Skip to main content
Log in

Substructure analysis of vibration-assisted drilling systems

  • ORIGINAL ARTICLE
  • Published:
The International Journal of Advanced Manufacturing Technology Aims and scope Submit manuscript

Abstract

Designing vibration-assisted drilling (VAD) toolholders requires accurately modeling the dynamics of the mechanical structure and its electromechanical coupling with the piezoelectric actuator. The overall dynamics of the VAD toolholder depends on the dynamics of its individual components (e.g., concentrator, piezoelectric disk, drill bit) and the interactions between them. Modelling by substructure analysis therefore provides an efficient framework for studying the effect of each component on the overall dynamics of the VAD system. Nonetheless, the existing substructure analysis methods are only applicable to holders with basic concentrator geometry, which limits their application in designing high-performance holders.

In this paper, Receptance Coupling Substructure Analysis (RCSA) is used to combine the dynamic models of the toolholder’s components to determine its vibration response to the electrical excitation of the piezoelectric actuator. This method allows for combining the 3D finite element models of the concentrator with the simplified and analytical models of the rest of the VAD components, enabling the inclusion of complex concentrator designs in modeling without imposing a large computational load. Furthermore, because this method couples the mechanical model of the structure with the electrical model of the piezo-actuator’s impedance, it can be used as a design tool to modify the toolholder’s natural frequencies by tuning the electrical impedance of its piezoelectric components. This possibility effectively converts the piezoelectric components to mechanical members with adjustable structural dynamics, while they also actuate vibrations. Both axial and axial-torsional VAD toolholders are studied in this work, and the accuracy of the presented modelling approach is verified experimentally.

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

Similar content being viewed by others

References

  1. Brehl D, Dow T (2008) Review of vibration-assisted machining. Precis Eng 32:153–172

    Article  Google Scholar 

  2. Neugebauer R, Stoll A (2004) Ultrasonic application in drilling. J Mater Process Technol 149:633–639

    Article  Google Scholar 

  3. M’Saoubi R, Axinte D, Soo S et al (2015) High performance cutting of advanced aerospace alloys and composite materials. CIRP Ann 64:557–580

    Article  Google Scholar 

  4. Astashev V, Babitsky V (2007) Ultrasonic processes and machines: dynamics, control and applications. Springer Science and Business Media, New York

    MATH  Google Scholar 

  5. Moriwaki T, Shamoto E, Inoue K (1992) Ultraprecision ductile cutting of glassby applying ultrasonic vibration. CIRP Ann 41:141–144

    Article  Google Scholar 

  6. Ma C, Shamoto E, Moriwaki T et al (2004) Study of machining accuracy in ultrasonic elliptical vibration cutting. Int J Mach Tools Manuf 44:1305–1310

    Article  Google Scholar 

  7. Shuyu L (1997) Sandwiched piezoelectric ultrasonic transducers of longitudinal-torsional compound vibrational modes. IEEE Trans Ultrason Ferroelectr Freq Control 44:1189–1197

    Article  Google Scholar 

  8. Ohnishi O, Myohga O et al (1993) Piezoelectric ultrasonic motor using longitudinal-torsional composite resonance vibration. IEEE Trans Ultrason Ferroelectr Freq Control 40:687–693

    Article  Google Scholar 

  9. Gao J, Altintas Y (2019) Development of a three-degree-of-freedom ultrasonic vibration tool holder for milling and drilling. IEEE/ASME Trans Mechatron 24:1238–1247

    Article  Google Scholar 

  10. Al-Budairi H, Lucas M, Harkness P (2013) A design approach for longitudinal–torsional ultrasonic transducers. Sensors Actuators Phys 198:99–106

    Article  Google Scholar 

  11. Harkness P, Lucas M, Cardoni A (2012) Coupling and degenerating modes in longitudinal–torsional step horns. Ultrasonics 52:980–988

    Article  Google Scholar 

  12. Asami T, Miura H (2011) Vibrator development for hole machining by ultrasonic longitudinal and torsional vibration. Jpn J Appl Phys 50:7–31

    Article  Google Scholar 

  13. Amini S, Soleimanimehr H et al (2008) FEM analysis of ultrasonic-vibration-assisted turning and the vibratory tool. J Mater Process Technol 201:43–47

    Article  Google Scholar 

  14. Abboud N, Wojcik G et al (1998) Finite element modeling for ultrasonic transducers in medical imaging. Ultrasonic Transducer Engineering. Int Soc Opt Photon 3341:19–42

    Google Scholar 

  15. Amini S, Soleimani M et al (2017) Effect of longitudinal-torsional vibration in ultrasonic-assisted drilling. Mater Manuf Process 32:616–622

    Article  Google Scholar 

  16. Sherrit S, Leary S et al (1999) Comparison of the mason and klm equivalent circuits for piezoelectric resonators in the thickness mode. IEEE Ultrason Symp Proc Int Symp 2:921–926

    Google Scholar 

  17. Lin S, Guo H, Xu J (2018) Actively adjustable step-type ultrasonic horns in longitudinal vibration. J Sound Vib 419:367– 379

    Article  Google Scholar 

  18. Schmitz T, Smith K (2001) Mechanical vibrations: modeling and measurement. Springer Science and Business Media, New York

    MATH  Google Scholar 

  19. Schmitz T (2010) Torsional and axial frequency response prediction by RCSA. Precis Eng 34:345–356

    Article  Google Scholar 

  20. Moheimani S, Fleming A (2006) Piezoelectric transducers for vibration control and damping. Springer Science and Business Media, New York

    MATH  Google Scholar 

  21. Rosen A (1983) Theoretical and experimental investigation of the nonlinear torsion and extension of initially twisted bars. ASME J Applied Mechanics 50:321-326

  22. Ogata K (2010) Modern control engineering 5th edn. Prentice hall United States

  23. Ewins D (2009) Modal testing: theory, practice and application. John Wiley & Sons, New York

    Google Scholar 

  24. Mohammadi Y, Azvar M, Budak E (2018) Suppressing vibration modes of spindle-holder-tool assembly through FRF modification for enhanced chatter stability. CIRP Ann 67:397–400

    Article  Google Scholar 

  25. Erturk A, Ozguven H, Budak E (2006) Analytical modeling of spindle–tool dynamics on machine tools using timoshenko beam model and receptance coupling for the prediction of tool point FRF. Int J Mach Tools Manuf 46:1901–1912

    Article  Google Scholar 

  26. Rao S (2005) Mechanical vibrations. Addison Wesley, Boston

    Google Scholar 

  27. Jin X, Koya N (2016) Investigation of warping effect on coupled torsional-axial vibration of drilling tool. Int J Adv Manuf Technol 86:2961–2974

    Article  Google Scholar 

Download references

Acknowledgements

The authors thank Mr. Joshua Columbus for designing the toolholder used in the experiments.

Funding

This research was financially supported by the National Sciences and Engineering Research Council of Canada (NSERC) through the Discovery Grant program and Canadian Network for Research and Innovation in Machining Technology Network (CANRIMT 2).

Author information

Authors and Affiliations

Authors

Contributions

V. Ostad Ali Akbari contributed to conceptualization, formal analysis, validation, data curation, visualization, and original draft. K. Ahmadi contributed to conceptualization, writing-review, editing, supervision, and funding acquisition.

Corresponding author

Correspondence to Keivan Ahmadi.

Ethics declarations

Conflict of interest

The authors have no competing interests.

Additional information

Publisher’s note

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

Appendix : 1: Modeling vibrations of VAD components for RCSA

Appendix : 1: Modeling vibrations of VAD components for RCSA

In this section, the analytical mode shapes of rods with uniform cross-section, the finite element model of rod vibrations, and the finite element model of twisted rod vibrations are explained.

1.1 1.1 Uniform cylindrical rod

Consider an elastic rod with length l, Young’s modulus E, density ρ, and cross-section area A. Assuming that the rod is rigid in lateral directions, and its torsional and axial deformations are uncoupled, the natural frequencies (ωn) and mass-normalized mode shapes (Un) of its axial vibrations in free-free end condition are obtained as follows [26]:

$$ \begin{array}{@{}rcl@{}} {\omega_{n}} &= &\frac{{n\pi }}{l}\sqrt {\frac{E}{\rho}} ;\\ {U_{0}}(x) &=& \frac{1}{{\sqrt {\rho Al} }};{U_{n}}(x) = \sqrt {\frac{2}{{\rho Al}}} \cos \frac{{n\pi x}}{l} \end{array} $$
(24)

where n = 0, 1, 2,…, and x is the distance from a free end of the rod. Using these mode shapes, the axial FRF between two points at xi and xj is synthesized as follows:

$$ h_{ij}^{xf}\left( {x_{i}},{x_{j}},\omega\right) =\frac{x_{i}}{f_{j}}= \sum\limits_{n = 0}^{N_{m}} {\frac{{{U_{n}}\left( {x_{i}}\right){U_{n}}\left( {x_{j}}\right)}}{{ - {\omega^{2}} + {\omega_{n}^{2}}}}} $$
(25)

where Nm is the number of modes considered in synthesizing the FRF. Now, consider the same elastic rod with modulus of rigidity of G and polar moment of inertia of J. The natural frequencies (ωn) and mass-normalized mode shapes (Θn) of its torsional vibrations are obtained as follows [26]:

$$ \begin{array}{@{}rcl@{}} {\omega_{n}} &= &\frac{{n\pi }}{l}\sqrt {\frac{G}{\rho }};\\ {{\varTheta}_{0}}(x) &=& \frac{1}{{\sqrt {\rho Jl} }};{{\varTheta}_{n}}(x) = \sqrt {\frac{2}{{\rho Jl}}} \cos \left( \frac{{\pi nx}}{l}\right) \end{array} $$
(26)

where n = 0, 1, 2,…. Similar to Eq. 25, the torsional FRF between points at xi and xj is also synthesized using the torsional mode shapes:

$$ h_{ij}^{\theta t}\left( {x_{i}},{x_{j}},\omega\right) =\frac{\theta_{i}}{t_{j}}= \sum\limits_{n = 0}^{N_{m}} {\frac{{{{\varTheta}_{n}}\left( {x_{i}}\right){{\varTheta}_{n}}\left( {x_{j}}\right)}}{{ - {\omega^{2}} + {\omega_{n}^{2}}}}} $$
(27)

1.2 1.2 Rod finite element model

Finite element method is used to model the vibrations of rods with non-uniform cross-section—for example, the tapered part of back mass in Section 2.1. A rod element with cross-section area A, length l, polar moment of inertia per unit length Ip, density ρ, modulus of elasticity E, modulus of rigidity G, and polar moment of area J is considered. Assuming uncoupled axial and torsional deformations, the mass and stiffness matrices of a 2-nod, 4 DOF, rod element with DOF vector \(\textbf {u}=\left [x_{1},\theta _{1},x_{2},\theta _{2}\right ]^{T}\), are obtained as follows [26]:

$$ \begin{array}{l} {\textbf{m}} = \frac{{\rho l}}{6}\left[ {\begin{array}{*{20}{c}} {2A}&0&A&0\\ 0&{2{I_{p}}}&0&{{I_{p}}}\\ A&0&{2A}&0\\ 0&{{I_{p}}}&0&{2{I_{p}}} \end{array}} \right]\\ {\textbf{k}} = \frac{1}{l}\left[ {\begin{array}{*{20}{c}} {EA}&0&{ - EA}&0\\ 0&{GJ}&0&{ - GJ}\\ { - EA}&0&{EA}&0\\ 0&{ - GJ}&0&{GJ} \end{array}} \right] \end{array} $$
(28)

The global mass and stiffness matrices of the non-uniform rod are obtained by assembling its element matrices; the desired FRFs are subsequently obtained based on the resulting global mass and stiffness matrices [26].

1.3 1.3 Coupled axial-torsional finite element to model the drill bit

Due to the pre-twisted geometry of the drill bit, its axial and torsional deflections are coupled. This coupling is considered in the presented work by using Rosen’s model [21] of the coupled non-linear deflection of pre-twisted rods. The drill bit is divided into finite rod elements with pre-twisted geometry as shown in Fig. 11. As shown in this figure, when an axial force (f ) is applied to the pre-twisted rod element, it simultaneously deflects in axial direction by Δx1 and twists through by Δ𝜃1. Similarly, when the element is subjected to torque t, it deflects axially by Δx2 and twists through by Δ𝜃2.

Fig. 11
figure 11

Deformations of a drill bit element under a an axial force and b a torsional torque

Rosen [21] presented a non-linear model to describe the mechanics of the torsional-axial deflections in pre-twisted rods. Nonetheless, Jin and Koya [27] showed that, for typical deflections during drilling, the values of the non-linear terms in Rosen’s model are less than 1% of the linear terms. Therefore, the effect of non-linearities can be neglected. Based on the uniform axial strain and twist per unit length assumptions in Rosen’s work [21], the following equation is developed to describe the coupling between the axial and torsional loads (f and t) and the resulting deflections (Δx and Δ𝜃) in the pre-twisted rod element shown in Fig. 11:

$$ f = k^{xf}{\varDelta} x + k^{\theta f}{\varDelta} \theta ; t = k^{xt}{\varDelta} x + k^{\theta t}{\varDelta} \theta $$
(29)

where the stiffness coefficients are expressed as follows:

$$ \begin{array}{l} {k^{xf}} = {EA}/{l};{k^{\theta f}} = {ES_{s}}/{l};\\ {k^{xt}} = {ES_{s}}/{l};{k^{\theta t}} = \left( {G{J_{s}} + EK_{s}}\right)/{l} \end{array} $$
(30)

and E, G, A, and l being the modulus of elasticity, shear modulus, cross-section area, and length of the element. The parameters Ss, Js, Ks, and Fs are section integral parameters [21] and depend on the geometrical dimensions of the cross-section. Various methods are available in the literature to obtain the section integral parameters. For example, a numerical method based on a 2D FE solution was used in [27]. In this work, assuming a standard twist drill cross-section geometry, a curve-fitting method is used to obtain these parameters as a function of drill diameter, d. A pre-twisted rod element is modelled in the commercial finite element software COMSOL Multiphysics, and its stiffness coefficients are obtained by applying axial force and torsional torque to the element and measuring the deformations, as shown in Fig. 11. This process is repeated for rods with various diameters. Subsequently, a cubic polynomial is fitted on the resulting stiffness coefficients, shown below:

$$ \begin{array}{@{}rcl@{}} {k^{xf}}& =& 765.6{d^{3}} + 1.125 \times {10^{5}}{d^{2}} + {1.49910^{5}}d\\ &&- 3.089 \times {10^{5}}\\ {k^{\theta f}} &=& 13.38{d^{3}} + 0.8516{d^{2}} - 0.6822d - 4.749\\ {k^{\theta t}} &= &0.1071{d^{3}} - 1.209{d^{2}} + 5.622d - 8.954\\ {k^{\theta t}} &=& 0.1071{d^{3}} - 1.209{d^{2}} + 5.622d - 8.954 \end{array} $$
(31)

where d is rod diameter in mm.

The material considered for obtaining the cubic polynomials is tungsten carbide with modulus of elasticity of EWC and for a unit length of the element. Since the stiffness coefficients are proportional to the modulus of elasticity E and inversely proportional to the element length l, the stiffness matrix for a material with the modulus of elasticity of E and the element length of l, is expressed as follows:

$$ \begin{array}{l} {\mathbf{m}} = \left[ {\begin{array}{*{20}{c}} {\frac{m}{3}}&0&{\frac{m}{6}}&0\\ 0&{\frac{I}{3}}&0&{\frac{I}{6}}\\ {\frac{m}{6}}&0&{\frac{m}{3}}&0\\ 0&{\frac{I}{6}}&0&{\frac{I}{3}} \end{array}} \right];\\ {\mathbf{k}} = \frac{{ E}}{{l{E_{WC}}}}\left[ {\begin{array}{*{20}{c}} {{k^{xf}}}&{{k^{\theta f}}}&{ - {k^{xf}}}&{ - {k^{\theta f}}}\\ {{k^{xt}}}&{{k^{\theta t}}}&{ - {k^{xt}}}&{ - {k^{\theta t}}}\\ { - {k^{xf}}}&{ - {k^{\theta f}}}&{{k^{xf}}}&{{k^{\theta f}}}\\ { - {k^{xt}}}&{ - {k^{\theta t}}}&{{k^{xt}}}&{{k^{\theta t}}} \end{array}} \right] \end{array} $$
(32)

where m and I are mass and moment of inertia of the element, respectively.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ostad Ali Akbari, V., Ahmadi, K. Substructure analysis of vibration-assisted drilling systems. Int J Adv Manuf Technol 113, 2833–2848 (2021). https://doi.org/10.1007/s00170-021-06777-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00170-021-06777-1

Keywords

Navigation