacheco.

Publisher's Note

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

Appendices

Appendix 1: Mathematical framework for VFIFE

In this study, we employed the vector form intrinsic finite element (VFIFE) method [36, 37] and incorporated it into the topology optimization process.

The VFIFE framework describes the motion of a continuous body by using a finite number of points (or particles) associated with the theory of vector analysis. As shown in Fig. 

Fig. 20
figure 20

Description of motion and the path of a particle

20, a set of points representing the configuration of a frame structure is displaced from one position to another, where m denotes the mass of an arbitrary point in the structure and x denotes the position of the point of interest. Assuming that the interaction among points is modeled by a beam element, the properties of each element can be characterized by nodal displacements and corresponding nodal forces.

1.1 Discrete path and governing equation

Assume the particle in the structure undergoes a motion from time ta to t whose path is called the path element and within which the deformation of the structure is infinitesimal. Then, applying infinitesimal strain and engineering stress to evaluate continuum stress and compute virtual work is feasible. The subsequent formulation involves displacement and deformation processes, calculating equivalent internal forces and external forces, and solving the equation of motion for each particle.

The VFIFE method begins by modeling the individual movement of every particle following Newton’s 2nd Law of Motion as its governing equation:

$${m\varvec{\ddot{d}}} = {\varvec{F}}_\text{ext} { - }\sum\limits_{i = 1}^{{n_{e} }} {{\varvec{f}}_{{\text{int}}}^{i} }$$
(A1)

where d is the displacement vector, including translational and rotational displacements of the particle, Fext is the external forces including the concentrated external force vector applied at the particle and the equivalent nodal force vector from the load on the element,\({\varvec{f}}_{{\text{int}}}^{i}\) is the equivalent nodal forces due to deformation in elements, ne is the number of elements incident with the particle of interest, and m is the nodal mass whose elements contain the mass and mass moment of inertia of the particle.

1.2 Separation of rigid body motion from total displacement

To obtain the deformation in an element, the displacement of rigid body motion should be first separated from the overall displacement. Figure 

Fig. 21
figure 21

Reverse rotation of a beam element

21 shows the positions of an element containing two nodes (j, k) at two consecutive time instances. If the positions of node j at two-time instances are fictitiously aligned, the angular displacement of the rigid body motion can be found as

$$\varphi = \sin^{ - 1} \left| {{\hat{\mathbf{e}}}_{x} \times {\mathbf{\hat{e}^{\prime}}}_{x} } \right|$$
(A2)

where \({\hat{\mathbf{e}}}_{x}\) is the x-axis of local coordinate system established on the element at time ta and \({\mathbf{\hat{e}^{\prime}}}_{x}\) is the x-axis of local coordinate system established at time tb. Meanwhile, the change of the element displacement follows

$${\mathbf{u}}_{{\mathbf{k}}} {\mathbf{ = u}}^{{\mathbf{r}}}_{{\mathbf{k}}} { + }{\mathbf{u}}^{{\mathbf{d}}}_{{\mathbf{k}}}$$
(A3)

where \({\mathbf{u}}_{k}^{d}\) and \({\mathbf{u}}_{k}^{r}\) are respectively the changes of displacement in the axial and lateral directions. Also, the angular deformation of the nodes (θ j, θk) can be written as

$$\theta^{j} = \beta^{j} - \varphi {\text{ and }}\theta^{k} = \beta^{k} - \varphi$$
(A4)

1.3 Internal force evaluation

The incremental internal forces in the element during two-time instances in a motion are raised by the pure deformation in the element. For detailed derivations of the formula, please refer to [34]. The resulting equation can be expressed as:

$$\Delta {\varvec{\hat{f}}}{ = }\frac{{{E}}}{{{{l}}_{{\text{a}}} }}\left[ {\begin{array}{*{20}c} {{A}} & 0 & 0 \\ 0 & {4{{I}}} & {2{{I}}} \\ 0 & {2{{I}}} & {4{{I}}} \\ \end{array} } \right]\left[ {\begin{array}{*{20}c} \Delta \\ {{\theta }^{{{j}}} } \\ {{\theta }^{{{k}}} } \\ \end{array} } \right]$$
(A5)

where \(\Delta \hat{\varvec{f}} = [\Delta \hat{{f}}_{{{j,x}}} {,}\;\Delta {m}_{{{j,z}}} {,}\;\Delta {m}_{{{k,z}}} ]^{{\text{T}}}\), Δ is the magnitude of \({\mathbf{u}}_{k}^{d}\), and E, I, A, and la are the Young’s modulus, second moment of area, area of cross section, and element length of the object, respectively. In addition, three unknown force components in each node can be obtained by the static equilibrium of the element:

$$\begin{array}{*{20}c} {\sum {F_{{\hat{x}}} } = 0} & : & {\Delta \hat{f}_{1x} = - \Delta \hat{f}_{2x} } \\ {\sum {F_{{\hat{y}}} } = 0} & : & {\Delta \hat{f}_{1y} = - \Delta \hat{f}_{2y} } \\ {\sum {M_{1a} } = 0} & : & {\Delta \hat{f}_{2y} = - \left( {\Delta m_{1z} + \Delta m_{2z} } \right)/l_{a} } \\ \end{array}$$
(A6)

Once the incremental internal forces for each node are known, the new nodal forces, \({\hat{\varvec{f}}}\), can be obtained as:

$${\hat{\varvec{f}}}= {\hat{\varvec{f}}}_{{a}} { + }\Delta \hat{f}$$
(A7)

Note that the hat indicates the internal forces are calculated in the local coordinate system at time ta. This force must be transformed to the global coordinate system first and then to the time state tb by the following equation:

$${\varvec{f}}_\text{int} = {\mathbf{R\Omega} \varvec{\hat{f}}}$$
(A8)

where Ω is the transformation matrix from local to global coordinate system and R is the rotation matrix from coordinate at time ta to coordinate at time tb.

1.4 Time integration

To ensure a rapid attainment of a stable static solution, we incorporate a DRM [38] with kinetic damping to minimize the system’s potential energy efficiently. Additionally, we can extract a steady reaction force corresponding to the input node’s displacement. Time integration techniques are required to solve the equation of motion for particles within a path element as described in Eq. (A1). In this work, we apply the central difference explicit scheme. Alternatively, if we opt for DRM, we need to derive the kinetic energy of the particle during the time segment. In this approach, the static equilibrium in a conservative system aligns with the condition where the potential energy reaches a minimal value. Let the displacement equation without damping be expressed as:

$$d_{n + 1} = \frac{{(\Delta t)^{2} }}{m}(F_{n}^{{{\text{ext}}}} - F_{n}^{{\text{int}}} ) + 2d_{n} - d_{n - 1}$$
(A9)

The nodal velocities for the next step can be represented by the half-interval method as

$$\dot{d}_{n - 1/2} = (d_{n} - d_{n - 1} )/\Delta t$$
(A10)
$$\dot{d}_{n + 1/2} = (d_{n + 1} - d_{n} )/\Delta t$$
(A11)

Then, the minimum potential energy can be detected by checking if the inequality below holds at each instance:

$$\sum {\frac{1}{2}(m_{i} \dot{d}_{n + 1/2}^{2} )} < \sum {\frac{1}{2}(m_{i} \dot{d}_{n - 1/2}^{2} )}$$
(A12)

If it holds, then set \(\dot{d}_{n - 1/2}\) to obtain \(d_{n} = d_{n - 1}\). This yields a new displacement:

$$d_{n + 1} = \frac{{(\Delta t)^{2} }}{m}(F_{n}^{{{\text{ext}}}} - F_{n}^{{\text{int}}} ) + d_{n}$$
(A13)

By iteratively applying Eq. (A9) and Eq. (A13) based on the condition of kinetic energy, we can attain the static solution for a dynamic system. Fig.

Fig. 22
figure 22

Flowchart of VFIFE algorithm for quasi-static nonlinear analysis

22 depicts the algorithm of VFIFE coupled with kinetic DRM for quasi-static nonlinear structural analysis.

In the explicit quasi-static nonlinear analysis, the time step is set to Δt = 10−5, and the convergence threshold is defined as the kinetic energy Ek < 10−15. The relationship between the applied displacement and time is given as:

$$d_{{t,{\text{control}}}} = \left\{ {\begin{array}{*{20}c} {\frac{t}{{t_{{{\text{set}}}} }}d_{{{\text{control}}}} { , }t < t_{{{\text{set}}}} } \\ {d_{{{\text{control}}}} { , }t \ge t_{{{\text{set}}}} } \\ \end{array} } \right.$$
(A14)

where dt,control is the displacement at time t, tset is the time at which the control displacement is reached, and dcontrol is the goal of displacement control value.

Appendix 2: Efficiency of the optimization

The GA toolbox in MATLAB was used in this study. The optimization was initiated with 80 individuals and terminated at a maximum generation limit of 70 iterations. The algorithm will stop if tolerance between two steps is less than 10−5. In the finite element analysis performed by VFIFE, the number of iterations for each displacement control point (1–30 mm) to reach a steady state for three types of structures, are shown in Fig.

Fig. 23
figure 23

Number of iterations for three types of synthesized structure

23. Calculating one individual of the data took approximately 25–30 s of CPU time, resulting in about 45 h to achieve one final successful result.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lo, CW., Chang, Y., Wang, ML. et al. Design of a large stroke compliant gripping mechanism for constant-force applications. J Braz. Soc. Mech. Sci. Eng. 46, 363 (2024). https://doi.org/10.1007/s40430-024-04929-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s40430-024-04929-4

Keywords

Navigation