Skip to main content
Log in

An analytical formulation of ZOA-based machining stability for complex tool geometries: application to 5-axis ball-end milling

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

Abstract

The paper describes a computationally convenient analytical formulation of the stability of the cutting process with respect to self-excited vibrations in the case of five-axis milling based on the commonly used zero-order approximation. In the case of five-axis milling with general milling cutters, it is difficult to calculate stable machining process conditions for two main reasons. The first reason is the difficulty of calculating the mean value of the cutting force Jacobian with respect to the regenerative displacement (closely related to a milling directional matrix) for a generally inclined tool, and the second reason is the nonlinearity of this Jacobian with respect to the process parameters, which means that the problem cannot be reduced to a linear eigenvalue problem as is usual for linear cases (e.g. cylindrical milling with respect to the axial depth of cut). In the first part, this paper presents a modification of the calculation of the machining stability limits for a nonlinearly dependent cutting force Jacobian. A new formulation of this Jacobian for a general tool based on the surface integral over the tool and workpiece engagement region is presented which leads directly to the mean value of the Jacobian of the cutting force (direction matrix) without the need to calculate it as a function of time and then calculate the mean value over one revolution. The advantage is that if we can analytically describe the engagement area, we also obtain an analytical relation for the cutting force Jacobian. This is presented with the practical example of a generally inclined ball-end mill. This analytical formulation of the force Jacobian allows the calculation of its derivatives with respect to the technological parameters (depth of cut, step over, tilt and lead angles), which is useful both for the calculation of stability diagrams and for solving optimization problems related to machining stability.

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

Similar content being viewed by others

Availability of data and material

Not applicable.

Code availability

Not applicable.

References

  1. Beudaert X, Franco O, Erkorkmaz K, Zatarain M (2020) Feed drive control tuning considering machine dynamics and chatter stability. CIRP Ann 69(1):345–348. https://doi.org/10.1016/j.cirp.2020.04.054. https://www.sciencedirect.com/science/article/pii/S0007850620300767

  2. Wu J, Yu G, Gao Y, Wang L (2018) Mechatronics modeling and vibration analysis of a 2-DOF parallel manipulator in a 5-DOF hybrid machine tool. Mech Mach Theory 121:430–445. https://doi.org/10.1016/j.mechmachtheory.2017.10.023. https://www.sciencedirect.com/science/article/pii/S0094114X17307103

  3. Tlusty J, Polacek M (1963) The Stability of Machine Tools against Self Excited Vibrations in Machining. International research in production engineering, ASME. pp. 465–474

  4. Budak E, Altintaş Y (1998) Analytical prediction of chatter stability in milling–part I: general formulation. Journal of Dynamic Systems, Measurement and Control, Transactions of the ASME 120(1):22–30. https://doi.org/10.1115/1.2801317

    Article  Google Scholar 

  5. Gradisek J, Kalveram M, Insperger T, Weinert K, Stepan G, Govekar E, Grabec I (2005) On stability prediction for milling. Int J Mach Tool Manuf 45(7–8):769–781. https://doi.org/10.1016/j.ijmachtools.2004.11.015

    Article  Google Scholar 

  6. Altintas Y, Lee P (1999) Analytical prediction of stability lobes in ball end milling. Transactions of the ASME 121(November 1999):586–592

    Google Scholar 

  7. Ozturk E, Budak E (2010) Dynamics and stability of five-axis ball-end milling. J Manuf Sci Eng 132(2):021003. https://doi.org/10.1115/1.4001038. http://manufacturingscience.asmedigitalcollection.asme.org/article.aspx?articleid=1469873

  8. Ozturk E, Tunc LT, Budak E (2011) Analytical methods for increased productivity in five-axis ball-end milling. Int J Mechatron Manuf Syst 4(3-4):238–265. https://doi.org/10.1504/IJMMS.2011.041471. https://www.inderscienceonline.com/doi/abs/10.1504/IJMMS.2011.041471

  9. Wojciechowski S, Twardowski P, Pelic M (2014) Cutting forces and vibrations during ball end milling of inclined surfaces. Procedia CIRP 14(December 2014):113–118. https://doi.org/10.1016/j.procir.2014.03.102

    Article  Google Scholar 

  10. Graham E, Park CI, Park SS (2014) Force modeling and applications of inclined ball end milling of micro-dimpled surfaces. Int J Adv Manuf Technol 70(1–4):689–700. https://doi.org/10.1007/s00170-013-5310-5

    Article  Google Scholar 

  11. Hao Y, Tang G, Zhang M, With I (2014) Calculation method of milling contact area for ball-end milling tool with tool inclination angle. UPB Scientific Bulletin, Series D: Mechanical Engineering 76(3)

  12. Tsai CL, Liao YS (2008) Prediction of cutting forces in ball-end milling by means of geometric analysis. J Mater Process Technol 205(1–3):24–33. https://doi.org/10.1016/j.jmatprotec.2007.11.083

    Article  Google Scholar 

  13. Shtehin OO, Wagner V, Seguy S, Landon Y, Dessein G, Mousseigne M (2016) Stability of ball-end milling on warped surface: semi-analytical and experimental analysis. Int J Adv Manuf Technol. https://doi.org/10.1007/s00170-016-9656-3

    Article  Google Scholar 

  14. Lazoglu I, Liang SY (2016) Modeling of ball-end milling forces with cutter axis inclination in. 122(February 2000):3–11

  15. Lazoglu I, Erdim H (2011) Five-axis milling mechanics for complex free form surfaces. CIRP Annals - Manufacturing Technology (December). https://doi.org/10.1016/j.cirp.2011.03.090

    Article  Google Scholar 

  16. Taner TL, Ömer Ö, Erhan B (2015) Generalized cutting force model in multi-axis milling using a new engagement boundary determination approach. pp 341–355. https://doi.org/10.1007/s00170-014-6453-8

  17. Chao S, Altintas Y (2016) Chatter free tool orientations in 5-axis ball-end milling. Int J Mach Tool Manu 106:89–97. https://doi.org/10.1016/j.ijmachtools.2016.04.007

    Article  Google Scholar 

  18. Li J, Kilic ZM, Altintas Y (2020) General cutting dynamics model for five-axis ball-end milling operations. J Manuf Sci Eng 142(12):1–13. https://doi.org/10.1115/1.4047625

    Article  Google Scholar 

  19. Ju G, Song Q, Liu Z, Shi J, Wan Y (2015) A solid-analytical-based method for extracting cutter-workpiece engagement in sculptured surface milling. Int J Adv Manuf Technol 80:1297–1310. https://doi.org/10.1007/s00170-015-7118-y

    Article  Google Scholar 

  20. Zhan D, Jiang S, Niu J, Sun Y (2020) Dynamics modeling and stability analysis of five-axis ball-end milling system with variable pitch tools. Int J Mech Sci 182(March):105774. https://doi.org/10.1016/j.ijmecsci.2020.105774

  21. Otto A, Rauh S, Kolouch M, Radons G (2014) Extension of Tlusty’s law for the identification of chatter stability lobes in multi-dimensional cutting processes. Int J Mach Tool Manuf 82–83(November 2017):50–58. https://doi.org/10.1016/j.ijmachtools.2014.03.007

    Article  Google Scholar 

  22. Montgomery D, Altintas Y (1991) Mechanism of cutting force and surface generation in dynamic milling. J Eng Ind 113(May 1991):160. https://doi.org/10.1115/1.2899673

    Article  Google Scholar 

  23. Insperger T, Stépán G (2011). Semi-discretization for time-delay systems—stability and engineering applications. https://doi.org/10.1007/978-1-4614-0335-7

  24. Smiley MW, Chun C (2001) An algorithm for finding all solutions of a nonlinear system. J Comput Appl Math 137(2):293–315. https://doi.org/10.1016/S0377-0427(00)00711-1. https://www.sciencedirect.com/science/article/pii/S0377042700007111

  25. Ozturk E, Tunc LT, Budak E (2009) Investigation of lead and tilt angle effects in 5-axis ball-end milling processes. Int J Mach Tools Manuf 49(14):1053–1062. https://doi.org/10.1016/j.ijmachtools.2009.07.013

    Article  Google Scholar 

  26. Dombovari Z, Munoa J, Kuske R, Stepan G, Kuske R, Dombovari Z, Munoa J, Kuske R, Stepan G, Stief P, Dantan JY, Etienne A, Siadat A (2018) Milling stability for slowly varying parameters. Procedia CIRP 77:110–113. https://doi.org/10.1016/j.procir.2018.08.233

    Article  MATH  Google Scholar 

  27. Shamoto E, Akazawa K (2009) Analytical prediction of chatter stability in ball end milling with tool inclination. CIRP Ann Manuf Technol 58:351–354. https://doi.org/10.1016/j.cirp.2009.03.087

    Article  Google Scholar 

  28. Altíntaş Y, Lee P (1998) Mechanics and dynamics of ball end milling. J Manuf Sci Eng 120(4):684–692. https://doi.org/10.1115/1.2830207

    Article  Google Scholar 

  29. Sanz-Calle M, Munoa J, Iglesias A, de Lacalle LL, Dombovari Z (2021a) The influence of radial engagement and milling direction for thin wall machining: a semi-analytical study. Procedia CIRP 102:180–185. https://doi.org/10.1016/j.procir.2021.09.031. https://www.sciencedirect.com/science/article/pii/S2212827121007745. 18th CIRP Conference on Modeling of Machining Operations (CMMO), Ljubljana, Slovenia, June 15-17, 2021

  30. Sanz-Calle M, Munoa J, Iglesias A, Lopezde Lacalle LN, Dombovari Z (2021b) Semi-analytical period-doubling chatter analysis in thin wall milling. MM Science Journal 2021-November:5126–5133. www.scopus.com

  31. Gradišek J, Kalveram M, Weinert K (2004) Mechanistic identification of specific force coefficients for a general end mill. Int J Mach Tools Manuf 44(4):401–414. https://doi.org/10.1016/j.ijmachtools.2003.10.001. https://www.sciencedirect.com/science/article/pii/S0890695503002682

  32. Altintas Y, Engin S (2001) Generalized modeling of mechanics and dynamics of milling cutters. CIRP Ann 50(1):25–30. https://doi.org/10.1016/S0007-8506(07)62063-0. https://www.sciencedirect.com/science/article/pii/S0007850607620630

  33. Engin S, Altintas Y (2001) Mechanics and dynamics of general milling cutters.: Part I: helical end mills. Int J Mach Tools Manuf 41(15):2195–2212. https://doi.org/10.1016/S0890-6955(01)00045-1. https://www.sciencedirect.com/science/article/pii/S0890695501000451

Download references

Funding

Supported from the EU Operational Programme Research, Development and Education, and from the Center of Advanced Aerospace Technology (CZ.02.1.01/0.0 /0.0/16_019/0000826), Faculty of Mechanical Engineering, Czech Technical University in Prague.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: Jiří Falta, Pavel Zeman, Matěj Sulitka; formal analysis: Jiří Falta; funding acquisition: Pavel Zeman; investigation: Jiří Falta; methodology: Jiří Falta, Pavel Zeman, Matěj Sulitka; project administration: Pavel Zeman; supervision: Pavel Zeman, Matěj Sulitka; visualization: Jiří Falta; writing—original draft: Jiří Falta, Pavel Zeman, Matěj Sulitka

Corresponding author

Correspondence to Jiří Falta.

Ethics declarations

Ethics approval

Not applicable.

Consent to participate

Not applicable.

Consent for publication

Not applicable.

Conflict of interest

The authors declare no competing interests.

Additional information

Publisher’s Note

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

Appendix

Appendix

1.1 A.1 Derivation of the formula for the calculation of the cutting force Jacobian

Below, a derivation of the transition from the standard method of calculating the mean value of the Jacobian to the calculation via surface integration over the engagement surface (Eq. 8) is outlined. Here, the engagement is assumed to be the same for all blades within a revolution and we can therefore replace the sum by N-times.

$$\begin{aligned}{}[\overline{\nabla _{\mathbf \Delta } \mathbf{F}}]&=\frac{1}{2\pi }\int \limits _0^{2\pi }\sum _{j=1}^N \int \limits _{s_{min}}^{s_{max}} g(s,j,\varphi )\, [T]\left. \frac{\partial {{ {f}}}}{\partial {h}}\right| _{\mathbf \Delta =\mathbf {0}} \mathbf {\hat{n}}^T \frac{d {w}}{d {s}}{\text {d}}{s} d\varphi \\&=\frac{N}{2\pi }\underbrace{\int \limits _0^{2\pi }\int \limits _{s_{min}}^{s_{max}}g(s,0,\varphi )}_{\int _{S_{eng}}}\frac{ [T]\left. \frac{\partial {{ {f}}}}{\partial {h}}\right| _{\mathbf \Delta =\mathbf {0}} \mathbf {\hat{n}}^T}{\rho (\mathbf{x})} \underbrace{\rho (\mathbf{x})\frac{d {w}}{d {s}}{\text {d}}{s} d\varphi }_{d S}. \end{aligned}$$

where \(s_{min}, s_{max}\) denote lower and upper limit of cutting edge parametrization. The merit of the formulation is that while the previous procedure works with a parametrization based on tool rotation and a curve parameter for the cutting edge, the reformulation into a surface integral no longer depends on the surface parametrization and it is therefore possible to choose the parametrization in which the calculation is the simplest.

The derivation of the formulation is based on a textbook formula for a surface element in the curvilinear coordinate system, which is

$$\begin{aligned} {\text {d}}{S}&=\left\| \frac{d\mathbf {x}}{{\text {d}}{s}}\times \frac{d\mathbf {x}}{d\varphi } \right\| {\text {d}}{s} d\varphi . \end{aligned}$$

The cross product can be conveniently written as

$$\begin{aligned} \frac{d\mathbf {x}}{d s}\times \frac{d\mathbf {x}}{d\varphi }&=\frac{d\mathbf {x}}{d s}\times (\mathbf {\hat{e}}_\Omega \times \mathbf {x})\\&= \left( \frac{d\mathbf {x}}{d s}\cdot \mathbf {\hat{e}}_b\right) \left\| \mathbf {x}-(\mathbf {x}\cdot \mathbf {\hat{e}}_\Omega ) \mathbf {\hat{e}}_\Omega \right\| \mathbf {\hat{n}}, \end{aligned}$$

which leads to formulation used in the integral above

$$\begin{aligned} {\text {d}}{S}&= \underbrace{\left\| \mathbf {x}-(\mathbf {x}\cdot \mathbf {\hat{e}}_\Omega ) \mathbf {\hat{e}}_\Omega \right\| }_{\rho (\mathbf {x})} \underbrace{\left| \frac{d {\varvec{\gamma }_j}}{d {s}}\cdot \mathbf {\hat{b}}\right| {\text {d}}{s}}_{{\text {d}}\!{w}} d\varphi , \end{aligned}$$

where \(\mathbf {\hat{e}}_\omega\) is the tool axis direction unit vector.

1.2 A.2 Material removal simulation SW MillVis

The SW implementation of the material removal simulation MillVis uses voxel and distance field representation of the workpiece. The voxel discretization is to divide the volume into cubic blocks, which are further divided into \(3\times 3 \times 3\) cells. Each block is labeled as either outer, surface, or inner, with only surface blocks intersecting the workpiece surface and the distance function values containing only vertices belonging to at least one intersected cell, see Fig. 10a. For all vertices of such a cell, the distance to the workpiece surface is reconstructed using trilinear interpolation — this approach is called a distance field, see Fig. 10b. For the vertices distant from the workpiece surface, the value of the distance function is set to plus or minus infinity for simplicity, since the actual value is not needed.

Fig. 10
figure 10

a Outer, surface and inner block for voxel representation of the workpiece, b distance of the workpiece boundary from cell vertices \(d_k\) for detailed workpiece representation by a trilinear surface

The workpiece represented by voxels and distance fields can be effectively visualized using ray tracing, which provides high fidelity detail of the machined surface.

In the material removal simulation, surface and internal blocks that are close to the tool are identified. Then, for each of their cells, it is checked whether it is intersected by the tool and, if so, the distance function values at the corresponding vertices are updated. In practice, the representation of the workpiece geometry by the distance field has been shown to be sufficiently accurate while maintaining reasonable memory requirements, as lower accuracy is usually required for larger workpieces and vice versa (see Fig. 11).

The tool envelope definition and cutting force formulation is based on the framework by [32, 33]. Over the course of the development of this software tool, the ability to predict cutting forces with sufficient accuracy was repeatedly confirmed by comparison with dynamometer data. Figure 11 shows a comparison of measured and simulated cutting forces during pocket machining (Fig. 12).

Fig. 11
figure 11

Comparison of real and simulated cutting force (active component) during pocket machining

Fig. 12
figure 12

Pocket machining for comparison of cutting forces: a virtual machining in material removal simulation software and b real-life machining on the same toolpaths

1.3 A.3 Ball-end milling local coordinate system basis

The LCS basis vectors defined in Eq. 4 in the spherical surface parametrization in the ECS are presented below. However, if possible, it is preferable to calculate the LCS basis formulas based on tool axis vector Eq. (22) and their definition which reduces possibility of an error. The tangential direction is dominantly given by circumferential velocity calculated from tool axis vector and point on the tool envelope (or engagement surface)

$$\begin{aligned} \mathbf {\hat{e}}_\Omega \times \mathbf {x}&=r\begin{pmatrix} \tan \psi _T \cos \theta +\cos \phi \sin \theta \\ -\tan \psi _L \cos \theta - \sin \phi \sin \theta \\ \sin \theta (\tan \psi _T \sin \phi -\tan \psi _L \cos \phi ) \end{pmatrix}. \end{aligned}$$

The LCS basis vectors are shown below

$$\begin{aligned} \mathbf {\hat{t}}&=\frac{\mathbf {\hat{e}}_\Omega \times \mathbf {x}}{\Vert \mathbf {\hat{e}}_\Omega \times \mathbf {x}\Vert },\\ \mathbf {\hat{n}}&=\begin{pmatrix} \sin \phi \sin \theta \\ \cos \phi \sin \theta \\ -\cos \theta \end{pmatrix},\\ \mathbf {\hat{b}}&=\frac{r}{\Vert \mathbf {\hat{e}}_\omega \times \mathbf {x} \Vert }\begin{pmatrix} \cos \phi \sin ^2\theta (\tan \psi _L \cos \phi -\tan \psi _T \sin \phi )+\tan \psi _L \cos ^2\theta +\sin \phi \sin \theta \cos \theta \\ \tan \psi _T \left( \sin ^2\phi \sin ^2\theta +\cos ^2\theta \right) +\cos \phi \sin \theta (\cos \theta -\tan \psi _L \sin \phi \sin \theta )\\ \sin \theta \cos \theta (\tan \psi _T \cos \phi +\tan \psi _L \sin \phi )+\sin ^2 \theta \end{pmatrix}. \end{aligned}$$

1.4 A.4 Script for cutting force Jacobian calculation

The calculation of the Jacobian for milling spherical ends leads to matrices that are too complex to be reasonably presented in text form. The code below in Wolfram Mathematica follows the calculation described in the paper. The division into the functions shown is made with readability in relation to the article in mind, not code efficiency. For faster computation, it is preferable to analytically precalculate some of the functions (e.g. matrix [T]).

figure a
figure b

Rights and permissions

Springer Nature or its licensor 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

Falta, J., Sulitka, M. & Zeman, P. An analytical formulation of ZOA-based machining stability for complex tool geometries: application to 5-axis ball-end milling. Int J Adv Manuf Technol 123, 1499–1519 (2022). https://doi.org/10.1007/s00170-022-10170-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00170-022-10170-x

Keywords

Navigation