Sunteți pe pagina 1din 20

29.

The Loop-Shaping Approach

George Papageorgiou 1, K e i t h Glover 1


and Rick A. Hyde 2

A b s t r a c t . The 7/oo loop-shaping design procedure (referred to


hereafter as LSDP) was used to design fixed gain controllers for
the HIRM. It was found that a single fixed gain controller designed
for the linearisation ±ni3005, performed adequately over the whole
flight envelope of the HIRM. As scheduling would most probably
be required for a real fighter aircraft with a much wider flight enve-
lope, the control law has been developed with a structure suitable
for scheduling. As the longitudinal and lateral dynamics are es-
sentially decoupled, two separate controllers were designed. The
lateral controller has an inner loop for stabilisation and an outer
loop for tracking. The longitudinal controller has one single loop
that provides both functions. A second fixed gain controller was
designed for the HIRM using the multi-model approach described
on pp. 67-72 in [247]. This method enables the designer to find a
fixed gain controller that makes the normalised coprime factor cost
function flat over the whole flight envelope. Details about how to
apply this method can be found in [191]. The design team found it
relatively easy, even with the limited time available, to satisfy the
specifications set out in Chapter 27.

29.1 Introduction
The LSDP is an intuitive method for designing fixed gain robust controllers. A
tutorial on 7-/00 loop-shaping is given in Chapter 7. A controller that has been
designed using 7-{oo loop-shaping provides robust stability to coprime factor
uncertainty. Coprime factor uncertainty is a general type of uncertainty much
in the same way that single-input single-output (SISO) gain and phase margins
are. Therefore, when there is little detailed knowledge about the uncertainty
present in a plant the LSDP is a good method for designing robust controllers.
The difference to gain and phase margins is that coprime factor uncertainty
can be used to directly address robustness in systems with multiple feedback
loops, i.e. multivariable systems (see pp. 240-244 in [266]).
The LSDP has been used in a variety of applications and studies. Most
relevant to the HIRM design challenge is the work in [120] where a flight con-
1Cambridge University Engineering Department, Cambridge CB2 1PZ, England
2Cambridge Control Limited, Cambridge CB4 4WZ, England

464
trol law was developed for the DRA Bedford Research Harrier XW175 and
subsequently flight tested. The control law performed well and is the subject
of on-going work. The experience of developing this control law showed that
7-/00 loop-shaping has a number of key attributes which make it particularly
suitable for this type of application. Perhaps the most important attribute
is that the resulting controller can be written as an exact plant observer plus
state feedback. This structure allows gain scheduling, as designs at different
operating points have the same state-space structure and hence the observer
and state feedback gains can be linearly interpolated. The observer also allows
the handling of input limitations (e.g. authority and rate limits) by driving
the observer with achieved plant inputs rather than demanded ones. A third
very important attribute of 7-too loop-shaping controllers is that the observer
structure adds visibility, in that physical units and interpretation can be ap-
plied to the controller coefficients and states. This may be advantageous with
respect to flight clearance. The advantages of the observer structure are fully
exploited for the HIRM control law design.
7/00 loop-shaping has much in common with the design approach currently
used by industry, sometimes referred to as "classical control". The loop-shaping
part of the design procedure is carried out in exactly the same way that classi-
cal design is carried out: inputs and outputs are matched up, and single loop
shaping is carried out to ensure that low frequency gain is large enough, roll-
off at cross-over is not excessive and that sufficient high frequency roll-off is
provided. Once this is done, then the optimal 7-/o0 loop-shaping controller is
synthesised for this so-called "weighted plant", see Section 7.2. Hence, it is
perfectly possible to take a classical design, and to augment it with a corre-
sponding 7-/oo controller which will then modify the feedback structure so as to
allow for the multivariable nature of the system.
When designing using classical control where the system has inherent cross-
coupling (as for example with most yaw - roll augmentation systems), corre-
sponding cross-terms are put into the controller. Design of these terms is not
always straightforward in that their effect on the feedback loops is not ad-
dressed directly, and some iteration may be required. With the LSDP, these
cross-terms can be left to the 7-/00 synthesis part of the design. In the authors'
view, one of the prime motivations for 7-/0o loop-shaping is the potential for de-
sign time reduction, particularly when dealing with multi-input multi-output
systems with strong cross-coupling.

29.2 T h e Controller A r c h i t e c t u r e
The first stage of the design process is to select the control law architecture.
Selection of the architecture is an essentially design method independent task,
and the reasoning used for the presented design is much the same as would be
used for a classical control law. However, "Ho0 loop-shaping and related robust
optimal control methods could have been used to help select the architecture by
looking at the robustness implications of each candidate architecture, e.g. ex-

465
amination of the robustness implications of different canard and taileron blend-
ing schemes would be possible. Limited time available precluded this type of
analysis. If a full control law design for a prototype or production aircraft was
undertaken, control law structure selection with reference to robustness impli-
cations is definitely recommended. In selecting the structure here, robustness
requirements are taken into account in a more heuristic way from knowledge
of the system to be controlled.
All of the actuators available for the design challenge are used with ex-
ception of the differential canards. This is because omitting any of them will
necessarily compromise performance in terms of achievable forces and moments.
The reasons for not using the differential canards are given in Section 29.2.2.
There are four primary feedback loops to design, three rotational ones and
airspeed. Multivariable control allows the designer to design all four simulta-
neously. However, the longitudinal motion - pitch and airspeed - is essentially
decoupled from the lateral loops - yaw and roll. What coupling there is be-
tween lateral and longitudinal motion is due to kinematic cross-coupling and/or
asymmetric aerodynamic forces due to, for example, different flow regimes over
each of the wings. The linearisations provided for wings level, steady flight do
not capture these effects. Benefits from designing on the complete 4-input 4-
output system are likely to be more prominent in a Linear Parameter Varying
(LPV) framework [259] within which these coupling terms could be modelled.
For example, parametric dependence on roll rate could be modelled and hence
designed for. However, given the time constraints, an LPV solution was not
investigated and hence the decision was taken to separate the longitudinal and
lateral control law designs.
Figure 29.1 shows the top level SIMULINK specification of the controller.
The two 7/0o loop-shaping controllers are contained within the lateral and lon-
gitudinal H - i n f i n i t y c o n t r o l l e r blocks. They are implemented in discrete
time observer form and hence have two sets of inputs, the measurements and
the achieved aircraft inputs. The pre-compensator weights, l o n g i t u d i n a l W1
and l a t e r a l W1, contain all the integrators, phase advance terms and roll-off
terms designed in the same way as for a classical control law. They are im-
plemented in a modified Hanus self-conditioned form (see Chapter 7 in [120]
and [110]). This is exactly the structure used for the Harrier control law de-
veloped in [120]. Note the two scaling blocks in the feedback paths. These are
used to trade-off the relative amounts of coupling which are to be tolerated
between outputs, e.g. scaling speed in knots and pitch rate in degrees implies
that a 1 knot variation in airspeed is as equally undesirable as l ° / s coupling
in pitch rate. The o u t p u t s block implements first order high frequency roll-off
filters on the p, q and r measurements. The cut-off frequency is 50 rad/s.

29.2.1 Longitudinal Controller

The primary feedback variables used to design the longitudinal controller are
pitch rate q and airspeed V. This choice is straightforward in that these are
the quantities the pilot wishes to control. Use of pitch attitude 8 to stabilise

466
-

Figure 29.1: The control law architecture

the aircraft would require phase advance, which in effect differentiates the mea-
surement over some frequency range. This would produce a noisier signal than
the measured q and a less robust design. Furthermore, the dynamics of the
pitch attitude sensor are slower than those of the q sensor and pitch attitude
can not be used at large roll angles.
The HIRM has tailerons and canards available for longitudinal control.
There are several different strategies which could be used to determine how
to apportion a required pitching moment between the surfaces. Before select-
ing a strategy a number of considerations must be taken into account:

• The canards have much faster dynamics, and hence can be used to higher
frequencies with less phase lag.

• The tailerons can generate much larger pitching moments.

• The tailerons generate a non-minimum phase flight path response, whereas


the canards produce a response in the commanded direction immediately.

• Producing countering pitching moments simultaneously from tailerons


and canards is inefficient.

• The architecture may have structural loading implications.

Consideration must be made of what happens when the surfaces rate or


authority limit. Rate limiting of the surfaces can lead to pilot-induced
oscillations.

467
Two possible schemes are:

t Frequency blending of the two actuators. Using a complementary filter,


the higher frequency component of the demand is fed to the canards, and
the lower frequency component to the tailerons. In this way the canards
are used to obtain a fast initial response. The canard pitching moment
is then transferred to the tailerons which take the low frequency part of
the demand. In this way the trim is taken by the tailerons which have
the moment generating power. This is energy efficient in that the trim is
taken only on one surface and hence the surfaces do not oppose each other.
Using the canards at higher frequencies may also allow more bandwidth
to be extracted from the system. Rate limiting can be addressed by cross-
feeding the demand not generated by the surface that has rate limited to
the other surface.

• Driving both surfaces in tandem. The inputs can be scaled such that the
demand is a percentage of total travel so that both surfaces saturate at the
same point. One of the motivations for this approach is that rate limiting
is less likely to occur. If both surfaces effect the demanded pitching
moment they both have less far to travel than if one surface had been
used. A disadvantage is that the extra agility of the canards is not being
exploited. Additionally, small high frequency disturbances drive the much
heavier tailerons which may be less energy efficient in terms of required
hydraulic power. A second advantage of this scheme is that a failure of
one of the surfaces still gives a system which generates pitching moments
across the required frequency range, albeit with reduced authority.

Either of the two above strategies could be employed. The first scheme was
chosen for this design. The complementary filter is of the form

s F~ilero~(s) = ~I
F c ~ , ~ r d ( S ) -- s + w I ' s + wf"

These transfer functions are implemented in discrete time. The canard demand
is also normalised with the gain Nc. This gain is such that the gain per unit
demand to pitch rate at open-loop cross-over frequency is the same for both
canard and taileron. Therefore if one surface saturates or rate limits, its de-
mand can be fed directly across to the other surface. All of the limiting and
cross-feeding occurs in the a c t u a t o r demands block, Figure 29.1.
A pitch attitude hold is implemented in the o u t p u t s block in Figure 29.1.
Although not listed as a specification of the control law in Section 27.3.2, some
kind of hold is required in practice. This enables the pilot to go stick-free if he
so wishes. Figure 29.2 shows the p i t c h a t t i t u d e h o l d block, the output of
which is a pitch rate demand. In effect the variable out_l becomes q + 0 . 2 5 8 ~ ,
where 8~,.~ = 8cu~r - 8pr~v. The robustness properties of the closed-loop are not
altered significantly by feeding back q + 0.258. This is because the 8 portion of
the signal does not modify the loop gain at cross-over too much. The attitude
hold is only engaged when the flag input, h o l d f l a g in Figure 29.2, is set to

468
out_l

Figure 29.2: The p i t c h a t t i t u d e hold block

1 and the roll angle ¢ is smaller than 0.1 rad. A pitch attitude hold is not
desirable at large bank angles.
The construction of the total pitch rate demand, command f i l t e r block in
Figure 29.1, consists of the following terms:

• Longitudinal stick demand. The stick commands normal acceleration in


g. This is converted into an equivalent pitch rate demand. The demand
is also fed through a pre-compensator block which is used to meet the
pitch drop-back requirement, see Figure 27.15.

• Normal acceleration limiting term. If the normal acceleration limits are


exceeded, the amount by which the limit is exceeded is converted into the
required pitch rate demand to remove the excess.

• Incidence limiting term. The excess incidence is turned into a pitch rate
demand that will restore incidence to within the specified limits.

The a t t i t u d e - h o l d engaged f l a g block is shown in Figure 29.3. This


block is contained within the command f i l t e r block mentioned above. The
output of this block engage f l a g , is the input h o l d f l a g of the block depicted
in Figure 29.2. The attitude hold is only engaged if the pitch rate is less than
l°/s, and the pitch rate demand is less than 0.1°/s (effectively zero). Once
engaged, it stays engaged whatever the pitch rate, until the pilot commands a
non-zero pitch rate.

29.2.2 Lateral Controller


The primary feedback variables used to design the lateral controller are roll
rate p and yaw rate r. Roll angle and heading angle were not used as primary
feedback variables for the same reason that pitch attitude was not used when
designing the longitudinal controller (see Section 29.2.1). In addition, note that
p and r are valid for any orientation in space, which is not the case for Euler
angles. A velocity vector roll rate and side-slip demand system is specified
in Section 27.3.2. The designer might therefore be tempted to use side-slip
as a primary feedback variable. This will almost certainly produce poorer

469
zero

q lch rate delay engage


/lag

0.017

Figure 29.3: engage f l a g logic

later~ stick degreesto radians


"I ,5]
demand p demand

[]
side-slip controller and
demand

alpha

COS

Figure 29.4: Lateral demands

performance in that the side-slip measurement is both more noisy and a slower
measurement than r. These two effects will mean that less bandwidth would
be extracted from the yaw loop. Designing a tight primary feedback controller
using p and r robustly stabilises the aircraft and provides an inner closed-loop
system around which an outer loop side-slip tracking system can be built. The
side-slip controller can be seen in the lower left-hand corner of Figure 29.1.
Its output is a yaw rate demand which enters the inner loop as illustrated in
Figure 29.4.
A yaw rate demand is also cross-fed from lateral stick to effect a velocity
vector roll - the required term is sin a times the lateral stick demand as is
illustrated in Figure 29.4. Omitting this cross-term would leave it to the side-
slip controller to reject the side-slip induced when a roll rate is commanded. As
the side-slip outer loop has a lower bandwidth than the inner loop, excessive
side-slip coupling will occur. The cross-term puts in a fast yaw rate demand
to achieve the velocity vector roll.
T h e HIRM has differential canards and tailerons available for roll control.
However, the canards are very ineffective in roll as they generate smaller forces,
and are located much closer to the centreline of the aircraft. Hence using
canards requires large surface deflections and gives little benefit. Furthermore,
it limits their availability for longitudinal control for which they have definite

470
benefits over the tailerons. By studying the aerodynamics of the HIRM it can
be deduced that the differential canards also have a very significant influence
on the effectiveness of the symmetrical canards and tailerons thus creating a
robustness issue. Hence, only the differential tailerons are used for roll control.
For yaw control, only the rudder is available.

29.3 Controller Design


This chapter discusses how the weighting functions and controller parameters
were selected to meet the design requirements.

29.3.1 Lateral Primary Feedback Controller


In designing the feedback controller the main objective is to push the cross-over
frequency as high as possible, whilst retaining an acceptable level of robustness.
No reference is made to the handling qualities required. This is because if the
feedback controller is as fast as possible given robustness constraints, design of a
pre-filter to meet handling qualities should be straightforward. A fast feedback
controller reduces closed-loop response lag, and gives a robust response i.e.
one which should not vary significantly with envelope operating point. If this
approach gives better handling qualities than specified, the high bandwidth is
still justified in terms of the achieved disturbance rejection.
The argument against incorporating handling quality requirements when
designing the feedback controller is that necessarily a trade-off between robust
stability and handling qualities will be effected. One could argue that the
primary purpose of the feedback controller is to maximise robustness for the
specified open-loop cross-over frequencies, i.e. disturbance rejection properties.
The design procedure can be divided into the following steps.

1. Scale the differential taileron and rudder inputs by 1/½PV 2. This nor-
malises the moment generated so as not to vary significantly with flight
condition.
2. Select the lateral states from the linearisation i.e. v, p, r and ¢. Append
this linearisation with the actuator models, full order sensor models, anti-
aliasing filters and computational delay. The full order sensor models are
used since the final controller is model reduced anyway.
3. Scale the outputs to reflect the coupling requirements. The scaling used
is (#-tools and MATLAB commands are used)

>> out_scaling = d i a g ( [ 0 . 3 1]);

A couple of design iterations were carried out before arriving at this


scaling. The scaling reflects the inherent coupling within the plant i.e.
that a unit coupling in roll happens a lot more easily than a unit coupling
it yaw. Attempting to scale the outputs to make the coupling into yaw

471
and roll approximately equal results in poorer robustness. In general,
trying to change the directionality of the plant is not good practice. The
roll rate p is augmented with the roll angle to give the output variable
p +A¢¢. This boosts the low frequency gain, and enables a roll angle
hold to be effected for zero lateral stick demand. The A~ term is removed
during a roll-rate demand. This can be justified (in terms of robust
stability) provided that A~ is chosen such that the open-loop cross-over
is entirely set by the p part of the constructed output variable.

. The desired cross-over frequency for both loops is 10 rad/s. This is the
highest the cross-over frequency can go before robustness margins are
necessarily reduced, due primarily to actuator roll-off. To verify this,
a few design iterations at slightly higher cross-over frequencies can be
carried out, and the resulting achieved robustness margin e, monitored.
Both loops have suitable roll-off rates at cross-over, and so all that is
required is to boost low frequency gain, and add high frequency roll-off
filters. The selection of the appropriate transfer functions is exactly as for
a classical design. Both loops are rolled off with the filter s-W~"
50 The filters
are discretised using a bilinear transformation with frequency warping to
match the filters at 10 rad/s.
Low frequency gain is boosted with the pre-compensator

>> W_p = nd2sys([l I],[1 0]);


>> W_r = nd2sys([l 2],[i 0]);
>> W_I = daug(W_p,W_r);

Figure 29.5: Singular values of the weighted plant

5. Multiply the two plant inputs with gains kw4 and kw6, an input scaling
which gives the required cross-over frequencies. Plot the singular values
of the shaped 2-input 2-output plant, Figure 29.5. Check that the desired
loop shapes have been achieved.
6. Design the 7/oo loop-shaping controller, Ko~, for the shaped plant. Check
that the resulting robustness margin is sufficient (typically e > 0.3 indi-

472
cates a robust design). Check the step responses. These are shown in
Figures 29.6 and 29.7.
Slop ~ p Slap ~ ¢

o., ..... : ........ i ......... i ....... : . . . . . . . . . . : ........ ~ ....... i ....... i ........

0.6 .............. i ....... : ....... ! ................. ".......... i ........ i ........ ~. . . . . . .

0.4 ." :. ' i. " i : • • ::

o.,?. .............. i ......... ~ ...................... i ....... i ......... : '~ ........

-0"20 0,2 04 06 08 ijml( ) 1"2 ~4 is ~8 2 0 0.2 04 O0 OJ ~ml() 1"2 14 ~O 15 2

Figure 29.6: Step on p demand Figure 29.7: Step on r demand

If gain scheduling is to be carried out, then the above procedure is applied


to each of the design points. Typically, the dynamic weighting functions will
not need to be changed between design points, and so all that changes are kw4,
kw6 and the controller Koo. The controller is then implemented in observer
form, and the state-space matrices gain scheduled using linear interpolation.
Details on how to do this can be found in [120].
For the HIRM design challenge, the flight envelope is not that wide and
hence a single fixed gain controller can be used. The fixed gain controller
used was the one designed for the operating point ini3005. This linearisation
corresponds to 12° incidence, and is somewhere in the middle of the flight
envelope. The corresponding values of kw4 and kw6 for this flight condition
were then used for all flight cases. The achieved e was 0.32. Should the level of
robustness have proven insufficient, then the loop-shaping exercise above could
have been re-run with less stringent performance requirements.

29.3.2 Side-slip Controller


An outer loop side-slip controller was designed using loop-shaping and nor-
malised coprime factor robust stabitisation. By examining the Bode plot of
yaw rate demand to sideslip, the following compensator was determined

4(s + 2)
wl=
s(0.1s + 1)

This has a first order roll-off at 10 rad/s to attenuate side-slip sensor noise, and
additional low frequency gain below 2.0 rad/s. The Bode plot of the shaped
loop is shown in Figure 29.8.
The resulting 7-/o~ controller was model reduced to 4 states by carrying out a
least squares matching of the gain and phase plots of the full order controller.
This can be done using the MATLAB function invfreqz.m. The achieved
robustness margin is e = 0.42. This guarantees a gain margin of at least 2.4,

473
o.e

a.a

2 a 4
~. (s)

Figure 29.8: Shaped loop Figure 29.9: Step on/3 response

and a phase margin of at least 45 ° (see Section 7.3 for the relevant formulae).
Figure 29.9 shows the side-slip step response which meets the specification set
out in Figure 27.16.

29.3.3 Longitudinal Primary Feedback Controller

The longitudinal controller initially had a multivariable structure with feed-


back variables q and speed. However, the difference in the required cross-overs
for these two loops (approximately a factor of 10) meant that there was little
or no advantage to having a multivariable structure. Hence, two SISO con-
trollers were designed given that the added complexity of the cross-terms in
the MIMO case could not be justified in terms of robustness or performance.
The q feedback was designed using the LSDP, and the velocity feedback using
classical loop-shaping. The closed-loop time response shows a good decoupled
response with no overshoot. Full details can be found in [191].

29.3.4 Pilot q-command Filter Design

The closed-loop step response gives little or no overshoot. In order to achieve


the drop-back requirement, a pilot command filter of the following form was
used:
Tis+l
ppilot(S) - T28 + 1

By putting Ti > T2 pitch rate overshoot is introduced. The values of 7'1 and
T2 were determined by performing a search over a range of values of Ti and T2,
and then selecting all pairs which give a drop-back of approximately 0.125 s.
The other handling quality criteria were then examined for these pairs, and a
particular solution selected. The values chosen were 7'1 = 0.336 and T2 = 0.144
which gave a drop-back of 0.1 s, an average phase rate of 57°/Hz and ]c = 1.7
Hz. From Figure 27.12, it can be seen that this gives Level 1 performance.
Figure 29.10 shows the time response of a pitch rate demand.

474
I.S

i
1.4 . . . . . . . . . . . . . . ~. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

O.8 ............................ : ................... ~ ........................

o.* ............................. i ........................ i . . . . . . . . . . . . . . . . . . . . . . . . . . . .

o.4 ....................... ! ........................... ~. ..........................

a 2 ......................... ~ ..................... .......................

0 0.5 1.5

Figure 29.10: Step on q demand

29.4 R e s u l t s from the E v a l u a t i o n P r o c e d u r e


This chapter contains results from the evaluation procedure. These results are
from the first application of the procedure, and indicate that very little modifi-
cation is required to the controller in order to meet all of the requirements. The
authors believe that this is in part due to one of the key strengths of the design
method in that loop-shaping gives good visibility as to what performance can
be achieved from the system. Hence, the design is likely to exploit the sys-
tem's inherent performance limits fully. At the same time, robustness must be
maintained so as to respect the Nichols exclusion regions. This is equivalent to
achieving a large enough value of e.
It should also be borne in mind that the controller designed for ±n±3005 is
used at all operating points in the analysis, whereas in practice it would be gain
scheduled with airspeed. Note that the control law tested here is non-linear in
that it contains models of actuator rate and authority limits, and is in discrete
time. Hence it is very realistic and implementable.
All of the Nichols plot tests were satisfied, with the exception of the one
shown in Figure 29.11 where the exclusion region is just clipped. A small
increase in the phase advance in the differential taileron weight is required.
The results indicate a suitably robust design across the envelope despite not
gain scheduling the controller.
Figure 29.12 shows the Gibson Nichols plot criterion for small perturbations
of the stick. It can be seen that the response is Level 1. The Nichols plot for
large stick demands has not been produced yet. The drop-back and phase rate
criteria are given in Table 29.1. All meet the Level 1 requirements. Table 29.2
gives data on the response to turbulence.
Figures 29.13, 29.14, 29.15 and 29.16 show responses resulting from com-
mands in pitch, roll, side-slip and speed. Different flight cases have been used
to demonstrate that performance is good across the flight envelope. In Figure
29.13 a 5°/s pitch rate command is made. Note the pitch rate overshoot used
to give the required drop-back. At 4 s the incidence limit is hit, and the alpha
limiter forces the pitch rate down, so as to maintain alpha on the limit. Note
also that the coupling into side-slip at high alpha is well controlled. In Figure

475
Flight condition Drop-back
Mach 0.20,1000 ft -0.17 s 1.32 -56.0°/Hz
Mach 0.24,20000 ft -0.23 s 1.40 -56.0°/Hz
Mach 0.30,5000 ft -0.10 s 1.58 -56.6°/Hz
Mach 0.40,10000 ft -0.07 s 1.40 -57.0°/Hz
Mach 0.50,15000 ft -0.12 s 1.35 -57.1°/Hz

Table 29.1: Drop-back and phase rate criterion

RMS roll rate due to turbulence 1.7272 °/s


RMS pitch rate due to turbulence 0.2470 °/s
RMS yaw rate due to turbulence 2.1431 °/s
RMS normal acceleration due to turbulence 0.1317 g

Table 29.2: Response to turbulence

29.14 a maximum roll-rate command is made. Performance is reasonable in


that this is a severe manoeuvre which results in the aircraft banking over well
past 90 °. Coupling to side-slip is small (about 1°), and coupling to q small in
comparison to the demand on p.
In Figure 29.15 a 10° step on side-slip is commanded. The response time
is well within the requirements. Decoupling from other controlled outputs is
also good. In Figure 29.16 a rapid speed demand change is made. Note how
full use of the engine is made, and that the control law respects the engine
rate limits. However there is some speed overshoot which should be reduced,
possibly by reducing the cross-over frequency of the speed loop, or by limiting
the dynamics of the speed demand.
Overall, reasonable performance was achieved from the design. As with any
design approach, a degree of refinement will be possible with some iteration.
However our design came close to meeting all requirements on the first attempt,
and this was with very limited time available to the authors.

29.5 Conclusions
A stability augmentation system has been developed for the HIRM using 7/~
loop-shaping. In the authors' view, the design problem was relatively simple,
particularly given the narrow flight envelope of the HIRM. However, the ap-
proach which has been taken will extend in an easy manner to a much wider
flight envelope. Key to this is the observer structure used to implement the
7/~ controller. The control law structure has much in common with a classical
control law, the main difference being the addition of the lateral and longitu-
dinal 7-/~ controllers to the feedback path. The observer structure is also used
to handle actuator authority and rate limiting in a systematic fashion. Ap-

476
Nominal flighL Nichols plot, Mach 0.24, 20000 ft
20
• ..:" . :".. Y.i /,,....' ..-":. '~I
: '"" ::
: ........ " i i '.-: ."
i i ':': '. i ... ...... :
15 _i .:':-..... i..i.:. .." '-. !i ~ .- ',. :-,.i..i.. .-::-. :: -dts
-" "-." .~ ! :"o. '. ii o ~..':-ii .... ..-' '.
• " .'" : : " ," ". ". ,'~ ."l " : : I .". "
- - dtd
10 • "
•: "
.'" ". " :i::
i:~ :" ". " . , .'.' : : ' ~::i ' ? ' ..", - ' ~ .'" II ":"-.!.!.i
:i'll ' ." .'" "'. '-. ":- o dcs


--.:...,.:~
: ' "". "
...":-. ~::
". '.~.: " " "" • ~'.
.:.':.. ..",::i:.- ..~.~i i// :".-..
"...'" ~..'"..'" ." "I" ?' '• ." ". -"
:....!"
'. • x dcd

;i : ...-........-
o.. :.:i:..i - I ..'%~ ,:: . "... '.. - ! + dr
!: : ." ..' ". '-. ':-'.'. : " :t -:'Y ." ."'-. ".. ': :
~.": :'...,..
" ." ....' " o... "..:.."~:.-"-:"i
.::.'~,....... ":.~'"".'....."
..~ ~- .. .... .....,..':
: ".!. * thr
" - "": .'":" ".".:'. : ""..'" ].''i'ii'".'"~.~iii: ;:7 .:',. :~'" : "

~-5
:! . ~ = 5 _ + o ~. "+- ; : . - . ~ ' ~,~. . + . . - . . : . . , £ - ¢,,..- .J...'~ .: : : ::
- 1 0 i~""":. .........:: o : : : ....~.... - .~ '.. -~..-...-"-".-"?~

-15
:: " : : O : : : ::
:i ! : : o: : " :~ : : : : ~ : :~
-2(
-350 -300 -250 -200 - 150 - 1O0 -50
Open-Loop Phase (deg)

Figure 29.11: Nichols plot of the nominal plant at Mach 0.24, 20000 ft

G i b s o n criterion, pitch: l i n e a r a n a l y s i s
20

16

10

~ 5

-5

-10
o ii
-180 -160 -140 -120 -100 -60 -60 -40
P h a s e (deg)

Figure 29.12: Gibson Level 1 boundaries

477
pw az
0 . 5

_1 I
0 5 10 0 5 10 0 5 10
time (s) time (s)
Va alpha beta
0.4
30 . . . . . . . . . . . . ..........

~25 Ac~-0.0"2
02~............... i ...........

0 5 10 0 10 0 5 10
time (s) time (s) time (s)
q cmd (:its dtd
4
11' i '
2 ............. i ...........

_::[ ..... ~ ...... : ....... [


0 ~
0 5 10 0 5 10 0 10
time (s) time (s) t i m e (s)

dc~ dcd dr
1

10 . . . . . . . . . . . . . . i. . . . . . . . . . . . . . . o.: , 1

-0' I I ....... .....

0 5 10 0 5 10 0 5 10
time (s) lime (s) time (s)

oof
50 ........
thrl the2

0
0 10 0 5 10
time (s) lime (a)

Assessment manoeuvre: pitch rate demand, Mach 0.2, 1000 ff

Figure 29.13: 5°/s pitch rate demand

478
........... ...

~_2~t ~.-4
0 2 4 8 0 2 4 6 0 2 4
time (s) time (e) ,me (e)

alpha beta

"[ /
Va
2 - •

1 .......... ! ....... i ........


tO . . . . . . . . . . :. . . . . . . . . . : . . . . . . . . . .
g'~ .......... i.......... i" $
: i

100r
0 2 4 0 4 6 0 2 4
time (s) Ume(e) time (s)

p cmd dtd
o : i
2o .......... i.......... i . . . . . . .
~ ~-5i i..........
~_ i ...............

0
. . .0. . . . . . .0. . . . . . .~. .

2 4 0 2 4
=[
0
.....
2 4
...... 1
6
time (s) time (s) time (s)

dcs dcd dr

o.s ......... i .......... :: ........


0 ......... ;....

0 2 4 6 0 2 0 2 4 6
time (s) time (s) time (s)

thrl thr2
6

E2o~!'
... t
4 ........ ;.......... : ........
g :

:: ......... :
. . . . .........
. . . . .

0 4 6 0 2 4
time (s) time (s)

Assessment manoeuvre: roll rate demand, Mach 0.3, 5000 ft

Figure 29.14: 70°/s roll rate d e m a n d

479
pw q az
1 0.05~ : ;
o.5 ............ i ................

-0.06
-9.6L i I
1 /
0 10
-0.10 5 10 0 10
time (s) time(s) time (s)
Va alpha beta
161,4~
10
61.2 7"1f i //
............. : . . . . . . . . . . .

161

160.8 69
0 10 0 10 0 10
time (s) UmeIs) time (s)
betcmd dts did
10
-6.6 ............... ::. . . . . . . . . . . .

~ 5 ...........................

-6.7

0~ ~ _ 1 0 ~ ........
0 5 10 0 5 10 0 5 10
time (s) time (s) time (s)
dcs dcd dr
1

or 1
0.1
tO . . . . . . . . . . . . . . . . . . .
0.05
ca
. . . . . . . . . . . . .

0
-0.5
-I [ . . . . . . . . . . . . .
-0.05 -5 ' ' r
0 5 10 0 5 10 0 5 10
time (s) time (s) time (s)
thrl thr2
14
12 . . . . . . . . . . . . . . : - - 12 . . . . . . . . . . . . . . . . . . . . . .

g, gt

0 10
61 0
. . . . . . . . . . .

5
. . . . . . . . . . . . . .

10
time (s) time (s)

Assessment manoeuvre: sideslip demand, Mach 0.5, 15000 ft

Figure 29.15:10 ° side-slip demand

480
10 x 10 -'0 pw 0.5 - q * -9 . az. .

..... ,o

0 10 20 30 40 0 10 20 30 40 0 10 20 30 40
time (s) time (s) time (s)

Va alpha x 10"~ beta

,oo,.. ri '~' 8[ ...... ~ ~ ...... )


0 10 20 30 40 0 10 20 30 40 0 10 20 30 40
time (s) time (s) time (s)

Va cmd dts x 10 "'° dtd

~'°I ..... ! ....... ! ....... i ....... - ...... i ....... i........ o ..

0 10 20 30 40 0 10 20 30 40 0 '~0 20 30 40
time (s) time (s) time (s)

dcs dcd x 10 -'0 dr

o.I ...... :,....... :, ....... ::........ ........................

-o.s ...... !....... ':...... :....... _:


-2
0 10 20 30 40 0 10 20 30 40 0 10 20 30 40
time (s) time {s) time (s)

~rl ~r2
100 - 100 - *

so ...... .............. : ........ ~ so ...... .............. i .......

v 0 0

0 10 20 30 40 0 10 20 30 40
time (s) time (s)

Assessment m a n o e u v r e : air s p e e d d e m a n d , M a c h 0.3, 5 0 0 0 ft

F i g u r e 29.16: A 50 m / s s p e e d c h a n g e

481
propriate handling of these non-linearities is essential for achieving the desired
handling qualities, and ensuring sensible apportioning of control power.
7-/oQ loop-shaping is a fairly intuitive controller design method that can be
picked up in a relatively short amount of time, particularly from someone with
a flight control background. The design strategy employed bypassed the step
of trying to express in detail, the time response requirements in the frequency
domain (a drawback of frequency domain based methods).
The order of the control law is not seen as an issue as regards implementa-
tion provided that the complexity can be justified in terms of desired robustness
and performance. Experimentation with model reduction showed that the lat-
eral 7Qo controller could be reduced to 10 states and the longitudinal to 8
states, both with minimal change to the robustness and performance. Bal-
anced truncation of a coprime factorisation of the controller was used for this.
However, some care would be required when model reducing a gain scheduled
design, as the physical interpretation of the model reduced states must be the
same for all designs to allow gain scheduling.
The type of uncertainty (coprime factor uncertainty) adopted, although
quite general, did not prove to be too conservative. Scheduling was avoided
partly because the open-loop HIRM was scaled with dynamic pressure, a well
known technique within industry.
The paradigm of 7-/00 loop-shaping is extremely powerful. There are other
extensions which have not been demonstrated here due to lack of time available
for the project. One exciting new area is that of self-scheduled design meth-
ods whereby a parameter dependent controller is synthesised in one step for a
parameter dependent plant. For the HIRM, the objective would be to find a
controller dependent on airspeed given a speed dependent model of the HIRM.
The synthesis of the controller relies on solving a set of linear matrix inequal-
ities for which there are numerous algorithms available. In the authors' view,
self-scheduled methods and linear parameter varying (LPV) plant descriptions
will provide very powerful and relevant tools for the aerospace industry. In
particular, they provide a framework in which to address aerodynamic non-
linearities and rate dependent effects. ~ r t h e r investigation of the best control
structure is also a possibility. Designing a single controller for roll, pitch and
yaw would be worth carrying out to see what the cross-terms between lateral
and longitudinal feedback loops are as a function of flight condition. This might
include looking at non-steady flight conditions such as a non-zero roll rate.
For new types of controllers to be used by industry, there are two major
pre-requisites. Firstly benefits need to be quantified so as to justify a change in
approach. Secondly, once the benefits have been established, the flight clear-
ance aspects need to be addressed. The HIRM design challenge was not set up
to address these issues directly. Its focus was to demonstrate the methods to
industry, and to highlight what the advantages might be. As regards potential
benefits of the method, there are two main ones worth mentioning. Firstly, for
complex multi-input multi-output systems, the ~oo loop-shaping method pro-
vides a way of synthesising a controller which reflects the cross-coupling in the
plant. As such, better performance and robustness may be obtained since se-

482
lecting controller diagonal terms with non-multivariable methods can be a trial
and error task. Secondly, the 7-/oo loop-shaping approach has great potential
to reduce design time. This is partly a result of its ability to handle com-
plex multivariable systems, and partly that given the nature of the robustness
optimisation, it is hard to design a bad controller.
Flight clearance of this type of control law may be easier than for many
other non-classical design methods. In the main, this is because the loop-
shaping aspect has direct connections to classical design. The main difference
to a classical design is the addition of the multivariable 7/00 feedback controller
element. However, the 7-/oo feedback element is implemented in an observer
form for which the structure is clear, and for which interpretation of the physical
units of individual gains are clear. However, this is still (in the UK and probably
many other countries) a new approach to have state-space elements within
the control law as opposed to SISO transfer functions. However, this will
probably not be an insurmountable problem, particularly given the numerical
and efficiency advantages of using state-space implementations. The next step
therefore has to be to quantify benefits in some meaningful manner.

483

S-ar putea să vă placă și