You are on page 1of 8

DSC TOC

Proceedings of 2001 ASME International Mechanical Engineering Congress and Exposition November 11-16, 2001, New York, NY

Proceedings of IMECE 2001 2001 ASME International Mechanical Engineering Congress & Exhibition November 11-16, 2001, New York, New York, USA

IMECE2001/DSC-24563

IMECE01/DSC-2B-1

DYNAMIC REDESIGN OF A FLOW CONTROL SERVO-VALVE USING A PRESSURE CONTROL PILOT

Perry Y. Li Department of Mechanical Engineering University of Minnesota 111 Church St. SE, Minneapolis, Minnesota 55455 Email: pli @me.umn.edu

ABSTRACT In this paper, the dynamic performance of an unconventional two-spool flow control servo valve using a pressure control pilot is analyzed. Such valves are less expensive than typical servovalves but also tend to be limited in their dynamic performance. Based on a previously developed eight state nonlinear model, we develop a simplified linear model which is able to capture the essential dynamics of the valve. Using root locus analysis method, the limitation in dynamic performance is shown to be due to a "zero" introduced by the structure of the interconnection of the subsystems. Design parameters that move the "zero" further to the left half plane, and do not adversely affect other steady state criteria are identified. The effectiveness of these parameters to improve the dynamic performance is demonstrated.

Introduction Most designs of servo flow control valves [4] consist of a single spool boost stage, a nozzle flapper pilot, and a feedback wire. These valves have very high performance but tend to be expensive because of the stringent manufacturing tolerances and the complicated assembly process. A less common, commercially available alternate design (Fig. 1) consists of a pressure control pilot stage and a boost stage that uses two separate spools to independently meter flow into and out of the valve. Since the critical dimensions are easier to adjust, post assembly and feedback wire is not used, such valves are easier to manufacture and to assemble. Consequently they tend to be cheaper. Readers are also referred to [1], in which an experimentally validated
I

complete physical model is presented, for a more detailed discussion of the advantages of the unconventional two-spool servo valve. Despite these advantages, the unconventional two-spool servo valve design tends to have lower dynamic performance in terms of bandwidth compared to the conventional servo valve design utilizing a single spool and a feedback wire. For example, the valves studied in [1], [3] have bandwidths between 15-40Hz whereas conventional servo-valves of similar rating can have bandwidths of over 100Hz. It would therefore be advantageous if dynamic response of the two-spool design can be improved. In this paper, we study the unconventional two-spool design so as to understand the nature of the performance limitation, and to suggest design modifications for potential performance improvements. The experimentally validated eight state nonlinear physical model derived in [1] consists of the interconnection of three subsystems. This model is similar to the one constructed by Lin and Akers previously [3]. Using this model, we develop a simplified five state linear model that retains the interconnection structure as well as the predominant dynamics. The reduced model reveals a puzzling aspect of the valve dynamics in that each of the three subsystems has bandwidth at least an order of magnitude higher than the bandwidth of the complete model. Using simple root locus arguments, it is found that the way in which the subsystems are interconnected creates a "zero" which causes the bandwidth of the interconnected system to be significantly lower than the individual subsystems. Based on this insight, several systems parameters that can potentially improve the dynamic performance without adversely affecting the steady state perforCopyright 2001 by ASME

Null 7 S

)rtag

Magi

dagnetic )r Plata

)n Plate

This tends to increase the pressure P2 and to decrease the pressure Pl. The differential pressure acts on the two ends of the two spools in the boost stage. Since the spools are spring centered, their equilibrium displacements will be roughly proportional to the differential pilot pressure and inversely proportional to the spring stiffness. Flow into and out of the valve are separately metered in and out according to the displacements of the two spools. 2.1
Fleview of full s t a t e m o d e l

Tran~

oolB

The serve valve can be considered an interconnection of three subsystems, l) the pilot subsystem whose states are the flapper displacement x f (left to right positive) and velocity ~f; 2) the pressure chambers whose states are the chamber pressures P] and/)2; and the boost stage spool dynamics whose states are the displacements and velocities of the two spools Xa, Yea and Xb, ~b. Therefore, the total number of states is eight. Following [1], the dynamics of the pilot subsystem can be represented by:

ng

Mp:if + Bt~ f -IKpxf = An(PI - P2) + 4~C2: [(Xfo+Xf)2Pl - (xfo-xf)2p2] +g(xf, i) (1)
....... a.v.

Figure 1. A two-spool flow control servo-valve using a pressure control pilot state.

mance (such as flow gain) are identified, and their effects demonstrated. The rest of the paper is organized as follows. In section 2, we formulate a simplified model of the two-spool flow control serve valve. The interconnections of the three linearized subsystems are studied using root locus techniques in section 3. Section 4 presents the effort to optimize the performance by applying the insights gained in section 3. Sections 5 and 6 contain discussion and concluding remarks respectively.

where Mp, Bp and ~:p are respectively the combined inertia, damping and mechanical stiffness of the flapper, Xfo is the null nozzle-flapper gap when x f = O, and Cdf is the discharge coefficient of the flapper-nozzle, An is the nozzle area. The first and second terms on the right hand side correspond respectively to the pressure and the flow induced forces at the nozzle, and g(xf, i), which is a highly nonlinear function (see [1] for details), represents the force on the flapper generated by the electromagnetic torque motor with input current i. The dynamics of the two boost stage spools are given by: MsXa "Jrnsfa "~ 2Ksxa = (P2 - Pl)As -

?f(xa:Pa)fCa-

:f(Xa,Pa)Xa

(2)

transient flow force

steady state flow force

Simplified model of the two-spool flow control servo valve The unconventional flow control serve valve shown in Fig. 1 uses a two-spool boost stage and a pressure control flappernozzle pilot stage. The two stages are separated by a simple transition plate and connected via two pressure chambers. The design philosophy of the valve is as follows. The pressure control pilot stage generates a differential pressure between the two fluid chambers adjacent to the flapper, determined by the current input to and the torque generated by the electromagnetic torque motor. For example, if the torque motor applies a counter clockwise torque to the flapper, the flapper displaces to the right. 2

MsJib + Bs:h, + 2Ksxb = (P2 - PI)As + (3)


transient flow force steady state flow force

where Xa and Xb represent the upward displacements of spools A and B respectively, Ms is the spool inertia, Bs is the viscous damping coefficient, 2Ks is the total stiffness of the two springs above and below the spools (Fig. 1), As is the spool area. The steady state and transient flow forces manifest themselves as spring forces and positive/negative damping forces with Kf(.,.) > O, and Bf(x,P) > 0 when x > 0, BI(x,P) < 0 when Copyright 2001 by ASME

x < 0, and not well defined when x = 0. Therefore, depending on the sign of the spool displacement, the transient flow force may introduce negative damping effects. The third subsystem is associated with the dynamics of the pressures Pl, P2 in respectively the upper and the lower fluid chambers connecting the pilot stage and the boost stage spool. QI (Pl,xf) VI (t)
f'l (t)

1. In the spool system, we consider only the total spool displacement (not the displacements of the individual spools) Z(t)
=

--(Xa(t)+Xb(t)).

2. In the pressure chamber system, we consider only the differential pressure AP(t) = Pl(t)
-

P1

; [~2 = ~ Q2(P2,xf) - f'2(t) V2(t)

(4)

P2(t).

Here, Ol(Pl,xf) and Q2(Pz,xy) are the total flows into the upper and lower chambers, Vl(t) and V2(t) are the volumes in the chambers, and 13 is the compressibility of the fluid. Q] (Pl,Xf) and QE(P2,xf) are comprised of the flows from the pilot supply orifice, leakage past the nozzle, and to a small extent, leakage past the spools: for i = 1 and 2,

Qi

:CdoAo~2(Psp-Pi)--CdfEOn(xfo-bxf)(-~Pi--leakagei

3. The transient component of the flow forces and leakage flows past the spools are ignored. 4. The pilot stage dynamics in (1), the spool dynamics in (2)(3) and the differential pressure dynamics obtained from (4) are linearized at the equilibrium condition given by flapper displacement xf = 0, chamber pressures Pl = P2 =:/~, spool displacements Xa = Xb = 0, spool velocities :~a = :~b = 0, chamber volumes V1 = V2 = (I/10+ V20)/2 =: I7', and work pressures Pa = Pb = Ps/2 where Ps is the boost stage supply pressure. The resulting reduced linear models for the pilot, chamber pressure and spool subsystems are respectively:

(5)
where "+" sign is used for i = 1 and " - " sign is used for i = 2, Psp is the pilot supply pressure (which is usually lower than the supply pressure for the boost stage), Ao is the area of the orifice to the supply pressure, Cao and Caf are the discharge coefficients of the orifice to the supply and the gap between the flapper and nozzle. The first two terms in (5) are monotonically decreasing functions of Pi. Thus, they provide at least local exponential stability for the pressure dynamics (4). Notice also the pilot stage communicates with the chamber pressures via Q I and Q2 since they depend on the flapper displacement xy. On the other hand, the pressure chambers are affected by the boost stage spool dynamics via V1(t), Vz(t), f'l (t) and f'2(/) in (4) since V1 = Vlo - Asxa - AsXb, V2 = V2o + Asxa + AsXb fzl = -- ~'2 = -AsYca -- as:~b (6) (7) (8)

Pilot subsystem
Mp2f + Bp~f + Kpxf = (An + B)AP + G. i (9)

Pressure chamber subsystem


[~xf + ~ - ~ . ( t ) A P = - otI~AP ~ - 2T~ (10)

Spool subsystem
Ms~,+BsY.+(2Ks+K/)E=

-2AsAP(t)

(11)

where

G:= 3g ~i xz=O,i=o

Ky := Ky(O,Ps/2)
~g

K# = ff(p - 16~C2fx fo # - ~xf x/=0,i=0 B:=4~C2f~f' T---- ~--'x'Tf


~QJ : , v =0 = _

where Vlo and V2o are the chamber volumes when the spools are centered (Xa = xb = 0). For details of the model, readers are referred to [1].
2.2 Reduced order linear model
In order to obtain meaningful design information, we

= ~xf :r~Cf=0

~Qz

consider a reduced 5th order linear model. This is achieved as follows: 3

B, T and a are all positive quantities. The expression for Kp shows that that nozzle flow forces and the magnetics tend to offset the mechanical stiffness Rp of the flapper. The term An + B in Copyright 2001 by ASME

Step responses of nonlinear and linear models

Chamber/Pilot

system

Pilot

:i
1_1

.........

i .........

.................................................................
t

IM,,~+B:+ 2K,+K:,V
Spool

2,,

...........................

i ........

0 !~ 2 1 -140 0.01 002

~
! : :

Figure 3. loop as
0.1

theinner loop.

Block diagram for root locus analysis upper feedback The actual valve dynamics are obtained when

with the

KI----K2= 1.
Root locus of the d i f l ~ t i a l pressure I pilot intecolnectlon ! !

0.03

I I 0.04 0.05 0.06 Time - sec

0.07

0.08

0,09

Figure 2.

Differential pressure response to a 20mA step input current:

500

full nonlinear model and simplified linearized model.

2ooo

(9) is the apparent nozzle area of the flapper-nozzle upon which the pressure and the nozzle flow forces act. From (10), a is the convergence rate of the pressure chamber normalized by the inverse of the chamber capacitance, ~/V. To verify that the linearized model in (9)-(11) indeed captures the dominant dynamics of the valve, the responses to the step current input are simulated for an input step size of i = 20mA (50% full range) using the complete nonlinear model in [1] and the linearized model in (9)-(11). The step responses of the differential pressure AP(t) are very close (Fig. 2). The 64% rise-time for the linearized and nonlinear models are 8.1 ms and 8.2ms respectively. The similarity among the responses of the model in (9)-(11) and of the full nonlinear model in [1] suggests that the dynamics of the system represented by the interconnection between the pilot, pressure and spool dynamics are well captured by the reduced order linear dynamics. Each of the pilot (9), differential pressure (10), and the boost spool (9) subsystems are stable. Using physical parameters of the valve that are verified in [1], it can be shown that the pilot subsystem has a natural frequency of (On,p = 3037 rad/s, and a damping ratio of ~v = 0.91; the differential chamber pressure subsystem has an eigenvalue of - a ~ = -1599rad/s; the boost spool subsystem has a natural frequency of (on: = 1861rad/s with a negligible damping ratio of ~s = 0.025. The eigenvalues of the combined system are at -137.2rad/s, -702 4- 16475jrad/s, -2837 4- 700jrad/s. We point out that the dominant pole is at - 137.2rad/s which is consistent with the fact that the 64% risetime of the fully linearized model is 8.2 ms. In order to improve the dynamic performance of the valve, the dominant pole must be moved further into the left half plane. 4

F o:

"

".................................i.........i

_500

-4000

-3500

-3000

-2500

-2000 -1500 Real Ads

-1000

-500

500

Figure 4. Root locus diagram of the pilot/ chamber differentMI pressure


subsystem as Kl increases from 0 ~ oo in Fig. 3.

A n a l y s i s for p e r f o r m a n c e

limitation Wenowproceed to analyze thelinearized model (9)-(11)

to understand why the bandwidth of the valve is relatively low, whereas the natural frequency of each individual subsystem is at least an order of magnitude higher. Is the limited performance due to the fact that the spools are too lightly damped (~ = 0.025)? Or, is the fluid capacitance in the pressure chamber the reason? As we shall see, neither the spool damping nor the chamber capacitance is important. The key turns out to be the structure of the interconnection between the pressure chamber, the pilot and the spool subsystems.
3.1 Full o r d e r r o o t l o c u s

The pilot, chamber pressure and the boost spool subsystems are connected in a closed loop manner as shown in Fig. 3 with Copyright 2001 by ASME

10~lOOI Ioctis OI the intolcomlectk)n between spool and differential pressure / pilol subsystems 2

x 104ROOtIOcue et the hlt rcnnectk~ I ~ t v v ~ I~1~ and cuff~'Ont~ I~auure / IpOet

! t.~ .......... , ..........

! i .~

! 0-

! :

' ;

' i i ;

'

: ..........

, ..........

: ..........

, K2. -~1 . . . .

T ........ i

i i

K1.1
i :: i

i : i

? ;

!i

i .....

: 1

......... i

K:i

.......... i .......... i .......... i .......... i ..........

~i

..........

_r
--o

~ ~
: ! i

'

.
!

!
! i

~
Op~oppo~

!
i .:

-1

i
K1=1

i
: : :

:.

-1.s -1.5

i :

i i
-1200

.... i
-1000

i "-800

: i
-200

! ! i

i ! 600

-'~
-4000

-1400

. -80O .

-2'

Real Axis

.-400

~o do

'

-3500

-3000

-2500

-2000 -15(X) Rtml Axis

-1000

-500

500

Figure 5.

Root locus diagram of the

spool subsystem and

the pilot /

locus diagram of outer loop with inner loop being the spool / chamber system with K2 = 1 as K] increases from 0 --> oo.
Figure 6. Partial root

chamber differential pressure subsystem as/(2 increases from 0 ~ ~,


and KI = 1 in Fig. 3.

Two other poles on the far left are not included,

Kl = K2 = 1. To understand the effect of the interconnection, we apply Evan's root locus technique [2] to investigate how the closed loop eigenvalues migrate as the parameters in the system are varied. Consider first the inner loop in Fig. 3 which is the interconnection between the pilot and the differential pressure subsystems. Fig. 4 shows the loci of the closed loop poles of the inner loop as the fictitious gain Kl is varied from 0 ~ oo. K1 = 1 corresponds to the gain in the actual loop in the present valve design. Figure 5 shows the locus of the closed loop poles of the outer loop system in Figure 3 as the fictitious gain K2 is increased. The set of poles at K2 = 1 are the actual poles in the valve. As expected, when K2 = 1, the dominant pole is at p = - 137.2rad/s (which is the reason why the dynamic performance is limited). Notice that the real parts of all the other eigenvalue locations are significantly more negative. Figure 5 shows that for the present valve design (i.e. K2 ~ 1), the pole locations are well approximated by the asymptotic behaviors of the root locus. These are governed by the "open loop" pole and zero configurations. In particular, the dominant pole at p = - 137.2 is being attracted to the "zero" at 0. Two other poles are close to the zeros at the pilot's open loop pole locations. The remaining two poles are also close to the asymptotes. Since the damping in the spools can only affect the asymptotes slightly, contrary to our initial speculation, the negligible damping of the spool does not contribute significantly to the relatively poor dynamic performance of the valve. Rather, the reason is due to the "zero" at the origin. This "zero" is present because the spool subsystem interacts with the chamber pressure dynamics via ~. As far as the loop gains KI and K2 are concerned, from Fig. 5, decreasing/(2 delays the migration of the dominant pole to the 5

zero at the origin. Similarly, if we had interconnected the differential pressure and the spool (lower loop) first before connecting the pilot system in Fig. 3 and derived the corresponding root loci, then we would have noticed that by increasing Kl in Fig. 3, the partial root locus is shown in Fig. 6. Notice that the dominant pole migrates from an open loop pole near the origin towards the left half plane as Kt increases.

3.2

Reduced

order root locus

The root locus analysis above indicate that the four complex poles of the valve design can be approximated by their asymptotic behaviors. In addition, the complex poles originate from the poles associated with the pilot and the spool subsystems. This suggests that we may approximate the behavior of the dominant eigenvalue of the valve by considering the pilot and the spool systems as quasi-static systems, i.e. by assuming that at each instant, the spools and the flapper are in static equilibria with the instantaneous differential pressure. The resulting configuration is given in Fig. 7. Indeed, the closed loop pole of the reduced order quasi-static system in Fig. 7 is p = -t27.7rad/s which is very close to the actual dominant pole (-137.2rad/s). The characteristic equation for the system in Fig. 7 is given by:

2r rp ) -v 2Ks+r/ s = 0 . s

(12)

Copyright 2001 by ASME

ISimplified sl~ol/pilot system

in Section 3.2, performance is limited by the open loop zero of the reduced order system. Since the dominant pole location of the valve will be close to the "zero" location, the "zero" location must be moved to the left if the dynamic performance is to improve. The expression for this blocking "zero" is: ~+2#p 2AsGs '

Figure7. Blockdiagramwiththe pilotand the spoolsubsystemsapproximated by their staticsystems.

(14)

pole location where Gp := ~ Kp and Gs := 2Ks+Kfs zar are respectively the steady state ratio of the flapper displacement x f to the differential pressure AP, and steady state ratio between the total spool displacement - Z and AP. Therefore, from (14), the key design parameters in the various subsystems are:

Zero at /4
_ ~+2-~p

2AsGs Figure 8. Root locus of the reduced order model in Fig.7.

Nozzle-flapper: The apparent nozzle area An + B, which can


The root locus for positive v~ is given in Fig. 8 which shows that the performance of the system would be limited by a "zero" at be modified by changing the physical size of the nozzle and the gap between nozzle-flapper. 7 is the sensitivity of the nozzle flow to flapper displacement which can also be modified by changing the nozzle diameter. The apparent flapper stiffness Kp is affected by the mechanical stiffness, the negative magnetic stiffness and the negative nozzle flow induced stiffness. Boost stage spools: The spool areaAs, and the centering spring stiffness Ks. Pressure Chambers: a, which is the convergence rate of the differential pressure normalized by the inverse chamber capacitance, ~/V'. Shifting the "zero" in (14) to the left can be achieved by 1) modifying the flapper nozzle design so as to increase "yGt, (increase An + B, decrease K v, increase 7); 2) modifying the boost stage spool design so as to decrease AsGs (decrease As, increase Ks); or 3) by modifying the open loop convergence rate of the pressure chamber dynamics so as to increase a.
4.2 Steady State Criteria In addition to their effects on the dynamic response of the valve, it is important also to evaluate the effects of these design parameters on the operating pressure and motion ranges, and ultimately the flow gain of the valve. We determine these from the D.C. components of the transfer functions assuming 2)'/ix is large.

a + 27 An + B

I t, 2K:-c s ) "

(13)

In the current valve design, the "zero" of the quasi-static model is at - 129.3 rad/s. Since the pole (of the quasi-static model) is already at - 127.7 rad/s, the performance cannot be significantly improved by increasing 13/IS'which is the ratio between the fluid compressibility and the chamber volume. Although the presence of finite pressure chamber dynamics is essential for the existence of the slow valve dynamics, the values of the chamber volume, the compressibility, or capacitance (i.e. ~'/13) do not matter significantly. Rather, the performance limitation is determined by the feedback structure itself. The performance limitation can be alleviated if the "zero" in (13) can be moved further to the left. Consistent with the previous analysis, this can be achieved by increasing the pilot loop gain and by decreasing the spool loop gain (i.e. Kl and K2 respectively in Fig. 3).

4 Dynamic Redesign 4.1 Dynamic performance Since the dominant pole of the system limits the performance of the valve, it must be moved further to the left half plane to improve dynamic performance. According to the root locus analysis in Section 3.1, this can be achieved if the loop gain in the upper loop is increased, and the loop gain in the lower loop is decreased. From the reduced order root locus analysis

Differential pressure gain:


AP(s) s=O : l(s) " 6
(2T/a)GP G G I + (2T/a)G p An +--'~B~' - An +----B

Copyright 2001 by ASME

Flapper displacement gain: at'


xy(S) s= (a) l(s) =-- "~ a G -~'~- s=O ~ 2T An + B
AP(s)
x:

Kp.[.
QL

(An+B) ~f T"f
~/

Gs.l. As.l.
-

aT
x

q
X

Spool displacement gain:


Z(s) s=0 AP(s) s=0 GsG ](s) = -Gs ~ ~" An + B

Figure 9. Consequences on the static criteria when various design parameters are used to improve the dynamic performance. Direction of the arrow indicates the direction of proposed change. "X" represents significant degradation, "~/" represents some improvement. For each column, the variables in all other columns are assumed to be constant.

Flow gain:

oL(s) ,=0
l(s)

Cdw Z(s),=0 Cdw GG, 2 V " ff 7 " ~ ~ 2 V-ff A---~B

where G = ~/for the torque motor, Ps is the supply pressure for the boost stage. For a given input current, it is generally preferable that AP and xf be small, and the flow QL be large. Because the pilot supply pressure Psp is limited, large AP excursion reduces the operating range. On the other hand, large xy generally requires large magnetic airgap, which can complicate the design of the torque motor. A large flow gain is desirable so that a small input current can be used to control large flows. The consequences on these criteria when the various parameters are used to improve the dynamics performance are summarized in Figure 9. If the apparent nozzle area A n + B is increased, for the same input current, the differential pressure and the steady state flapper displacement will be decreased. Unfortunately, since the flow gain has also been proportionately reduced, more force is required from the torque motor to achieve the same flow. Similarly, decreasing Gs of the spool system will decrease the flow gain significantly. Increasing (x will have the adverse effect of increasing flapper motion. The design parameters that do not adversely affect the steady state criteria significantly are the apparent flapper stiffness Kp, the spool area As, (while maintaining Gs constant) and T of the nozzle-flapper. Of these three parameters, in fact only Kp and As can be used independently to improve dynamic performance without affecting the steady state performance. T, which is the sensitivity of the chamber flow to flapper displacement, cannot be varied to any significant degree without affecting the apparent nozzle area An + B. Indeed, even a 5% increase in ~/necessitates an increase in the actual nozzle area An by over 140%! This can potentially increase the apparent nozzle area An + B which in turn decreases the steady state flow gain.

bandwidth of the system should double and the flow gain should remain the same. Ignoring Kfs (spring constant due to steady state flow force) in Gs, Ks is halved. Figure 10a) shows that the rise time of the modified 20mA step response has been reduced from 8.2ms to 4.2ms. The quasi-static flow-current relationship in Fig. 10b), generated using a 5Hz sinusoidal current input, shows that the flow-gain of the modified valve is only slightly smaller than in the original design. The small decrease is due to the fact that Gs is not exactly kept constant by ignoring Kfs. The modified valve also shows a decreased hysteresis, which is consistent with improved dynamic response.
4.4 R e d u c i n g t h e f l a p p e r s t i f f n e s s Kp Kp is the apparent flapper stiffness given by:

~g Kp = Kp - 16nC~fx foP - ~xy x/=0,i=0"

It can be reduced by reducing the mechanical stiffness Kp or by increasing the nozzle flapper gap Xfo, or by increasing the magnitude of the negative stiffness due to the permanent magnet in the torque motor. We consider reducing the mechanical stiffness Kp --+ 0.75Rp so that Kp is reduced to 54% of the original value. The location of the "zero" is expected to migrate from -129.3 rad/s to -222.3 rad/s. Figure 11 shows that the 64% rise time has indeed been reduced from 8.2ms to 4.9ms.

Discussion

Reducing the Spool Area As According to the expression of the "zero" in (14), if spool area As is reduced by 50% and Gs is maintained constant, the
7

4.3

The original design philosophy of the two spool flow control servo-valve using a pressure control pilot is that the pilot stage first establishes a differential pressure, which in turn positions the two boost stage spools according to the stiffness of the centering spring. This assumes that the pilot stage and the boost spool stage are in a cascade configuration. In the actual configuration (Fig. 3), the pilot stage and the boost stage are in fact in a closed loop configuration. As the differential pressure moves the spools, the differential pressure itself is changed. This upsets Copyright 2001 by ASME

F------r--~

......... i.......... i.......... i.......... :. . . . . . . . .

===,,.

I i
-0.0 ~iili

it

....
...... i.........::........ .........:: .........:........ i....... i.......

::

i i~\s/i~ \ :'~

i /

i\ /, ( i

::

:--..__~_ :
i i
0.03 0.04

i
-30 0.0l

'i-0.05 "nine - s

..........................
0.00 0.07 0.0e 0.09 0.1

0.(~

Curmnt Ilow relaUonshdpswhe~ A= ia halved

the effectiveness of the pilot stage to establish the intended differential pressure. As this feedback effect from the boost stage to the pilot stage is reduced, we recover the original intention for a cascade configuration. This is exactly what reducing Gs (e.g. by decreasing the spool area As, or increasing Ks) does. Similarly, the relative importance of the feedback effect of the spool motion is ameliorated if the effectiveness of the pilot stage is improved. This is achieved by increasing 7 or Gt,. The "zero" in Fig. 7 (which limits the dynamic performance) exists because the open-loop spool subsystem itself has a zero at the origin (see Fig. 3). An interesting avenue of future investigation is to remove this "zero" in the spool system. Since the spools affect the pressure dynamics via the spool velocities, removing the "zero" implies that the spools should be damped, not spring loaded as is presently the situation. In this case, the current input would control the spool speed, and hence the flow acceleration (aa-~t) rather than the flow rate itself. Such a device would be consistent with current controlled electromechanical motors which are typically considered torque or acceleration devices.

Conclusions

Figure 10. Step response (left) and current-flow relationship of the valve modified by As ~ As/2, Ks +-"Ks/2.
! i j , ,

:
........

!i

!
i......... ! :

! ......... i i ..... ! ........ i ........ i i i ! i i i i i i i : i

i.......... i ......... ~......... i ........ i ! : i i i i i ..... ~........ ! ! ........ i

The dynamic response of a unconventional two-spool flow control serve-valve using a pressure control pilot stage has been analyzed. Using a reduced order linear model and simple root locus analysis, the limited dynamic performance of the valve has been shown to be related to the existence of a "zero" in the interconnection of the subsystems. Based on the analysis, reducing spool area or the apparent flapper stiffness have been identified and shown to be effective in improving dynamic performance, while maintaining steady state performance such as flow gain. In this paper, simple techniques taught at the undergraduate level, such as local linearization and root locus are used. Although the valve model is highly nonlinear, these simple techniques are still useful in providing fundamental insights into the design of the valves, especially how the structure of subsystem interconnection can affect the system performance. These techniques should be amenable to the analysis and optimization of other hydraulic components as well.
References
[1] Randall T. Anderson and Perry Y. Li. Mathematical modeling of a two spool flow control servovalve. In Proceedings of the ASME Dynamic Systems and Control Division, IMECE Orlando, FL, volume DSC-Vol. 69-1, pages 321328, 2000. Also submitted to the ASME Journal of Dynamic Systems, Measurement and Control. [2] Gene E Franklin, J. David Powell, and Abbas Emami-Naeini. Feedback control of dynamic systems. Addison Wesley, third edition, 1995. [3] S-C. J. Lin and A. Akers. Modeling and analysis of the dynamics of a flow control servovalve that uses a two-spool configuration. In Proceedings of the ASME Winter Annual Meeting, volume WA90/FPST-3, 1990. [4] Hebert E. Merritt. Hydraulic Control Systems. John Wiley and Sons, 1967.

........ ~ ............. i ,

:
i
0.01 o,G2

..... _..... _..... _ ......................


i
0.04

i
0.06 Time-see

i
0.00 0.07 0.00 o.o9 0,1

0.03

Figure 11. Step response of the valve modified by Kp <-- 0.75/~p


(Kp +-- 0.54Kp). Note that 64% rise time has been reduced from 8.2ms
to 4.9ms.

Copyright 2001 by ASME

You might also like