Next Article in Journal
Fault Diagnosis in a 2 MW Wind Turbine Drive Train by Vibration Analysis: A Case Study
Previous Article in Journal
The Innovative Design and Performance Testing of a Mobile Robot for the Automated Installation of Spacers on Six-Split Transmission Lines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dynamic Error Compensation for Ball Screw Feed Drive Systems Based on Prediction Model

by
Hongda Liu
,
Yonghao Guo
,
Jiaming Liu
and
Wentie Niu
*
Key Laboratory of Mechanism Theory and Equipment Design of Ministry of Education, Tianjin University, Tianjin 300350, China
*
Author to whom correspondence should be addressed.
Machines 2025, 13(5), 433; https://doi.org/10.3390/machines13050433
Submission received: 20 April 2025 / Revised: 15 May 2025 / Accepted: 15 May 2025 / Published: 20 May 2025
(This article belongs to the Section Automation and Control Systems)

Abstract

The dynamic error is the dominant factor affecting multi-axis CNC machining accuracy. Predicting and compensating for dynamic errors is vital in high-speed machining. This paper proposes a novel prediction-model-based approach to predict and compensate for the ball screw feed system’s dynamic error. Based on the lumped and distributed mass methods, this method constructs a parameterized dynamic model relying on the moving component’s position for electromechanical coupling modeling. Using Latin Hypercube Sampling and numerical simulation, a sample set containing the input and output of one control cycle is obtained, which is used to train a Cascade-Forward Neural Network to predict dynamic errors. Finally, a feedforward compensation strategy based on the prediction model is proposed to improve tracking performance. The proposed method is applied to a ball screw feed system. Tracking error simulations and experiments are conducted and compared with the transfer function feedforward compensation. Typical trajectories are designed to validate the effectiveness of the electromechanical coupling model, the dynamic error prediction model, and the feedforward compensation strategy. The results show that the prediction model exhibits a maximum prediction deviation of 1.8% for the maximum tracking error and 13% for the average tracking error. The proposed compensation method with friction compensation achieves a maximum reduction rate of 76.7% for the maximum tracking error and 63.7% for the average tracking error.

1. Introduction

A ball screw feed system is composed of rotary motors, linear guides, and ball screws, functioning together to transform the motor’s rotational motion into precise linear displacement along the guide rail. Ball screws demonstrate high efficiency, extended service life, superior load-bearing capacity, excellent stiffness, long travel ranges, and minimal heat generation, enabling consistent acceleration performance even under significant variations in workpiece inertia. Multiple technical parameters collectively affect feed system accuracy–mechanical vibration in screw drives, control system design, environmental noise, and sensor resolution all contribute to final positioning results. Monitoring and compensating for tracking errors in ball screw feed systems facilitates error reduction and accuracy enhancement.
The prerequisite for dynamic error prediction and compensation is establishing an accurate electromechanical coupling model that accounts for positional variations. Numerous simulation and experimental studies confirm that ball screw feed drives display substantial dynamic behavior variations correlated with position, stemming mainly from changes in the active screw length during motor–worktable displacement. This phenomenon results in position-dependent dynamic behavior during system operation. Current approaches for developing position-dependent dynamic models include dictionary function libraries [1], observer-based methods [2], and linear parameter-varying (LPV) techniques [3]. These position-dependent characteristics substantially influence motion control precision. Integrating worktable position effects into motion control strategies can significantly improve dynamic tracking accuracy.
Constructing dynamic models for ball screw feed systems represents a fundamental aspect of studying their dynamic characteristics. An accurate dynamic model provides theoretical foundations for analyzing system dynamic behavior and establishes a critical technical foundation for digital design and simulation, dynamic optimization design, and dynamic error modeling and control. Existing mainstream modeling strategies encompass lumped parameter models [4,5,6,7], distributed parameter frameworks [6,7], and finite element-based techniques [8,9,10]. The lumped mass method has been widely adopted due to its computational efficiency and low resource requirements. However, its limitations include potential inaccuracies in capturing mass distribution effects on structural dynamic responses, which may compromise model prediction accuracy under certain operating conditions. In comparison, the distributed mass method facilitates the construction of more precise dynamic models, albeit with substantially increased computational complexity. While the finite element method offers superior modeling accuracy, its computational resource requirements significantly exceed those of the other techniques, consequently limiting its practical engineering applications to some extent.
Dynamic error modeling forms the theoretical foundation for dynamic error prediction and compensation research, where the rationality of the model architecture directly determines the scientific validity and reliability of related studies. Based on the diversity of modeling principles, existing dynamic error modeling approaches can be systematically categorized into three main classes: analytical modeling based on dynamic systems, statistical modeling based on characteristic parameters, and mechanism modeling based on generation principles. The dynamics-based modeling method utilizing system transfer functions achieves error characterization by establishing mathematical relationships between control elements [11,12]. While this method demonstrates computational efficiency advantages, its prediction accuracy is inherently limited by model simplification assumptions. In contrast, the integrated modeling approach based on multi-physics coupling incorporates servo control and mechanical structure models through commercial simulation platforms [13,14]. The improved accuracy comes at the expense of computational burden, revealing the intrinsic conflict between model precision and processing efficiency. Under the data-driven modeling paradigm, characteristic parameter-based dynamic error modeling reconstructs errors by extracting feature parameters from measurement signals and establishing parametric functional relationships [15,16]. Notably, the multi-source nature of dynamic error generation mechanisms induces prevalent multi-band coupling phenomena in measured signals, significantly challenging the identification accuracy of characteristic parameters. The mechanism-based modeling approach follows a methodology combining mechanism decomposition and linear superposition: first analyzing error generation mechanisms, then quantifying operational patterns of key influencing factors, and ultimately constructing system-level error models [17,18]. Although this approach demonstrates strong physical interpretability, current research still faces two critical limitations: (1) incomplete theoretical descriptions of dynamic error generation mechanisms and (2) unestablished quantitative mapping relationships between influencing factors and resultant error patterns. These limitations constitute fundamental bottlenecks hindering the engineering implementation of this methodology.
Current error compensation strategies can be classified into offline pre-compensation [19,20] and online feedforward compensation [21,22]. Feedforward compensation methods exhibit superior computational efficiency and require minimal modifications to the original control system architecture, making them highly valuable for practical engineering applications.
The tracking accuracy of ball screw feed systems is significantly influenced by friction, and their tracking performance can be further enhanced through friction compensation. The LuGre model [23] integrates key features of the Coulomb, Stribeck, and viscous friction models while incorporating frictional hysteresis. Notably, it retains the state variable z from the Dahl model to describe pre-sliding dynamics. It simplifies the complexity of the Bristle model, thereby offering greater engineering applicability due to its structural simplicity and ease of implementation.
The accuracy of the dynamics-based modeling method improved with computational burden growth. The accuracy of the characteristic parameter-based dynamic error modeling method depends on the identification accuracy of characteristic parameters. The mechanism-based modeling approach’s accuracy relies on complete theoretical descriptions of dynamic error generation mechanisms and established quantitative mapping relationships between influencing factors and resultant error. Considering the disadvantages of these methods, the Cascade-Forward Neural Network (CFNN) prediction model was proposed to describe dynamic error with high precision and low computational burden. With adequate sample points, CFNN can quantitatively map relationships between influencing factors and resultant responses without identifying characteristic parameters. To obtain samples, a dynamic modeling framework that accounts for coupled rigid–flexible interactions in ball screw feed drives with explicit consideration of moving element positioning is formulated. The computational efficiency and accuracy are balanced by integrating the lumped and distributed mass methods. An electromechanical coupling model is developed based on this foundation. The sampling space was populated using Latin Hypercube Sampling (LHS) techniques, with generated parameter combinations serving as inputs to the coupled electromechanical simulation model. These samples train a CFNN prediction model to predict worktable position and dynamic error within a single control cycle, comprehensively accounting for position variation and friction effects. With sufficient sample points, the prediction model achieves high-fidelity emulation of the original system behavior while substantially reducing computational demands. At the start of each control cycle, the dynamic error is calculated by subtracting the predicted worktable position from the reference position. The predicted dynamic error is compensated to the position loop for reducing real-time tracking errors, effectively improving motion accuracy. The main contributions of this work are as follows:
(1) A dynamic modeling method relying on moving components’ positions is proposed based on the lumped and distributed mass methods. The resulting coupled electromechanical system model is systematically validated via simulations and experiments.
(2) Based on the proposed electromechanical coupling model and LHS, a prediction model is constructed to predict the table position and dynamic error within a single control cycle.
(3) The ball screw feed system’s dynamic error is compensated by applied the constructed prediction model and a friction model.
(4) The methodological effectiveness is substantiated through comprehensive testing, incorporating both computational simulations and physical experiments performed on a purpose-built ball screw feed system platform.
The remainder of this paper is organized as follows. Section 2 describes the methodology of dynamic error prediction and feedforward compensation. Section 3 details the electromechanical coupling modeling process. Section 4 systematically presents the dynamic error modeling process and feedforward compensation strategy. Section 5 describes the simulations and experiments. Finally, Section 6 provides conclusions and discusses future work.

2. Methodology

The schematic representation of the methodological framework is presented in Figure 1. To begin with, a position-dependent dynamic model was established based on the distributed mass method and lumped mass method. Combined with Field-Oriented Control (FOC), an electromechanical coupling model of the ball screw feed system was developed. Subsequently, input parameters and design space were determined. Different input combinations were obtained through the LHS experiment design. Numerical simulations using the electromechanical coupling model and the input combinations generated sample points. These sample points were then employed to train a CFNN for dynamic error prediction. Following this, a feedforward compensation strategy was proposed based on the dynamic error prediction model. Finally, experimental validation was conducted to verify the accuracy of the dynamic error prediction model and the effectiveness of the dynamic error compensation strategy.

3. Electromechanical Coupling Model of the Ball Screw Feed System

In this section, two complementary mass modeling methods–lumped and distributed–are combined in this section to formulate a dynamic model with explicit position dependency for ball screw feed systems. Subsequently, the system’s friction is identified based on the LuGre model. Finally, an electromechanical coupling model is developed through FOC.

3.1. Position-Dependent Dynamics Modeling Based on the Lumped Mass Method and Distributed Mass Method

3.1.1. Lumped Mass Model of Two Shafts and Worktable

Figure 2a displays the ball screw feed system’s configuration. Figure 2b is a real ball screw feed system. The bed, the permanent magnet synchronous motor (PMSM), the couplings, the torque sensor, the bearings, the bearing housings, the nut, the guide rails, the sliders, the ball screw, and the worktable. Components with relatively small deformations in the ball screw feed system are treated as lumped masses to balance computational efficiency and accuracy, including the shaft of the PMSM, the shaft of the torque sensor, and the worktable, numbered sequentially as 1, 2, and 3. The bed, bearing housings, and guide rails are neglected and have minimal influence on axial motion. The ball screw, which exhibits relatively high flexibility, is modeled using the Timoshenko beam theory [24] to establish a distributed mass dynamic model in Section 3.1.2. The generalized coordinates for lumped masses 1–3 are denoted as q1, q2, and q3, where q1 and q2 represent the rotation of lumped masses 1 and 2, respectively, and q3 represents the translation of lumped mass 3.

3.1.2. Position-Dependent Distributed Mass of the Screw

The screw shaft in ball screw feed mechanisms exhibits substantial elastic deformation characteristics, necessitating the adoption of Timoshenko beam theory [24] for developing a distributed-parameter dynamic model. The calculation of Timoshenko beam element matrices can be found in the Appendix A. Based on the principles underlying the derivation of the Timoshenko beam element matrices, a distributed mass model related to the position of the lead screw is established. Define L as the total screw length divided into N discrete segments. When the axial displacement between drive and load interfaces is x, the output position resides in segment nS:
n S = N L x
where ⌈⋅⌉ denotes the ceiling function. The remaining length xR falls within the nS-th segment. The remaining dimension xR is positioned within the nS-th section. Through modification of these two adjacent segments, the nS-th section’s length becomes equal to xR. The subsequent segment (nS+1) length, denoted as lns+1, can be determined using Equation (2). This adjusted nS+1 node subsequently functions as the lead screw’s output terminal, with the modified configuration visually represented in Figure 3. The remaining portion of the lead screw is modeled as a conventional Timoshenko beam. Therefore, both the stiffness and mass distribution matrices characterizing the lead screw system exhibit (6N + 6) × (6N + 6) dimensionality, where all constituent elements dynamically depend on the spatial coordinate x.
l n S + 1 = 2 L N x R

3.1.3. Position-Dependent Dynamic Model of Ball Screw Feed System

The matrix elements about the RX-direction DOF are extracted to derive the screw’s RX-direction distributed mass stiffness matrix KRX and mass matrix MRX based on the distributed mass model presented in Section 3.1.2. These matrices’ dimensions are (N + 1) × (N + 1). The model of joints must also be established with the lumped masses described in Section 3.1.1. Based on the lumped mass assumption in Section 3.1.1, Lumped Mass 1 and Lumped Mass 2 are connected by a single degree of freedom (SDOF) spring KR1, representing rotational stiffness along the RX-direction, with a magnitude equal to the rotational stiffness of the shaft coupling between the motor shaft and the torque sensor. Similarly, Lumped Mass 2 and the screw input end are connected by an SDOF spring KR2, corresponding to rotational stiffness along the RX-direction, with a magnitude equal to the rotational stiffness of the shaft coupling between the torque sensor and the screw. According to [25], the ball screw nut generates axial thrust simultaneously with torque. The functional dependence of axial thrust F on torque Tn is mathematically defined through Equation (3):
T n = F R tan ϕ
where R is the screw radius and ϕ is the screw contact angle. Let the generalized coordinate vector between the screw output end and Lumped Mass 3 be defined as Qscn = [qscnRx qwt], where qscnRx represents the degree of freedom of the screw output end along the RX-direction, and qwt is the worktable DOF along the X-direction. The corresponding screw–nut pair’s stiffness matrix Kscn is then given by
K s c n = r s c n l l e a d tan ( ϕ ) 2 π r s c n tan ( ϕ ) l l e a d 2 π 1 k s c n
where rscn is the screw radius, llead is the lead of the screw, and kscn is the axial stiffness of the screw–nut pair. Let the ball screw feed system’s mass matrix M and stiffness matrix K be matrices of size (N + 4) × (N + 4). The stiffness matrix K’s elements are defined as follows:
K ( 1 , 1 ) = K ( 2 , 1 ) = K ( 1 , 2 ) = K R 1
K ( 2 , 2 ) = K R 1 + K R 2
K ( 2 , 3 ) = K ( 3 , 2 ) = K R 2
K ( 3 , 3 ) = K R 2
K ( n S + 1 , n S + 1 ) = K s n c ( 1 , 1 )
K ( n S + 1 , N + 4 ) = K s n c ( 1 , 2 )
K ( N + 4 , n S + 1 ) = K s n c ( 2 , 1 )
K ( N + 4 , N + 4 ) = K s n c ( 2 , 2 )
All other elements are set to zero. Based on this framework, the elements of KRX are added to the corresponding positions of the submatrix K [{3, N + 3}, {3, N + 3}], thereby obtaining the position-dependent stiffness matrix K(x) associated with the coordinate x. The mass matrix M’s elements are defined as follows:
M ( 1 , 1 ) = j m
M ( 2 , 2 ) = j T S
M ( N + 4 , N + 4 ) = m w t
All other elements are set to zero. Based on this framework, the elements of MRX are added to the corresponding positions of the submatrix M [{3, N + 3}, {3, N + 3}], thereby obtaining the position-dependent mass matrix M(x) associated with the coordinate x. The ball screw feed system’s position-dependent dynamic model is expressed as follows:
M ( x ) Q ¨ + K ( x ) Q = F
Here, Q corresponds to the node-level generalized displacement vector, while F signifies the per-node force components. The ball screw feed system’s position-dependent dynamic model is illustrated in Figure 4.

3.2. Friction Identification Based on the LuGre Model

The LuGre model constitutes a generalized friction modeling framework incorporating the Stribeck effect, Coulomb friction, and static friction maxima, enabling precise characterization of both gross sliding regimes and pre-sliding. This model quantifies frictional disturbances through elastic bristle deflection: microscopic bristles (modeled as damped springs) undergo pseudo-deflection, whose temporal evolution generates frictional torque. The equations are expressed as
T f = σ 0 z + σ 1 d z d t + σ 2 ω
d z d t = ω ω g ( ω ) z
σ 0 g w = T C + T S T C e ω Ω S 2
where ω denotes the angular speed, z denotes the bristle deformation, σ0 is the bristle stiffness coefficient, σ1 corresponds to the bristle damping coefficient, σ2 signifies the viscous friction coefficient, ΩS represents the critical Stribeck velocity, TC denotes the Coulomb friction torque, and TS represents the maximum static friction torque. The LuGre model requires identifying six parameters, including four static and two dynamic parameters. In ball screw feed systems, friction primarily occurs at the motion–contact interfaces. A systematic friction identification methodology was employed, with constant-velocity experiments specifically designed. When the system operates at a constant velocity, the system acceleration becomes zero. According to Newton’s second law,
T m = m e a e + T f
Here, Tm denotes the motor output torque, me corresponds to the system’s equivalent inertia, and ae represents the system’s equivalent acceleration. Consequently, Tm equals Tf. The identification results are presented in Table 1.

3.3. Electromechanical Coupling Modeling of the Ball Screw Feed System

The electromechanical coupling model based on FOC is constructed using the dynamic model proposed in Section 3.1. The core of FOC lies in its approach, akin to that of DC motor control, utilizing coordinate transformation theory to decouple the armature current and excitation current in a PMSM into two independent components for individual control. FOC methods primarily include maximum output power control method, maximum torque/current control method, field weakening control method, the Id = 0 control method, and transverse flux linkage control method, depending on the application. The Id = 0 control method is employed to regulate the PMSM [26].
The electromechanical coupling model is illustrated in Figure 5. The input to the electromechanical coupling model is the reference table position Swr. As shown in Figure 5, Sw denotes the table position, ΔSw represents the tracking error, ωRr is the reference velocity, ωR is the actual motor velocity, ΔωR is the velocity error, Ir is the reference current, I is the actual current, ΔI is the current error, and Ur is the reference voltage. By employing space vector pulse width modulation (SVPWM), Ur is input to the PMSM to generate torque Tm. SVPWM further smoothens the motor’s output torque [27]. The LuGre model provides the friction force Tf. Sw is the generalized coordinate of the worktable’s lumped mass along the y-axis.
The operational specifications of the PMSM and the ball screw feed mechanism are in Table 2 and Table 3, respectively. Controller gains for the PI and P regulators were optimized following methodologies in [28]. Position–velocity control loop parameters, designated as PSw, PωR, and IωR, govern the motion tracking dynamics, while current regulation in the rotating reference frame is implemented through q- and d-axis PI coefficients Piq, Iiq, Pid, and Iid, as detailed in Table 4. A 10 kHz sampling rate was implemented throughout the experimental framework, with inverter dead-time effects and viscous damping forces intentionally excluded from the system modeling.

4. Dynamic Error Modeling of One Control Cycle and Feedforward Compensation

This section presents the specific workflow of the dynamic error feedforward compensation method based on the prediction model. First, the fundamental principles are elaborated. Subsequently, sample points are generated through LHS combined with an electromechanical coupling model of the ball screw feed system. These sample points are then utilized to train a CFNN-based prediction model. Finally, the prediction model feedforward compensation is proposed for the dynamic error.

4.1. Basic Principles and Processes

Assuming the feed system is at time point k, the position of the workbench is Sw(k), and the velocity is S ˙ w k The acceleration is S ¨ w k , and the motor’s reference q-axis current command is Iqr(k). Under the command of Iqr(k), the motor drives the feed system to operate. At time point k + 1, the position of the workbench is Sw(k+1), as shown in Figure 6.

4.2. Latin Hypercube Sampling Experimental Design

LHS is a widely utilized experimental design method in computer simulations. Compared with classical Monte Carlo sampling approaches, which tend to produce clustered sample distributions, LHS optimizes Monte Carlo sampling by ensuring more uniform and efficient sampling across the design space of multiple variables. The fundamental principle of LHS parallels stratified sampling, incorporating both stratification and random permutation. The procedure of LHS is described here. Let (v1, v2, …, vm) represent m input variables, where the range of vj is defined as Rj = [Aj, Bj] for j = 1, 2, …, m. The cumulative distribution function of vj is denoted as V = G(vj), as documented in reference [29].
Step 1: Partition Rj into n stratified intervals, denoted as Rj1, Rj2, …, Rjn, each possessing equal probability. These subdivisions satisfy the following conditions:
R j 1 R j 2 R j m R j k R j l =
P ( v j R j k ) = 1 n ( k , l = 1 , 2 , n )
Step 2: The cumulative probability value of the input variable vj for the k-th interval is derived via Equation (23):
G v j R j k = k 1 n + u j k n , u j k U ( 0 , 1 )
Here, ujk constitutes a stochastic variable following uniform distribution across the unit interval [0, 1]. The cumulative probabilities of the input variables (v1, v2, …, vm) can be formalized as an n×m matrix Q = (Pkj)n×m, where Pkj is the cumulative probability value associated with the k-th stratified interval of the j-th variable.
Step 3: Utilizing the inverse function of G (·), the probability values from Equation (23) are transformed into sampled values vkj, as expressed in Equation (24):
G v j R j k = k 1 n + u j k n , u j k U ( 0 , 1 )
The sample matrix V can be derived.
V = v 11 v 12 v 1 m v 21 v 22 v 2 m v n 1 v n 2 v n m
Step 4: Randomly combine the variables into n sampling groups as follows:
V = v 11 v 12 v 1 m v 21 v 22 v 2 m v n 1 v n 2 v n m = v 1 m , v 2 m , , v n m T
Each row within the matrix encapsulates a distinct grouping of spatial or temporal sampling coordinates, reflecting the structured organization of collected data points. The design space comprises Sw(k), S ˙ w k , S ¨ w k , and Iqr(k). The ball screw has a length of 0–1500 mm; consequently, the worktable position Sw(k) is constrained within the range of 0–1500 mm. S ˙ w k directly correlates with the motor’s nominal angular velocity of 2000 RPM. Given the ball screw lead of 10 mm, the range of S ˙ w k is determined to be −333 to 333 mm/s. The motor’s rated current is 10 A, and considering that the equivalent current equals 0.707 times the instantaneous current, the range of Iqr(k) is set to −15 A to 15 A. With a worktable mass of 515 kg and a maximum acceleration capability of 30,000 mm/s2, the stator current may trigger the motor driver’s overcurrent protection when the initial acceleration reaches 4000 mm/s². Therefore, accounting for the acceleration characteristics of the S-shaped velocity profile, the worktable acceleration S ¨ w k is bounded within −1000 to 1000 mm/s2. The design ranges for these four parameters are summarized in Table 5. The parts of 5000 sample points are illustrated in Table 6.

4.3. Dynamic Error Modeling Based on a CFNN Prediction Model

4.3.1. Cascade-Forward Neural Network

The CFNN shares structural similarities with conventional feedforward networks, as shown in Figure 7 [30]. The CFNN shares architectural similarities with feedforward networks employing backpropagation, adopting gradient-based parameter optimization through error backpropagation. This network’s distinctive architectural paradigm, however, emerges from its topological configuration, in which hidden layers maintain cross-tier linkages through progressive connectivity pathways [31]. Within the CFNN architecture, as with other feedforward networks, one or more hidden layers exist with mutual connections and nonlinear activation functions. Neurons possess individual biases, while interconnections carry specified weights. The calibration of artificial neural networks necessitates optimization of synaptic connection strengths to achieve prediction accuracy within predefined tolerance thresholds, as documented in reference [32].
The network architecture comprises three principal components: output neurons yk (k = 1, 2, …, K), input units xi (i = 1, 2, …, I), and hidden layer nodes nj (j = 1, 2, …, J). While exhibiting structural parallels with conventional feedforward backpropagation architectures in employing error backpropagation for synaptic weight adjustments, the CFNN’s key differentiator resides in implementing full cross-layer connectivity, in which each neuronal layer maintains direct synaptic linkages with all antecedent processing stages [33]. The hidden layer employs the tansig nonlinear operator as its core computational transformation mechanism.
f ( x ) = 2 1 + e 2 x 1
The output layer implements the purelin linear transfer operator as its transformation mechanism.
f ( x ) = x

4.3.2. Training of the CFNN Prediction Model of Worktable’s Dynamic Error

To train the CFNN, the Levenberg–Marquardt (LM) algorithm is utilized, which effectively integrates the optimization capabilities of the Gauss–Newton technique with the swift convergence characteristic of gradient descent [34]. The BPNN’s learning process can be essentially viewed as the solution to Equation (29).
min S R n F ( S ) = 1 2 r S T r S = 1 2 i = 1 m r i ( S ) 2 m n
where n denotes the sample dimension, m represents the sample’s number, and i indicates the i-th sample. r(S) is the residual function of S.
r S = r 1 S , r 2 S , , r m S T
The matrix S is obtained by combining the weights and the bias vector.
S = p 11 p 12 p 1 n q 1 p 21 p 22 p 2 n q 2 p n 1 p n 2 p n n q n
where p denotes the weights, and q represents the bias. The Jacobian matrix of r(S) is Ja(S).
J a S = r 1 S , r 2 S , , r m S T = r 1 S 1 r 1 S 2 r 1 S n r 2 S 1 r 2 S 2 r 2 S n r m S 1 r m S 2 r m S n T
G(S) is the gradient of F(S).
G S = i = 1 m r i ( S ) r i ( S ) = J a S T r S
At the k-th optimization stage, the gradient can be expressed mathematically as
Δ k = J a S K T J a S K + u k I J a S K T r S K
where I denotes the identity matrix and uk is the step size parameter. A larger uk causes the Levenberg–Marquardt (LM) algorithm to approximate gradient descent, as the diagonal terms dominate the update. In contrast, a smaller uk shifts the LM method toward the Gauss–Newton approach [35]. During the training of the prediction model, performance assessment is essential. For this purpose, the absolute error (AE) metric is employed to evaluate the model’s accuracy.
A E i = y ^ i y i
where y ^ i represents the output of the CFNN for the i-th sample point, and yi denotes the output of the i-th sample. When training the CFNN, the number of neurons Mh in the h-th hidden layer, the number of hidden layers N, the number of iterations epochs, and the learning rate u influence the performance of the CFNN. A total of 1000 samples were randomly generated using LHS and the electromechanical coupling model to evaluate the performance of the CFNN. Table 7 shows the worktable position’s maximum and average absolute errors when the learning rate u is set to 0.3, 0.5, and 0.7, respectively (N = 1, M1 = 3, epochs = 100). Table 8 presents the worktable position’s maximum and mean absolute errors when the number of hidden layers N is set to 1, 2, and 3, respectively (M1 = M2 = M3 = 3, epochs = 100, u = 0.3). Table 9 displays the workbench position’s maximum and mean absolute errors when the number of neurons Mh in the hidden layer is set to 3, 4, and 5, respectively (N = 3, epochs = 100, u = 0.3). Table 10 illustrates the workbench position’s maximum and mean absolute errors when the number of iterations epochs is set to 100, 150, and 200, respectively (N = 3, M1 = M2 = M3 = 3, u = 0.3). Table 7, Table 8, Table 9 and Table 10 demonstrate that the accuracy of the CFNN improves as N, Mh, and epochs increase and as u decreases. It is worth noting that excessively large or small values of N, Mh, u, and epochs may lead to overfitting or underfitting of the CFNN, resulting in degraded performance.
The steps for constructing a neural network are summarized in reference [36]. The final parameters are N = 3, M1 = 34 M2 = 19, M3 = 8, u = 0.97, and epochs = 500. A CFNN with an input layer consisting of four nodes and an output layer with one node was trained. The inputs to the CFNN are Sw(k), S ˙ w ( k ) , S ¨ w ( k ) , and Iqr(k), while the output is S ^ w ( k + 1 ) . The input range for Sw(k) is 0 to 1500 mm, for S ˙ w ( k ) it is −333 to 333 mm/s, for S ¨ w ( k ) it is −1000 to 1000 mm/s2, and for Iqr(k) it is −15 to 15 A.

4.3.3. Validation and Testing of CFNN

An additional 1000 random sample points were generated using LHS and the electromechanical coupling model to test and validate the CFNN. The validation samples were input into the electromechanical coupling model and the CFNN, and the results from the electromechanical coupling model and the CFNN were obtained, respectively. Table 11 presents the maximum and mean absolute errors of the trained CFNN. The CFNN model demonstrates sufficient accuracy in predicting S ^ w ( k + 1 ) . The predicted tracking error of the worktable, E ^ w ( k + 1 ) , can be predicted by subtracting the predicted worktable position S ^ w ( k + 1 ) from the reference position Swr(k+1) at time k, as shown in Equation (36).
E ^ w k + 1 = S w r ( k + 1 ) S ^ w k + 1

4.4. Feedforward Compensation Structure Based on the Prediction Model

The feedforward compensation structure is illustrated in Figure 8. Firstly, the prediction model is employed to calculate S ^ w . Subsequently, the predicted tracking error E ^ w is obtained using Equation (36). Finally, the tracking error E ^ w is applied to the position loop. When the feedforward compensation is incorporated, the input to the position loop P controller becomes ΔSw = SwrSw + E ^ w . The remaining control structure remains consistent with that depicted in Figure 5. The Lugre model was also used to compensate for friction, where KT represents the torque constant.

5. Simulation and Experiment

In this section, simulations and experiments using typical trajectories validate the effectiveness of the electromechanical coupling model, the dynamic error prediction model, and the feedforward compensation strategy. The electromechanical coupling model is compared with the experimental results. The dynamic error prediction model is compared with the transfer function approach [37]. A sampling rate of 10 kHz is used for all the simulations and experiments.

5.1. Ball Screw Feed System Experimental Test Platform

Figure 9 illustrates the experimental setup of the ball screw feed system. The setup comprises a HEIDENHAIN LC185ML1840 grating scale, a worktable equipped with a rotary stage, two BK30 bearing blocks with two bearings, two HIWIN HGH45CA linear guides with four sliders, a HIWIN FDI40-10T3 ball screw with a nut, a HAIBOHUA HCNJ-101 dynamic torque sensor, two MJC-65-BL couplings, a machine bed, and a Hilectro PMSM. An encoder is mounted on the PMSM to provide feedback on the angular position and velocity of the PMSM rotor. The grating scale measures the position of the worktable. The current signals of the PMSM are acquired by the PMSM driver DY2H. The OP4510v2 collects feedback signals from the grating scale, the PMSM encoder, and the motor driver DY2H. Based on these feedback signals, the OP4510v2 generates control signals for the PMSM in each control cycle, which are then transmitted through the DY2H to drive the PMSM. A computer is employed to control the OP4510v2. Simulation and experimental programs are developed and compiled on the PC using RT-LAB software and downloaded to the OP4510v2 for execution.

5.2. Simulation and Experimentation of Typical Trajectories

Six S-shaped velocity interpolation trajectories were designed to validate the compensation effectiveness, each characterized by distinct velocity, acceleration, and jerk profiles. Table 12 enumerates the parameters of these six S-shaped velocity interpolation trajectories.
The distinction between Trajectory 1 and Trajectory 2 lies in their jerk values, while Trajectory 3 and Trajectory 4 differ in acceleration, and Trajectory 5 and Trajectory 6 vary in velocity. Figure 10 graphically presents these six S-shaped velocity interpolation trajectories.

5.2.1. Electromechanical Coupling Model Simulation and Experimentation

Figure 11 presents the tracking errors obtained from simulations based on the electromechanical coupling model under the six trajectories alongside the experimentally measured tracking errors from the ball screw feed system test platform. In Figure 11, the green line represents the experimental results, the blue line corresponds to the position-dependent electromechanical coupling model, and the purple line indicates the deviations of the position-dependent (PD) model. Table 13 lists the corresponding maximum tracking errors and average tracking errors.
As shown in Figure 11 and Table 13, the maximum and average tracking errors of the position-dependent electromechanical coupling model are closer to the experimental results. The maximum acceleration of Trajectory 2 resulted in the most significant tracking error. Regardless of variations in velocity or acceleration, their impact on the tracking error of the ball screw feed system is relatively minor. In contrast, the influence of jerk on tracking error is more pronounced.

5.2.2. Dynamic Error Prediction

Figure 12 illustrates the tracking errors obtained from simulations based on the prediction model and transfer function under the six trajectories alongside the experimentally measured tracking errors from the ball screw feed system test platform. In Figure 12, the green line represents the experimental results; the blue line corresponds to the prediction model (PM); the red line represents the transfer function (TF); the purple line indicates the deviations of the SM; and the orange line denotes the deviations of the TF. Table 14 lists the corresponding maximum tracking errors and average tracking errors.
Figure 12 and Table 14 indicate that the prediction model exhibits a maximum prediction deviation of 1.8% for the maximum tracking error and 13% for the average tracking error. In contrast, the transfer function demonstrates a maximum prediction deviation of 21% for the maximum and 19% for the average tracking error. These comparisons indicate that the prediction model provides significantly higher accuracy in predicting tracking errors.

5.2.3. Dynamic Error Compensation

Figure 13 presents the experimental tracking errors for the six trajectories under five conditions: without compensation, the proposed PM feedforward compensation method with LuGre friction compensation, and transfer function feedforward compensation with LuGre friction compensation, the proposed PM feedforward compensation method without LuGre friction compensation, and transfer function feedforward compensation without LuGre friction compensation. In Figure 13, the green line represents the experimental results without compensation; the blue line corresponds to the proposed feedforward compensation method and LuGre friction compensation; the red line represents the transfer function feedforward compensation method and LuGre friction compensation; the brown line represents the proposed feedforward compensation method without LuGre friction compensation; the faint yellow line corresponds to the transfer function feedforward compensation method without LuGre friction compensation; the purple line indicates the errors compensated (EC) after applying the proposed compensation method and LuGre friction compensation; the orange line denotes the errors compensated after applying the transfer function feedforward compensation and LuGre friction compensation; the dark green line indicates the errors compensated after applying the proposed compensation method without LuGre friction compensation; and the indigo line denotes the errors compensated after applying the transfer function feedforward compensation without LuGre friction compensation. Table 15 lists the corresponding maximum tracking errors and average tracking errors; Table 16 lists the corresponding maximum tracking errors and average tracking errors without LuGre friction compensation.
Figure 13 and Table 15 demonstrate that the proposed method with LuGre friction compensation achieves a maximum reduction rate of 76.7% for the maximum tracking error and 63.7% for the average tracking error. The transfer function feedforward compensation method with LuGre friction compensation attains a maximum reduction rate of 56.9% for the maximum and 40% for the average tracking error. These comparisons indicate that the proposed compensation method is more effective in reducing tracking errors. As can be seen from Figure 13, Table 15 and Table 16, the friction compensation further reduces the maximum and average tracking errors for the proposed method and the transfer function. Specifically, the maximum tracking error of the proposed method is reduced by 56.1% and the average tracking error by 44.1%, while the transfer function demonstrates a 42.2% reduction in maximum tracking error and a 17.9% decrease in average tracking error.

6. Conclusions

This paper presents a prediction-model-based approach for predicting and compensating dynamic errors in ball screw feed systems. Initially, the lumped mass method and distributed mass method were employed to establish a dynamic model of the ball screw feed system relying on the moving components’ positions, which is then extended to construct an electromechanical coupling model. LHS and the electromechanical coupling model were utilized to generate sample points for training a CFNN prediction model to predict the worktable position and dynamic errors within a control cycle. The simulation and experimental results demonstrate the dynamic error prediction model’s deviations of 1.8% for the maximum tracking error and 13% for the average tracking error, outperforming the transfer function method’s 21% and 19%, respectively. The simulation and experimental results demonstrate the high accuracy of the proposed position-dependent electromechanical coupling model and the dynamic error prediction model. Subsequently, a compensation strategy was proposed. The experimental results indicate that the proposed compensation method with friction compensation achieves maximum reduction rates of 76.7% for the maximum tracking error and 63.7% for the average tracking error, which are better than the transfer function compensation with friction compensation’s 56.9% and 40%, respectively. Comparisons with the transfer function feedforward compensation experiments and uncompensated experiments confirmed the effectiveness of the proposed compensation method.
Future work includes applying transfer learning to the prediction model using experimental data to enhance the accuracy further; extending the prediction model to multi-axis feed systems; and accounting for the thermal error by simultaneously compensating for both thermal and dynamic errors, thereby improving tracking precision.

Author Contributions

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

Funding

This research received no external funding.

Data Availability Statement

The original data presented in the study are openly available in Science Data Bank at https://doi.org/10.57760/sciencedb.23972, accessed on 14 May 2025.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

The single Timoshenko beam element is depicted in Figure A1. l denotes its length; D is the diameter of the beam element’s cross-section. The coupled Timoshenko beam structure is depicted in Figure A2. Here, the X-axis corresponds to the ball screw’s axial direction, while the Y- and Z-axes define its radial plane. Each Timoshenko beam element contains two nodes, with each node possessing six translational displacements (u1, u2, u3, u4, u5, u6) along the X, Y, and Z axes and six rotational displacements (θ1, θ2, θ3, θ4, θ5, θ6) about these axes, totaling 12 DOFs per element. The stiffness matrix component for the beam element, denoted as ktim(i,j), corresponds to the entry in the i-th row and j-th column and is defined as follows:
Figure A1. The single Timoshenko beam element.
Figure A1. The single Timoshenko beam element.
Machines 13 00433 g0a1
Figure A2. Structure and degrees of freedom of the coupled Timoshenko beam element.
Figure A2. Structure and degrees of freedom of the coupled Timoshenko beam element.
Machines 13 00433 g0a2
k t i m 1 , 1 = k t i m 7 , 7 = k t i m 7 , 1 = k t i m 1 , 7 = E A l
k t i m 2 , 2 = k t i m 8 , 8 = 12 E I z 1 + b l 3
k t i m 3 , 3 = k t i m 9 , 9 = 12 E I y 1 + b l 3
k t i m 4 , 4 = k t i m 10 , 10 = G J x l
k t i m 5 , 5 = k t i m 11 , 11 = 4 + b E I y 1 + b l
k t i m 6 , 6 = k t i m 12 , 12 = 4 + b E I z 1 + b l
k t i m 6 , 2 = k t i m 2 , 6 = k t i m 12 , 2 = k t i m 2 , 12 = k t i m 8 , 6 = k t i m 6 , 8 = k t i m 12 , 8 = k t i m 8 , 12 = 6 E I z 1 + b l 2
k t i m 8 , 2 = k t i m 2 , 8 = 12 E I z 1 + b l 3
k t i m 5 , 3 = k t i m 3 , 5 = k t i m 9 , 5 = k t i m 5 , 9 = k t i m 11 , 9 = k t i m 9 , 11 = 6 E I y 1 + b l 2
k t i m 9 , 3 = k t i m 3 , 9 = 12 E I y 1 + b l 3
k t i m 11 , 3 = k t i m 3 , 11 = 6 E I y 1 + b l 3
k t i m 10 , 4 = k t i m 4 , 10 = G J x l
k t i m 11 , 5 = k t i m 5 , 11 = 2 b E I y 1 + b l
k t i m 12 , 6 = k t i m 6 , 12 = 2 b E I z 1 + b l
where G denotes the shear modulus.
G = E 2 1 + u
where E denotes the elastic modulus, u denotes the Poisson ratio, and A denotes the Timoshenko beam element’s cross-sectional area.
A = π D 2 2
J x = I y + I z
Iy and Iz can be calculated using Equation (A18).
I y = I z = π D 4 64
v = 6 1 + u 7 + 6 u
b = 12 E I y v G A l 2
All the other elements are zero. The element in the i-th row and j-th column of the mass matrix of the beam element, denoted as mtim(i,j), is given as follows:
m t i m 1 , 1 = m t i m 7 , 7 = 1 3 ρ A l
m t i m 2 , 2 = m t i m 8 , 8 = A Z ρ A l
m t i m 3 , 3 = m t i m 9 , 9 = A y ρ A l
m t i m 4 , 4 = m t i m 10 , 10 = J x 3 A ρ A l
m t i m 5 , 5 = m t i m 11 , 11 = E y ρ A l
m t i m 6 , 6 = m t i m 12 , 12 = E z ρ A l
m t i m 7 , 1 = m t i m 1 , 7 = 1 6 ρ A l
m t i m 6 , 2 = m t i m 2 , 6 = m t i m 12 , 8 = m t i m 8 , 12 = C z ρ A l
m t i m 8 , 2 = m t i m 2 , 8 = B z ρ A l
m t i m 12 , 2 = m t i m 2 , 12 = m t i m 8 , 6 = m t i m 6 , 8 = D z ρ A l
m t i m 5 , 3 = m t i m 3 , 5 = C y ρ A l
m t i m 9 , 3 = m t i m 3 , 9 = B y ρ A l
m t i m 11 , 3 = m t i m 3 , 11 = m t i m 9 , 5 = m t i m 5 , 9 = D y ρ A l
m t i m 10 , 4 = m t i m 4 , 10 = J x 6 A ρ A l
m t i m 11 , 5 = m t i m 5 , 11 = F y ρ A l
m t i m 12 , 6 = m t i m 6 , 12 = F z ρ A l
All the other elements are zero, where ρ is the density.
A z = 13 35 + 7 b 10 + b 2 3 + 6 5 r z l 2 1 + b 2
A y = 13 35 + 7 b 10 + b 2 3 + 6 5 r y l 2 1 + b 2
B z = 9 70 + 3 b 10 + b 2 6 6 5 r z l 2 1 + b 2
B y = 9 70 + 3 b 10 + b 2 6 6 5 r y l 2 1 + b 2
C z = 11 210 + 11 b 120 + b 2 24 + 1 10 b 2 r z l 2 l 1 + b 2
C y = 11 210 + 11 b 120 + b 2 24 + 1 10 b 2 r y l 2 l 1 + b 2
D z = 13 420 + 3 b 40 + b 2 24 1 10 b 2 r z l 2 l 1 + b 2
D y = 13 420 + 3 b 40 + b 2 24 1 10 b 2 r y l 2 l 1 + b 2
E z = 1 150 + b 60 + b 2 120 2 15 + b 6 + b 2 3 r z l 2 l 2 1 + b 2
E y = 1 150 + b 60 + b 2 120 2 15 + b 6 + b 2 3 r y l 2 l 2 1 + b 2
F z = 1 140 + b 60 + b 2 120 + 1 30 + b 6 b 2 3 r z l 2 l 2 1 + b 2
F y = 1 140 + b 60 + b 2 120 + 1 30 + b 6 b 2 3 r y l 2 l 2 1 + b 2
where
r z = I z A
r y = I y A
Let Ktim and Mtim denote the stiffness matrix and mass matrix of the Timoshenko beam element, respectively. Then,
K t i m = F ( E , u , ρ , l ) M t i m = G ( E , u , ρ , l )
where Ktim and Mtim are 12 × 12 square matrices. As illustrated in Figure A2, their mass and stiffness matrices are 18 × 18 square matrices for the connected two Timoshenko beam elements, corresponding to 18 degrees of freedom. To solve for the elements within these matrices, one may determine each beam element’s stiffness and mass matrices separately. Subsequently, the elements of each beam element’s stiffness and mass matrices are populated into the 18 × 18 square matrix according to the degrees of freedom. For nodes shared by the elements, the principal diagonal elements corresponding to the nodes from each matrix are summed, and the resultant values are then inserted into the 18 × 18 square matrix, as demonstrated in Equations (A52) and (A53).
K t i m = k t i m n ( 1 , 1 ) k t i m n ( 1 , 6 ) k t i m n ( 1 , 7 ) k t i m n ( 1 , 12 ) 0 0 k t i m n ( 6 , 1 ) k t i m n ( 6 , 6 ) k t i m e n ( 6 , 7 ) k t i m n ( 6 , 12 ) 0 0 k t i m n ( 7 , 1 ) k t i m n ( 7 , 6 ) k t i m n ( 7 , 7 ) + k t i m n + 1 ( 1 , 1 ) k t i m n ( 7 , 12 ) + k t i m n + 1 ( 1 , 6 ) k t i m n + 1 ( 1 , 7 ) k t i m n + 1 ( 1 , 12 ) k t i m n ( 12 , 1 ) k t i m n ( 12 , 6 ) k t i m n ( 12 , 7 ) + k t i m n + 1 ( 6 , 1 ) k t i m n ( 12 , 12 ) + k t i m n + 1 ( 6 , 6 ) k t i m n + 1 ( 6 , 7 ) k t i m n + 1 ( 6 , 12 ) 0 0 k t i m i + 1 ( 7 , 1 ) k t i m e n + 1 ( 7 , 6 ) k t i m n + 1 ( 7 , 7 ) k t i m n + 1 ( 7 , 12 ) 0 0 k t i m i + 1 ( 12 , 1 ) k t i m n + 1 ( 12 , 6 ) k t i m n + 1 ( 12 , 7 ) k t i m n + 1 ( 12 , 12 )
M t i m = m t i m n ( 1 , 1 ) m t i m n ( 1 , 6 ) m t i m n ( 1 , 7 ) m t i m n ( 1 , 12 ) 0 0 m t i m n ( 6 , 1 ) m t i m n ( 6 , 6 ) m t i m e n ( 6 , 7 ) m t i m n ( 6 , 12 ) 0 0 m t i m n ( 7 , 1 ) m t i m n ( 7 , 6 ) m t i m n ( 7 , 7 ) + m t i m n + 1 ( 1 , 1 ) m t i m n ( 7 , 12 ) + m t i m n + 1 ( 1 , 6 ) m t i m n + 1 ( 1 , 7 ) m t i m n + 1 ( 1 , 12 ) m t i m n ( 12 , 1 ) m t i m n ( 12 , 6 ) m t i m n ( 12 , 7 ) + m t i m n + 1 ( 6 , 1 ) m t i m n ( 12 , 12 ) + m t i m n + 1 ( 6 , 6 ) m t i m n + 1 ( 6 , 7 ) m t i m n + 1 ( 6 , 12 ) 0 0 m t i m i + 1 ( 7 , 1 ) m t i m e n + 1 ( 7 , 6 ) m t i m n + 1 ( 7 , 7 ) m t i m n + 1 ( 7 , 12 ) 0 0 m t i m i + 1 ( 12 , 1 ) m t i m n + 1 ( 12 , 6 ) m t i m n + 1 ( 12 , 7 ) m t i m n + 1 ( 12 , 12 )
where ktimn(i,j) and mtimn(i,j) denote the elements in the i-th row and j-th column of the stiffness and mass matrices, respectively, for the n-th Timoshenko beam element.

References

  1. Liu, X.; Li, Y.G.; Cheng, Y.G.; Cai, Y. Sparse identification for ball-screw drives considering position-dependent dynamics and nonlinear friction. Robot. Comput.-Integr. Manuf. 2023, 81, 102486. [Google Scholar] [CrossRef]
  2. Zhu, M.; Bao, D.; Huang, X. The Design of a Tracking Controller for Flexible Ball Screw Feed System Based on Integral Sliding Mode Control with a Generalized Extended State Observer. Robot. Actuators 2023, 12, 387. [Google Scholar] [CrossRef]
  3. Zhang, L.; Liu, J.H.; Zhuang, C.B.; Yao, M.Q.; Chen, F.H.; Zhang, C.Y. Gain scheduling control of ball screw feed drives based on linear parameter varying model. Int. J. Adv. Manuf. Technol. 2023, 124, 4493–4510. [Google Scholar] [CrossRef]
  4. Liu, Z.D.; Li, C.Y.; Wang, G.D.; Yang, X.D.; Xu, M.T.; Li, Z.Y.; Li, J.J. Modeling and investigation on dynamic behaviors of ball screw feed system with bolt looseness. Int. J. Nonlinear Mech. 2025, 170, 104996. [Google Scholar] [CrossRef]
  5. Gong, X.X.; Zhao, J.S.; Zhao, C.Y.; Li, C.Y.; Hao, J.; Xu, M.T. Dynamic modeling and nonlinear characteristic analysis of the cross-axis ball screw feed system considering thermal deformation. Nonlinear Dyn. 2025, 113, 14515–14560. [Google Scholar] [CrossRef]
  6. Zhang, H.J.; Zhang, J.; Liu, H.; Liang, T.; Zhao, W.H. Dynamic modeling and analysis of the high-speed ball screw feed system. Proc. Inst. Mech. Eng. Part B-J. Eng. Manuf. 2014, 229, 870–877. [Google Scholar] [CrossRef]
  7. Vicente, D.A.; Hecker, R.L.; Villegas, F.J.; Flores, G.M. Modeling and vibration mode analysis of a ball screw drive. Int. J. Adv. Manuf. Technol. 2012, 58, 257–265. [Google Scholar] [CrossRef]
  8. Lu, H.J.; Rui, X.T.; Ma, Z.Y.; Ding, Y.Y.; Chen, Y.H.; Chang, Y.; Zhang, Y.P. Hybrid multibody system method for the dynamic analysis of an ultra-precision fly-cutting machine tool. Int. J. Mech. Syst. Dyn. 2022, 2, 290–307. [Google Scholar] [CrossRef]
  9. Miao, H.H.; Wang, C.Y.; Hao, J.; Li, C.Y.; Xu, M.T.; Liu, Z.D. Dynamic analysis of the column-spindle system considering the nonlinear characteristics of kinematic joints. Mech. Mach. Theory 2022, 174, 104922. [Google Scholar] [CrossRef]
  10. Hu, Y.J.; Wang, Y.Y.; Zhu, W.D.; Li, H.L. Dynamic Modeling of a Ball-Screw Drive and Identification of Its Installation Parameters. J. Dyn. Syst. Meas. Control-Trans. ASME 2021, 143, 041003. [Google Scholar] [CrossRef]
  11. He, W.W.; Wang, L.P.; Guan, L.W. A novel approach to calculating the dynamic error reflected on an S-shaped test piece. Proc. Inst. Mech. Eng. Part C-J. Eng. Mech. Eng. Sci. 2021, 235, 3273–3298. [Google Scholar] [CrossRef]
  12. Lyu, D.; Liu, Q.; Luo, S.Y.; Wang, D.W.; Liu, H. The influence of dynamic error outside servo-loop on the trajectory error. Int. J. Adv. Manuf. Technol. 2021, 113, 1517–1525. [Google Scholar] [CrossRef]
  13. Yang, L.; Zhang, X.; Wang, L.; Zhao, W.H. Dynamic error of multi-axis machine tools considering position dependent structural dynamics and axis coupling inertial forces. Proc. Inst. Mech. Eng. Part B-J. Eng. Manuf. 2022, 236, 281–295. [Google Scholar] [CrossRef]
  14. Hu, J.J.; Yang, D.Z.; Sun, Z.C. Torque Fluctuation Suppression Strategy of Integrated Electric Drive System Based on the Principle of Minimization of Instantaneous Current Tracking Error. IEEE Access 2022, 10, 124673–124684. [Google Scholar] [CrossRef]
  15. Jiang, J.L.; LI, B.R.; Lin, F.Y.; Zhang, H.; Ye, P.Q. Prediction and compensation strategy of contour error in multi-axis motion system. Int. J. Adv. Manuf. Technol. 2022, 119, 163–175. [Google Scholar] [CrossRef]
  16. Zhang, J.; Li, B.; Zhou, C.X.; Zhao, W.H. Positioning error prediction and compensation of ball screw feed drive system with different mounting conditions. Proc. Inst. Mech. Eng. Part B-J. Eng. Manuf. 2016, 230, 2307–2311. [Google Scholar] [CrossRef]
  17. Liu, Q.; Lu, H.; Yonezawa, H.; Yonezawa, A.; Kajiwara, I.; Jiang, T.; He, J.J. Geometric and dynamic error compensation of dual-drive machine tool based on mechanism-data hybrid method. Mech. Syst. Signal Proc. 2025, 224, 112041. [Google Scholar] [CrossRef]
  18. Liu, C.; Zhao, C.Y.; Wen, B.C. Dynamics analysis on the MDOF model of ball screw feed system considering the assembly error of guide rails. Mech. Syst. Signal Proc. 2022, 178, 109290. [Google Scholar] [CrossRef]
  19. Li, X.W.; Zhang, J.; Zhao, W.H.; Lu, B.H. A zero phase error tracking based path precompensation method for high-speed machining. Proc. Inst. Mech. Eng. Part C-J. Eng. Mech. Eng. Sci. 2016, 230, 230–239. [Google Scholar] [CrossRef]
  20. Zhang, D.L.; Chen, H.C. A data-driven iterative pre-compensation method of contouring error for five-axis machine tools. Int. J. Adv. Manuf. Technol. 2024, 135, 1669–1684. [Google Scholar] [CrossRef]
  21. Wu, H.; Meng, Y.X.; Zhao, Z.Y.; Zhu, Z.W.; Ren, M.J.; Zhang, X.Q.; Zhu, L.M. Surrogate model-based tool trajectory modification for ultra-precision tool servo diamond turning. Precis. Eng.-J. Int. Soc. Precis. Eng. Nanotechnol. 2025, 93, 46–57. [Google Scholar] [CrossRef]
  22. Li, H.C.; Li, Y.F.; Huang, S.M. Optimization strategy for the control of permanent magnet synchronous motors in numerical control machine tools based on intelligent feedforward self-learning. Rev. Adhes. Adhes. 2023, 11, 744–757. [Google Scholar] [CrossRef]
  23. Wang, Z.G.; Feng, X.Y.; Yang, H.T.; Jin, H.W. Research on low-speed characteristics of differential double-drive feed system. Mech. Sci. 2021, 12, 791–802. [Google Scholar] [CrossRef]
  24. Jin, L.; Li, C.H.; Wang, X.; Xie, L.M. A dynamic stiffness model for high-speed ball screw pair with the mass center of screw nut considered. Mech. Ind. 2023, 24, 21. [Google Scholar] [CrossRef]
  25. Li, F.; Jiang, Y.; Li, T.; Ehmann, K.F. Compensation of dynamic mechanical tracking errors in ball screw drives. Mechatronics 2018, 55, 27–37. [Google Scholar] [CrossRef]
  26. Li, X.T.; Xi, D.L.; Zhang, Y.L. Permanent magnet synchronous motor vector control based on MATLAB/Simulink. In Proceedings of the 2017 First International Conference on Electronics Instrumentation & Information Systems (EIIS), Harbin, China, 3–5 June 2017. [Google Scholar] [CrossRef]
  27. Lascu, C.; Trzynadlowski, A.M. A sensorless hybrid DTC drive for high-volume low-cost applications. IEEE Trans. Ind. Electron. 2004, 51, 1048–1055. [Google Scholar] [CrossRef]
  28. Preitl, S.; Precup, R.E. On the algorithmic design of a class of control systems based on providing the symmetry of open-loop Bode plots. Sci. Bull. UPT 1996, 41, 47–55. [Google Scholar]
  29. Meng, D.B.; Li, Y.; He, C.; Guo, J.B.; Lv, Z.Y.; Wu, P. Multidisciplinary design for structural integrity using a collaborative optimization method based on adaptive surrogate modelling. Mater. Des. 2021, 206, 109789. [Google Scholar] [CrossRef]
  30. Jesús, O.D.; Hagan, M.T. Backpropagation algorithms for a broad class of dynamic networks. IEEE Trans. Neural Netw. 2007, 18, 14–27. [Google Scholar] [CrossRef]
  31. Shohda, A.M.A.; Ali, M.A.M.; Ren, G.F.; Kim, J.G.; Mohamed, M.A. Application of Cascade Forward Backpropagation Neural Networks for Selecting Mining Methods. Sustainability 2022, 14, 635. [Google Scholar] [CrossRef]
  32. Nami, F.; Deyhimi, F. Prediction of activity coefficients at infinite dilution for organic solutes in ionic liquids by artificial neural network. J. Chem. Thermodyn. 2011, 43, 22–27. [Google Scholar] [CrossRef]
  33. Yavuz, M.; Iphar, M.; Once, G. The optimum support design selection by using AHP method for the main haulage road in WLC Tuncbilek colliery. Tunn. Undergr. Space Technol. 2008, 23, 111–119. [Google Scholar] [CrossRef]
  34. Zhang, J.S.; Tian, J.L.; Alcaide, A.M.; Leon, J.I.; Vazquez, S.; Franquelo, L.G.; Luo, H.; Yin, S. Lifetime extension approach based on the Levenberg–Marquardt neural network and power routing of DC-DC converters. IEEE Trans. Power Electron. 2023, 38, 10280–10291. [Google Scholar] [CrossRef]
  35. Wang, S.W.; Yuan, Y.; Zhou, A.H.; Yin, C. Environmental Study on Analysis of Characteristic Parameters of Rockfall Movement Based on Field Riprap Test and Establishment of SVM and LM-BPNN Prediction Models. Ekoloji Dergisi 2019, 28, 3319–3326. [Google Scholar]
  36. Dong, P.; Li, J.Q.; Guo, W.; Lai, J.B.; Mao, F.H.; Wang, K.F.; Xu, X.Y.; Wang, S.H. Surrogate-assisted multi-objective optimization for control parameters of adjacent gearshift process with multiple clutches. Control Eng. Practice. 2023, 135, 105519. [Google Scholar] [CrossRef]
  37. Cui, Y.X.; Zhong, B.; Qian, W.K.; Zheng, C.J. Identification of Time Series Transfer Function Parmeter. In Proceedings of the 2014 International Conference on Mechatronics Electronic industrial and Control Engineering (MEIC 2014), Shenyang, China, 15–17 November 2014. [Google Scholar] [CrossRef]
Figure 1. Overview of dynamic error prediction and compensation methods for the ball screw feed system.
Figure 1. Overview of dynamic error prediction and compensation methods for the ball screw feed system.
Machines 13 00433 g001
Figure 2. Ball screw feed system: (a) ball screw feed system’s configuration; (b) real ball screw feed system.
Figure 2. Ball screw feed system: (a) ball screw feed system’s configuration; (b) real ball screw feed system.
Machines 13 00433 g002
Figure 3. Two-element Timoshenko beam after length adjustment.
Figure 3. Two-element Timoshenko beam after length adjustment.
Machines 13 00433 g003
Figure 4. Position-dependent dynamic model of the ball screw feed system.
Figure 4. Position-dependent dynamic model of the ball screw feed system.
Machines 13 00433 g004
Figure 5. Electromechanical coupling model of the ball screw feed system.
Figure 5. Electromechanical coupling model of the ball screw feed system.
Machines 13 00433 g005
Figure 6. The principle of predicting the position of the worktable of one control cycle.
Figure 6. The principle of predicting the position of the worktable of one control cycle.
Machines 13 00433 g006
Figure 7. The Cascade-Forward Neural Network.
Figure 7. The Cascade-Forward Neural Network.
Machines 13 00433 g007
Figure 8. Feedforward compensation structure.
Figure 8. Feedforward compensation structure.
Machines 13 00433 g008
Figure 9. The ball screw feed system experimental platform.
Figure 9. The ball screw feed system experimental platform.
Machines 13 00433 g009
Figure 10. S-curve velocity interpolation trajectories: (a) Velocity; (b) acceleration; (c) jerk.
Figure 10. S-curve velocity interpolation trajectories: (a) Velocity; (b) acceleration; (c) jerk.
Machines 13 00433 g010
Figure 11. Electromechanical coupling simulation and experimental tracking errors: (a) Trajectory 1; (b) Trajectory 2; (c) Trajectory 3; (d) Trajectory 4; (e) Trajectory 5; (f) Trajectory 6.
Figure 11. Electromechanical coupling simulation and experimental tracking errors: (a) Trajectory 1; (b) Trajectory 2; (c) Trajectory 3; (d) Trajectory 4; (e) Trajectory 5; (f) Trajectory 6.
Machines 13 00433 g011
Figure 12. Tracking error prediction simulation and experimental tracking error: (a) Trajectory 1; (b) Trajectory 2; (c) Trajectory 3; (d) Trajectory 4; (e) Trajectory 5; (f) Trajectory 6.
Figure 12. Tracking error prediction simulation and experimental tracking error: (a) Trajectory 1; (b) Trajectory 2; (c) Trajectory 3; (d) Trajectory 4; (e) Trajectory 5; (f) Trajectory 6.
Machines 13 00433 g012
Figure 13. Tracking error compensation experiment: (a) Trajectory 1; (b) Trajectory 2; (c) Trajectory 3; (d) Trajectory 4; (e) Trajectory 5; (f) Trajectory 6.
Figure 13. Tracking error compensation experiment: (a) Trajectory 1; (b) Trajectory 2; (c) Trajectory 3; (d) Trajectory 4; (e) Trajectory 5; (f) Trajectory 6.
Machines 13 00433 g013
Table 1. Identification results.
Table 1. Identification results.
ParametersTC (N·mm)TS (N·mm)ΩS (rpm)σ0(N)σ1 (N·mm/rpm)σ2 (N·mm/rpm)
value0.67570.765325.91471.58750.15930.0003
Table 2. The operational specifications of the PMSM.
Table 2. The operational specifications of the PMSM.
Operational SpecificationsValue
Rotor inertia0.0053 (kg·m2)
Torque constant1.641 (N·m/A)
Stator resistance0.75 (ohm)
The inductance of the quadrature axis13.2 (mH)
The inductance of the direct axis4.11 (mH)
Rated torque24 (N·m)
Rated current10 (A)
Rated speed2000 (RPM)
Coil pitch4
Silicon steel gradesM270-35A
Permanent magnet gradesN42UH
Table 3. The operational specifications of the ball screw feed system.
Table 3. The operational specifications of the ball screw feed system.
Operational SpecificationsValue
Ball screw manufacturerHIWIN
Ball screw catalogFDI 40-10T3
Bed material gradeHT200
Worktable material gradeHT200
Rotary table material gradeQT450-10
Torque sensor shaft stiffness1 × 105 (N·m/rad)
Coupling torsional stiffness2800 (N·m/rad)
Screw–nut pair axial stiffness7.45 × 107 (N/m)
1.04 × 107 (N/m)
Table 4. Control parameters.
Table 4. Control parameters.
PIdIIdPIqIIqPωRIωRPSw
2.776626.28.8051003.30.06780.7122280.2
Table 5. The design ranges of the four parameters.
Table 5. The design ranges of the four parameters.
ParametersLower BoundUpper Bound
Sw(k) (mm)01500
S ˙ w k (mm/s)−333333
S ¨ w k (mm/s2)−10001000
Iqr(k) (A)−1515
Table 6. The parts of the sample point set.
Table 6. The parts of the sample point set.
TrialSw(k) S ˙ w k S ¨ w k Iqr(k)Sw(k+1)
141.71−270.9−690.10.17141.68
21.5276.9−830.1−14.821.472
4997349.5104.4330.4.707349.6
4998151.5−292.3350.18.938151.6
4999954.5161.7390.16.604954.4
5000407.115.54270−13.94407.2
Table 7. Maximum and average AE of worktable position for different u.
Table 7. Maximum and average AE of worktable position for different u.
u = 0.3u = 0.5u = 0.7
Maximum AE (mm)118.940.3925.62
Average AE (mm)15.085.5592.622
Table 8. Maximum and average AE of worktable position for different N.
Table 8. Maximum and average AE of worktable position for different N.
N = 1N = 2N = 3
Maximum AE (mm)118.926.6215.23
Average AE (mm)15.081.6740.631
Table 9. Maximum and average AE of worktable position for different Mh.
Table 9. Maximum and average AE of worktable position for different Mh.
Mh = 3Mh = 4Mh = 5
Maximum AE (mm)15.238.5585.545
Average AE (mm)0.6310.3090.112
Table 10. Maximum and average AE of worktable position for different epochs.
Table 10. Maximum and average AE of worktable position for different epochs.
epochs = 100epochs = 150epochs = 200
Maximum AE(μm)15.232010.76957.0424
Average AE (mm)0.63140.48800.3277
Table 11. Maximum and average AE of the trained CFNN.
Table 11. Maximum and average AE of the trained CFNN.
Maximum AE (μm)Average AE (μm)
9.549 × 10−92.572 × 10−9
Table 12. The parameters of the six S-curve velocity interpolation trajectories.
Table 12. The parameters of the six S-curve velocity interpolation trajectories.
TrajectoriesDisplacementMaximum
Velocity
Maximum
Acceleration
Maximum Jerk
13001002251200
23001002251400
33001001201000
43001001401000
53001101601000
63001201601000
Table 13. Maximum/average tracking errors of electromechanical coupling simulation and experiment.
Table 13. Maximum/average tracking errors of electromechanical coupling simulation and experiment.
TrajectoriesExperimentPosition-Dependent
Maximum
Tracking Error (μm)
Average
Tracking Error (μm)
Maximum
Tracking Error (μm)
Average
Tracking Error (μm)
154.866.75354.755.757
258.357.01357.35.848
351.985.05651.644.482
451.895.27951.644.771
551.086.3851.645.552
651.916.71951.646.052
Table 14. Maximum/average tracking error prediction simulation and experiment.
Table 14. Maximum/average tracking error prediction simulation and experiment.
TrajectoriesExperimentPrediction ModelTransfer Function
Maximum
Tracking Error (μm)
Average
Tracking Error (μm)
Maximum
Tracking Error (μm)
Average
Tracking Error (μm)
Maximum
Tracking Error (μm)
Average
Tracking Error (μm)
154.866.80454.756.57464.757.573
258.357.09857.36.17367.757.694
351.985.05651.644.70961.795.878
451.895.27951.645.09861.796.261
551.086.41451.645.86261.797.313
651.916.71951.646.40761.797.995
Table 15. Maximum/average tracking error for tracking error compensation experiment.
Table 15. Maximum/average tracking error for tracking error compensation experiment.
TrajectoriesNo CompensationPM+ LuGreTF+ LuGre
Maximum
Tracking Error (μm)
Average
Tracking Error (μm)
Maximum
Tracking Error (μm)
Average
Tracking Error (μm)
Maximum
Tracking Error (μm)
Average
Tracking Error (μm)
154.866.80417.072.70525.214.33
258.357.09815.322.78127.924.63
351.985.05617.192.27523.393.28
451.895.27919.072.33122.353.22
551.086.41411.882.31725.34.076
651.916.71912.792.51224.524.032
Table 16. Maximum/average tracking error for the tracking error compensation experiment without friction compensation.
Table 16. Maximum/average tracking error for the tracking error compensation experiment without friction compensation.
TrajectoriesPMTF
Maximum
Tracking Error (μm)
Average
Tracking Error (μm)
Maximum
Tracking Error (μm)
Average
Tracking Error (μm)
130.094.44343.675.177
233.214.40441.225.183
327.393.33736.503.666
427.163.43236.013.872
527.104.15238.204.860
628.594.34038.844.912
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, H.; Guo, Y.; Liu, J.; Niu, W. Dynamic Error Compensation for Ball Screw Feed Drive Systems Based on Prediction Model. Machines 2025, 13, 433. https://doi.org/10.3390/machines13050433

AMA Style

Liu H, Guo Y, Liu J, Niu W. Dynamic Error Compensation for Ball Screw Feed Drive Systems Based on Prediction Model. Machines. 2025; 13(5):433. https://doi.org/10.3390/machines13050433

Chicago/Turabian Style

Liu, Hongda, Yonghao Guo, Jiaming Liu, and Wentie Niu. 2025. "Dynamic Error Compensation for Ball Screw Feed Drive Systems Based on Prediction Model" Machines 13, no. 5: 433. https://doi.org/10.3390/machines13050433

APA Style

Liu, H., Guo, Y., Liu, J., & Niu, W. (2025). Dynamic Error Compensation for Ball Screw Feed Drive Systems Based on Prediction Model. Machines, 13(5), 433. https://doi.org/10.3390/machines13050433

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