Next Article in Journal
Kinematics, Speed, and Anthropometry-Based Ankle Joint Torque Estimation: A Deep Learning Regression Approach
Next Article in Special Issue
Improved Haptic Transparency of Bilateral Control Using Torque-Measured Magnetic Coupling
Previous Article in Journal
Synchronization Control of a Dual-Cylinder Lifting Gantry of Segment Erector in Shield Tunneling Machine under Unbalance Loads
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Iterative Parameter Optimization for Multiple Switching Control Applied to a Precision Stage for Microfabrication

1
Department of Mechanical Engineering, National Taiwan University, Taipei 10617, Taiwan
2
Department of Mechanical Engineering, National Taiwan University of Science and Technology, Taipei 10617, Taiwan
*
Author to whom correspondence should be addressed.
Machines 2021, 9(8), 153; https://doi.org/10.3390/machines9080153
Submission received: 9 July 2021 / Revised: 21 July 2021 / Accepted: 30 July 2021 / Published: 3 August 2021
(This article belongs to the Special Issue Design and Control of Electrical Machines)

Abstract

:
This paper proposes an iteration procedure to derive optimal parameters for a multiple switching control architecture. Control design is usually a compromise between various performance requirements; therefore, switching between multiple controllers that achieve a particular performance under different conditions can potentially improve the overall system behavior. In this paper, we consider a control-switching mechanism that can automatically switch controllers based on the prediction of future responses, and we develop an iteration procedure that can optimize the mechanism parameters, such as the number of controllers and the prediction horizon. We then implement the proposed mechanism in a long-stroke precision stage, and demonstrate the effectiveness of switching robust control with simulations and experiments. Lastly, we integrate the stage with a two-photon polymerization system to fabricate microlenses. The optical properties confirm that the proposed iterative parameter optimization procedure is effective in improving the performance of microfabrication employing multiple switching control.

1. Introduction

Control design is a compromise between various performance requirements. For example, controllers that provide fast transient responses tend to have large overshoots, while controllers that provide smooth responses tend to have long rise times. Therefore, switching controllers can potentially achieve better performance than can be achieved using a single controller. For instance, Solihin et al. [1] designed a fuzzy-tuned proportional integral derivative (PID) controller for a gantry crane. They updated the PID parameters according to the errors to accomplish smaller settling times and overshoots than were obtained with the original PID control. Bashash et al. [2] designed two controllers for a piezoelectric transducer (PZT) stage and switched them based on the derivatives of position errors. Qin et al. [3] designed a fuzzy adaptive PID controller for a fuel cell power plant and updated the PID parameters using fuzzy logic algorithms. Xu et al. [4] designed a fuzzy PID controller for marine vessels, where the PID coefficients were automatically adjusted by positioning accuracy. Armaghan et al. [5] designed two PID controllers and a switching logic for a magnetically driven system. Asl et al. [6] proposed a fuzzy switching control, which fused a PID controller and a linear quadratic regulator, for a unicycle robot. Rana et al. [7] applied model predictive control to improve the high-speed imaging performance of an atomic force microscope. Wang et al. [8] applied an integral control with tunable gains to a PZT stage, where the integrator eliminated steady-state errors and the varying gains provided better performance than were obtained using a fixed gain. The idea was extended in [9] by switching two robust controllers based on the prediction of future responses. Wang et al. [10] further proposed multiple-switch robust control, which considered multiple control sequences and obtained more performance benefits than were attained with a single switching control. In this paper, we propose an iterative algorithm to tune the mechanism parameters for further improvement of system performance, because the parameters of the multiple control mechanism can significantly influence system performance and computing loads.
The stability and performance issues caused by the initial value problem in switching controls have been considered in much research. For example, Nichols et al. [11] interpolated the linear controller parameters among four H robust controllers to remove the hidden coupling terms. Yamaguchi et al. [12] proposed initial value compensation to control a hard disk driver. Because control reduction can reduce the initial value problems, Wang et al. [10] simplified high-order robust controllers into robust PID controllers that could achieve similar responses to the original controllers. In this paper, we further reduce the robust controllers to robust proportional–integral (PI) controllers, thereby reducing the computing loads in the multiple switching control mechanism.
The iterative parameter optimization procedures were then applied to a long-stroke precision positioning stage employing multiple switching control. This stage consists of a PZT stage and a motor stage, where the PZT stage achieves nanometer-level precision and the motor stage provides large displacement of up to 10 cm. The PZT stage was frequently applied for precision positioning because of its high resolution and large driving forces. However, the displacements of PZTs were usually limited. Therefore, large-travel actuators, such as linear motors, were normally integrated with the PZT stages to increase their working ranges. For instance, Wang et al. [8] integrated a motor stage and a PZT stage to achieve precision positioning for large travels. Hossain and Rahman [13] designed a triple stage, which consisted of a voice coil motor, a PZT, and a thermal flying-height control. Because the hysteretic effects of PZTs might degrade system performance, advanced control methods are usually applied to improve system performance. For example, Zhu and Rui [14] designed a PID control with an inverse generalized Bouc–Wen model to improve the precision positioning of a PZT actuator. Wang et al. [8] proposed an integral control with gain scheduling to a PZT stage. Saleem et al. [15] applied a Bouc–Wen model and particle swarm optimization (PSO) to compensate for the hysteretic characteristics. Gan and Zhang [16] developed a generalized Bouc–Wen model to characterize the rate-dependent hysteresis of PZTs. Fang et al. [17] developed a modified Bouc–Wen model and designed a fuzzy PID control with an inverse model feedforward compensator for a PZT stage. Wang et al. [10] proposed an automatic switching control for a PZT stage. Zhang et al. [18] designed an autonomous switching control to suppress the vibration of a double-beam system employing piezoelectric shunt damping. In this paper, we designed robust controllers for the PZT stage and optimized the switching parameters via the proposed iterative procedures. We then applied similar approaches to the motor stage. Because the motor stage model is a first-order type, its switching control can be further simplified to gain scheduling with feedforward control to reduce the computing loads. The two stages were then combined to demonstrate the system’s capacity for long-stroke precision positioning. Lastly, we integrated the combined stage with a two-photon polymerization (TPP) system, and showed the benefits of multiple switching control employing iterative parameter optimization in microfabrication.
The rest of this paper is arranged as follows: Section 2 introduces the multiple switching control structure and proposes an iterative procedure to optimize the structure parameters. Section 3 applies the iterative parameter optimization method to a large-stroke precision stage. The simulation and experimental results demonstrate the effectiveness of the proposed iterative parameter tuning. We further integrate the stage with a TPP system to fabricate microlenses and demonstrate the effectiveness of the proposed iterative parameter optimization in microfabrication. Lastly, we draw conclusions in Section 4.

2. Multiple Switching Control with Iterative Parameter Optimization

The multiple switching control architecture is shown in Figure 1, consisting of a feedback control loop and a response predictor. The former implements the selected controller to the system, while the latter picks the best controller that provides the optimal system responses among all potential control sequences. We propose an iterative algorithm to optimize the structure parameters, such as the number of controllers, the prediction horizon, and the switching steps.
Because control design is usually a compromise between various performance requirements, we can combine the merits of different controllers by switching them at appropriate instances. For example, consider a standard second-order system G ( s ) = ω n 2 / ( s 2 + 2 ξ ω n s + ω n 2 ) with ξ = 1   and   ω n = 6 ; the controller design is a compromise between various specifications, such as the rise time, the overshoot, and the steady-state error. For instance, a controller C 1 ( s ) = 3 / s can provide a short rise time with zero steady-state error, but with a large overshoot, while a controller C 2 ( s ) = 0.8 / s can achieve small overshoot and zero steady-state error, but with a long rise time. Therefore, we can combine the advantages of these two controllers by switching them at the right moments. For example, we can define the system cost as follows:
J = 1 H P k k + H p ( r ( k ) y ( k ) ) 2
where k is the current step and Hp is the future horizon. That is, the predictor calculates the future system responses from step k to step k + Hp, and the corresponding cost J, by all potential control sequences. Then, the optimal control sequence that minimizes J is selected to control the system. For example, we set Hp = 40 and derive the system response, as shown in Figure 2, where the switching control achieves short rising time, small overshoot, and small steady-state error simultaneously.
Because the performance of the switching control structure can be significantly influenced by the structure parameters, we propose a method to tune these parameters. As shown in Figure 1, the response predictor estimates future system responses by all possible control sequences, and selects the optimal control sequence that minimizes certain system costs. Suppose that there are N controllers, and the number of controller switching steps is Sp; the number of possible control sequences is then N S P . For instance, suppose that two controllers—C1 and C2 (i.e., N = 2)—have two switching steps (i.e., Sp = 2); there will be 22 = 4 possible control sequences, as follows: (C1, C1, C1, …, C1), (C1, C2, C2, …, C2), (C2, C1, C1, …, C1), (C2, C2, C2, …, C2). Therefore, the predictor needs to calculate the system responses employing these four control sequences, and then choose the optimal control sequence that minimizes certain performance indices.
Although the control parameters can significantly influence the system performance, simultaneous optimization of these parameters is impractical because of the hardware computing loads. For example, the predictor needs to estimate the future system responses in the next Hp steps with N S P control sequences. Therefore, we propose an iterative method that can tune these parameters one by one at each iteration until they converge. For example, suppose that we have three structure parameters ( N , H P , S P ) , where N is the number of controllers, Hp is the prediction horizon, and Sp is the number of switching steps. The proposed iteration procedures are illustrated as follows (see Figure 3):
  • Set the default parameters ( N , H P , S P ) ;
  • Apply ( N , H P , S P ) to derive an optimized N, labelled as N o p t , which can improve system performance without exceeding hardware computing limits;
  • Apply ( N o p t , H P , S P ) to derive an optimized H P , labelled as H P o p t , which can improve system performance without exceeding hardware computing limits;
  • Apply ( N o p t , H P o p t , S P ) to derive an optimized S P , labelled as S P o p t , which can improve system performance without exceeding hardware computing limits;
  • If ( N o p t , H P o p t , S P o p t ) = ( N , H P , S P ) , then the iteration is terminated, and the optimal parameters ( N o p t , H P o p t , S P o p t ) can be implemented by the multiple control structure. Otherwise, set ( N , H P , S P ) = ( N o p t , H P o p t , S P o p t ) and return to step 1.

3. Iterative Parameter Optimization for the Long-Stroke Precision Stage Employing Multiple Switching Control

This section applies the proposed iterative parameter tuning method to the long-stroke precision stage employing multiple switching control. This stage consists of a PZT stage and a motor stage, as illustrated in Figure 4. The specifications of the stage are illustrated in Table 1. We apply the switching control mechanism to both stages, and integrate them for simulations and experiments.

3.1. Multiple Switching Control for the PZT Stage

The PZT stage [19] has a travel range of 100 μm and is equipped with an encoder with a resolution of 1.22 nm. We applied PCI-6221 and PCI-6229 data acquisition (DAQ) cards [23] to measure the stage displacements, and to transmit the control signals within ±5 V. Because the operation voltage of the PZT was −20 to 120 V, we used an SVR/150/3 amplifier [20] to magnify the voltage signals 10-fold and offset 50 V to control the stage displacements between 0 and 100 μm.
The system model was derived from the following experiments: we applied a swept sinusoidal voltage signal V P and measured the stage displacements X P from the encoder. Given the system variation and uncertainties during operations, we repeated the procedures 10 times and obtained the following transfer functions:
G P i ( s ) = T v P X P ,    i = 1 ,   2 ,   ,   10
A nominal plant was selected for the controller design. We assume that a nominal plant G P o has the following left coprime factorization [24]:
G P o = M ˜ 1 N ˜ ,  
where M ˜ , N ˜ R H and M ˜ M ˜ * + N ˜ N ˜ * = I . Suppose that a perturbed plant G P Δ can be represented as:
G P Δ = ( M ˜ + Δ M ˜ ) 1 ( N ˜ + Δ N ˜ ) ,  
in which Δ M ˜ , Δ N ˜ R H . The gap between the nominal plant G P o and the perturbed plant G P Δ is defined as in [25]. The smallest value of [ Δ M ˜ Δ N ˜ ] that perturbs G P o into G P Δ is denoted as δ ( G P o , G P Δ ) . Based on the gap analyses, the following G P 1 was selected as the nominal plant G P o , because it minimized the maximum gaps between models:
G P o = arg { ε = min G P o max G P i δ g ( G P 0 , G P i ) } ,   i = G P 1 ( s ) = 1517 s 3 + 1.75 10 5 s 2 + 6.053 10 8 s + 1.778 10 10 s 4 + 292.6 s 3 + 4.142 10 5 s 2 + 8.282 10 7 s + 2.064 10 9
We designed the following three robust loop-shaping controllers [26] for G P o :
C F = 67.3 s 6 + 4.19 × 10 4 s 5 + 1.53 × 10 10 s 3 + 2.73 × 10 12 s 2 + 1.97 × 10 14 s + 3.79 × 10 15 s 7 + 923 s 6 + 6.50 × 10 5 s 5 + 3.59 × 10 8 s 4 + 7.48 × 10 10 s 3 + 5.45 × 10 12 s 2 + 1.04 × 10 14 s C M = 34.8 s 6 + 1.81 × 10 4 s 5 + 1.67 × 10 7 s 4 + 6.21 × 10 9 s 3 + 7.40 × 10 11 s 2 + 3.26 × 10 13 s + 4.63 × 10 14 s 7 + 883.2 s 6 + 6.19 × 10 5 s 5 + 3.36 × 10 8 s 4 + 6.35 × 10 10 s 3 + 3.60 × 10 12 s 2 + 5.92 × 10 13 s C S = 12.1 s 6 + 5912 s 5 + 5.77 × 10 6 s 4 + 2.00 × 10 9 s 3 + 2.42 × 10 11 s 2 + 1.07 × 10 13 s + 1.52 × 10 14 s 7 + 697 s 6 + 5.49 × 10 5 s 5 + 2.56 × 10 8 s 4 + 4.20 × 10 10 s 3 + 2.34 × 10 12 s 2 + 3.87 × 10 13 s
where C F ,   C M ,   and   C S provide fast, intermediate, and smooth responses, respectively, for the PZT stage. The control design processes are illustrated in Appendix A, in which the stability margins of all controllers are greater than the system gap [ Δ M ˜ Δ N ˜ ] ; therefore, internal stability can be guaranteed during operation.
Because the designed robust controllers are seventh-order controllers, this might increase the computing loads and cause initial value problems. Therefore, we also apply the PSO algorithms [27] to approximate these controllers as the following robust PI controllers:
C ¯ F = 0.05 + 36.97 s ,   C ¯ M = 0.08 + 8.01 s ,   C ¯ S = 0.05 + 3.94 s  
The design of the robust PI controllers is shown in Appendix B. The system responses by these controllers are illustrated in Figure 5 and Table 2, where the robust PI controllers provide system responses similar to those of the high-order robust controllers, but with much simpler forms. That is, the fast controllers C ¯ F provide faster responses (i.e., shorter rise time and settling time), but with larger overshoots than are achieved with the other controllers. Conversely, the smooth controllers C ¯ S give smooth responses (i.e., without overshoot), but with larger rise times and settling times than are seen for the other controllers. The behaviors of the intermediate controllers C ¯ M fall between the fast controllers and the smooth controllers. We further add an extra controller C ¯ 4 = 0.07 + 23.01 / s with characteristics between C ¯ F   and   C ¯ M . We will demonstrate the possibility of achieving the merits of these controllers simultaneously with a switching control employing the iterative optimization procedures.
We applied the iteration procedures (see Figure 3) to derive the optimal parameters for the switching control mechanism. The processes are shown in Appendix C, where the optimal parameters were ( N o p t = 3 ,   H P o p t = 20 ,   S P o p t = 2 ) . Therefore, we implemented these parameters in the PZT stage employing the multiple control structure, and we illustrate the system responses in Figure 6. At t = 1 s, the stage began to track a step command of 10 μm with the fast controller C ¯ F . Based on the response prediction, the switching mechanism switched the controller to C ¯ S at t = 1.003 s, and to C ¯ M at t = 1.004 s. Finally, the fast controller C ¯ F was chosen at t = 1.02 s to track the step command. As shown in Figure 6, the switching control combined the merits of these controllers to achieve a fast and smooth response.

3.2. Switching Control for the Motor Stage

The motor stage [21,22] has a travel range of 10 cm, and is equipped with an encoder with a resolution of 0.1 μm. Similarly, we derived its transfer functions from experiments, and selected the nominal plant, as follows:
G M o = 0.1 s  
for control design. Because G M o is first-order, we only need to apply a zero-order control K P to achieve arbitrary pole placement. Therefore, we designed a gain-scheduling control with command feedforward, as shown in Figure 7, for the motor stage.
The gain scheduling control is a stepless switching control based on tracking errors, while the command feedforward can reduce tracking errors for varying inputs, such as ramp or sinusoidal inputs. The gain-scheduling control can be represented as C(s) = Kp, where Kp is adjusted by the following equation:
  K P = { 1600 ,   if   | e M | 50 μ m 30 | e M | + 100 ,   if   10 < | e M | < 50 μ m 400 ,   if   | e M | 10 μ m  
We applied the motor stage for tracking the ramp and sinusoidal inputs. The results are shown in Figure 8 and Table 3. First, Figure 8a,b shows the tracking responses to the ramps of 100 μm/s and 500 μm/s, respectively. The gain-scheduling control was set as Kp = 400, because the tracking errors were less than 10 μm. Compared to the control used in [9], the RMSE was decreased from 2.4797 μm to 0.2238 μm for the ramp of 100 μm/s, and from 8.7794 μm to 0.7851 μm for the ramp of 500 μm/s. Second, the tracking responses for sinusoidal inputs of 0.1 Hz and 1 Hz are shown in Figure 8c,d, respectively. Similarly, the gain-scheduling control was set as Kp = 400, because the tracking errors were less than 10 μm. Compared to the control used in [9], the RMSE was reduced from 0.5609 μm to 0.1355 μm for r ( t ) = sin ( 0.2 π t ) , and from 5.5460 μm to 0.4161 μm for r ( t ) = sin ( 2 π t ) .

3.3. The Combined Stage

We integrated the PZT and the motor stage with the control structure, as shown in Figure 9, where the PZT stage applies the multiple switching control with ( N = 3 ,   H P = 20 ,   S P = 2 ) , while the motor stage employs the gain-scheduling control with command feedforward. The response predictor also applies a similar layout to estimate the stage responses by all possible control sequences. The predictor then calculates the corresponding system costs and selects the optimal control sequence for the PZT stage.
The motor stage applies gain-scheduling control based on the current positioning error e M ( k ) = r ( k ) - X M ( k ) , where r ( k ) is the command and X M ( k ) is the motor stage position. Considering that the PZT stage has a travel limit of 50 μm, we set the following anti-lock function:
e P ( k ) = { 0 ,   if   | e ( k ) | 50   μ m r ( k ) - y M ( k ) - y P ( k ) ,   if   | e ( k ) | < 50   μ m  
where   e P ( k ) is the error of the PZT stage and e ( k ) = r ( k ) - X M ( k ) - X P ( k ) is the error of the combined stage. The PZT stage provides precision positioning using multiple switching control to regulate the position errors of the combined stage. Because the optimal parameters for the PZT stage are N = 3 and Sp = 2, the predictor estimates the system responses and calculates the corresponding system costs with all nine ( N S P = 9 ) possible control sequences; it then chooses the optimal one to regulate the PZT stage.
For experimental verification, we applied the combined stage to track the word “NTU” with a scale of 100 mm × 50 mm, as shown in Figure 10a. The experimental results are illustrated in Figure 10 and Table 4. Compared with previous works [8,9], the RMSE was reduced, because the multiple switching control can quickly adjust the tracking errors at the turning points, as shown in Figure 10c,d. For example, the switching mechanism switched the x-axis controller of the PZT stage from C ¯ F to C ¯ M at t = 331.990 s, and from C ¯ M to C ¯ S at t = 332.060 s. Finally, the intermediate controller C ¯ M was chosen at t = 332.080 s to complete the turning. Similarly, the switching mechanism switched the controller from C ¯ F to C ¯ S at t = 564.008 s, from C ¯ S to C ¯ M at t = 564.023 s, and from C ¯ M to C ¯ S at t = 564.033 s to complete tracking the turning point. Compared with previous studies [8,9], parameter optimization of the switching control led to significantly improved system performance.

3.4. Microfabrication by Two-Photon Polymerization

We integrated the combined stage with a TPP system, as shown in Figure 11 to fabricate a microlens. An adaptor with a microscope slide was connected to the PZT stage, and the laser was projected to fabricate microstructures by hardening the materials (OrmoComp) on the microscope slide. Note that the adaptor and the microscope slide were put on the stage when we identified the PZT stage models in Equation (2). Furthermore, the model variation caused by the loading effects can be neglected because the adaptor is lightweight. We demonstrate the effectiveness of the proposed control method using the images and optical properties of the microlens.
The Fresnel zone plate (FZP) is a planar microlens designed to focus light. We designed an FZP with a diameter of 128 μm, as follows [28]:
r n = n f λ + 1 4 n 2 λ 2
where n is the number of circles, rn is the radius of the n-th circle, λ is the wavelength in μm, and f is the focal length of the lens in μm. For example, setting n = 13, λ = 632.8 nm, and f = 500 μm, the radius is calculated as r n = 18 ,   25 ,   ,   64   μ m when n = 1 ,   2 ,   ,   13 . The design and fabrication results are shown in Figure 12.
We evaluated the optical quality of the FZP by the light intensity and sharpness detected on the CMOS images. The FZP focused light onto a CMOS camera, as shown in Figure 13a, where the CMOS image was gray, scaled as black and white. Every pixel had a value of 0–255 to represent its brightness, as shown in Figure 13b, where 0 and 255 represented an image that was completely black and completely white, respectively. We analyzed the intensity of the image brightness along the red line, as shown in Figure 13c. The light sharpness was defined as the derivative of the intensity, as illustrated in Figure 13d. The tracking RMSE of the combined stage and the optical properties of the microlens are shown in Table 5. Compared with previous reports [8,9], the optical qualities of the microlens were greatly improved. The proposed multiple switching control employing iterative parameter tuning is therefore deemed effective in improving microfabrication.

4. Conclusions

This paper proposed an iteration procedure that optimizes the structure parameters for multiple switching control. Because control design is a compromise between various performance requirements, the merits of different controllers can potentially be achieved by switching them at appropriate moments. We introduced a multiple switching control structure and proposed an iteration method to optimize the structure parameters, such as the number of controllers and the prediction horizon. The iterative method began with a default set of parameters, which were iteratively tuned until they converged. We then applied the proposed iterative optimization method to a long-stroke precision stage, and demonstrated its effectiveness via simulations and experiments. We further integrated the stage with a TPP system to fabricate a microlens and evaluate its optical properties. The results show that the proposed iterative method for multiple switching control is effective in improving the performance of microfabrication systems. The developed iteration parameter optimization method for multiple switching control is a general control technique, and can also be applied to other control systems—such as atomic force microscopes [29,30], air-bearing planar stages [31], and visual-servo systems [32]—to improve system performance.

Author Contributions

Conceptualization, F.-C.W.; methodology, F.-C.W. and J.-F.L.; software, J.-F.L.; validation, F.-C.W.; formal analysis, F.-C.W. and J.-F.L.; investigation, F.-C.W. and J.-F.L.; resources, F.-C.W., T.-T.C. and J.-Y.Y.; data curation, F.-C.W. and J.-F.L.; writing—original draft preparation, F.-C.W. and J.-F.L.; writing—review and editing, F.-C.W.; visualization, F.-C.W. and J.-F.L.; supervision, F.-C.W.; project administration, F.-C.W., T.-T.C. and J.-Y.Y.; funding acquisition, F.-C.W. and J.-Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Ministry of Science and Technology of Taiwan with Grand MOST 109-2221-E-002-146.

Institutional Review Board Statement.

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

This work was financially supported by the Ministry of Science and Technology of Taiwan with Grand MOST 109-2221-E-002-146. This research was also financially supported in part by the Ministry of Science and Technology of Taiwan (Grands MOST 107-2634-F-002-018, MOST 108-2634-F-002-016-, MOST 109-2634-F-002-027-) and National Taiwan University, Center for Artificial Intelligence and Advanced Robotics. The authors would like to thank Ming-Hsiang Chang for helping with paper submission.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study, in the collection, analyses, or interpretation of data, in the writing of the manuscript, or in the decision to publish the results.

Appendix A. Robust Control Design for the PZT Stage

We applied loop-shaping techniques (see Figure 2 of [10]), to design robust controllers for the PZT stage. The principles of loop shaping are:
  • Increasing the loop gains at low frequencies for disturbance rejection;
  • Decreasing the loop gains at high frequencies for noise attenuation;
  • Smoothing the magnitude slopes near the crossover frequency for stability consideration.
The weighting functions were iteratively adjusted and verified by the system performance by simulations and experiments. Finally, we selected the following weighting functions for the PZT stage:
W F = 50 ( s + 40 π ) s ( s + 30 π ) ,   W M = 20 ( s + 15 π ) s ( s + 30 π ) ,   W S = 15 ( s + 10 π ) s ( s + 40 π )
where the subscripts F, M, and S represent fast, intermediate, and smooth, respectively. The corresponding robust controllers were designed as follows:
K F = 1.773 s 5 + 792.5 s 4 + 8.101 s 3 + 2.597 10 8 s 2 + 2.464 10 10 s + 5.213 10 11 s 5 + 874.1 s 4 + 5.619 10 5 s 3 + 3.212 10 8 s 2 + 4.03 10 10 s + 9.243 10 11 , K M = 1.773 s 5 + 792.5 s 4 + 8.101 s 3 + 2.597 10 8 s 2 + 2.464 10 10 s + 5.213 10 11 s 5 + 874.1 s 4 + 5.619 10 5 s 3 + 3.212 10 8 s 2 + 4.03 10 10 s + 9.243 10 11 , K S = 1 . 194 s 5 + 466 . 6 s 4 + 5 . 247 10 5 s 3 + 1 . 47 10 8 s 2 + 1 . 056 10 10 s + 1 . 96 10 11 s 5 + 575 . 3 s 4 + 4 . 67 10 5 s 3 + 1 . 964 10 8 s 2 + 1 . 333 10 10 s + 2 . 341 10 11 .
Then, the weighted controllers C F = W F K F ,   C M = W M K M ,   and   C S = W S K S were implemented to the plant.

Appendix B. Robust PI Control Design by the PSO Algorithms

We applied PSO techniques to derive robust PI controllers, because the standard robust controllers are high-ordered, and might increase difficulties in implementation.
PSO is a stochastic optimization technique [27] that simulates the social behavior of birds and fish in order to effectively solve multidimensional optimization problems, especially for nonlinear systems. Figure A1 shows the concept of a particle swarm optimization algorithm, where X i k and V i k represent the particle’s current velocity and position, respectively. Pbest i is the historical best position of the i-th particle, while Gbest is the best position among the swarms. Each particle updates its position and velocity through the following equations:
V i k + 1 = w × V i k + c 1 × rand 1 × ( Pbest i X i ) + c 2 × rand 2 × ( Gbest X i ) , X i k + 1 = X i k + V i k + 1 .
where w is the inertia weighting function for the velocity, c 1 and c 2 are learning factors, and rand1 and rand2 are random numbers between 0 and 1.
Figure A1. Particle swarm optimization.
Figure A1. Particle swarm optimization.
Machines 09 00153 g0a1
We applied the PSO techniques to derive the following PI controllers:
C P I ( s ) = K P + K I s
which can provide similar responses to the standard robust controllers. We define the fitness function as follows:
F ( K P , K I ) = i = 1 n ω i J ¯ i = i = 1 n ω i ( J i P I J i R o b u s t 1 ) 2
where J i P I and J i R o b u s t represent the performance indices employing the PI and robust controllers, respectively. J i ¯ indicates the similarity between the PI controller and the robust controller in terms of J i , while ω i represents the weighting of J i . We set n = 5 and consider the following five indices:
  • Stability margin: J 1 = b ( G P 0 , C P I ) ;
  • Root-mean-square error: J 2 = ( ( 1 1.2 | e ( t ) | 2 d t ) / 0.2 ) 1 / 2 ;
  • Settling time: J 3 = the settling time to a step input;
  • Overshoot: J 4 = percentage overshoot of a step response;
  • Rising time: J 5 = rising time to a step input.
We applied 100 particles with random initial values and set w = 0.4 ,   c 1 = 0.4 ,   c 2 = 0.4 with 50 iterations. The PSO algorithms iteratively update the parameters. Finally, the fast robust PI controller is designed as:
C ¯ F ( s ) = K P F + K I F s = 0.05 + 36.97 s
Similarly, the intermediate and the smooth PI controllers are derived as follows:
C ¯ M ( s ) = 0.08 + 8.01 s ,   C ¯ S ( s ) = 0.05 + 3.94 s
These controllers ( C ¯ F ,   C ¯ M ,   and   C ¯ S ) are implemented for multiple switching control. We compare the system responses by these robust PI controllers with the standard robust controllers ( C F ,   C M ,   and   C S ), as illustrated in Table A1. First, all J ¯ i 0 except for J ¯ 1 , indicating that the robust PI controllers provide similar time-domain characteristics to the higher-order robust controllers. Second, the robust PI controllers cannot achieve the original stability margins, because these robust controllers were designed to provide the maximum stability margins. However, system stability can still be guaranteed, because the stability margins of robust PI controllers are much larger than the system gap (0.0039).
Table A1. The performance costs for robust PI controller design.
Table A1. The performance costs for robust PI controller design.
J 1 ¯ J 2 ¯ J 3 ¯ J 4 ¯ J 5 ¯
C F C ¯ F 0.4194.476 × 10−52.100 × 10−23.000 × 10−35.100 × 10−5
C M C ¯ M 0.5204.300 × 10−32.000 × 10−35.371 × 10−51.200 × 10−2
C S C ¯ S 0.6018.000 × 10−36.573 × 10−404.000 × 10−2

Appendix C. Iterative Parameter Optimization for the PZT Stage

First, we set ( N , H P , S P ) = ( 4 , 40 , 4 ) at step 1. At step 2, we apply H P = 40   and   S P = 4 to optimize N; the results are shown in Figure A2a, where J = 1.557, 1.489, and 1.477 (μm) when N = 2, 3, and 4, respectively. Therefore, we set N o p t = 3 , because it provides similar performance to N = 4, but with less computing load. At step 3, we apply N o p t = 3 and S P = 4 to optimize H P ; the results are shown in Figure A2b, where J = 1.617, 1.489, and 1.489 (μm) when H P = 5, 20, and 40, respectively. Hence, we set H P o p t = 20 , as this provides the best performance with a lesser prediction horizon. At step 4, we apply N o p t = 3 and H P o p t = 20 to optimize S P ; the results are shown in Figure A2c, where J = 1.860, 1.489, and 1.489 (μm) when S P = 1, 2, and 3, respectively. Therefore, we set S P o p t = 2 , because it provides the best performance with less computing load. At step 5, we notice that ( N o p t = 3 ,   H P o p t = 20 ,   S P o p t = 2 ) ( N = 4 ,   H P = 40 ,   S P = 4 ) ; therefore, we set ( N = 3 ,   H P = 20 ,   S P = 2 ) and return to step 1 to repeat the procedures. At the second iteration, the optimal parameters are found to be ( N o p t = 3 ,   H P o p t = 20 ,   S P o p t = 2 ) , which is the same as ( N = 3 ,   H P = 20 ,   S P = 2 ) , as illustrated in Table A2.
Figure A2. Iterative parameter tuning: (a) N optimization; (b) HP optimization; (c) SP optimization.
Figure A2. Iterative parameter tuning: (a) N optimization; (b) HP optimization; (c) SP optimization.
Machines 09 00153 g0a2
Table A2. Iterative parameter tuning for the PZT stage.
Table A2. Iterative parameter tuning for the PZT stage.
Initial Setting ParametersOptimal ParametersCosts
Fitst iteration H P = 40 ,   S P = 4 N o p t = 3 N = 2, J = 1.557 μm
N = 3, J = 1.489 μm
N = 4, J = 1.477 μm
S P = 4 ,   N o p t = 3 H P o p t = 20 Hp =5, J = 1.617 μm
Hp = 20, J = 1.489 μm
Hp = 40, J = 1.489 μm
N o p t = 3 ,   H P o p t = 20 S P o p t = 2 Sp = 1, J = 1.860 μm
Sp = 2, J = 1.489 μm
Sp = 3, J = 1.489 μm
Second iteration H P = 20 ,   S P = 2 N o p t = 3 N = 2, J = 1.557 μm
N = 3, J = 1.489 μm
N = 4, J = 1.477 μm
S P = 2 ,   N o p t = 3 H P o p t = 20 Hp =5, J = 1.617 μm
Hp =20, J = 1.489 μm
Hp =40, J = 1.489 μm
N o p t = 3 ,   H P o p t = 20 S P o p t = 2 Sp = 1, J = 1.860 μm
Sp = 2, J = 1.489 μm
Sp = 3, J = 1.489 μm

References

  1. Solihin, M.I.; Wahyudi; Legowo, A. Fuzzy-tuned PID anti-swing control of automatic gantry crane. J. Vib. Control 2010, 16, 127–145. [Google Scholar] [CrossRef]
  2. Bashash, S.; Saeidpourazar, R.; Jalili, N. Tracking control of time-varying discontinuous trajectories with application to probe-based imaging and nanopositioning. In Proceedings of the 2009 American Control Conference, St. Louis, MO, USA, 10–12 June 2009. [Google Scholar]
  3. Qin, Y.; Sun, L.; Hua, Q.; Liu, P. A fuzzy adaptive PID controller design for fuel cell power plant. Sustainability 2018, 10, 2438. [Google Scholar] [CrossRef] [Green Version]
  4. Xu, S.; Wang, X.; Yang, J.; Wang, L. A fuzzy rule-based PID controller for dynamic positioning of vessels in variable environmental disturbances. J. Mar. Sci. Technol. 2019, 25, 1–11. [Google Scholar] [CrossRef]
  5. Armaghan, S.; Moridi, A.; Sedigh, A.K. Design of a switching PID controller for a magnetically actuated mass spring damper. In Proceedings of the World Congress on Engineering, London, UK, 6–8 June 2011; Volume III. [Google Scholar]
  6. Asl, R.M.; Mahdoudi, A.; Pourabdollah, E.; Klančar, G. Combined PID and LQR controller using optimized fuzzy rules. Soft Comput. 2019, 23, 5143–5155. [Google Scholar]
  7. Rana, M.; Pota, H.R.; Petersen, I.R. MPC in high-speed atomic force microscopy. In Proceedings of the 2016 Australian Control Conference (AuCC), Newcastle, NSW, Australia, 3–4 November 2016. [Google Scholar]
  8. Wang, K.-A.; Peng, Y.-K.; Wang, F.-C. The Development and Control of a Long-Stroke Precision Stage. Smart Sci. 2017, 5, 85–93. [Google Scholar] [CrossRef]
  9. Wang, F.-C.; Peng, Y.-K.; Lu, J.-F.; Chung, T.-T.; Yen, J.-Y. Micro-lens fabrication by a long-stroke precision stage with switching control based on model response prediction. Microsyst. Technol. 2019, 3, 1–14. [Google Scholar] [CrossRef]
  10. Wang, F.-C.; Lu, J.-F.; Su, W.-J.; Yen, J.-Y. Precision positioning control of a long-stroke stage employing multiple switching control. Microsyst. Technol. 2020, 1, 1–14. [Google Scholar] [CrossRef]
  11. Nichols, R.A.; Reichert, R.T.; Rugh, W.J. Gain scheduling for H-infinity controllers: A flight control example. IEEE Trans. Control. Syst. Technol. 1993, 1, 69–79. [Google Scholar] [CrossRef] [Green Version]
  12. Yamaguchi, T.; Shishida, K.; Tohyama, S.; Hirai, H. Mode switching control design with initial value compensation and its application to head positioning control on magnetic disk drives. IEEE Trans. Ind. Electron. 1996, 43, 65–73. [Google Scholar] [CrossRef]
  13. Hossain, A.; Rahman, M. Comparative Analysis among Single-Stage, Dual-Stage, and Triple-Stage Actuator Systems Applied to a Hard Disk Drive Servo System. Actuators 2019, 8, 65. [Google Scholar] [CrossRef] [Green Version]
  14. Zhu, W.; Rui, X.-T. Hysteresis modeling and displacement control of piezoelectric actuators with the frequency-dependent behavior using a generalized Bouc–Wen model. Precis. Eng. 2016, 43, 299–307. [Google Scholar] [CrossRef]
  15. Saleem, A.; Al-Ratrout, S.; Mesbah, M. A fitness function for parameters identification of Bouc-Wen hysteresis model for piezoelectric actuators. In Proceedings of the 5th International Conference on Electrical and Electronic Engineering (ICEEE), Istanbul, Turkey, 3–5 May 2018. [Google Scholar]
  16. Gan, J.; Zhang, X. Nonlinear hysteresis modeling of piezoelectric actuators using a generalized Bouc–Wen model. Micromachines 2019, 10, 183. [Google Scholar] [CrossRef] [Green Version]
  17. Fang, J.; Wang, J.; Li, C.; Zhong, W.; Long, Z. A compound control based on the piezo-actuated stage with bouc–wen model. Micromachines 2019, 10, 861. [Google Scholar] [CrossRef] [Green Version]
  18. Zhang, F.; Liu, J.; Tian, J. Analysis of the Vibration Suppression of Double-Beam System via Nonlinear Switching Piezoelectric Network. Machines 2021, 9, 115. [Google Scholar] [CrossRef]
  19. Physik Instrumente P-517 • P-527 Multi-Axis Piezo Scanner. Available online: https://www.physikinstrumente.com/en/products/nanopositioning-piezo-flexure-stages/multi-axis-piezo-flexure-stages/p-517-p-527-multi-axis-piezo-scanner-201500/ (accessed on 4 July 2021).
  20. Piezomechanik Amplifier SVR/150/3. Available online: https://www.piezomechanik.com/products/ (accessed on 4 July 2021).
  21. Chuo Precision Industrial Stepper Motor ALS-510-H1P. Available online: https://www.chuo.co.jp/contents/hp0543/list.php?CNo=543&ProCon=7065 (accessed on 4 July 2021).
  22. Chuo Precision Industrial Stepper Motor ALV-104-HP. Available online: https://www.chuo.co.jp/contents/hp0058/list.php?CNo=58&ProCon=94 (accessed on 4 July 2021).
  23. National Instruments DAQ Card PCI-6221 and PCI-6229. Available online: https://www.ni.com/zh-tw.html (accessed on 4 July 2021).
  24. Glover, K.; McFarlane, D. Robust stabilization of normalized coprime 457 factor plant descriptions with H/sub infinity/-bounded uncertainty. IEEE 458 Trans. Autom. Control 1989, 34, 821–830. [Google Scholar] [CrossRef]
  25. Georgiou, T.T.; Smith, M.C. Optimal robustness in the gap metric. In Proceedings of the 28th IEEE Conference on Decision and Control, Tampa, FL, USA, 13–15 December 1989. [Google Scholar]
  26. McFarlane, D.; Glover, K. A loop-shaping design procedure using H/sub infinity/synthesis. IEEE Trans. Autom. Control 1992, 37, 759–769. [Google Scholar] [CrossRef]
  27. Kennedy, J.; Eberhart, R. Particle swarm optimization. In Proceedings of the ICNN’95-International Conference on Neural Networks, Perth, WA, Australia, 27 November–1 December 1995. [Google Scholar]
  28. Meyrán, G.J. Augustin-Jean Fresnel. Rev. Mex. Oftalmol. 2008, 2, 82. [Google Scholar]
  29. Zhang, Y.; Fang, Y.; Zhou, X.; Dong, X. Image-based hysteresis modeling and compensation for an AFM piezo-scanner. Asian J. Control 2009, 11, 166–174. [Google Scholar] [CrossRef]
  30. Lu, H.; Fang, Y.; Ren, X.; Zhang, X. Improved direct inverse tracking control of a piezoelectric tube scanner for high-speed AFM imaging. Mechatronics 2015, 31, 189–195. [Google Scholar] [CrossRef]
  31. Kuo, F.-C.; Hsu, C.; Hsieh, M.-R.; Yen, J.-Y.; Chen, L.-C.; Chung, T.-T.; Wang, F.-C. Study on the transient response to the point-to-point motion controls on a dual-axes air-bearing planar stage. Int. J. Adv. Manuf. Technol. 2020, 111, 2759–2772. [Google Scholar] [CrossRef]
  32. Liu, L.-C.; Yang, P.-H.; Liao, S.-C.; Li, B.-P.; Wang, F.-C.; Lin, P.-C. Development of a dual-stage and visual-servo filming robot. Proc. Inst. Mech. Eng. Part I J. Syst. Control Eng. 2021, 235, 0959651821991363. [Google Scholar]
Figure 1. The multiple switching control structure.
Figure 1. The multiple switching control structure.
Machines 09 00153 g001
Figure 2. Response improvement by the switching control.
Figure 2. Response improvement by the switching control.
Machines 09 00153 g002
Figure 3. The procedures of iterative parameter tuning.
Figure 3. The procedures of iterative parameter tuning.
Machines 09 00153 g003
Figure 4. The long-stroke precision stage.
Figure 4. The long-stroke precision stage.
Machines 09 00153 g004
Figure 5. Comparison of the standard robust controllers and the robust PI controller: (a) simulation; (b) experimental results.
Figure 5. Comparison of the standard robust controllers and the robust PI controller: (a) simulation; (b) experimental results.
Machines 09 00153 g005
Figure 6. Step response of PZT stage employing the optimal switching control.
Figure 6. Step response of PZT stage employing the optimal switching control.
Machines 09 00153 g006
Figure 7. Gain-scheduling control with command feedforward for the motor stage.
Figure 7. Gain-scheduling control with command feedforward for the motor stage.
Machines 09 00153 g007
Figure 8. Tracking responses of the motor stage: (a) ramp input (100 μm/s); (b) ramp input (500 μm/s); (c) sinusoidal input (0.1 Hz); (d) sinusoidal input (0.1 Hz).
Figure 8. Tracking responses of the motor stage: (a) ramp input (100 μm/s); (b) ramp input (500 μm/s); (c) sinusoidal input (0.1 Hz); (d) sinusoidal input (0.1 Hz).
Machines 09 00153 g008
Figure 9. The switching control structure for the combined stage.
Figure 9. The switching control structure for the combined stage.
Machines 09 00153 g009
Figure 10. Experimental responses for tracking characters.: (a) character tracking; (b) x-axis responses; (c) zoom-in responses at the turning point around 332 s; (d) zoom-in responses at the turning point around 564 s.
Figure 10. Experimental responses for tracking characters.: (a) character tracking; (b) x-axis responses; (c) zoom-in responses at the turning point around 332 s; (d) zoom-in responses at the turning point around 564 s.
Machines 09 00153 g010aMachines 09 00153 g010b
Figure 11. Integration of the combined stage and the TPP system.
Figure 11. Integration of the combined stage and the TPP system.
Machines 09 00153 g011
Figure 12. The design and fabrication of the FZP: (a) the lens design; (b) lens fabrication.
Figure 12. The design and fabrication of the FZP: (a) the lens design; (b) lens fabrication.
Machines 09 00153 g012
Figure 13. Optical properties of the micro-lens: (a) light spot; (b) brightness; (c) image intensity; (d) image sharpness.
Figure 13. Optical properties of the micro-lens: (a) light spot; (b) brightness; (c) image intensity; (d) image sharpness.
Machines 09 00153 g013
Table 1. Specifications of the long-stroke precision stage.
Table 1. Specifications of the long-stroke precision stage.
P-517.RCD PZT Stage [19]
Active axis x, y
Maximum stroke −50 to 50 μm
Mass 1.4 kg
Resolution 1 nm
SVR/150/3 amplifier [20]
Output voltage range −30 to 150 V
Max gain 30 (tunable)
ALS-510-H2 P stepper [21]
Active axis x, y
Maximum stroke 100 mm
Resolution 0.1 μm
Maximum loading 40 kgf
Maximum command 80,000 pulse/sec
ALV-104-HP stepper [22]
Active axis z
Maximum stroke 40 mm
Resolution 0.1 μm
Maximum loading 10 kgf
Maximum command 40,000 pulse/sec
Table 2. Statistical comparison of the controllers.
Table 2. Statistical comparison of the controllers.
Robust ControllerRobust PI Controller
C F C M C S C ¯ F C ¯ M C ¯ S
Sim.Rise time (sec)0.00430.01710.06220.00510.03270.0692
Settling time (sec)1.04151.06121.13311.05631.06541.1283
Overshoot (%)41.16000039.27920.01010
RMSE (μm)1.68342.05112.64341.68631.87332.5951
Exp.Rise time (sec)0.00430.01580.06220.00380.02870.0654
Settling time (sec)1.06471.05791.12691.04951.05741.1183
Overshoot (%)54.82000.1100047.43000.21000.0500
RMSE (μm)1.88722.14192.70211.69341.90342.4406
Sim.: simulation; Exp.: experiments.
Table 3. Tracking performance of the motor stage.
Table 3. Tracking performance of the motor stage.
InputsRampSinusoidal
Sizes100 μm/s500 μm/s0.1 Hz1 Hz
Sim.Phase lag (°)--00
Maximum error (μm)0.33241.66220.00170.171
RMSE (μm)0.04010.20070.00130.2939
Exp.Phase lag (º)--00
Maximum error (μm)0.40002.50000.30001.1267
RMSE(μm)0.22380.78510.13550.4161
Sim.: simulation; Exp.: experiments.
Table 4. Tracking performance of the combined stage.
Table 4. Tracking performance of the combined stage.
RMSE
Control method in [8]203.9 nm
Control method in [9]140.5 nm
Multiple switching control136.8 nm
Table 5. Comparison of the lens performance.
Table 5. Comparison of the lens performance.
Control in [8]Control in [9]Proposed Control
Intensity164255255
Sharpness9.1511.715.8
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, F.-C.; Lu, J.-F.; Chung, T.-T.; Yen, J.-Y. Iterative Parameter Optimization for Multiple Switching Control Applied to a Precision Stage for Microfabrication. Machines 2021, 9, 153. https://doi.org/10.3390/machines9080153

AMA Style

Wang F-C, Lu J-F, Chung T-T, Yen J-Y. Iterative Parameter Optimization for Multiple Switching Control Applied to a Precision Stage for Microfabrication. Machines. 2021; 9(8):153. https://doi.org/10.3390/machines9080153

Chicago/Turabian Style

Wang, Fu-Cheng, Jun-Fu Lu, Tien-Tung Chung, and Jia-Yush Yen. 2021. "Iterative Parameter Optimization for Multiple Switching Control Applied to a Precision Stage for Microfabrication" Machines 9, no. 8: 153. https://doi.org/10.3390/machines9080153

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop