6
Sliding-Mode Control of a Hydrostatic Drive Train with Uncertain Actuator Dynamics Hao Sun and Harald Aschemann Abstract— A sliding-mode approach with disturbance com- pensation is proposed in this paper for the tracking control of a hydrostatic drive train, which is commonly used in off-road vehicles. A control-oriented modelling results in a system of four nonlinear differential equations which are subject to un- certain actuator dynamics with saturation effects and unknown disturbances – a leakage volume flow and a disturbance torque acting on the hydraulic motor. The disturbances are estimated by a nonlinear reduced-order disturbance observer and used for a disturbance compensation. Thereby, the switching part in the sliding-mode control law and chattering phenomena can be reduced. Simulation results point out that the proposed approach leads to an excellent tracking performance despite the given uncertainties in the actuator dynamics. I. I NTRODUCTION A hydrostatic transmission uses hydraulic oil to transmit power from the power source to the drive mechanism [1]. A simple hydrostatic transmission consists of a hydraulic pump and a hydraulic motor, of which at least one must have a variable displacement, operating together in a closed circuit. The fluid flows directly from the motor outlet to the pump inlet, without returning to the tank. On the pump side the mechanical torque from the engine is transformed by the hydraulic pump into a pressurised fluid flow which is transformed back to mechanical torque by the hydraulic motor. By varying the tilt angle (changing the displacement) of either pump or motor, any desired transmission ratio can be obtained within a predefined boundary. Hydrostatic transmissions are widely used as a charac- teristic component of drive trains in practically all types of working machines like harvesters, wheel loaders, excavators, telehandlers and agricultural tractors. They are typically operated in combination with diesel engines for mobile applications and offer a variety of advantages in comparison to pure mechanical transmissions. Besides the capability of a continuously variable transmission with high power density and the generation of large traction forces at low speeds, hydrostatic transmissions allow for reversing the direction of rotation without changing the gear. Moreover, it is possible to perform breaking manoeuvres without mechanical wear. Due to the significant advantages in comparison with mechanical gearboxes, hydrostatic transmissions have also attracted the attention of engineers from other industrial fields like wind energy technology in [2], [3], and [4]. Hydrostatic transmission systems are subject to several nonlinearities due to input saturation, friction, directional Hao Sun and Harald Aschemann are with the Chair of Mechatronics, University of Rostock, Germany. {Hao.Sun, Harald.Aschemann}@uni-rostock.de change of the valve opening, unknown load and leakage, which complicates a model-based control of such a system. Even so, gain-scheduled PID-controllers are still currently the choice of industrial practice for controlling hydrostatic transmissions [5]. In order to improve energy efficiency and to consider environmental aspects in mobile applications, a model-based nonlinear control approach has been proposed in [6], [7]. Both the simulation and experimental results show that the approach can lead to higher tracking accuracy and active damping of pressure oscillations. Additional uncer- tainty due to an unknown load torque acting on the motor side is considered and counteracted by an observer-based disturbance compensation. The control approach discussed in [6], [7] is based on the assumption that an approximate feedforward compensation of the known actuator dynamics can be achieved, which leads to a model order reduction. In this case, only the difference pressure and motor angular velocity are used for feedback control. Fig. 1. Drive train with a closed-circuit hydrostatic transmission. Considering both uncertainty in the actuator dynamics – imperfectly known time constants in the actuator dynamics – and unknown disturbances – a leakage volume flow as well as a load torque acting on the hydraulic motor – a sliding-mode controller with disturbance observer is proposed in this paper. Sliding-mode control uses a high frequency switching control action to bring the trajectories in finite time to a predefined sliding surface and keep them within the vicinity of this surface thereafter. Sliding-mode control has been known as a standard approach to tackle parametric uncertainties, although it has been subject to problems in applications regarding chattering phenomena [8], [9] and [10]. The paper is organised as follows: in Sec. II, the control-oriented mod- elling of the hydrostatic drive train system depicted in Fig. 1 is addressed. A sliding-mode controller is designed for the 2013 European Control Conference (ECC) July 17-19, 2013, Zürich, Switzerland. 978-3-952-41734-8/©2013 EUCA 3216

Sliding-Mode Control of a Hydrostatic Drive Train with ...folk.ntnu.no/skoge/prost/proceedings/ecc-2013/data/papers/0840.pdf · perform breaking manoeuvres without mechanical wear

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Sliding-Mode Control of a Hydrostatic Drive Train with ...folk.ntnu.no/skoge/prost/proceedings/ecc-2013/data/papers/0840.pdf · perform breaking manoeuvres without mechanical wear

Sliding-Mode Control of a Hydrostatic Drive Trainwith Uncertain Actuator Dynamics

Hao Sun and Harald Aschemann

Abstract— A sliding-mode approach with disturbance com-pensation is proposed in this paper for the tracking control ofa hydrostatic drive train, which is commonly used in off-roadvehicles. A control-oriented modelling results in a system offour nonlinear differential equations which are subject to un-certain actuator dynamics with saturation effects and unknowndisturbances – a leakage volume flow and a disturbance torqueacting on the hydraulic motor. The disturbances are estimatedby a nonlinear reduced-order disturbance observer and usedfor a disturbance compensation. Thereby, the switching part inthe sliding-mode control law and chattering phenomena canbe reduced. Simulation results point out that the proposedapproach leads to an excellent tracking performance despitethe given uncertainties in the actuator dynamics.

I. INTRODUCTION

A hydrostatic transmission uses hydraulic oil to transmitpower from the power source to the drive mechanism [1].A simple hydrostatic transmission consists of a hydraulicpump and a hydraulic motor, of which at least one musthave a variable displacement, operating together in a closedcircuit. The fluid flows directly from the motor outlet tothe pump inlet, without returning to the tank. On the pumpside the mechanical torque from the engine is transformedby the hydraulic pump into a pressurised fluid flow whichis transformed back to mechanical torque by the hydraulicmotor. By varying the tilt angle (changing the displacement)of either pump or motor, any desired transmission ratio canbe obtained within a predefined boundary.

Hydrostatic transmissions are widely used as a charac-teristic component of drive trains in practically all types ofworking machines like harvesters, wheel loaders, excavators,telehandlers and agricultural tractors. They are typicallyoperated in combination with diesel engines for mobileapplications and offer a variety of advantages in comparisonto pure mechanical transmissions. Besides the capability of acontinuously variable transmission with high power densityand the generation of large traction forces at low speeds,hydrostatic transmissions allow for reversing the direction ofrotation without changing the gear. Moreover, it is possible toperform breaking manoeuvres without mechanical wear. Dueto the significant advantages in comparison with mechanicalgearboxes, hydrostatic transmissions have also attracted theattention of engineers from other industrial fields like windenergy technology in [2], [3], and [4].

Hydrostatic transmission systems are subject to severalnonlinearities due to input saturation, friction, directional

Hao Sun and Harald Aschemann are with the Chair ofMechatronics, University of Rostock, Germany. {Hao.Sun,Harald.Aschemann}@uni-rostock.de

change of the valve opening, unknown load and leakage,which complicates a model-based control of such a system.Even so, gain-scheduled PID-controllers are still currentlythe choice of industrial practice for controlling hydrostatictransmissions [5]. In order to improve energy efficiency andto consider environmental aspects in mobile applications, amodel-based nonlinear control approach has been proposedin [6], [7]. Both the simulation and experimental results showthat the approach can lead to higher tracking accuracy andactive damping of pressure oscillations. Additional uncer-tainty due to an unknown load torque acting on the motorside is considered and counteracted by an observer-baseddisturbance compensation. The control approach discussedin [6], [7] is based on the assumption that an approximatefeedforward compensation of the known actuator dynamicscan be achieved, which leads to a model order reduction.In this case, only the difference pressure and motor angularvelocity are used for feedback control.

Chair of Mechatronics

Hydrostatic Transmissionsin Mobile Applications

Fig. 1. Drive train with a closed-circuit hydrostatic transmission.

Considering both uncertainty in the actuator dynamics –imperfectly known time constants in the actuator dynamics –and unknown disturbances – a leakage volume flow as well asa load torque acting on the hydraulic motor – a sliding-modecontroller with disturbance observer is proposed in this paper.Sliding-mode control uses a high frequency switching controlaction to bring the trajectories in finite time to a predefinedsliding surface and keep them within the vicinity of thissurface thereafter. Sliding-mode control has been knownas a standard approach to tackle parametric uncertainties,although it has been subject to problems in applicationsregarding chattering phenomena [8], [9] and [10]. The paperis organised as follows: in Sec. II, the control-oriented mod-elling of the hydrostatic drive train system depicted in Fig. 1is addressed. A sliding-mode controller is designed for the

2013 European Control Conference (ECC)July 17-19, 2013, Zürich, Switzerland.

978-3-952-41734-8/©2013 EUCA 3216

Page 2: Sliding-Mode Control of a Hydrostatic Drive Train with ...folk.ntnu.no/skoge/prost/proceedings/ecc-2013/data/papers/0840.pdf · perform breaking manoeuvres without mechanical wear

hydrostatic drive train considering explicitly the uncertaintiesin the actuators in Sec. III. In Sec. IV, a reduced-orderdisturbance observer is introduced. The estimated leakagevolume flow and the disturbance torque can then be used forcompensation measures in the control scheme. At controlallocation, the displacement limitation due to physicallybounded tilt angle at both pump and motor is also consideredin this paper. The simulation results presented in Sec. Vshow excellent tracking performance for both the differencepressure and the motor angular velocity.

II. CONTROL-ORIENTED MODELLING

Dynamic system modelling plays a dominant role inmodern control technology. An accurate system model isthe key to improving the overall system performance. Theconsidered mechatronic system is divided into a hydraulicsubsystem and a mechanical subsystem, which are coupledby the torque generated by the hydraulic motor [11].

A. Hydraulic subsystem

The hydraulic pump is connected to the rotor shaft, andits main function is to convert mechanical energy intohydraulic energy. Pumps are characterized by a volumetricdisplacement Vp in m3/rev, which describes the fluid volumedisplaced per rotation of the rotor shaft. The pump flow qp isdetermined by a nonlinear function qP = VP (αP )nP . Thecalculation of the pump volumetric displacement dependson the specific mechanical design. For the widely used axialpiston pump with a tiltable swashplate, VP can be expressedas

VP (αP ) = NPAP dP tan(αP,max · αP ). (1)

Here, αP = αP,max · αP denotes the swashplate angle ofthe pump, AP the effective piston area, dP the diameter ofthe piston circle, and NP the number of pistons. With thereasonable assumption of a small swashplate angle |αP | ≤18◦ [12], and

VP =NPAP dPαP,max

2π, (2)

the pump volume flow can be described by qP = VP αPωP .Here, the normalized swashplate angle αP ∈ (−1, 1) and therotational speed of pump ωP are introduced. Accordingly,the volume flow rate qM = VM (αM )nM into the hydraulicmotor can be formulated as a nonlinear function of the bentaxis angle, with

VM (αM ) = NMAMdM sin(αM,max · αM ). (3)

As before, the assumption of a small tilt angle αM ≤ 20◦

andVM =

NMAMdMαM,max

2π(4)

allow for a simplified relationship qM = VM αMωM , withthe normalized angle αM ∈ (εM , 1), εM > 0, and themotor angular velocity ωM . Without consideration of thepressure loss in the transmission lines, the high- and low-pressure side can be described by one pressure value pA andpB , respectively. The pressure dynamics of the hydrostatic

Fig. 2. Axial piston hydraulic pump (left, swashplate design) as well ashydraulic motor (right, bent axis design).

transmission in closed-loop configuration follows from amass balance in combination with an oil model [12]

pA =βAVA

(qP − qM − qIL − qELA), (5)

pB =βBVB

(−qP + qM + qIL − qELB). (6)

Here, βk, k ∈ {A,B}, denotes the effective bulk modulus ofthe fluid, and Vk describes the volume of the correspondingtransmission line. The internal leakage oil flow qIL is mod-elled as a laminar flow resistance [1], which depends linearlyon the difference pressure

qIL = kIL · (pA − pB) (7)

and is characterized by the leakage coefficient kIL. Theexternal leakage can be described by [12]

qELA = kELA · pA, (8)qELB = kELB · pB . (9)

Here, kELA and kELB represent the external leakage coef-ficients. A model order reduction becomes possible, if somereasonable symmetry assumptions are made. Consideringidentical capacitances CA = CB =: CH , see [1] and [13],with

CH =Vkβk, k ∈ {A,B} (10)

leads to an order reduction by one. As a result, the differentialequation for the difference pressure is given by

∆p =2

CH

(VP αPωP − VM αMωM −

qU2

), (11)

with a resulting leakage volume flow

qU = 2kIL∆p+ kELA pA − kELB pB . (12)

as a lumped disturbance. The dynamics of the displacementunits for both pump and motor is characterised by a first-order lag behaviour, respectively, according to

Tui ˙αi + αi = ui, i ∈ {P,M}, (13)

where the time constants Tui are not exactly known anduncertain. The input voltages ui of the corresponding pro-portional valves for the displacement units serve as physicalcontrol inputs. As the tilt angles of the diplacements units

3217

Page 3: Sliding-Mode Control of a Hydrostatic Drive Train with ...folk.ntnu.no/skoge/prost/proceedings/ecc-2013/data/papers/0840.pdf · perform breaking manoeuvres without mechanical wear

are limited, saturation functions satab (αi) are introduced asfollows

satab (αi) =

αi,max αi ≥ aαi for b < αi < a ,

αi,min αi ≤ b(14)

where a = αi,max and b = αi,min represent the upper andlower output limits determined by the mechanical design:{εM , 1} for the hydraulic motor and {−1, 1} for the hy-draulic pump. In the simulation model, (13) is implementedwith limited integrators for αP and αM .

B. Mechanical subsystemThe longitudinal dynamics of the working machine is

governed by the equation of motion. The vehicle with thedrive train (vehicle mass mv , wheel radius rw, gear boxtransmission ratio ig , rear axle transmission ratio ia, dampingcoefficient dg at the drive shaft), see also Fig. 3, can bedescribed by the following first order differential equation(JM +

Jg

ig2 +

Ja +mV r2w

i2a i2g

)︸ ︷︷ ︸

JV

ωM +dgi2g︸︷︷︸dV

ωM = (15)

VM∆p αM︸ ︷︷ ︸τM

−(τMf tanh(

ωMε

) + τgf tanh(ωMigε

) +τLiaig

)︸ ︷︷ ︸

τU

,

where τM is the torque of the hydraulic motor. JM , Jg andJa are the mass moments of inertia of the hydraulic motor,gear box and rear axle, respectively. The maximum valuesτMf and τgf characterise the friction models of the hydraulicmotor and the gear box, whereas ε << 1 represents a smallnumber.

C. State-space descriptionThe overall system model involves four first order differ-

ential equations. Introducing the normalized angles of thedisplacement units αi, i ∈ {P,M}, the difference pressure∆p, and the motor angular velocity ωM as state variables,the state vector results in x = [αP , αM ,∆p, ωM ]T . Thecorresponding state-space representation becomes

˙αP˙αM∆pωM

=

− 1TuP

αP + 1TuP

uP− 1TuM

αM + 1TuM

uM2VPωP

CHsat 1

−1(αP )− 2VMωM

CHsat 1

εM (αM )− qUCH

− dVJV ωM + VM

JV∆p sat 1

εM (αM )− τUJV

.(16)

, ,

Fig. 3. Kinematical structure of the drive train with gear boxes anddisturbance torque.

III. SLIDING-MODE CONTROL DESIGN

A sliding-mode controller is derived in this section. Thecontrol scheme of the whole dynamic system is shown inFig 4. For the hydrostatic transmission system, the proposedsliding-mode controller provides a systematic approach to theproblem of maintaining stability and consistent performancedespite modelling imprecisions like unknown time constantsof the actuators. Here, the leakage volume flow qU andthe resulting disturbance torque τU are determined by anonlinear reduced-order disturbance observer and employedfor a correction of the inverse dynamics, see Sec. IV. For

text

Disturbance Observer

Desired Tajectrory

Hydrostatic Transmission/Drive ChainActuator P

Actuator MSliding Mode Control

Fig. 4. Block diagram of the observer-based sliding-mode control.

the control design, the the first two time derivatives of thedifference pressure and motor angular velocity are calculated.In both cases, the second time derivative depends directlyon the control input, leading to a relative degree of four thatcomplies with the system order. By inserting the actuatordynamics, the system dynamics in form of two second-ordernonlinear differential equations becomes

∆p = fP1(x, ωP ) +fP2(uM , αM )

TuM

+gP (ωP )

TuP(uP − αP ) + dPτ (τU ) + dPq(qU ), (17)

ωM = fM (x, ωP )

+gM (∆p)

TuM(uM − αM ) + dMτ (τU , τU ) + dMq(qU ).

(18)

Here, fP1,P2and gP,M are nonlinear functions of the system

states, whereas dPi,Mi, i ∈ {τ, q} denote nonlinear functionsof the disturbance variables. The vector e = [eP , eM ]comprises the tracking error of the difference pressure andthe motor angular velocity according to

eP = ∆pd −∆p, (19)eM = ωMd − ωM . (20)

As starting point for the sliding-mode controller, the time-varying switching functions si(t), i ∈ {P,M}, are definedas follows

sP (t) = eP + 2λeP + λ2∫ t

0

eP dτ, (21)

sM (t) = eM + 2λeM + λ2∫ t

0

eMdτ. (22)

Here, λ > 0 is a strictly positive constant. As can be seenfrom (21) and (22), the dynamics on the sliding surfacesi(ei) = 0 is of second order with respect to the tracking

3218

Page 4: Sliding-Mode Control of a Hydrostatic Drive Train with ...folk.ntnu.no/skoge/prost/proceedings/ecc-2013/data/papers/0840.pdf · perform breaking manoeuvres without mechanical wear

errors ei, i ∈ {P,M}. Calculating the first time derivativesof 1

2s2i , i ∈ {P,M}, and inserting (17), (18), (21) and (22),

results in1

2

d

dts2P = sP (eP + 2λeP + λ2eP ),

= sP (∆pd + 2λeP + λ2eP − fP1 −fP2

TuM

− gPTuP

(uP − αP )− dPτ − dPq), (23)

1

2

d

dts2M = sM (eM + 2λeM + λ2eM ),

= sM (ωMd + 2λeM + λ2eM − fM− gMTuM

(uM − αM )− dMτ − dMq). (24)

The following design step consists in determining a controllaw such that the switching functions according to (21) and(22) converge to zero and remain on the sliding surfacessi(ei) = 0 afterwards. This objective is achieved by design-ing an appropriate control input ui, i ∈ {P,M}, so that thecorresponding sliding-condition is satisfied [14]. As a result,the control inputs can be stated as

uP =TuPgP

[uP + φP ] + αP , (25)

uM =TuMgM

[uM + φM ] + αM , (26)

with

φp = ∆pd + 2λeP + λ2eP − fP1 −fP2

TuM− dPτ − dPq,

(27)

φM = ωMd + 2λeM + λ2eM − fM − dMτ − dMq. (28)

Here Tui, i ∈ {P,M}, represent the estimated actuator timeconstants. Substituting the designed controller (25) and (26)into (23) and (24) results in

1

2

d

dts2P = −uP sP +

(ψpP

TuPTuP

+ ψpMTuMTuM

)︸ ︷︷ ︸

ψP

sP , (29)

1

2

d

dts2M = −uMsM + ψM

TuM

TuMsM , (30)

where ψpP,pM,M are time-varying nonlinear functions. Theassumption that the actuator time constants are bounded by

0 < Ti,min ≤ Ti ≤ Ti,max, i ∈ {P,M} (31)

leads to

‖ TuiTui‖ ≤ Ti, i ∈ {P,M}. (32)

By introducing Ti > 0, i ∈ {P,M}, as positive bounds, thetime-varying terms on the right-hand side of the equation(29) and (30), respectively, are also bounded by a positivevalue

‖ψi‖Ti ≤ Fi, i ∈ {P,M}. (33)

The switching control action ui, i ∈ {P,M}, can be chosenas

ui = kisgn(si), i ∈ {P,M}, (34)

where ki = ηi + Fi denote time-varying but positive gainsand sgn(si) represent the sign-functions according to

sgn(si) =

1 si > 0

0 si = 0

−1 si < 0

, i ∈ {P,M}. (35)

By using (34) instead of ui in the equations (29) and (30), aconvergence towards the sliding surfaces si = 0 is achievedin finite time

1

2

d

dts2i ≤ −ηisgn(si)si = −ηi|si|, i ∈ {P,M}, (36)

with constant scalar values ηi, i ∈ {P,M}

IV. NONLINEAR REDUCED-ORDER DISTURBANCEOBSERVER

Disturbance behaviour and tracking accuracy in viewof model uncertainties can be significantly improved byintroducing a compensating control action provided by anonlinear reduced-order disturbance observer as describedin [15]. The observer design is based on the state equations.The key idea for the observer design is to extend the stateequations with an integrator as disturbance model

ym = f(ym, τS , τM ) , τS = 0 , (37)

where ym = [∆p, ωM ]T denotes the measurable differencepressure and motor angular velocity, τM = [αP , αM ]T isthe vector of the normalised actuator outputs and τS =[qU , τU ]T represents the vector of system disturbances. Theestimated leakage volume flow qU and disturbance torque τUfollow from

τS = H · ym + z, (38)

where H represents the observer gain matrix. The vector ofstate equations for z is given by

z = Φ (ym, τS , τM ) . (39)

The observer gain H = diag(h1, h2) and the vector Φare chosen such that the steady-state observer error τS =τS − τS converges to zero. Thus, Φ can be determined byconsidering a vanishing steady-state estimation error

˙τS = 0 = τS −H · ym − Φ (ym, τS − 0, τM ) . (40)

In view of τS = 0, equation (40) yields

Φ (ym, τS − 0, τM ) = −H · ym

= −H ·

2qPCH− 2qM

CH− qU

CH

−dV ωM

JV+ τM

JV− τU

JV

.(41)

3219

Page 5: Sliding-Mode Control of a Hydrostatic Drive Train with ...folk.ntnu.no/skoge/prost/proceedings/ecc-2013/data/papers/0840.pdf · perform breaking manoeuvres without mechanical wear

The linearised error dynamics ˙τS must be asymptoticallystable. All eigenvalues of the Jacobian are, hence, placed inthe left complex half-plane according to

∂Φ (ym, τS , τM )

∂τS=

[−h1

CH0

0 h2

JV

]!=

[−sB1 0

0 −sB2

],

(42)with sB1 > 0 and sB2 > 0. This directly leads to theobserver gains h1 = sB1 CH and h2 = −sB2 JV . Asdepicted in block diagram Fig. 4, disturbance compensa-tion is achieved by substituting the leakage volume flowand the disturbance torque in the sliding-mode control bytheir estimated values, i.e., qU = qU and τU = τU . Theasymptotic stability of the closed-loop control system hasbeen investigated thoroughly by simulations.

V. SIMULATION RESULTS

Taking into account both the measurement noise of thepressure sensors and the quantization effects in the encoder,the tracking performance as well as the steady-state accuracyw.r.t. the motor angular velocity and the difference pressureshall be investigated by simulation. Therefore, the desiredtrajectories of the pressure difference and motor angularvelocity are shown in Fig. 5 as well as their first and secondtime derivatives.

0 5 10 15 200

2

4

6x 106

t in s

Δp d

inN

/m2

Δp in N/m2

0 5 10 15 20

−2

0

2

4 x 106

t in sΔp d

inN

/(m

2s)

Δp in N/(m2 s)

0 5 10 15 20−4

−2

0

2x 106

t in s

Δp

dN

/(m

2s2

)

Δp in N/(m2 s2)

0 5 10 15 200

100

200

t in s

ωM

dra

d/s

ωMd in rad/s

0 5 10 15 20

0

50

100

t in s

ωM

dra

d/s

2

ωMd in rad/s2

0 5 10 14 0−50

0

50

t in s

ωM

dra

d/s

3

ωMd in rad/s3

Fig. 5. Desired trajectories for the pressure difference and motor angularvelocity.

The desired angular velocity ωMd is formulated by a7th order polynomial, and the desired trajectory of ∆pd iscalculated with a desired value for the normalized motorinput uMd = 0.8, which can be adapted to the given drivesituation. This leads to the following relationships for ∆pdand the first two time derivatives

∆pd =JV ωMd + dV ωMd + τU

0.8VM, (43)

∆pd =JV ωMd + dV ωMd + ˙τU

0.8VM, (44)

∆pd =JV

...ωMd + dV ωMd + ¨τU

0.8VM. (45)

Note that ∆pd depends on both ωMd and estimated dis-turbance τU . The disturbance torque that is employed inthe simulation consists of two terms: 1) a static nonlinearfriction model and 2) a vehicle mass increased by 10%. Inthe implementation, the time derivatives of τU , ˙τU and ¨τU ,are neglected in the equations (44) and (45). The estimationresults concerning the disturbance torque shown in Fig. 6(upper part) are quite good and, hence, used for a disturbancecompensation. The same holds for the estimated leakagevolume flow (proportional to difference pressure), which isshown in Fig. 6 (lower part).

0 5 10 15 20−5

0

5

10

15

20x 10−5

q Uin

m3/s

t in s

0 5 10 15 200

5

10

15

τ uin

Nm

t in s

simulatedestimated

simulatedestimated

τU = 0.1JV ωMd + 7tanh(ωMd

0.01)

qU = (2kIL + kEL)Δp

Fig. 6. Estimation of the considered disturbance torque (upper part) andleakage volume flow (lower part).

Fig. 7 depicts the tracking performance regarding thedifference pressure and the motor angular velocity as wellas the tracking errors with only imperfectly known timeconstants (−50 % deviation from their nominal values) inthe actuator dynamics. It indicates that the sliding-modecontroller with the disturbance observer works quite wellin the case of uncertainty in the actuator dynamics: themotor angular velocity is tracked with an error boundedby 0.5 rad/s, whereas the error of the difference pressureis bounded by 0.03 bar. Fig. 8 shows the correspondingnormalized tilt angles of the hydraulic pump and motor αi,i ∈ {P,M}

The estimated leakage volume flow and tracking errorsdepicted in Fig. 9 emphasise that the combination of thesliding-mode controller and the observer-based disturbancecompensation guarantees an outstanding tracking perfor-mance despite an unexpected additional leakage. This addi-tional leakage oil flow could be caused by a pressure reliefvalve that is not perfectly closed.

VI. CONCLUSIONS

In this paper, a sliding-mode trajectory tracking control incombination with a disturbance observer is presented for ahydrostatic transmission with uncertainties. The modelling ofthis mechatronic system leads to a system of four nonlineardifferential equations. Two sources of model uncertainty areconsidered in the paper: 1) unknown time constants in theactuator dynamics and 2) unknown disturbances in formof a leakage volume flow and a disturbance torque acting

3220

Page 6: Sliding-Mode Control of a Hydrostatic Drive Train with ...folk.ntnu.no/skoge/prost/proceedings/ecc-2013/data/papers/0840.pdf · perform breaking manoeuvres without mechanical wear

0 5 10 15 200

1

2

3

4

5

6

7 x 106

Δp

inN

/m2

t in s

0 5 10 15 200

50

100

150

200

250

ωM

inra

d/s

t in s

ωMdωM

Δpd

Δp

0 5 10 15 20−0.2

0

0.2

0.4

0.6

e ωM

inra

d/s

t in s

0 5 10 15 20−1.5

−1

−0.5

0

0.5

1

1.5

2 x 105

e Δp

inN

/m2

t in s

Fig. 7. Tracking performance using sliding-mode control for the hydrostaticdrive train with only imperfectly known time constants in the actuatordynamics.

0 5 10 15 200

0.2

0.4

0.6

0.8

1

αP

,M

t in s

αP

αM

Fig. 8. Normalised tilt angles of both the hydraulic pump and the motor.

on the hydraulic motor. Reasonable assumptions on theboundedness of the time constants in the actuator dynamicsenables to select proper gains for the sliding-mode control.The control approach guarantees the robust stability of theclosed-loop system. The simulation results show that theproposed control leads to a high tracking performance evenin the case of imperfect knowledge of the actuator dynamics.

REFERENCES

[1] M. Jelali and A. Kroll, Hydraulic Servo-systems: Modelling, Identifi-cation and Control. London, UK: Springer-Verlag, 2003.

[2] N. Diepeveen and A. Laguna, “Dynamics modelling of fluid powertransmissions for wind turbine,” in EWEA Offshore (2011), Amster-dam, Netherland, 2011.

[3] B. Dolan and H. Aschemann, “Control of a wind turbine with ahydrostatic transmission - an extended linearisation approach,” in 17thInternational Conference on Methods and Models in Automation and

0 5 10 15 20−0.5

0

0.5

1

1.5

2

2.5x 10−4

q Uin

m3/s

t in s

simulatedestimated

0 5 10 15 20−0.2

0

0.2

0.4

0.6

e ωin

rad/s

t in s

0 5 10 15 20−1.5

−1

−0.5

0

0.5

1

1.5

2 x 105

e Δp

inN

/m2

t in s

qU = (2kIL + kEL)Δp

ΔqU = ΔkΔp

Fig. 9. Estimation of the leakage volume flow with an additional leakage at4.5 s as well as the corresponding tracking errors of motor angular velocityand difference pressure.

Robotics (MMAR) (2012), Miedzyzdroje, Poland, 27-30 Aug 2012, pp.445–450.

[4] A. Pusha, A. Izadian, S. Hamzehlouia, N. Girrens, and S. Anwar,“Modelling of gearless wind power transfer,” in IECON 2011-37thAnnual Conference on IEEE Industrial Electronics Society, 7-10 Nov2011, pp. 3176–3179.

[5] S. Stoll, M. Kliffken, M. Behm, and X. Wang, “Regelungskonzepte furhydrostatische Antriebe in mobilen Arbeitsmaschinen (in German),” at– Automatisierungstechnik, p. 5 (2007) 2, 19-21 Aug. 2007.

[6] H. Aschemann, J. Ritzke, and H. Schulte, “Model-based nonlineartrajectory control of a drive chain with hydrostatic transmission,” in14th International Conference on Methods and Models in Automationand Robotics (2009), vol. 14, Miedzyzdroje, Poland, 19-21 Aug. 2009.

[7] J. Ritzke and H. Aschemann, “Design and experimental validationof nonlinear trajectory control of a drive chain with hydrostatictransmission,” in 12th Scandinavian International Conference on FluidPower, Tampere, Finland, 18-21 May 2011.

[8] A. Koshkouei, K. Burnham, and A. Zinober, “Dynamic Sliding ModeControl Design,” in IEE Proc.-Control Theory Appl., vol. 152 No.4,July. 2005.

[9] J.-J. E. Slotine and W. Li., Applied Nonlinear Control. Prentice Hall,1991.

[10] K. D. Yang, V. I. Utkin, and U. Ozguner., “A Control Engineer’s Guideto Sliding Mode Control.” in IEEE Transactions on Control SystemsTechnology., vol. 7 No.3, May. 1999.

[11] H. Schulte, “Control-oriented modelling of hydrostatic transmissionsusing Takagi-Sugeno fuzzy systems,” in IEEE International Confer-ence on Fuzzy Systems (2007), London, UK, 2007, pp. 2030–2035.

[12] A. Kugi, K. Schlacher, H. Aitzetmueller, and G. Hirmann, “Mod-elling and simulation of a hydrostatic transmission with variable-displacement pump,” Mathematics and Computers in Simulation,vol. 53, pp. 409–414, 2000.

[13] P. Rohner, Industrial Hydraulic Control: A Textbook for Fluid PowerTechnicians. John Wiley & Sons, 2004.

[14] J. E. Slotine and W. Li, Applied Nonlinear Control. Prentice Hall,1991.

[15] B. Friedland, Advanced Control System Design. Prentice Hall, 1996.

3221