Abstract
In the machining process that material removal involves, the time-delay effect due to the regenerative cutting is the root cause of tool chatter, which is a severe nonlinear vibration that leads to system failures. The Chebyshev collocation method (CCM) can be applied to the stability analysis of the delay-affected machining systems. However, when the degree-of-freedom (DOF) of the system is high, the computational efficiency of the Chebyshev collocation method is far lower than the commonly used semidiscretization and full-discretization methods (FDMs). In this article, a robotic milling model allowing an arbitrarily high degree-of-freedom is proposed as a benchmark to evaluate the computation performance for different stability analysis algorithms. An improved algorithm named the “fast Chebyshev collocation method” (FCCM) is proposed to handle the delay differential equation (DDE) with a high degree-of-freedom. The proposed fast Chebyshev collocation method accelerates the traditional Chebyshev collocation method in two approaches: one is the inversion of the matrix when constructing the transition matrix, and the other is the reduction of the dimension of the transition matrix by applying the Sherman–Morrision–Woodbury formula. Subsequently, both the full-discretization method and the proposed method are applied to the robotic milling system to show their convergence rate, computation efficiency, and accuracy. The results demonstrate that the proposed method is overall advantageous to the full-discretization methods in convergence, efficiency, and accuracy even when the degree-of-freedom is sufficiently high, implying that the proposed fast Chebyshev collocation method can be a potential alternative tool to deal with the stability analysis for time-delay systems.
1 Introduction
Regenerative chatter [1,2] is known as the main factor causing poor surface quality and tool wear in the field of high-speed machining [3–5]. To avoid chatter, stability analysis of the dynamic system describing the machining system is important. Typically, the dynamic of the machining structure is modeled by modal parameters obtained via experimental modal analysis [6] or finite element analysis (FEA) [7,8]. As for the cutting force model, the linear force model proportional to the chip thickness is commonly used. By considering the wave left on the machined surface by the vibration of the tool in the previous time, the time delay is introduced to the cutting force model. Combining the structure dynamic model and the cutting force model, the governing equation of the machining process is formulated as the delay differential equation (DDE) with constant or time-varying periodic coefficients.
As for the stability analysis of the machining system based on the DDE, since a closed-form analytical solution of DDE is usually impossible to find, many efforts have been taken to find the numerical method to predict the stability of DDE with both high computation efficiency and accuracy. These methods explore the stability of DDE around the equilibrium or periodic orbit under different combinations of parameters, resulting in the stability lobe diagram (SLD). SLD divides the parameter space, usually, the two-dimension parameter space including the spindle speed and the axial depth of the cutting tool, into the stable (chatter-free) and unstable (chatter) zones, then one can easily choose the appropriate parameter to avoid chatter during the practical machining process. Nowadays, many methods have been proposed to predict the SLD of a variety of DDEs. For DDE with constant coefficients, the D-subdivision method [9] can be used by finding the set of parameters under which the characteristic equation has imaginary roots. Sadath and Vyasarayani [10] proposed Galerkin approximations for stability analysis of DDE with periodic coefficients by transforming the original DDE into an equivalent partial differential equation with a time-periodic boundary condition. By implementing the frequency-domain transformation of the DDE, Altintas and Budak [11] first proposed the zero-order approximation method, or single frequency method, to predict the SLD of the milling process, where the critical axial depth has an analytical solution, which greatly enhances the computation efficiency. Then, the multifrequency method [12] was proposed to obtain more accurate results in low radial immersion milling.
Different from the analytical, frequency-domain method mentioned above, Insperger and Stépán [13] proposed the commonly used semidiscretization method (SDM) as a time-domain method to predict the stability of DDE, where only the delay term of the dynamic equation is approximated by zero-order [14] or first-order interpolation polynomials [15], discretizing the original DDE into piecewise ordinary differential equations (ODE). Then, the stability is predicted by calculating the critical eigenvalue of the transition matrix, which is an approximation of the infinite-dimension monodromy operator of the original DDE. Subsequently, Ding et al. [16] proposed the full-discretization method (FDM) as a modified version of SDM, where the coefficients of the state term are decomposed into the time-invariant part and the periodic part, and the state term and the delay term are discretized simultaneously. This modification significantly reduces the calculation times of the matrix exponent, enhancing the computation efficiency. To improve the prediction accuracy of the critical eigenvalue, several high-order SDMs [17] and FDMs [18] were proposed using high-order interpolation polynomial to discretize the state and the delay terms. On the other hand, to improve the computation efficiency, Henninger and Eberhard [19] summarized the numeric techniques to reduce the computation time of SDM, including reducing the dimension of the transition matrix and utilizing the sparsity of the corresponding matrix when implementing matrix multiplying.
In some situations such as interrupted milling with low radial immersion, the tool leaves the workpiece periodically, causing the delay term to vanish during some part of the period. In these time intervals, the original DDE becomes ODE. Based on this observation, Mann et al. [20] proposed the time finite element analysis (TFEA) method to predict the SLD and the surface location error simultaneously. TFEA method separates the DDE into the cutting state and the free state, at the free state, the DDE is replaced by an ODE, improving the prediction accuracy. However, the computation efficiency is low due to the symbolic operation used in the construction of the transition matrix. To overcome this difficulty, Khasawneh and Mann [21] proposed the spectral element method, which can achieve the spectral convergence rate compared to the linear convergence rate in SDM and FDM. Butcher et al. [22] proposed the Chebyshev collocation method (CCM) which is also able to achieve spectral convergence rate. Breda et al. [23,24] proposed a pseudospectral method for the stability analysis of autonomous and periodic linear delayed dynamical systems, which has a fast convergence rate. Given that one of the main drawbacks of the above spectral element method and CCM is that the spectral convergence rate is lost when the coefficients of DDE have discontinuities within the whole period, which limits their use for the machining process when multiple teeth of the tool engaged in cutting simultaneously, Khasawneh et al. [25] proposed the multi-interval Chebyshev collocation method to handle DDE with discontinuities. A comprehensive comparison of the above-mentioned semi-analytical methods can be found in Refs. [26] and [27].
It should be noted that in most literature, only the most flexible mode is chosen to model the structure dynamic of the machining system. But, Wan et al. [28] pointed out that this choice is only valid when the multiple modes are separated far away enough in terms of the modal stiffness, and proposed the lowest envelope method to construct the SLD when two or more modes with the almost modal stiffness co-exist. Moreover, in some cases such as the rotary drilling system with a long, flexible drill pipe, it is convenient to apply finite element model to discretize the original continuous system [7,8]. All these modeling methods result in a DDE with high degrees-of-freedom (DOF). There is little literature discussing and comparing the performance of different semi-analytical methods when applied to DDE with high DOFs, and among these methods, CCM seems to be a potential choice to predict the SLD due to its spectral convergence rate and relatively simple construction of the transition matrix. Totis et al. [29] used both CCM and SDM to predict the SLD in milling with spindle speed variation and compared the performance of those two methods in detail. It was concluded that CCM is superior to SDM both in computation accuracy and efficiency. However, they pointed out in their recent work [30] that the computation performance is not satisfactory especially when the dimension of the transition matrix is large.
Inspired by the work in Ref. [19], this article proposes a fast Chebyshev collocation method (FCCM) which greatly reduces the dimension of the transition matrix. Since it has been shown that FDM computes faster than SDM when the dimensions of the transition matrix are the same [16], this article uses FDM as the reference and compares the computation performance of the proposed method and the FDM. The remainder of this article is organized as follows. In Sec. 2, the dynamic model of a robotic milling system is built; this model allows the use of a finite element model of the cutting tool, providing arbitrarily high DOFs by increasing the number of elements. Combining this dynamic model with the classic orthogonal cutting force model, the DDE of the full system is derived which will be used in Secs. 3 and 4 to test the computation performance of stability prediction algorithms. In Sec. 3, the original algorithm of CCM is briefly introduced, and the computation efficiency of CCM and FDM is compared. In Sec. 4, an acceleration algorithm named fast Chebyshev collocation method is proposed in detail, including the reduction of the dimension of the transition matrix and the acceleration of the matrix inversion. The convergence rate and the computation efficiency of the proposed method are then analyzed comparing to the FDM. Finally, the main conclusions are drawn in Sec. 5.
2 Modeling of the Robotic Milling System
To evaluate the performance of the SLD prediction method under dynamic systems with high DOFs, a robotic milling model allowing arbitrarily high DOFs is proposed in this section. The modeling process of the structural dynamic of the system and the cutting force is introduced in Secs. 2.1 and 2.2, respectively.
2.1 Structural Dynamics of the Robotic Milling System.
A robot milling system consists of three parts, the robotic arm, the spindle, and the cutting tool, as shown in Fig. 1. Since the spindle is fixed on the last joint of the robotic arm, it can be considered as a part of the robotic arm. The modeling contains two parts: the robot-spindle assembly's dynamic model and the cutting tool's finite element model.
where is the Jacobi matrix of size mapping the angular velocity of the joints into the velocity and angular velocity in the end coordinate frame [31]. denote the error between the actual and expected displacement and rotation angle in the end coordinate frame, respectively: , , where x, y, z denote the displacement and denote the rotation angle. The subscript “d” denotes “expected,” i.e., displacement and rotation angle along the ideal feed trajectory. Scalars denote the control parameters used in the RMAC method. The deriving process of Eq. (2) in detail is given in Appendix A.
where , , and denote the mass, damping, and stiffness matrix of the whole tool. The vector F is the external force related to the cutting process, refer to Sec. 2.2. Hence, time-delay effects due to the regenerative cutting are involved in this system. Assuming that the structural damping of the tool is Rayleigh damping, the damping matrix is directly obtained by the linear combination of and : , where are the corresponding mass and stiffness coefficients. X denotes the vector consisting of all the nodal displacements and rotation angles. Assuming that the control (the RMAC method) applied to the joints has already been stable, and the rotation angle of the joints and the deformation of the tool are all small. Then, X can approximately have the following expression: , where denote the error of the displacement and rotation angle of the ith node, respectively. n is the number of elements. Specifically, the subscript “1” denotes the node at the end of the tool ( in Fig. 1, where the spindle and the tool are connected), subscript “” denotes the node at the tool tip ( in Fig. 1).
Here, we assume that during the milling process, changes in the elements of the matrix and due to the changes in q are small and can be ignored; therefore, the matrices , and remain to be constant. Here, assuming that the two-dimensional cutting model is applied, then F is formulated as: , where denote the milling force applied on the tool tip, i.e., the node , in x and y direction, whose expression in detail is derived in Sec. 2.2.
2.2 Cutting Force Model.
A classic two-dimensional cutting force model applied on the tool tip [33] is shown in Fig. 2. The coordinate frame , i.e., the tool coordinate frame, is fixed on the tool tip, i.e., the node , where denote the axis along the feed direction, perpendicular to the feed direction, and along the tool axis, respectively. a denotes the axial depth of cut, denotes the radial depth of cut, and v denotes the feed velocity of the workpiece relative to the cutting tool. is the diameter of the tool. denotes the spindle speed (unit: rotations per minute (rpm)). represents the angular position of the tooth j. For simplicity, assuming that the cutting tool has equally distributed cutting teeth, resulting in a constant pitch angle, then .
represent the tangential and normal cutting force applied on the jth tooth, and denotes the chip thickness of the jth tooth; by using the linear cutting force model and the classic circular tool path model [34], the cutting force components are formulated as: , , and , where and are tangential and normal cutting force coefficients, respectively. T denotes the tooth passing period: , which is also the delay of the dynamic system.
In Eq. (8), , and are constant matrices of size , and are periodic matrices of size : , and is periodic vector of length : . is the delay of the system which is equal to the period T.
As the number of elements n increases, the DOF of Eq. (8) can be arbitrarily high, allowing the evaluation of the performance of the SLD prediction algorithm applied to DDE with high DOFs. All the parameters used to derive Eq. (8), including the parameters of the robotic arm, the flexible cutting tool, and the cutting force are all listed in Appendix B.
3 Chebyshev Collocation Method and the Evaluation of Computational Efficiency
which is a DDE only corresponding to the perturbation .
3.1 Chebyshev Collocation Method.
where is the index of the cutting tooth such that the absolute value reaches the minimum. Then, let , which represents the ratio of the arc length of the cutting zone to the pitch angle. When , the coefficients are sufficiently smooth over , which means that the original CCM is valid, so only cases when need to be considered. (For convenience, the situation when is not considered in this article, but according to Ref. [25], generalization to that case is straightforward.)
where denotes the block of including the first N rows and columns, while denotes the block of including the first N rows and the last column. Notice that is a matrix that only depends on the geometry of the cutting tool, and the period T, which only depends on the spindle speed .
In practical application, the inversion of the matrix can be avoided by solving the generalized eigenvalue problem, i.e., finding such that the dominant .
3.2 Comparison Between Chebyshev Collocation Method and Full-Discretization Method.
and represent the number of discrete points along the axis of a and , respectively.
The algorithm of construction of SLD using CCM is shown in Algorithm 1. After performing the algorithm, the spectral radius of the transition matrix is obtained on all the mesh points. Then, the stability border where can be found, dividing the parameter space into stable and unstable zones.
for to do |
Calculate |
for to do |
Calculate , |
Solve the generalized eigenvalue problem |
end for |
end for |
for to do |
Calculate |
for to do |
Calculate , |
Solve the generalized eigenvalue problem |
end for |
end for |
To evaluate the computation efficiency of CCM, the commonly used FDM is chosen as the reference method in this article. Define the time spent on constructing the transition matrix and calculate the spectral radius of in single point when using CCM and FDM as and , respectively. Assuming that and remain the same regardless of the and , we do not use the time spent on constructing the whole SLD to assess the computation efficiency of CCM and FDM but and instead.
The DDE model built in Sec. 2 is adopted to apply the algorithm. The other necessary parameters are . All codes of the algorithm are written with matlab 2023a and implemented on the same computer (12th Gen Intel(R) Core(TM) i5-12400 CPU @2.50 GHz 16.0 GB RAM, Windows 11 OS). When applying FDM, the method to improve the computation efficiency proposed in Ref. [12] is adopted including reducing the dimension of the transition matrix and utilizing the sparse matrix multiplying. For reliability of the results, we run the algorithm proposed here and in the remains of this article ten times and use the average value of the running time as the final result.
The result of and in different system DOF l and the number of discrete segments N are shown in Fig. 3. The ratio of the computation time of the single point and distribution of whether CCM is faster in the parameter space is also shown in Fig. 4. The computation time for both FDM and CCM increases with the DOF l or number of segments N. In most cases except for , the CCM computes slower than FDM. The ratio of the computation time also increases with the increase of l or N, when , the ratio is 54.639, meaning that CCM is almost 50 times slower than FDM. The main reason for this inefficiency lies in the enormous difference in the dimension of the transition matrix when applying two kinds of methods. In FDM, according to Ref. [19], the final dimension of the transition matrix is , while in CCM it is . When the DOF of the system l is high, i.e., , then , resulting in a huge computation cost when solving the (generalized) eigenvalue problem compared to the FDM.
Hence, to improve the computation efficiency of CCM, the first attempt is to reduce the dimension of the transition matrix , which is one of the main topics of Sec. 4.
4 The Accelerated Algorithm: Fast Chebyshev Collocation Method
Based on the traditional CCM, an accelerated algorithm named FCCM is proposed in this section. The acceleration approach includes the reduction of the dimension of the transition matrix and the acceleration of the matrix inversion. The convergence rate and the computation efficiency of the proposed FCCM are then analyzed and compared to the FDM, referring to Secs. 4.1–4.4.
4.1 Dimension Reduction of the Transition Matrix.
Now is an size matrix that has almost the same dimension as the transition matrix constructed in SDM or FDM when the number of the discrete segments N are equal. The new algorithm constructing the SLD is shown as below.
for to do |
Calculate |
for to do |
Calculate |
Calculate the spectral radius |
end for |
end for |
for to do |
Calculate |
for to do |
Calculate |
Calculate the spectral radius |
end for |
end for |
Define the time spend on constructing and calculating for a single parameter pair is , similar to Sec. 3.2, then the results of and are shown in Fig. 5. The spindle speed , the depth of cut a, and other parameters used here and in the remainder of this section are all the same as in Sec. 3.2 unless otherwise noted. Although the dimension of the transition matrix has been reduced significantly, the time spent on calculating the spectral radius of is longer than that when using the original to calculate. When using Eq. (23) to calculate the spectral radius, the inversion of the matrix is avoided by solving the generalized eigenvalue problem instead. However, using Eq. (31) reintroduces the inversion. The time spent on inversion and solving the generalized eigenvalue problem has the same order corresponding to the dimension , which prevents the improvement of the computation efficiency. Therefore, effort must be taken to further accelerate the matrix inversion, which is discussed in Sec. 4.2.
4.2 Acceleration of Matrix Inversion.
Notice that , and only depend on the spindle speed , meaning that the matrices , , , , and can all be calculated only once outside the loop of the axial depth of cut a. Within the loop, the dimension of the matrix that needs to be inversed changes from lN to 2N, which is a significant reduction. The above improvement then leads to the following FCCM algorithm.
for to do |
Calculate , , , , and |
for to do |
Calculate |
|
Calculate the spectral radius |
end for |
end for |
for to do |
Calculate , , , , and |
for to do |
Calculate |
|
Calculate the spectral radius |
end for |
end for |
Now for every fixed spindle speed , the algorithm is divided into two parts: outside the loop of the depth of cut a, a decomposition for the matrix is performed once; while inside the loop, a decomposition of the matrix is performed for the construction of the transition matrix of dimension then following the calculation of the spectral radius for every a. We denote the time for the decomposition of as and the time spent on the spectral radius as . The comparison of and is shown in Fig. 6.
the derivation in detail is provided in Appendix C. Now, the inversion of follows two steps: first, calculate ; next, calculate . The main source of the computation cost changes from the decomposition of to that of , which is only half the size of .
Define the time for decomposing the matrix using Eq. (42) as , the comparison of and is shown in Fig. 7. Because the dimension of the matrix needed to be decomposed is only half of , the decomposition time should be approximately a quarter of that of , i.e., . From Fig. 7, however, the actual decomposition time is shorter than predicted, meaning a better efficiency improvement. The possible reason for this enhancement is the rearrangement of the elements in the sparse matrix (denoted as in the remaining of this article) compared to . The distribution of the elements of and is shown in Fig. 8. In , most blocks are identity matrices, however, along the main diagonal are the dense matrix , where almost all the elements are no-zero. On contrast, all the blocks of are the linear combination of the matrices , , and , which retain the sparse band structure due to discretization using the finite element analysis. The density of the original is in a sense “equally distributed” to the other blocks in , possibly easing the burden of the sparse matrix solver applied in matlab, resulting in a further improvement of the computation efficiency not only due to the reduction of the dimension.
4.3 Convergence Analysis of Fast Chebyshev Collocation Method.
One of the important advantages of CCM is its spectral convergence rate, to demonstrate this superiority, the spectral radius of the transition matrix at the point is calculated (other parameter remains the same as Sec. 3.2) under different system DOF l and number of the discretized segments N. The absolute error of the spectral radius and the reference value is shown in Fig. 9. Where for every fixed l, is the value of when (for the bigger value, the computation meets the memory usage problem, so in this article, we just regard it as the reference value).
It is clear that when N is the same, the magnitude of the error of FCCM is often several orders higher than FDM. And when , the FCCM shows the apparent spectral convergence rate: when N reaches 90, the error shows a rapid decrease with the exponential rate, when , the error order reaches , near the machine precise, and when N gets larger, this order remains unchanged. However, when l gets larger, this property seems to gradually disappear, and the order that the error can finally reach get lower, when , the error order can only reach at last. Two possible reasons may be able to explain this decay of the spectral convergence rate: one is that when l gets larger, the numeric error introduced by the decomposition of the sparse matrix of quite a huge dimension significantly disturbs and influences the numeric property of the collocation method; therefore, the spectral convergence property itself disappears under extremely high DOF; another possibility is that the spectral convergence property still exists even with very high DOF, but N used in this article is not large enough to observe it. But anyhow, FCCM is still proven to be superior to FDM in terms of computation accuracy.
4.4 Comparison Between Fast Chebyshev Collocation Method and Full-Discretization Method.
where denotes the smallest integer that is greater than or equal to the value “.” The FCCM is faster than FDM for the fixed when exceeds , meaning that FCCM is more computationally efficient only when the mesh of the SLD is sufficiently dense. The value of under different is shown in Fig. 11. Fortunately, for most of the , is below 100 except for an abrupt peak at the up-right corner of the space, where the maximum reaches 382. By increasing , as shown in Fig. 12, the number of the points where FCCM is slower gets lower. In most situations, the FCCM shows superiority to FDM both in computation accuracy and efficiency even with high DOF.
5 Concluding Remarks
A novel fast Chebyshev collocation method is proposed to perform the stability analysis of delay differential equations with linear, periodic coefficients of the system with high DOF. As an application, a robotic milling model allowing arbitrarily high DOF is used to validate the performance of the proposed method. By reducing the dimension of the transition matrix and acceleration of the inversion of the corresponding matrix component when constructing the transition matrix, the proposed method reaches high computation efficiency under high DOF. Numeric experiments are implemented to evaluate the computation efficiency and accuracy, and a comparison with the widely used FDM is performed. The following conclusions are drawn:
In most cases considered in this article, the CCM is slower than FDM in terms of constructing the SLD. In the worst case, CCM is nearly 50 times slower than FDM, proving that CCM is not suitable for the stability analysis of systems with high DOF.
In all the cases considered in this article, the error of the spectral radius of the transition matrix is smaller than FDM, usually several orders higher. FCCM retains the spectral convergent of the original CCM under low DOF (). When DOF is higher, this spectral convergent seems to disappear. However, given that FCCM still has higher computation accuracy than FDM, it is a potential choice for the stability analysis under high system DOF.
In most cases considered in this article, the FCCM performs better than the original CCM and the FDM in computation efficiency. The values and which denote the time spent on constructing the SLD at a fixed spindle speed are used as the comparing criteria. With different DOF l and the number of the discrete segments N, the ratio is below one except in several cases when l and N are both high (, ). Increasing , the number of mesh points along the axis of the depth of cut a in the SLD does reduce the occurrence of cases where FCCM is slower than FDM.
Acknowledgment
The authors gratefully acknowledge the support received from the NSFC (12272218) and the Ministry of Science and Technology of China (2023YFF0713400).
Funding Data
NSFC (12272218; Funder ID: 10.13039/501100001809).
Ministry of Science and Technology of China (2023YFF0713400; Funder ID: 10.13039/501100002855).
Data Availability Statement
The datasets generated and supporting the findings of this article are obtainable from the corresponding author upon reasonable request.
Appendix A: Derivation of Eq. (4)
according to Ref. [32], when the absolute value of the elements of is small, we approximately have .
Appendix B: Parameters Used to Derive Eq. (8)
Number of joint | d (mm) | a (mm) | α (deg) |
---|---|---|---|
1 | 815 | 0 | 0 |
2 | 0 | 350 | −90 |
3 | 0 | 850 | 0 |
4 | 820 | 145 | −90 |
5 | 0 | 0 | 90 |
6 | 170 | 0 | −90 |
Number of joint | d (mm) | a (mm) | α (deg) |
---|---|---|---|
1 | 815 | 0 | 0 |
2 | 0 | 350 | −90 |
3 | 0 | 850 | 0 |
4 | 820 | 145 | −90 |
5 | 0 | 0 | 90 |
6 | 170 | 0 | −90 |
Number of joint | Mass m (kg) | Center of mass (xc, yc, zc) (mm) | Center inertia tensor (Ixx, Iyy, Izz) (kg m2) |
---|---|---|---|
1 | 121.09 | (140.890, 15.739, −123.659) | (2.577, 6.170, 6.725) |
2 | 182.99 | (401.195, −0.501, −248.760) | (17.503, 1.640, 17.070) |
3 | 126.74 | (60.468, −4.014, 13.270) | (5.235, 2.031, 4.968) |
4 | 87.06 | (−3.556, 0.12, −169.963) | (0.430, 0.430, 0.487) |
5 | 39.72 | (0.059, 32.403, 0.013) | (0.183, 0.120, 0.174) |
6 | 44.25 | (−3.177, 0.801, 113.311) | (0.247, 0.175, 0.136) |
Number of joint | Mass m (kg) | Center of mass (xc, yc, zc) (mm) | Center inertia tensor (Ixx, Iyy, Izz) (kg m2) |
---|---|---|---|
1 | 121.09 | (140.890, 15.739, −123.659) | (2.577, 6.170, 6.725) |
2 | 182.99 | (401.195, −0.501, −248.760) | (17.503, 1.640, 17.070) |
3 | 126.74 | (60.468, −4.014, 13.270) | (5.235, 2.031, 4.968) |
4 | 87.06 | (−3.556, 0.12, −169.963) | (0.430, 0.430, 0.487) |
5 | 39.72 | (0.059, 32.403, 0.013) | (0.183, 0.120, 0.174) |
6 | 44.25 | (−3.177, 0.801, 113.311) | (0.247, 0.175, 0.136) |
Quantities | Symbol | Value | Units |
---|---|---|---|
Coordinate of the end of the spindle in the frame of the last link of the robot | pend | (−200, 0, 200) | mm |
The direction cosine matrix of the end frame relative to the frame of the last link | Aend | — | |
The RMAC control parameter | kp | 108 | s−2 |
kd | s−1 |
Quantities | Symbol | Value | Units |
---|---|---|---|
Coordinate of the end of the spindle in the frame of the last link of the robot | pend | (−200, 0, 200) | mm |
The direction cosine matrix of the end frame relative to the frame of the last link | Aend | — | |
The RMAC control parameter | kp | 108 | s−2 |
kd | s−1 |
Quantities | Symbol | Value | Units |
---|---|---|---|
Number of the cutting teeth | Ntool | 4 | — |
Diameter of the cutting tool | Dtool | 25 | mm |
Length of the cutting tool | L | 121 | mm |
Young's modulus of elasticity | E | 231.932 | GPa |
Poisson ration | μ | 0.324 | — |
Shear correction factor | k | — | |
Density of the tool | ρ | 8563.55 | kg/m3 |
The mass coefficient of the Rayleigh damping | αM | 35.372 | — |
The stiffness coefficient of the Rayleigh damping | αK | 2.061 × 10−10 | — |
The direction cosine matrix of the local frame of the element relative to the end frame | T | — | |
Tangential coefficient of the cutting force | Kt | 1835 | MPa |
Normal coefficient of the cutting force | Kr | 1137 | MPa |
Quantities | Symbol | Value | Units |
---|---|---|---|
Number of the cutting teeth | Ntool | 4 | — |
Diameter of the cutting tool | Dtool | 25 | mm |
Length of the cutting tool | L | 121 | mm |
Young's modulus of elasticity | E | 231.932 | GPa |
Poisson ration | μ | 0.324 | — |
Shear correction factor | k | — | |
Density of the tool | ρ | 8563.55 | kg/m3 |
The mass coefficient of the Rayleigh damping | αM | 35.372 | — |
The stiffness coefficient of the Rayleigh damping | αK | 2.061 × 10−10 | — |
The direction cosine matrix of the local frame of the element relative to the end frame | T | — | |
Tangential coefficient of the cutting force | Kt | 1835 | MPa |
Normal coefficient of the cutting force | Kr | 1137 | MPa |