Robotic tasks often exceed the scope of steady-state or periodic behavior, which necessitates generally-applicable models of actuators intended to generate transient or aperiodic motion. However, existing electromechanical models of servomotors typically omit consideration of the switching power converter circuits required for directional, speed, or torque control. In this study, a multi-domain framework is established for switched electromechanical dynamics in servomotor systems for their analysis and control in general aperiodic tasks including transient phases. The switched electromechanical dynamics is derived from the individual models of the internal DC motor, gear train, and H-bridge circuit. The coupled models comprehensively integrate all possible distinct switching configurations of on-state, off-state, and dead time. A combination of cycle averaging with piecewise analytical solutions of the non-smooth dynamics is introduced to handle different temporal scales from high-frequency electrical to low-frequency mechanical variables. System parameters were estimated from experimental data using a dual-servomotor test platform. The model was validated for predictive accuracy against measured data in two distinct tasks—dynamic braking of a pendulum system and sinusoidal trajectory following. The model was also used to formulate the servomotor power consumption, which was implemented for optimal control demonstration and energy analysis. In particular, the servomotor power consumption model provided true optimality (minimization) when compared with the squared rotor torque and the positive rotor mechanical power that are commonly used as proxy models. While the focus of this work is on permanent-magnet, armature-controlled brushed DC servomotors, the approach is applicable to general electromechanical systems with switching-based control.
Generally-applicable models of robot actuators are needed to consider the full range of robotic tasks including transient or aperiodic motion. So far, models of servomotors commonly used in robots typically do not include the switching circuits needed for directional, speed, or torque control. Here, the complete switched electromechanical dynamics of permanent-magnet, armature-controlled brushed DC servomotors are derived from the individual models of their key components, which are the internal DC motor, gear train, and H-bridge circuit, while including the effects of switching by considering all possible distinct switching configurations during operation. The non-smoothness and different temporal scales within the coupled dynamics, from high-frequency electrical to low-frequency mechanical variables, are handled with a combination of cycle averaging and piecewise analytical solutions. The model's predictions were validated in two distinct tasks—dynamic braking and sinusoidal trajectory following. The comprehensive model also allowed for accurate prediction of the servomotor power consumption and demonstrated its merits as a cost function for minimization in optimal control when compared with other common proxy models.
I. INTRODUCTION
In typical robot operations, the direct control input to actuators is a varying voltage profile, while the input from the external environment is provided as position/velocity or force/torque feedback. This combination of electromagnetic and mechanical variables, which each exhibits their own characteristic dynamics, requires special consideration in modeling and control. However, typical modeling approaches for electromechanical dynamics consider steady-state and/or periodic phases or use system parameters that are estimated for such conditions.1,2 As robotic tasks often exceed the scope of steady-state or periodic behavior,3,4 generally-applicable models are needed specifically for actuators intended to generate transient or aperiodic motion.
Modeling electromechanical systems that cross multiple domains as part of their operation requires the consideration of all their relevant components. Yet, existing electromechanical models of robot actuators typically consider only the DC motor5–10 because its dynamics is well understood.11 However, a simple DC motor has no inherent mechanism for directional, speed, or torque control. In practice, such control mechanisms are provided by power converter circuits that modulate the voltage applied across the motor terminals. Phase-controlled converters are used to control brushless DC motors, while chopper converters are commonly used to control brushed DC motors through pulse-width modulation (PWM), in which the mean voltage applied at the motor terminals is adjusted through rapid switching of the DC input voltage from the power supply.12,13
The combination of a motor, gear train, power converter, and feedback control mechanism is referred to as a servomotor. Due to their simple design and dynamic performance capable of rapid acceleration and deceleration, DC servomotors with chopper converters are used as actuators in a broad range of applications.11 The ubiquity of chopper converters and similar switching devices for control in electronics allows for ease of implementation but also introduces non-smooth switched dynamics at the electromagnetic level. Existing electromechanical models of the DC servomotor derived for use in control1,14,15 typically do not consider all possible switching configurations of the chopper converter during operation. Within mechanical components, such as the gear train, there are the direction-dependent effects of back-drivability16 and friction,17,18 which are additional non-smooth aspects.
Analytical approaches to electromechanical modeling typically assume that the switched electromagnetic dynamics are sufficiently fast relative to the mechanical dynamics, in terms of response to varying inputs, such that the electromagnetic and mechanical response can be fully or partially decoupled, as can be demonstrated by singular perturbation theory.19 The switched dynamics of the power converter is often modeled with an averaging approach (as opposed to data-driven methods, such as model-free reinforcement learning20) that obtains averaged electrical state trajectories from mechanical state trajectories with harmonics of much lower frequencies than the switching frequency. For sufficiently high switching frequencies, such as those used for PWM, the characteristics of the averaged dynamics are similar to those of the full-order switched dynamics. For example, the control designs that stabilize an averaged system will also stabilize the original non-averaged system.21,22
State-space averaging is often applied to obtain the transfer functions of switching power converters. By taking a weighted average of the state equations of each switching configuration according to their relative time durations, small-signal and steady-state models of the circuit can be obtained under the assumption that the state variables are perturbed around some steady-state operating point.12 Variations of this method include corrected full-order state-space averaged modeling,23 which is the most accurate method for both low and high switching frequencies, and parametric average-value modeling,24,25 which is a computational method that extends state-space averaging to complex systems that are difficult to solve analytically.
In circuit averaging, the input and output variables of each circuit element are averaged across one switching cycle to derive its cycle-averaged input–output relations instead of averaging the state equations directly. The switching power converter is then modeled as an equivalent circuit composed of averaged circuit elements.26 Other circuit-averaging methods include the injected-absorbed-current method, which is mostly suitable for low-frequency, small-signal phenomena.27
Averaging approaches are sufficient in applications where only low-frequency behavior is considered, such as typical trajectory tracking control. Modeling averaged rather than full-order switched dynamics also reduces the computational effort because less dense discretization of the solution time interval is required for convergence. However, any high-frequency oscillatory effects in the state variables, such as ripple current, must be recovered from the averaged model with estimation techniques that result in some degree of error.28,29 For instance, for the circuit-averaging approach, additional relations, such as the energy conservation principle, must be introduced to recover the ripple current from the cycle-averaged current profile in order to model power consumption.30–32 The use of linearization in deriving averaged dynamics12,33,34 also results in discrepancies against the full-order nonlinear dynamics. Thus, there remains a need for accurate multi-domain modeling of switched electromechanical systems integrated with averaging approaches to pursue more sophisticated control, especially for transient motion profiles, despite the well-established use of existing models.1,35
Averaging switched electromechanical dynamics also allows for the modeling of power consumption in general complex tasks. Energy usage is one of the most important criteria for mobile3,36 or legged robots5,8,14,37–43 (e.g., to conserve energy through real-time task management44) as well as for manipulators4,6,7,10,15,45 to resolve mechanical redundancy. Despite the need for accurate mathematical models to predict and minimize power consumption, broadly applicable (including transient phases and aperiodic tasks), physically-based mathematical models of power consumption are lacking. While bond-graph techniques37,46,47 provide tools for analyzing interconnections among the multi-domain elements of a system, the specific form of each model term is still required for the quantification of power consumption. Mathematical models of robot power consumption are usually required to be functions of kinematic (joint angles and velocities) and dynamic (joint torques, external forces) variables in the mechanical domain so that they can be used in prediction, design, control, and optimization. The current state of knowledge still relies on ad hoc, empirically-based models calibrated to specific instances, and incomplete proxies like mechanical energy or work,38,40 square of actuator torques or forces,3,45,48–50 or their combinations with other terms36,37,51 are often used to model power consumption instead. While the mechanical power alone may provide some implications about power consumption under periodic or steady-state conditions,38 and to some extent, during positive mechanical-work phases,43,52 it is different from the true power consumption of a system.16
The aforementioned proxies of power consumption, which are overly simplified, generally cannot differentiate between the significance of positive, zero, and negative output mechanical work or the power consumption of different switching configurations in a controlled system. Actuator torque or its squared value, for instance, lacks the velocity-dependence expected for power consumption and does not imply the actual quantity. When a more complete model of electrical power consumption is considered in the literature,10,15,39 there is often a condition imposed that omits the energetic effect of negative mechanical or electrical work,5,16,43 which may be inconsistent with experimental results.6,14 Data-driven approaches,36 which include, for instance, the use of reinforcement learning for improving locomotion efficiency,42 simply sidestep the complexities of multi-domain interactions by collecting more experimental data. While some of the existing approximations may provide acceptable trends for well-defined tasks that are periodic or steady-state, generally applicable models are required for general complex tasks for broader robot applications. Most recent works47,53–56 along this line are focused on predicting the power consumption of electric vehicles, for which the models are relatively successful in integrating the drive cycle of actual vehicles. However, they present similar issues as described above, i.e., lack of generality and accuracy due to over-simplification.
In this study, a multi-domain framework is established for switched electromechanical dynamics in servomotor systems for their analysis and control, in general, aperiodic tasks, including transient phases. While the approach is applicable to general electromechanical systems with switching-based control, the focus of this work is on permanent-magnet, armature-controlled brushed DC servomotors due to their ubiquity as actuators in robotics, in addition to their simple designs without field windings,57 a wide range of torque and speed controllability with identifiable torque-speed characteristics, and compatibility with various control methods.58 The complex system-of-systems challenge is addressed in a principled, step-by-step approach where the switched electromechanical dynamics is derived from the individual models of the internal DC motor, gear train, and H-bridge circuit. The coupled models comprehensively integrate all possible distinct switching configurations of the H-bridge and are cycle-averaged to handle non-smoothness and different temporal scales between electrical and mechanical variables. Experiments were conducted to estimate the system parameters and validate the model's prediction accuracy in two distinct tasks—dynamic braking and sinusoidal trajectory following. The model was also used to formulate the servomotor power consumption, which was implemented for optimal control and energy analysis and demonstrated for its merits against other common proxy models.
II. SWITCHED DYNAMICS OF SERVOMOTOR ELECTROMECHANICS
The main components of the DC servomotor (Fig. 1) are the armature DC motor, gear train, full-bridge inverter (referred to here as the H-bridge), microcontroller, and sensors.11–13 The system model and the corresponding system parameters are obtained here for the DYNAMIXEL MX-28 servomotor (ROBOTIS Co. Ltd, South Korea),59 which is used in many robotic systems,60–63 but the proposed models apply to general permanent-magnet, armature-controlled brushed DC servomotors due to their common operating principle. To obtain the full state model, the state equations are derived separately for each component and then integrated in terms of electrical and mechanical state variables, where the H-bridge and the DC motor are modeled with equivalent circuits. The electrical power supply is modeled as an ideal voltage source with constant input voltage Vin, where the input current Iin from the power supply is positive in the direction of positive Vin. Due to the constant input voltage, the capacitor typically in parallel with Vin in the power supply does not affect the servomotor dynamics. The electrical current and power consumption drawn by the sensors and microcontroller16 are not considered here and are assumed to be constant or otherwise motion- and actuation-independent. It is further assumed that the entire system of interest is isolated from any external electromagnetic field effects.
Servomotor system diagram: components (DC motor, gear train, and H-bridge) to be modeled, and other internal (microcontroller and sensors) and external systems (power supply and computer).
Servomotor system diagram: components (DC motor, gear train, and H-bridge) to be modeled, and other internal (microcontroller and sensors) and external systems (power supply and computer).
A. DC motor and gear train
The classic equivalent circuit model of the permanent-magnet, armature-controlled brushed DC motor consists of the inductor, resistor, and back electromotive force (back emf) elements in series7,8,9,11 (Fig. 1). The voltage VPWM,H(t) at time t applied by the H-bridge to the DC motor is
where L is the motor inductance, Ra is the armature resistance, (from Faraday's law) is the back emf voltage with motor torque constant K and rotor shaft angle qrot, Ia is the motor armature current, and the dot indicates time-derivative. The voltage drop of magnitude Vbr due to brush-contact resistance associated with slip rings and commutators, which changes sign when the motor current is reversed,14 is also included in series with the other elements as an additional non-smooth term. Ferromagnetic coupling losses due to eddy current or hysteresis that are usually neglected in control57,64 and a possible capacitor in parallel with the DC motor in certain designs to suppress electrical noise1 are not considered here.
The DC motor is attached to a compound reduction gear train of NG gears, where the ith gear shaft angle qG,i for i = 1 and i = NG are the rotor shaft angle qrot and the servomotor output shaft angle q, respectively. Under a rigid-body assumption, the equivalent moment of inertia inclusive of all rotating mass (rotor inertia, gear train, and output shaft) as reflected on the servomotor output shaft angle is
where is the signed gear ratio,11 with for brevity, and Ji is the axial moment of inertia of the gear shaft corresponding to qG,i.
The equation of motion of the combined system of the DC motor and the gear train is
where (from Lorentz's law) is the rotor torque applied by the stator on the rotor of the motor (where K is assumed to be identical to that used for the back emf), is the effective frictional torque (resultant applied by the servomotor housing on gear shaft axles and due to gear contacts) as reflected to the output shaft, and is the output torque applied by the servomotor on the external environment through the output shaft. The non-smooth model of frictional torque is for and for , where b0 and b1 are the Coulomb and viscous friction coefficients, respectively, and are negative in sign. Both the load-dependence of and the break-away friction are not considered here.
B. H-bridge switching configurations
The H-bridge, which consists of four switches (Fig. 2), allows for bidirectional control through PWM. By only closing switches S1 and S4, S2 and S4, and S2 and S3, a voltage of approximately Vin, 0, −Vin is applied to the DC motor, respectively. These three switching configurations are the positive on-state, off-state, and negative on-state, respectively (Table I). By varying the switch timing, a desired variable voltage profile for VPWM,H(t) within the interval [−Vin, Vin] is approximated with a binary or ternary voltage profile with the values Vin, 0, or −Vin that resembles a series of variable-width pulses occurring at a fixed frequency 1/TPWM, where TPWM is the fixed pulse period.65
H-bridge shown with numbered switches (left). Timing diagram depicts the quantized signed duty cycle D(t) and the carrier signal c(t) (center), along with a magnified view indicating the on-state, off-state, and dead time for unipolar leading-edge PWM in the case of Ia(t) < 0 during [t0, t1) and Ia(t) > 0 during [t2, t3) (right). The voltage shown is approximated.
H-bridge shown with numbered switches (left). Timing diagram depicts the quantized signed duty cycle D(t) and the carrier signal c(t) (center), along with a magnified view indicating the on-state, off-state, and dead time for unipolar leading-edge PWM in the case of Ia(t) < 0 during [t0, t1) and Ia(t) > 0 during [t2, t3) (right). The voltage shown is approximated.
H-bridge switching configurations. The switching configuration of D(t) = 0 and the off-state of non-zero D(t) are identical.
. | D(t) > 0 . | D(t) = 0 . | D(t) < 0 . | ||||||
---|---|---|---|---|---|---|---|---|---|
. | . | . | . | . | . | ||||
Switch . | [t0, t1) Dead Time . | [t1, t2) On-State . | [t2, t3) Dead Time . | [t3, t4) Off-State . | [t0, t4) Off-State . | [t0, t1) Dead Time . | [t1, t2) On-State . | [t2, t3) Dead Time . | [t3, t4) Off-State . |
S1 | OFF | ON | OFF | OFF | OFF | OFF | OFF | OFF | OFF |
S2 | OFF | OFF | OFF | ON | ON | ON | ON | ON | ON |
S3 | OFF | OFF | OFF | OFF | OFF | OFF | ON | OFF | OFF |
S4 | ON | ON | ON | ON | ON | OFF | OFF | OFF | ON |
. | D(t) > 0 . | D(t) = 0 . | D(t) < 0 . | ||||||
---|---|---|---|---|---|---|---|---|---|
. | . | . | . | . | . | ||||
Switch . | [t0, t1) Dead Time . | [t1, t2) On-State . | [t2, t3) Dead Time . | [t3, t4) Off-State . | [t0, t4) Off-State . | [t0, t1) Dead Time . | [t1, t2) On-State . | [t2, t3) Dead Time . | [t3, t4) Off-State . |
S1 | OFF | ON | OFF | OFF | OFF | OFF | OFF | OFF | OFF |
S2 | OFF | OFF | OFF | ON | ON | ON | ON | ON | ON |
S3 | OFF | OFF | OFF | OFF | OFF | OFF | ON | OFF | OFF |
S4 | ON | ON | ON | ON | ON | OFF | OFF | OFF | ON |
The direct control input to the DC servomotor is represented by the signed duty cycle function , where the sign and magnitude indicate the direction and the ratio of the on-state duration to TPWM of each pulse period, respectively11 (Fig. 2). In digital PWM implementations, D(t) is quantized. The microcontroller outputs the PWM voltage to drive the H-bridge, where VM is the microcontroller supply voltage (constant), H is the Heaviside unit step function defined such that H(0) = 0, and is the carrier signal for the selected PWM mode. The signal c(t), which is commonly a sawtooth or triangular wave, has the same frequency as D(t) such that H(|D(t)| − c(t)) is a series of pulses of width |D(t)|TPWM.12 The duty cycle D(t) then determines the switching state of the H-bridge as a function of time, regulated by VPWM,M(t) as communicated through the microcontroller. Although the state equations are derived here specifically for a unipolar leading-edge PWM mode, where an on-state of variable width precedes the off-state with a zero voltage,65 the same modeling process can be followed when considering a general servomotor that uses any variations of switching configurations for control.
In addition to the on- and off-states, there exist intermediate switching states where only switch S2 or switch S4 is closed (Table I). These intermediate switching states occur during dead time, also known as blanking time, and are introduced by design to avoid shoot-through when a short-circuit is formed by accidentally closing switches S1 and S2 or S3 and S4 together while toggling between the on- and off-states.13 Dead time occurs twice during a pulse period with a fixed dead-time duration TDT each instance. Thus, for a given pulse beginning at t = t0, the pulse period interval [t0, t4) is divided into the first dead time [t0, t1), on-state [t1, t2), second dead time [t2, t3), and off-state [t3, t4), where t1 = t0 + TDT, t2 = t0 + |D(t0)|TPWM, t3 = t0 + |D(t0)|TPWM + TDT, and t4 = t0 + TPWM (Fig. 2).
C. State models of coupled electromechanical dynamics
Ideal instantaneous switching is assumed, where the time to turn the switch on and off, which are the rise and fall time, respectively,12 is assumed to be zero. Commonly in H-bridges, each switch is a metal–oxide–semiconductor field-effect transistor (MOSFET), which is modeled as a switch with an on-state resistance Rtrans in parallel with the body diode of the transistor (Fig. 1). The diode, which is referred to as a freewheel or catch diode in the context of the H-bridge, provides a path for Ia to flow during dead time in order to avoid voltage spikes due to the motor inductance and to protect the transistors and is modeled as an ideal diode in series with the forward voltage VD and diode resistance RD.13
The H-bridge state equations during the on-state, off-state, and dead time are derived from Kirchhoff's voltage and current laws (KVL and KCL) considering the equivalent circuit model of the H-bridge in each switching configuration (Table I and Fig. 3).
H-bridge switching configuration conduction paths. For dead time, the direction of Ia is indicated with an arrow.
H-bridge switching configuration conduction paths. For dead time, the direction of Ia is indicated with an arrow.
For on-state () and off-state ():
For dead time ():
During dead time, the sign of Ia also determines the path of Ia. For instance, when D(t) > 0 and Ia(t) < 0, current is conducted through switch S4, the diode in parallel with switch S1, and the power supply with non-zero Iin(t) (Fig. 3). If D(t) > 0 and Ia(t) > 0, current is conducted through switch S4 and the diode in parallel with switch S2 but not through the power supply [Iin(t) = 0] because the diode in parallel with switch S3 is reverse-biased.13
The coupled state equations of the servomotor electromechanical dynamics are the DC motor and gear train equation of motion and the H-bridge circuit equations, with the voltage in terms of Vin and D(t).
On-state () and off-state ():
Dead time ():
At the boundary of the servomotor system of interest for this study, the mechanical variables are q(t), , and their time-derivatives, and the electrical variables are D(t), Iin(t) [dependent on Ia(t) and D(t)], and their time-derivatives, while Vin is constant.
III. CYCLE-AVERAGING OF SWITCHED ELECTROMECHANICAL DYNAMICS
In the proposed approach, the instantaneous state variables are cycle-averaged,12,13,26,57 as in a typical circuit-averaging approach, and their relations are derived. Given a function f(t) of time , which may represent a state or control, its cycle-averaged value can be regarded as a constant function over the given pulse period (Fig. 4). Here, the set of instantaneous and cycle-averaged f(t) values sampled at discretized time points across various pulse periods, which can be non-consecutive or unevenly spaced, are denoted by f and , respectively. For sufficiently high 1/TPWM, the mechanical state trajectories , , , and because their variations over the pulse period are negligible relative to the variations of the electrical state variables, which typically have a much smaller time constant than that of the mechanical state variables.58,64 Due to the quantized nature of D(t) in digital control, exactly (Fig. 2).
A sample trajectory of typical Ia(t) resulting from switching and a magnified view for a selected pulse period.
A sample trajectory of typical Ia(t) resulting from switching and a magnified view for a selected pulse period.
In this framework, the discretized mechanical state variable trajectories q (and its derivatives) and are mapped to the corresponding electrical state variable trajectories and D. First, the set of instantaneous coupled dynamics [Eqs. (3) and (6)] are cycle-averaged to obtain the state equations for , which are derived without the consideration of dead time due to its short [<5% TPWM by design in order to minimize the dead-time distortion of VPWM,H(t) during operation13] duration relative to TPWM:
where the approximation is used to interchange the order of differentiation and cycle-averaging operations.27 Given and D(t) = D(t0) for , it can be shown that and . In these cycle-averaged coupled state equations for , the effect of ripple current appears only in the term .
The state equation for is derived in terms of to include the effects of ripple current. Here, the dead time is included to account for instances of low |D(t) |, in which case TDT is not negligible with respect to the on-state duration . For the case of Vbr = 0, the instantaneous coupled dynamics [Eqs. (6) and (7)] are those of switched RL circuits and the corresponding analytical solution for Ia(t) for is a piecewise function during the on-state, off-state, and dead time.
For dead time ():
For on-state ():
For dead time ():
For off-state ():
with the values of Ia(t1), Ia(t2), and Ia(t3) satisfying the continuity between the piecewise solutions and written in terms of Ia(t0).
Due to the brief duration of dead time, the signs of Ia(t) during and are assumed to remain constant and equal to their values at t0 and t2, respectively, resulting in four possible combinations for the signs of Ia(t0) and Ia(t2). The correct combination of signs is identified when the signs of Ia(t0) and Ia(t2) resulting from Ia(t) agree with their pre-assumed combination.
Using the analytical forms of Ia(t), for can be written as a first-order form with respect to Ia(t0):
where and are functions of D(t0), , and system parameters (G, K, L, Ra, RD, Rtrans, Vin, VD, TDT, and TPWM), obtained through analytical integration depending on pre-assumed signs of Ia(t0) and Ia(t2).
From the instantaneous relations [Eqs. (4) and (5)] between Iin(t) and Ia(t) during the on-state and dead time and the analytical solution of Ia(t), for :
where and are functions of D(t0), , and the system parameters, and depend on pre-assumed signs of Ia(t0) and Ia(t2). At discretized time points across pulse periods, using Eqs. (8) and (14),
For the case of non-zero Vbr, the process of obtaining Ia(t) requires solving for the times and/or satisfying Ia(t2−) = 0 and/or Ia(t2+) = 0, respectively, when the sign of Ia(t) changes. If no sign change occurs during and/or , then these are set as t2− = t0 and/or t2+ = t4, respectively. Note that if the sign of Ia(t) never changes within the period, then t2+−t2− = TPWM. For each possible case of t2− and t2+, the proper piecewise form of Ia(t) is used. For example, if the sign of Ia(t) changes both during the on- and off-states (i.e., and ), Ia(t) consists of six segments defined over the time intervals [t0, t1), [t1, t2−), [t2−, t2), [t2, t3), [t3, t2+), and [t2+, t4) (Fig. 4). Solving for Ia(t) with the assumption of Vbr = 0 provides an initial approximation for t2− and t2+ as the times where Ia(t) = 0. The approximated t2− and t2+ are used to re-evaluate the cycle-averaged state equations with non-zero Vbr [Eqs. (8) and (9)] to update and D(t0) in order to obtain Ia(t). Since sgn(Ia(t)) is equal to sgn(Ia(t2)) for , zero for t = t2− and t = t2+, and −sgn(Ia(t2)) for the remainder of , for . If needed, t2− and t2+ of the updated Ia(t) can be used to update and D(t0), and the process can be repeated until a desired accuracy is reached.
In contrast to typical circuit-averaging approaches, the cycle-averaging approach here comprehensively considers the effect of all switching configurations, including dead time. In addition, the effect of ripple current is incorporated by analytically obtaining from at any set of discretized time points, which requires less computational effort than numerically solving the high-frequency switched dynamics at a denser time-discretization, without requiring additional relations (e.g., energy conservation) or approximate ripple reconstruction algorithms. On the other hand, the analytical solution Ia(t) may not be continuous for consecutive pulse periods as a result of the lack of continuity conditions imposed and the simplifying assumptions applied to obtain the approximated from the cycle-averaged state equations. Nevertheless, the effect of this potential discontinuity is negligible when the current is cycle-averaged.
IV. SERVOMOTOR POWER CONSUMPTION MODELING AND MINIMIZATION
The complete switched models of electromechanical dynamics allow for the formulation of servomotor power consumption (SPC) as the input electrical power in terms of mechanical state variables. In the subsequent energy analysis, the servomotor system (excluding the microcontroller and sensors) is identified as the system of interest and SPC is the input electrical power IinVin from the power supply. The changes of system energy due to mass transfer across the system boundary are assumed to be negligible during motor operation, and thermal dissipation is the primary form of dissipation considered. From the set of coupled state equations, SPC is derived as
where the thermal dissipation rate depends on the switching configuration:
For on- or off-state ():
For dead time ():
with . Consequently, SPC is a function of only the mechanical state and control of the servomotor. The SPC model is consistent with and can be analyzed with respect to the first law of thermodynamics for a closed system.66–68 The first law of thermodynamics provides an independent viewpoint through the overall breakdown analysis and identification of the system's various energy terms (potentials, work, and dissipation) in the SPC model and the transfer and transformation between them. Its terms are the rate of change of total kinetic energy of the system, including the rotor, gear train, and output shaft; output mechanical power applied by the servomotor on its environment; and the system's internal magnetic potential energy storage rate for the inductor. Any change in the system's potential energy due to external conservative forces, such as gravity or elasticity, will be reflected by the output mechanical power term. The thermal dissipation term is composed of losses due to the (negated) friction work, Joule resistive heat, brush contact, and diode voltage, each of which is always positive to indicate the unidirectional power loss as outbound irreversible dissipation. Accordingly, while all terms in the SPC model are coupled with each other through the mechanical state and control variables, the direct effects of dissipated heat on the other model terms and parameters are assumed to be negligible, which can be ensured by monitoring the internal thermal energy rates through motor temperature in experiments (see below). Unlike existing DC motor electrical power models in the literature,16,15,43 the inclusion of the entire servomotor components within the system of interest and the incorporation of their dynamic interactions across switching configurations provide more accurate and comprehensive energy terms and their cause-and-effect relationships.
The SPC model can serve as the cost function to resolve the redundancy in robot tasks while minimizing energy usage. The constrained nonlinear optimal control (or trajectory optimization) problem for minimum SPC with the solution time interval is formulated as
subject to system dynamics that are cycle-averaged [Eqs. (8), (9), and (16)] and other constraints based on design and task requirements. The optimization variables vector x may include state variables and/or control inputs, depending on the given problem, and is bounded such that the feasible set is realizable by the actual servomotor. Each variable and its time-derivatives, if needed, can be time-parameterized for numerical implementation and solutions.
V. EXPERIMENTS FOR IDENTIFICATION, VALIDATION, AND DEMONSTRATION
The system parameters of a typical servomotor can be obtained from a combination of manufacturer specifications, experimental measurements, and estimation techniques. To ensure consistency in the estimated parameter values, experiments were conducted from fixed initial conditions to reduce parameter variation resulting from history dependence on factors, such as current and temperature. For the MX-28 servomotor considered in this study, the parameters obtained from manufacturer specifications include the inductance L of the RE-17 DC motor (Maxon Motor AG, Switzerland; 17 mm diameter, precious metal brushes, 4 W, part number 214897);69 the motor torque constant K;69 the diode forward voltage VD and on-state transistor resistance Rtrans of the FDS6990A N-Channel Logic Level MOSFETs (Fairchild Semiconductor, Inc., USA) used as switches within the H-bridge;70 the dead-time duration TDT programmed in the IR2104 half-bridge drivers (Infineon Technologies, Inc., USA);71 and the magnitude of the total gear ratio G,59 where G < 0 due to the even number of gears, NG = 6.
The pulse period TPWM was measured with an oscilloscope. The diode resistance RD was chosen as equal to Rtrans such that the transistor resistance remains constant throughout its operation, as is typically assumed.13 The DC motor's armature resistance Ra, servomotor's capacitance C, and capacitance in parallel with the DC motor, which was confirmed to be negligible (<10 nF), were measured with a digital multimeter. All gear masses, diameters, and tooth counts were measured to obtain Gi and Ji to compute the equivalent moment of inertia J.
A. Dual-servomotor test platform
For the verification and estimation of the remaining system parameters, experiments were conducted where controlled q(t) or trajectories were externally imposed on the servomotor. A dual-servomotor test hardware platform was developed, similar to an existing design14 but with modifications, where the MX-28 (tested) servomotor was connected to the DYNAMIXEL XM-430 (driving) servomotor through a shaft coupler (Fig. 5). The driving servomotor was mounted on a linear stage to allow for manual alignment of both servomotor output shafts prior to testing. All experiments were conducted at room temperature, and the motor temperature was monitored with the servomotor's built-in temperature sensor to ensure steady-state thermal dissipation conditions and no overheating of the motor. The test platform was clamped to the table to reduce vibrations such that the imposed driving or tested servomotor trajectories were followed with minimal disturbance.
Dual-servomotor test platform hardware setup (modified from an existing design14) with a diagram illustrating the use of a cantilever beam to estimate the tested servomotor's output torque with a load cell.
Dual-servomotor test platform hardware setup (modified from an existing design14) with a diagram illustrating the use of a cantilever beam to estimate the tested servomotor's output torque with a load cell.
The realized trajectory q(t), which may deviate from the imposed trajectory due to coupler effects, was measured by the tested servomotor's built-in encoder. The current was measured with an INA219 current sensor (Adafruit Industries, LLC, USA), placed in series between the power supply and the servomotor, connected to an Arduino Uno microcontroller (Arduino AG, Italy). The current of interest Iin(t) was obtained from the measured power supply current by subtracting the idle current that represents the sum of actuation-independent sources of current flow in the system, including the built-in encoder, the temperature sensor, and the microcontroller that were connected to the same power supply as the servomotor. The idle current was measured while no external torque was applied, and the servomotor was stationary; i.e., the servomotor torque was disabled, which is equivalent to D(t) = 0 so that . The control input D(t) to the tested servomotor was recorded and synchronized with the other state variable measurements in post-processing. The local proportional-integral-derivative (PID) controller of the servomotor quantized the desired D(t) trajectories in increments of 1/885.
For the estimation of , the tested servomotor was mounted on a cantilever beam with a pivot point at one end and pressing on a load cell at its other end (Fig. 5). The load cell was connected to a USB-interfaced Wheatstone bridge (Phidgets Inc., Canada). Using the normal force measured by the load cell, was obtained from moment equilibrium analysis. All experimental conditions for the tested and driving servomotor trajectories were selected such that there was no lifting of the cantilever beam during testing. To account for the effect of weight and any unavoidable coupler misalignment, a five-term Fourier fitted model of the load cell force, obtained at 0.015 rad increments of q throughout the servomotor's range of motion under static and unactuated conditions, was subtracted from the measured load cell force.
The brush-contact voltage drop Vbr was measured by stalling the DC motor within the tested servomotor. The DC motor terminals were desoldered from the H-bridge circuit and directly connected to a benchtop power supply, and the stall current was measured for Vin from 0.3 to 1.5 V. The gear train friction coefficients b0 and b1 were obtained for both directions of rotation by disconnecting the DC motor from the gear train [i.e., ] and measuring the average exerted by the gear train during constant trajectories imposed by the driving servomotor for various values of . In these experiments, the motor temperature sensor was used to ensure that there were no significant thermal effects18 on the measured brush-contact and gear train friction properties.
B. Single pendulum system
To illustrate the proposed approach in a robotic system, in which the mechanical power is transferred from each servomotor to a link through the output shaft, a pendulum was attached to the servomotor output shaft and used for additional prediction and optimal control tasks. The equation of motion of the pendulum is , where Jp is the moment of inertia of the pendulum about the servomotor output shaft axis, M is the pendulum mass, g is the gravitational acceleration, dp is the distance between the center-of-mass of the pendulum and the axis of the output shaft, and q = 0 corresponds to the downward vertical position of the pendulum. For optimal control, the SPC is evaluated directly from and is time-integrated for the cost function. The electrical duty cycle D rather than the mechanical actuator torque is chosen as the control input to solve for D during evaluation of the system dynamics, which requires the inclusion of in x as an additional state variable. Based on the equation of motion and the cycle-averaged equation for D for the case of Vbr = 0, the pendulum system dynamics when does not change its sign is
For cases where changes sign, a differentiable sigmoid function approximation of should be used in the derivation.
VI. RESULTS AND DISCUSSION
The modeling assumptions were evaluated using the estimated values for the electromechanical system parameters of the MX-28 servomotor (Table II). The PWM frequency 1/TPWM = 40 kHz (TPWM = 25 μs) is sufficiently large relative to the no-load angular velocity of the servomotor (5.76 rad/s = 0.92 Hz59) such that the use of the cycle-averaged state equations is warranted.12 According to the manufacturer specifications, the dead-time duration TDT varies within the range 400–650 ns and has a typical value of 520 ns.71 That the total dead time 2TDT is 4.16% of TPWM warrants its omission when computing D(t) and from the cycle-averaged state equations. The state equations derived with Vbr = 0 are used here, as in common control schemes,64 because its estimated value (−0.0076 V from the linear fit of Vin as a function of Iin) is too low to be accurately measured, relative to the power supply input voltage Vin = 12.17 V. The filtering capacitance measured across the DC motor terminals while connected to the H-bridge was minimal, justifying its omission in the model. Using the estimated system parameters, the accuracy and merits of the models are experimentally illustrated and validated through predictions and optimal control tasks. Here, all trajectories in the figures were denoised using a Gaussian filter with a window of 0.16 s.
Estimated values of the servomotor system parameters. Nominal values obtained from manufacturer specifications are shaded.
Component . | Parameter (unit) . | Symbol . | Value . |
---|---|---|---|
DC motor | Brush-contact voltage drop (V) | Vbr | 0 |
Armature resistance () | Ra | 8.9 | |
Motor torque constant (Nm/A) | K | 0.0107a) | |
Motor inductance (mH) | L | 0.206 | |
Gear train | Coulomb friction coefficient , (Nm) | b0 | −0.0113, −0.0177 |
Viscous friction coefficient , (Nm s) | b1 | −0.024, −0.037 | |
Equivalent moment of inertia (g m2) | J | 3.3003 | |
Signed gear ratio | G | −193 | |
H-bridge | Pulse period (μs), PWM frequency (kHz) | TPWM, 1/TPWM | 25, 40 |
Diode resistance () | RD | 0.011 | |
Dead-time duration (ns) | TDT | 520 | |
On-state transistor resistance () | Rtrans | 0.011 | |
Diode forward voltage (V) | VD | 0.7 | |
… | Servomotor capacitance (μF) | C | 55b), 82.9b) |
Capacitance across DC motor terminals (nF) | … | 5.7–8.0 |
Component . | Parameter (unit) . | Symbol . | Value . |
---|---|---|---|
DC motor | Brush-contact voltage drop (V) | Vbr | 0 |
Armature resistance () | Ra | 8.9 | |
Motor torque constant (Nm/A) | K | 0.0107a) | |
Motor inductance (mH) | L | 0.206 | |
Gear train | Coulomb friction coefficient , (Nm) | b0 | −0.0113, −0.0177 |
Viscous friction coefficient , (Nm s) | b1 | −0.024, −0.037 | |
Equivalent moment of inertia (g m2) | J | 3.3003 | |
Signed gear ratio | G | −193 | |
H-bridge | Pulse period (μs), PWM frequency (kHz) | TPWM, 1/TPWM | 25, 40 |
Diode resistance () | RD | 0.011 | |
Dead-time duration (ns) | TDT | 520 | |
On-state transistor resistance () | Rtrans | 0.011 | |
Diode forward voltage (V) | VD | 0.7 | |
… | Servomotor capacitance (μF) | C | 55b), 82.9b) |
Capacitance across DC motor terminals (nF) | … | 5.7–8.0 |
The motor constants for rotor torque and back emf are given as identical in the manufacturer specifications.
Measurements from two separate servomotor controller circuit boards to assess manufacturer variability.
A. Predictions and validation
Selected state variables were predicted for given desired input trajectories and validated against experimental measurements. The first task considered is dynamic braking (by dissipation through armature resistance11) with the pendulum system (M = 0.214 kg; dp = 0.06928 m; Jp = 1.221 g m2) driven by gravity, where the pendulum falls from rest with initial conditions q(0) = 4.0 rad, , and , and the time duration T = 21.7 s (Fig. 6). The H-bridge remained in the off-state without switching. Thus, the trajectory of q(t) was predicted through the models using the given conditions of D(t) = 0 and Iin(t) = 0 at all times. Nevertheless, was non-zero due to the back emf from . The magnitude of exhibited the same pattern as that of , and its braking effect resulted in for t > 7.3 s and q = 5.7 rad at the final time. In comparison, the pendulum would reach the same angle at a much shorter time (t = 0.17 s) if the H-bridge was disconnected such that the DC motor circuit would be open (i.e., ), resulting in during the falling trajectory due to the positive gravitational torque overcoming the friction (Fig. 6). Since , was always negative, and its difference from the zero SPC was mostly dissipated as heat due to the non-zero (discussed in more detail later).
Predicted vs measured q(t) trajectories of the pendulum during the dynamic braking task along with motion snapshots and the predicted q(t) trajectory if the H-bridge was disconnected. The cycle-averaged motor current and servomotor output torque trajectories obtained from the measured q(t) are also shown.
Predicted vs measured q(t) trajectories of the pendulum during the dynamic braking task along with motion snapshots and the predicted q(t) trajectory if the H-bridge was disconnected. The cycle-averaged motor current and servomotor output torque trajectories obtained from the measured q(t) are also shown.
The root-mean-square deviation of the predicted q(t) from the measurement is 0.127 rad or 7.2% of the range of the measured q(t) over the total time duration. Since the dynamic braking task does not involve switching, this discrepancy may be due to the propagation of uncertainties and errors in the estimated system parameters. This may include the typical use of identical K for both Vemf and , which is based on the common assumption of ideal power conversion within a DC motor.57,64 Also, in numerical investigations with a combination of different velocities and friction coefficients, the model predictions showed that, due to the low pendulum angular velocity involved, the q(t) trajectory is highly sensitive to slight variations in the friction coefficients b0 and b1, which affect both the acceleration and the final steady-state angle of the pendulum.
The second prediction task is a sinusoidal trajectory following, where a set of sinusoidally varying D(t) trajectories were applied on the tested servomotor as the driving servomotor imposed the sinusoidally varying velocity trajectory corresponding to with PID position feedback control. The two desired D(t) trajectories tested were , which correspond to net negative and positive SPC, respectively (Fig. 7). The frequencies of both D(t) trajectories were equal to that of , while their phases were equal and opposite to that of , respectively. These D(t) trajectories were chosen to allow for a periodic continuous testing over a range of accelerations, while avoiding sudden changes in the direction of the force applied to the load cell that may cause vibration or lifting of the cantilever beam. With these desired D(t) trajectories, the trajectories of were predicted through the models using the measured trajectory of q(t) and the estimated . Due to the low magnitudes of the D(t) trajectories chosen, the magnitude of Iin(t) was significantly less than that of (Fig. 7). The root-mean-square deviation of the predicted from the measured Iin(t) is 13.4% (0.0049 A) and 16.4% (0.0035 A) for the net negative and positive SPC trajectories, respectively, of the range of the corresponding measured Iin(t) over the total time duration. The causes of these discrepancies may include fluctuations in the idle current drawn by the microcontroller and sensors, and the assumption of ideal switching, which neglects switching losses in the transistors of the H-bridge and the microcontroller, nonlinearity of TDT, and variation of Ia in the DC motor during non-instantaneous switching.12 Also, since Iin during the experiments was relatively low (<0.05 A) as compared with the stall current 1.4 A of the servomotor,59 the assumption of constant parameters without considering the nonlinearity of the DC motor64,72 and the H-bridge11 across the full operating range of current may be another source of error. In addition, the variation of reaction forces applied by the coupler on the output shaft in the dual-servomotor test platform may have contributed to errors in the estimates.
Predicted cycle-averaged vs measured input current trajectories for the sinusoidal trajectory following for net negative and positive SPC using the dual-servomotor test platform. The cycle-averaged motor current and output mechanical power trajectories obtained from the measured q(t) and the estimated output torque under the desired D(t) trajectories are also shown. Note that SPC is proportional to Iin by the constant power supply voltage Vin.
Predicted cycle-averaged vs measured input current trajectories for the sinusoidal trajectory following for net negative and positive SPC using the dual-servomotor test platform. The cycle-averaged motor current and output mechanical power trajectories obtained from the measured q(t) and the estimated output torque under the desired D(t) trajectories are also shown. Note that SPC is proportional to Iin by the constant power supply voltage Vin.
For both tasks, the overall similar trends and the calculated errors between the predicted and measured results confirm the validity of the proposed comprehensive models of servomotor electromechanical dynamics with switching, the cycle-averaging approach, and the estimated system parameters. The results also validate the bi-directional use of the cycle-averaging approach, i.e., mapping from electrical to mechanical (for the dynamic braking) and from mechanical to electrical (for the sinusoidal trajectory following) state trajectories. The use of general electromechanical system parameters, as opposed to custom parameters in specialized models meant to fit a specific state variable (e.g., power consumption) or an operating condition, results in their validity across a wide range of task conditions, including cases of positive, zero, and negative SPC during , which is often ignored in the literature.5,43
B. Optimal control of a pendulum for minimum servomotor power consumption
The optimization variables used for the optimal control tasks with the pendulum system were . The initial and final conditions q(0) = 0 rad, , , and were imposed as constraints, along with (for monotonic motion) and , over the time duration T = 10 s. The final velocity and acceleration were not specified. The optimal control algorithm was implemented with the open-source trajectory optimization library OptimTraj73 in MATLAB (The MathWorks, Inc., USA). Orthogonal collocation with Chebyshev polynomials was used to discretize the optimization variables, and the cost and constraint function gradients were computed numerically with built-in finite-difference methods.
For the cost function of the time-integrated SPC, the cycle-averaged [Eq. (16)] in terms of the optimization variables was used as a low-frequency function approximation. In addition, other cost function models that are commonly used as proxies for power consumption43,45,48–50,52 are also considered for comparison: the square of the rotor torque (SRT) and the positive rotor mechanical power (PMP) , which were cycle-averaged as and , respectively, and time-integrated. (Here, the PMP was calculated at the rotor, rather than the output shaft as is typical,43,52 so that the gear train friction directly affects its value.) The resulting optimal q(t) from each optimization was imposed as the desired trajectory for the PID position feedback control in experiments, where the gains were tuned manually such that the measured D(t) trajectory matches that from optimization. The resulting D(t) trajectories were similar in pattern to and SPC (Fig. 8) and had peak values of 0.20, 0.36, and 0.16 for the SPC, SRT, and PMP optimal trajectories, respectively. The results are analyzed with respect to two perspectives of the use of the proposed model—its merit in optimal control for power consumption minimization and its accuracy for further validation.
Measured SPC for optimal trajectories of the selected cost function models along with motion snapshots of the SPC optimal trajectory. For each optimal trajectory, the directly measured SPC is also compared with the SPC predicted from the measured q(t) and D(t) and from optimization.
Measured SPC for optimal trajectories of the selected cost function models along with motion snapshots of the SPC optimal trajectory. For each optimal trajectory, the directly measured SPC is also compared with the SPC predicted from the measured q(t) and D(t) and from optimization.
The time-integrated model SPC predicted from measured q(t) and D(t) was smaller than those directly measured by 9.1%, 6.0%, and 8.5% for optimal SPC, SRT, and PMP, respectively. The deviation of the measured D(t) from that of the optimization solution, which was due to the tracking error inherent in the PID position feedback control, contributed to the discrepancy between the SPC from experiments [directly measured or from measured q(t) and D(t)] and from optimization. The aforementioned issues of switching losses and other switching nonlinearities may also have contributed to the difference between the predicted and measured SPC, especially for (near) zero Iin(t). For all cost functions, the predicted reached zero (<1 mA)—for t > 7.4 s for optimal SPC, for 5.4 s < t < 8.8 s for optimal SRT, and near the final time for optimal PMP—while the measured Iin(t) did not (>5.6 mA for optimal SPC, >2.2 mA for optimal SRT, and >5.0 mA for optimal PMP). Overall, as opposed to applying D(t) = 0 for the continuous off-state without switching in the aforementioned dynamic braking task, which resulted in the predicted and measured Iin(t) to be exactly zero, the proposed models resulted in zero during falling as regulated by optimization (Figs. 8 and 9). The overall accuracy illustrated serves as another validation of the proposed cycle-averaging approach and the SPC model across different tasks.
The measured pendulum angle trajectories were used to predict the cycle-averaged motor current, servomotor output torque, and output mechanical power profiles for the selected cost function models.
The measured pendulum angle trajectories were used to predict the cycle-averaged motor current, servomotor output torque, and output mechanical power profiles for the selected cost function models.
The measured SPC integrated over was the lowest for the SPC optimal trajectory, which validates that the SPC predicted from the model best reflects the true SPC, as opposed to the other proxies pursued in the literature (Table III). For all cost functions, the measured SPC and trajectories peaked approximately when , where the gravitational torque was the largest (Figs. 8 and 9). While depends mainly on the angle (due to the low moment of inertia of the pendulum) and its peak values are almost identical across all cost functions, the variations of the SPC and trajectories (including the timing of each peak) reflect each cost function model's distinct behavior. The optimal SRT stayed near the initial equilibrium (within 0.1 rad for t < 1.3 s), then quickly reached the upright equilibrium with steep changes of the velocity where is near zero, and remained there for most of the time duration (within 0.1 rad for 3.4 s < t < 7.4 s) until around the final time before falling suddenly starts. In this way, the optimal SRT minimized the time spent near angles and where peaked (Fig. 9). The lack of velocity-dependence in the SRT cost function resulted in mostly static poses punctuated by sudden motion, as opposed to the smoother trajectories for the other cost function models. As a result, the output mechanical power of the optimal SRT suddenly peaked with relatively large values and then stayed zero during most of the time, which contributed to the distinct time-profile of the measured SPC that is mostly (near) zero with significant increases around and toward . The optimal SPC resulted in much larger actuations during most of the time but smaller time-integrated SPC when compared with that of the optimal SRT, which is additional evidence of the distinct behaviors between the two cost function models. The similarity of the SPC and PMP optimal trajectories reflects the similarity in pattern between their SPC and , which is approximately proportional to PMP ( due to the low moment of inertia and friction in the servomotor) during (t < 6.0 s) before falling (Figs. 8 and 9). On the other hand, by construction, PMP is unaffected by variations in SPC when during the falling phase, and consequently, the time-integrated SPC of the optimal PMP was greater than that of the optimal SPC.
Comparison of optimal trajectory results for the selected cost function models.
Cost function model . | Measured time-integrated SPC (J) . | % increase . |
---|---|---|
SPC | 3.111 | … |
SRT | 3.502 | +12.6% |
PMP | 3.619 | +16.3% |
Cost function model . | Measured time-integrated SPC (J) . | % increase . |
---|---|---|
SPC | 3.111 | … |
SRT | 3.502 | +12.6% |
PMP | 3.619 | +16.3% |
The falling phases () demonstrated further notable differences. The measured SPC during falling significantly increased for the optimal SRT, gradually reached (near) zero only at the end of falling for the optimal PMP, and stabilized at zero shortly after falling began for the optimal SPC. That the time t = 7.4 s at which SPC reaches nearly zero after passing the upright position [ and measured SPC < 0.1 W] for the optimal SPC was in between that of the optimal SRT and optimal PMP may reflect the SPC formulation as being mainly the sum of and , which are loosely associated with the SRT and PMP models, respectively. Since falling can be achieved solely by gravity given sufficient time, it is unnecessary to have any positive SPC during falling from the perspective of minimizing SPC, which was noticeably demonstrated by the optimal SPC but not by the other two cost function models.
Overall, for optimal SRT, the attributes of the SPC time-profile were significantly distinct from those for optimal SPC (Fig. 8). For optimal PMP, the time-integrated SPC was the largest (Table III), partially due to its lack of capability to take advantage of the gravity during the falling phase. These results indicate that, with respect to their patterns and magnitudes, these existing proxy models are not valid as representations of the actual power consumption. For the purpose of power consumption minimization, the proposed SPC model provides true optimality, both instantaneous and time-integrated, as compared with the proxy models.
Another merit of the proposed mathematical models of the electromechanical system, along with their thermodynamic analysis,46 is that they provide the breakdown of SPC into its constituent components (e.g., work vs heat, Joule heating from armature resistance vs frictional heating) and their distribution across time (Fig. 10), which is usually not possible through experiments (e.g., calorimetry74). These types of thermodynamic analyses are also not available in other approximation approaches used mainly for control. The inductor's magnetic potential storage energy rate was approximated as through a finite-difference method.27 The thermal dissipation, which is always non-negative, was obtained from as the remainder of the SPC after accounting for all non-dissipative components. [Calculating directly from was avoided since it does not account for the effect of ripple current on the resistor and diode power losses12,13 and may also introduce further error due to the lack of continuity conditions in deriving the piecewise analytical solutions of Ia(t).] For all cost functions, the relatively low proportion of the kinetic energy change rate is due to the choice of the servomotor as the system of interest, in which the rotor, gear train, and output shaft are the only moving parts. The energy stored or released by the inductor was also minimal, in line with other prior works where the motor inductance is typically neglected.5,7,8,16,15,43 The output mechanical work was relatively small (<5%) as a share of the time-integrated SPC, with the remainder mostly dissipated as heat (Table IV). While exhibited a similar trend as that of during the rising phase of the pendulum trajectory, they had opposite trends during the falling phase, which resulted in exceeding the measured SPC. The opposite trend is mainly due to , not , that is (re-)generated by the back emf of the DC motor during the negative mechanical-work phase such that . The measured SPC was zero or positive (Fig. 8) even when both and were negative during falling (Fig. 9), which implies the distinct roles of vs . [Note that it is possible to achieve (near) zero even when Ia(t) can still fluctuate to result in Joule heating from armature resistance or with non-zero D(t) by externally imposing q(t) and trajectories to induce (near) zero .] It can be seen (Fig. 10) that this aspect was exploited by the optimal SPC during the falling phase when the measured SPC was near zero, which was not noticeably demonstrated by the other cost function models. On the other hand, when zero measured SPC was maintained during the upright equilibrium of the optimal SRT, and were also (near) zero due to zero .
Component breakdown of the measured SPC of optimal trajectories for the selected cost function models.
Component breakdown of the measured SPC of optimal trajectories for the selected cost function models.
Component breakdown of SPC of optimal trajectories for the selected cost function models. The ratio of each component (integrated over the entire time duration) to the measured time-integrated SPC is indicated as a percentage.
Cost function model . | Thermal dissipation (%) . | Output mechanical work (%) . | Kinetic energy change (%) . | Inductor magnetic potential energy storage (%) . |
---|---|---|---|---|
SPC | 94.68 | 4.87 | 0.45 | 6.54 × 10−7 |
SRT | 95.32 | 3.94 | 0.74 | 1.82 × 10−5 |
PMP | 96.23 | 3.34 | 0.44 | 3.77 × 10−7 |
Cost function model . | Thermal dissipation (%) . | Output mechanical work (%) . | Kinetic energy change (%) . | Inductor magnetic potential energy storage (%) . |
---|---|---|---|---|
SPC | 94.68 | 4.87 | 0.45 | 6.54 × 10−7 |
SRT | 95.32 | 3.94 | 0.74 | 1.82 × 10−5 |
PMP | 96.23 | 3.34 | 0.44 | 3.77 × 10−7 |
In all of the experiments conducted, the measured power supply current was always positive even during negative SPC due to the measured current being the sum of Iin and the idle current. In general, the case of negative SPC requires special consideration because, while the DC servomotor is inherently a bi-directional energy converter, the power supply is not.11,16 Power supplies, such as a battery, can be either rechargeable or non-rechargeable, and even in the case of a rechargeable battery, the efficiency of a battery in absorbing power flows depends on battery-specific parameters.
VII. CONCLUSIONS
The integration of the coupled models for switched electromechanical dynamics of a typical DC servomotor in the proposed approach, which were experimentally validated for their predictive accuracy and demonstrated for their merits against common proxy models of power consumption, allowed for the analysis and control of general aperiodic tasks, including transient phases. These enhancements of the model were provided by its comprehensiveness in including the switching configurations of on-state, off-state, and dead time, and the combination of cycle averaging with piecewise analytical solutions to handle the non-smooth dynamics and different temporal scales from high-frequency electrical to low-frequency mechanical variables. Future studies will investigate switching losses, idle current, non-ideal power supplies (e.g., batteries), variations of the system parameters, and the scalability of the model for large current/power and multi-degree-of-freedom robotic systems. In addition, validity of the cycle-averaging method under different conditions needs to be further evaluated.
ACKNOWLEDGMENTS
This work was supported in part by the U.S. National Science Foundation (Nos. CMMI-1436636 and IIS-1427193) and a Mitsui USA Foundation scholarship.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
William Z. Peng: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Hyunjong Song: Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (equal); Validation (equal); Visualization (equal); Writing – review & editing (equal). Dariusz Czarkowski: Data curation (equal); Investigation (equal); Supervision (equal); Writing – review & editing (equal). Joo H. Kim: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Software (equal); Supervision (equal); Validation (equal); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available within the article.