Pole-placement Predictive Functional Control for over ...

27
Pole-placement Predictive Functional Control for over-damped systems with real poles J.A. Rossiter, R. Haber, K. Zabet b a J.A. Rossiter is at Dept. of Automatic Control and Systems Eng., University of Sheffield, S1 3JD, UK. email: j.a.rossiter@sheffield.ac.uk b R. Haber and K. Zabet are with Univ. of Applied Science Cologne, Dept. of Plant and Process Engineering, D-50679 Kln, Betzdorfer Str. 2, Germany. email: [email protected] Abstract This paper gives new insight and design proposals for Predictive Functional Control. Specifically it shows that contrary to the common practice and re- quirement of selecting a coincidence horizon greater than one for high order systems, by using parallel first order models to form an independent pre- diction model, it is possible to use a coincidence horizon of one and obtain precisely the desired closed-loop dynamics. This greatly simplifies coding, tuning, constraint handling and implementation. The paper derives the key results for high order and non-minimum phase processes and also demon- strates the flexibility and potential industrial utility of the proposal. Keywords: PFC; tuning; uncertainty. Preprint submitted to ISA Transactions April 4, 2016

Transcript of Pole-placement Predictive Functional Control for over ...

Page 1: Pole-placement Predictive Functional Control for over ...

Pole-placement Predictive Functional Control for

over-damped systems with real poles

J.A. Rossiter, R. Haber, K. Zabetb

aJ.A. Rossiter is at Dept. of Automatic Control and Systems Eng., University ofSheffield, S1 3JD, UK. email: [email protected]

bR. Haber and K. Zabet are with Univ. of Applied Science Cologne, Dept. of Plant andProcess Engineering, D-50679 Kln, Betzdorfer Str. 2, Germany. email:

[email protected]

Abstract

This paper gives new insight and design proposals for Predictive FunctionalControl. Specifically it shows that contrary to the common practice and re-quirement of selecting a coincidence horizon greater than one for high ordersystems, by using parallel first order models to form an independent pre-diction model, it is possible to use a coincidence horizon of one and obtainprecisely the desired closed-loop dynamics. This greatly simplifies coding,tuning, constraint handling and implementation. The paper derives the keyresults for high order and non-minimum phase processes and also demon-strates the flexibility and potential industrial utility of the proposal.

Keywords: PFC; tuning; uncertainty.

Preprint submitted to ISA Transactions April 4, 2016

Page 2: Pole-placement Predictive Functional Control for over ...

1. Introduction

Predictive functional Control (PFC) has been very successful in industry(e.g. Changenet et al. (2008); Fallahsohi et al (2010)) and yet surprisinglyreceived very little interest in the academic literature Richalet et al. (1978,2009); Haber et al (2011). A likely reason for this is that researchers inpredictive control have focussed on proofs of issues such as guarantees ofstability Kouvaritakis et al. (1992); Scokaert and Rawlings (1998) and feasi-bility, robust stability Kerrigan (2000), parametric methods Kvasnica (2009)and more recently robust feasibility in the presence of bounded disturbancesMayne (2005). It should be noted that PFC techniques are much simplerto code and implement then conventional predictive control methods Ma-ciejowski (2001); Rossiter (2003); Haber et al (2011) and thus PFC is bestviewed as an alternative to PID or other low level control law where the costper control law is necessarily small but nevertheless one may desire attributessuch as systematic constraint handling. As a consequence, comparisons withconventional predictive control are not appropriate as those, by definition,can give better performance, but of course at much higher cost.

A main selling point of PFC is that the design is done by choosing a targetbehaviour (equivalently closed-loop pole/time constant). If there is a stronglink between the user choice and the behaviour that results, this is an intuitiveand easy design technique, as compared to say PID. Moreover, PFC has acritical advantage over PID approaches, that is, constraint handling can beembedded systematically and with minimum coding/computation. However,the literature has given little attention to theoretical a priori guarantees ofstability, feasibility or robustness for PFC; of course some results do existRichalet et al. (2004) and industrial users always do practical assessments.This paper seeks to redress the balance slightly by demonstrating some usefulnew theoretical results for PFC which also extend the efficacy of tuning ofthe approach.

1.1. Background on PFC

A conventional PFC has two tuning parameters: the position of the coin-cidence point in the future (the coincidence horizon) and the desired settlingtime t95%, also called TRBF. It is implicitly assumed that the closed-loopresponse will become approximately first-order with the target settling time,although in fact this can only be assured if the coincidence horizon is one.

2

Page 3: Pole-placement Predictive Functional Control for over ...

For higher-order aperiodic processes Richalet et al. (2009) it is commonto recommend the coincidence point to be near to the inflection point of theprocess step response, which is the place of maximum value of the impulseresponse. The idea is based on the fact that at this point, the manipulatedvariable change has maximum effect. However, in such a case the actualsettling time will often not match the desired settling time t95 so tuningbecomes more challenging as the direct link between designer choice andeffect is lost. One reason is that in case of higher-order processes, the closed-loop response will not approximate a first-order response closely and hencethe relation to the settling time is not straightforward Rossiter et al. (2014).

1.2. Paper contributions

This paper will propose an alternative way to tune PFC for systems withreal poles. It will be shown that within the proposed PFC design, the closed-loop poles can be selected as free parameters and this is a major advance onconventional PFC where the closed-loop poles are selected through trial anderror, but without a guarantee of what is achievable. It is interesting to notethat the proposed method, in the unconstrained case, has some equivalencewith pole placement design, but a critical point is that it is still based onprediction and allows systematic constraint handling which is not the casefor pole-placement designs! It is also noteworthy that the proposed PFCmethod works with a coincidence horizon of just one, which is contrary tothe conventional advice and also enables a significant reduction in computingcomplexity.

Section 2 gives some background on PFC concepts, a conventional lawand demonstrates the tuning challenges. Section 3 introduces the proposedpole placement PFC approach for second-order systems along with someanalysis of the properties. Section 4 then generalises the approach to higherorder models and emphasises the additional degrees of freedom which enablemore flexible tuning. The paper finishes with numerous examples whichdemonstrate the attributes of the proposed algorithm and how these comparewith conventional PFC.

2. Background information on PFC

2.1. Target behaviour

PFC is based on the assumption that it is realistic to achieve closed-loopbehaviour close to a 1st order system with a delay τ (or h samples), time

3

Page 4: Pole-placement Predictive Functional Control for over ...

0 1 2 3 4 5 6 7 80

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Match prediction and target

Seconds

ry

Figure 1: Target step response r with pole of 0.8 and illustration of coincidence withoutput prediction y at a coincidence horizon of 5.

constant Tr and unit gain, for example:

r∗(s) =e−sτ

Trs+ 1; r∗(z) =

z−h(1− λ)

1− λz−1(1)

where r∗(s) is a Laplace representation and r∗(z) the z-transform represen-tation. A typical desired step response, with no delay, is plotted in figure 1where the pole is set at λ = 0.8 and the sample period is T = 1. Equivalently,industrial users use the notation of target closed-loop response time (CLTR),that is about 3 time constants, where λ = e−T/Tr with T the sample periodand Tr = CLTR/3.

2.2. Coincidence point and degrees of freedom

Assuming the desired closed-loop behaviour is ’known’ (as illustrated infigure 1), then the objective is for output predictions yp(k+i|k) (the predictedvalue of yp at sample k + i with prediction made at sample k) to follow thistarget exactly. Assuming a non-zero initial condition of yp(k), a first orderresponse with a known asymptotic value r and decay rate λ can be writtendown explicitly as follows:

yp(k + i|k) = r − [r − yp(k)]λi; i > 0 (2)

4

Page 5: Pole-placement Predictive Functional Control for over ...

PFC is not able to make all future predicted output values satisfy (2) andso instead chooses a single sample instant in the future, the so called co-incidence horizon ny, and ensures that the output prediction matches thetarget response (2) at that point only (as illustrated in figure 1 for ny = 5).Consequently the PFC control law reduces to (conceptually), enforcing thesingle equality:

yp(k + ny|k) = r − [r − yp(k)]λny (3)

In order to manipulate the predictions, some degrees of freedom (d.o.f.)are needed and these are conventionally the values of the future inputs,u(k), u(k + 1), · · ·. Within PFC, the coding and computation requirementsare deliberately very simple and thus, the predicted future input is taken tobe a constant, that is:

u(k) = u(k + 1|k) = u(k + 2|k) = · · · (4)

Thus the only d.o.f. is the proposed value u(k).

Remark 1. The target behaviour (2) and control law requirement (3) can becoded by inspection. Where the system has a delay ’h’, the target behaviourshould be modified slightly to:

yp(k + ny|k) = r − [r − yp(k + h|k)]λny−h.

A potential weakness of PFC is the simplicity of control law (3). Theuser needs to be sure that matching a single point implies the rest of theresponse is also closely matched to the target behaviour. It has been shownRichalet et al. (2009); Rossiter et al. (2014) that this is the case for first-order systems only. Hence this paper proposes a modified PFC algorithm,such that the result can be extended to some higher-order systems wherecommon understanding is that the best value for ny can only be found bytrail and error (and indeed that assumes a good choice exists). Indeed it isnotable that for non-minimum phase systems it is intuitively obvious thatny must be greater than the time of the inverse response part but how muchgreater is not obvious. This paper shows how such a requirement can beavoided thus enabling more systematic tuning.

2.3. Independent prediction model and disturbance estimation

Within PFC it is conventional to use an independent model for predictionas these are known to have low sensitivity to measurement noise. This means

5

Page 6: Pole-placement Predictive Functional Control for over ...

Gp

Gm

d

yp

ym

u

-

+

Figure 2: Independent model simulation structure.

running a model Gm in parallel with the process Gp (see figure 2). Thedifference d = yp−ym between the model output ym and process output yp isused as an estimate for the system disturbance and in fact can also be usedto account for parameter uncertainty. Replacing process output predictions(which are not known explicitly) with unbiased model output predictions,the PFC control (3) law is restated as:

r − [r − yp(k)]λny = yp(k + ny|k) = ym(k + ny|k) + d(k) (5)

An alternative and equivalent representation of this which is slightly easierto code is:

[r − yp(k)](1− λny) = ym(k + ny|k)− ym(k)) (6)

2.4. PFC for 1st-order models

A first-order model Gm is presented in discrete-time as:

ym(k) =bz−1

1 + az−1u(k) = Gm(z)u(k) (7)

The predicted model output ny steps ahead assuming a constant future u(k)is:

ym(k + ny|k) = (−a)nyym(k) +Km[1− (−a)ny ]u(k); Km =b

1 + a(8)

Substituting this into control law (6) gives:

[r − yp(k)](1− λny) = (−a)nyym(k) +Km[1− (−a)ny ]u(k)− ym(k) (9)

6

Page 7: Pole-placement Predictive Functional Control for over ...

This can be re-arranged to give the control law:

u(k) =[r − yp(k)](1− λny) + [1− (−a)ny ]ym(k)

Km[1− (−a)ny ](10)

Remark 2. Earlier work Rossiter et al. (2014) showed that for first-orderprocesses a coincidence horizon of ny = 1 ensures that the closed-loop dy-namic is precisely as desired, that is, the closed-loop pole is exactly λ. The keysteps are given here for completeness. Assuming nominal case ym = yp = y,the control law is summarised as:

u(k) =r(1− λ)− (1 + a− 1 + λ)y(k)

Km(1 + a)=

r(1− λ)− (λ+ a)y(k)

b(11)

Subsituting the u(k) from (11) into the process model (7) one finds the closed-loop dynamic is given from:

y(k) =bz−1

1 + az−1

[r(1− λ)− (λ+ a)y(k)]

b(12)

or(1 + az−1 − (a+ λ)z−1)︸ ︷︷ ︸Closed−loop pole polynomial

y(k) = (1− λ)r (13)

The closed-loop pole is therefore defined as: pole = −a+ (a+ λ) = λ, hence,one achieves the desired pole exactly if ny = 1.

2.5. PFC for higher order models having real roots

For higher order systems it is recommended to use coincidence horizonslarger than one Richalet et al. (2004); Rossiter et al. (2014) due to the in-herent lag in the response. However, one cannot easily write down a simpleexpression for the ny step ahead prediction Rossiter (1993). In order to main-tain simple coding, PFC overcomes the complexity of prediction algebra byusing partial fractions to express a model as a sum of first order modelsRichalet et al. (2009); Khadir et al. (2008); Haber et al (2011) and hence:

Gm(z) =b1z

−1

1 + a1z−1︸ ︷︷ ︸G1

+ · · ·+ b2z−1

1 + awz−1︸ ︷︷ ︸Gw

+ · · · (14)

Consequently the model output can be constructed as:

ym(k) = G1(z)u(k) + · · ·+Gw(z)u(k) = y(1)m + · · ·+ y(w)m (15)

An equivalent block diagram is indicated in figure 3.

7

Page 8: Pole-placement Predictive Functional Control for over ...

Gp

G1d

yp

ym

u

-

+

Gw

+

++

Figure 3: Independent model structure with partial fraction expansion.

Theorem 1. Assuming the future input is constant, the predictions for mod-els expressed in the form of (14) are trivial and can be written down explicitly.

Proof: This is obvious, but for completeness, the predictions for a singlemodel y

(i)m (k + 1) + aiy

(i)m (k) = biu(k) are:

y(i)m (k + ny|k) = (−ai)nyy(i)m (k) +Ki[1− (−ai)

ny ]u(k); Ki =bi

1 + ai(16)

and ym(k + ny|k) = y(1)m (k + ny|k) + y

(2)m (k + ny|k) + · · · �

Corollary 1. For a coincidence horizon ny = 1, the predictions are:

ym(k + 1|k) = −a1y(1)m (k)− · · · − a(w)

w y(w)m + (

w∑i

bi)u(k) (17)

Theorem 2. A PFC algorithm for a higher-order model using a coincidencehorizon of ny and a target pole of λ can be defined explicitly without compli-cated coding.

Proof: The prediction of (16) is substituted directly into control law (6) todetermine the desired input signal.

(1− λny)(r − yp(k)) =∑

i[(−ai)nyy

(i)m (k) +Ki[1− (−ai)

ny)uk − y(i)m (k)]

(18)

8

Page 9: Pole-placement Predictive Functional Control for over ...

Rearrange to determine the input as:

uk =((1− λny)(r − yp(k)) +

∑i[1− (−ai)

ny ]y(i)m (k)∑

iKi(1− (−ai)ny)(19)

It is clear that the terms in this law are simple to compute. �

Corollary 2. For a coincidence horizon ny = 1, the PFC law reduces to:

uk =([r − yp(k)](1− λ)−

∑i(1 + ai)y

(i)m (k)∑

i bi(20)

Remark 3. Using parallel forms for prediction requires the partial fractionexpansion to be computed. However, for cases with say 2 to 3 poles, thisis trivial and this minor extra computation is dwarfed by the gains in thesimplification of the prediction algebra.

2.6. The efficacy of λ as a tuning parameter with conventional PFC

It was shown in section 2.4 that for a first-order model and a choiceof ny = 1, then the expected closed loop pole is precisely λ. However,one can equally show Rossiter et al. (2014) that for higher choices of ny,the closed-loop is not equal to λ and this alone demonstrates that PFCtuning parameters have a weak connection with the behaviour that results.Moreover, with 2nd-order models tuning and coding is not as straightforwardand some trial and error is required; indeed there is no guarantee that thereexists a suitable choice of ny such a given λ will be achieved in the closed loopand as such tuning of PFC becomes less systematic than desirable. A pointof specific note is that almost always one needs to use ny > 1 and implicitlythis forces the behaviour towards open-loop dynamics Rossiter et al. (2014).

It is worthwhile giving some examples of this. Consider two systemsH1, H2 and determine the closed-loop poles for different choices of λ, ny andplot the slowest pole in figures 4 and 5 respectively.

H1 =0.02(4z−2 − z−1)

(1− 0.5z−1)(1− 0.9z−1); H2 =

10−4(3.3239z−1 + 0.313z−2 − 3.009z−3)

(1− 0.98z−1)(1− 0.967z−1)(1− 0.951z−1)(21)

9

Page 10: Pole-placement Predictive Functional Control for over ...

0 5 10 15 200.4

0.5

0.6

0.7

0.8

0.9

1

Coincidence horizon

Slo

wes

t pol

e po

sitio

n

λ=0.6λ=0.7λ=0.8λ=0.9λ=0.95

Figure 4: Relationship between tuning parameters and closed-loop poles for conventionalPFC on H1

It is immediately clear that the actual dominant closed-loop pole is welllinked to the choice of λ only if ny is small. What is less clear is that forthese cases a choice of ny < 10 may lead to significantly over active inputsignals (or even instability) and have output responses which are far fromideal. Consequently, for processes with significant lag, conventional PFCcan only deliver relatively slow closed-loop dynamics with potentially poorlinkage to the choice of λ.

0 5 10 15 200.8

0.85

0.9

0.95

1

Coincidence horizon

Slo

wes

t pol

e po

sitio

n

λ=0.8λ=0.85λ=0.9λ=0.95λ=0.98

Figure 5: Relationship between tuning parameters and closed-loop poles for conventionalPFC on H2.

10

Page 11: Pole-placement Predictive Functional Control for over ...

2.7. Summary

The background on PFC has established a few core insights which willbe used as the foundations of the proposal in this paper.

• With first-order models, the use of a horizon of one ensures a stronglink between the tuning parameter λ and the closed loop behaviourwhich results.

• With higher order models, there is not expected to be a strong linkbetween the tuning parameters and closed-loop behaviour and thustuning is less systematic, especially if small values of ny cannot beused, such as with non-minimum phase or lagged processes Rossiteret al. (2014).

• Using parallel independent models (figure 3) allows for much simplerprediction algebra and thus simpler coding.

3. Easy tuning by using a coincidence horizon of one with 2nd-order models

There is potential to improve the tuning of PFC for higher-order modelsand ideally to enable a stronger link between the tuning parameter λ and theresulting closed-loop dynamics. This section uses the special case of an over-damped second order process to show how a small change in the PFC controllaw can achieve this while not increasing complexity or affecting constrainthandling facilities. Specifically, the core concept is to exploit results for 1storder models for which the control law (11) can be written down explicitlyand for which one can fix the precise closed-loop pole with ny = 1.

Remark 4. For the purposes of closed-loop pole analysis this section willconsider the nominal case of ym = yp and d(k) = 0. This is useful because itindicates the efficacy of the tuning method; a formal sensitivity analysis con-stitutes future work although the examples demonstrate that good robustnessis expected.

3.1. Independent PFC designs with two parallel paths

The concept used here is to design separate PFC control laws for eachindependent model in figure 3 and then use a linear combination of these asthe actual control law applied to the system. For now the focus is on modelswith two poles.

11

Page 12: Pole-placement Predictive Functional Control for over ...

Lemma 3. The contribution to the model output steady-states of each par-allel model can be captured with variables γ1, γ2, γ1 + γ2 = 1 so that:

limk→∞

ym = r ⇒

{limk→∞ y

(1)m = γ1r

limk→∞ y(2)m = γ2r

(22)

Proof: This is obvious as limk→∞(y(1)m + y

(2)m ) = r. Moreover, γ can be

determined by using the model steady-state gains so that:

g1 =b1

1 + a1; g2 =

b21 + a2

; γ1 =g1

g1 + g2; γ2 = 1− γ1 (23)

Readers may note there is no need for g1 > 0, g2 > 0 nor indeed for γi > 0. �

Corollary 3. A number of simple corollaries fall out from dividing the targetbetween the two parallel paths:

1. Equivalent to meeting the target of ym → r, one can aim for separatetargets for each independent model:

y(1)m → γ1r = r(1), y(2)m → γ2r = r(2), (24)

2. The impact of the disturbance estimate on the required model predictionscan be divided between the subsystems using the steady-state gains sothat:

y(1)m → r(1) − γ1d(k); y(2)m → r(2) − γ2d(k) (25)

3. The ’process output’ can be divided in an equivalent ratio so that yp =

y(1)p + y

(2)p and:

y(1)p (k) = y(1)m (k) + γ1d(k), y(2)p (k) = y(2)m (k) + γ2d(k). (26)

Lemma 4. Assume for now that each subsystem has an independent inputand the disturbance is shared between the subsystems as above, then PFC laws(5) with a coincidence horizon ny = 1 for each parallel subsystem are definedas:

b1u(1)(k) = r(1)(1− λ) + λy

(1)p (k)− γ1d(k) + a1y

(1)m (k)

b2u(2)(k) = r(2)(1− λ) + λy

(2)p (k)− γ2d(k) + a2y

(2)m (k)

(27)

12

Page 13: Pole-placement Predictive Functional Control for over ...

Proof: This follows directly from the standard PFC law for 1st-order systemsassuming the disturbance and measurement are divided between the parallelloops in ratios γ1, γ2. �

If one were able to apply just u(1) to path 1 and u(2) to path 2, then themodel would of course reach the desired steady-state and with the desiredtime constant. However, the obvious inconsistency is that the actual processhas a single input and the independent model formulation of figure 3 assumesthat the same input enters both parallel paths. Consequently, the main roleof the control laws in (27) is to identify ideal choices; knowledge of theseideal choices can be used to find a compromise which balances the needs ofboth loops.

3.2. Proposed control law and nominal analysis

Algorithm 1. Determine u(1), u(2) using the control law definitions of (27)and then form the actual PFC system input as a linear combination of thesetwo.

u(k) = βu(1)(k) + (1− β)u(2)(k); (28)

The rationale for the choice of β will become clear next.

Lemma 5. For the nominal case, the control laws of (27) can be expressedin z-transforms as follows:

u(1)(z) = γ11−λb1

r(z) + (λ+a1)b1

y(1)m (z)

u(2)(z) = γ21−λb2

r(z) + (λ+a2)b2

y(2)m (z)

(29)

Proof: First replace y(1)p (k) = y

(1)m (k), y

(2)p (k) = y

(2)m (k) and d = 0, hence:

biu(i)(k) = (1− λ)r(i) + λy(i)m (k) + aiy

(i)m (k) (30)

Rearranging this one finds:

biu(i)(k) = (λ+ ai)y

(i)m + (1− λ)r(i) (31)

From which, with r(1) = γ1r and using a similar statement for loop 2, theresult is obvious. �Corollary 4. The proposed PFC control law combining (28) and (29) isgiven as:

u(z) = β(λ+ a1)

b1y(1)m (z) + (1− β)

(λ+ a2)

b2y(2)m (z) +Kr(z) (32)

where K = (1− λ)(βγ1/b1 + (1− β)γ2/b2).

13

Page 14: Pole-placement Predictive Functional Control for over ...

Theorem 6. The proposed PFC law of (32) achieves the target closed-looppole of λ and thus is an effective design tool for achieving the desired closed-loop dynamics, irrespective of the choice of β.

Proof: The first thing to note is that, direct from figure 3, one can deducethe following z-transform relationships:

y(i)m (z) =biz

−1

1 + aiz−1u(z) (33)

Substitute from (33) into (32) and one finds:

u(z) = β(λ+ a1)

b1

b1z−1

1 + a1z−1u(z) + (1− β)

(λ+ a2)

b2

b2z−1

1 + a2z−1u(z) +Kr(z)

(34)Next, combine all terms containing u(z):

pc(z)u(z) = a(z)Kr(z)pc(z) = [(1 + a1z

−1)(1 + a2z−1)− β(1 + a2z

−1)(λ+ a1)z−1 − (1− β)(1 + a1z

−1)(λ+ a2)z−1]

a(z) = (1 + a1z−1)(1 + a2z

−1)(35)

Validate the closed-loop pole polynomial has a root at λ by showing thatpc(λ) = 0. �

3.3. Exploiting the flexibility in β

The previous section proved that the proposed control law guaranteesthat one of the closed-loop poles will be at λ. This section investigates howalternative choices for β impact on the remaining poles and thus gives someinsights into how this parameter could be chosen.

Theorem 7. The root of p(z) given in (35) which is not at λ, is independentof the choice of λ and depends solely on the model parameters and the choiceof β.

Proof: Assume that the 2nd root is given as δ, then:

pc(z) = (1− λz−1)(1− δz−1) = 1− (λ+ δ)z−1 + λδz−2 (36)

Using eqn.(35) to define the coefficients of p(z) and match up the coefficientsof z−1, z−2 with (36).

a1 + a2 − β(λ+ a1)− (1− β)(λ+ a2) = −λ− δa1a2 − βa2(λ+ a1)− (1− β)a1(λ+ a2) = λδ

(37)

14

Page 15: Pole-placement Predictive Functional Control for over ...

Eliminating and simplifying these both reduce to:

(1− β)a1 + βa2 = −δ (38)

That is, δ has an explicit dependence upon β; conceptually δ is a linearcombination of the open-loop poles and one can define β from a desired δ orvice versa. �

Remark 5. In practice the choice of δ will be such that λ is the dominantclosed-loop pole.

4. Generalisation to higher-order models: Pole placement PFC

The previous section used a 2nd order example to demonstrate a simpleprinciple, that a PFC approach could be used to give precise closed-looppoles akin to a pole placement design while retaining core PFC propertieswhich enable systematic constraint handling. This section shows how thisapproach can be extended to higher-order examples.

Lemma 8. For multiple models in parallel such as in figure 3, the contribu-tion to the overall model G output steady-states of each parallel model can becaptured with variables γi, so that:

limk→∞

ym = r ⇒{

limk→∞ y(i)m = γir

γi = Gi(1)/G(1)(39)

The proof is obvious. Note that∑

i γi = 1 as∑

iGi(1) = G(1).

Algorithm 2. [Pole placement PFC] The proposed PFC control law (nom-inal case) combining (28) and (29) for multiple parallel models is given as:

u(k) =∑i

βi(λ+ a1)

b1y(i)m (k) +Kr (40)

for suitable K and βi variables to be chosen subject to∑

i βi = 1.

Lemma 9. The closed-loop pole polynomial for Algorithm 2 is given from:

pc(z) = 1−∑i

βi(λ+ ai)z−1

1 + aiz−1(41)

15

Page 16: Pole-placement Predictive Functional Control for over ...

Proof: Substitute the independent model equations y(i)m = [biz

−1/(1+aiz−1)]u(k)

into the control law of (32). Hence:

u(z) =∑i

βi(λ+ ai)

bi

biz−1

1 + aiz−iu(z) +Kr(z) (42)

The definition of the closed-loop pole polynomial is now obvious. �

Corollary 5. The closed-loop pole polynomial of (41) has at least one rootat λ as long as

∑i βi = 1. This follows easily by substituting z = λ:

pc(λ) = 1−∑i

βi(λ+ ai)

λ+ ai= 1−

∑βi (43)

Hence pc(λ) = 0 if 1 =∑

i βi.

Theorem 10. The degrees of freedom within βi are sufficient to place all thepoles of pc(z).

Proof: Let the desired poles be ρi, i = 1, 2, ... and let ρ1 = λ. Then, theseare closed-loop poles iff pc(z) can be expanded as follows:

1−∑i

βi(λ+ ai)

bi

biz−1

1 + aiz−i=

∏i

1− ρiz−1

1 + aiz−1=

∑i

ηi1 + aiz−1

(44)

Using a cover-up rule to determine suitable values for ηi gives:

ηj =

∏i(aj − ρi)∏

i,i ̸=j(−aj + ai)= −βj(ρ1 + aj) (45)

The key point here is that values of ηi can always be computed and there isa simple dependence of βi upon ρi, that is one can choose ρi and then findthe required βi. �

Corollary 6. If a desired pole ρi is specified as equal to an open-loop pole−ai , then this implies βi = 0 . This is obvious from (45) and also makes goodintuitive sense - the corresponding open-loop dynamic is in effect uncontrolledif the independent model output y

(i)m plays no part in the control law.

16

Page 17: Pole-placement Predictive Functional Control for over ...

5. Numerical examples

This section gives some simulation examples of the proposed pole place-ment PFC law [denoted PP-PFC] (32) and compare with conventional PFC[denoted CPFC]. These examples demonstrate that:

• the closed-loop dynamic of PP-PFC is directly affected by the choiceof λ whereas this is difficult with CPFC.

• the PP-PFC method is robust to non-zero disturbances and some pa-rameter uncertainty.

• the PP-PFC method allows systematic constraint handling.

• the PP-PFC method can deal with non-minimum phase processes easilywith ny = 1 (CPFC is far less flexible).

5.1. Illustrations of the impact of λ on closed-loop behaviour

This section will consider 2nd and 3rd order examplesH3, H2 respectively.

H2 =10−4(3.3239z−1 + 0.313z−2 − 3.009z−3)

(1− 0.98z−1)(1− 0.967z−1)(1− 0.951z−1); H3 =

0.1z−1 + 0.4z−1

(1− 0.5z−1)(1− 0.9z−1)(46)

The figures will compare the closed-loop behaviour with the conventionalPFC control law of (19) with the proposed pole placement PFC law of (40).The simulations include an additive output disturbance (around sample 31in figure 6) to demonstrate that both methods are robust to disturbances.

5.1.1. Example H3

A challenging scenario would be a substantial speed up of the process togive a closed-loop pole at ρ = 0.3. With CPFC a search of various possibletuning parameters demonstrates that although one can speed the process up,the actual closed-loop poles cannot be selected precisely, even with trial anderror search over λ, ny as seen in section 2 and here:

{λ = 0.3, ny = 2} ⇒ ρ1,2 = 0.25, 0.06{λ = 0.3, ny = 3} ⇒ ρ1,2 = 0.42± j0.16{λ = 0.2, ny = 3} ⇒ ρ1,2 = 0.42± j0.018

(47)

17

Page 18: Pole-placement Predictive Functional Control for over ...

10 15 20 25 30 35 400

0.5

1

Samples

PFC RESPONSES with ρ = 0.3

y (CPFC)y (PPPFC)

10 15 20 25 30 35 40

0

0.5

1

1.5

Samples

u (CPFC)u (PPPFC)

Figure 6: Simulation of model H3 with a target pole of λ = 0.3 using CPFC and PP-PFC.

Clearly however PP-PFC can deliver the desired pole precisely - in this casethe user choices are:

ρ1,2 = 0.3, 0.3; a1 = 0.9, a2 = 0.5, β1 = 1.5, β2 = −0.5 (48)

The closed-loop simulations are given in figure 6 and demonstrate that al-though the responses are similar, PP-PFC has given behaviour much closerto the target dynamic. A further demonstration of the efficacy of PP-PFCis given in figure 7 which shows that the algorithm reliably gives the desireddynamic as ρ1, ρ2 are changed.

5.1.2. Example H2

In this case the aim is to place a closed-loop pole at ρ = 0.8, that is toroughly double the speed. With CPFC a search of various possible tuningparameters demonstrates that although one can speed the process up, thedominant closed-loop poles cannot be selected precisely, even with trial anderror search over λ, ny as seen in section 2 and here:

{λ = 0.8, ny = 1} ⇒ ρ1,2,3 = 0.8, 0.9, 0.999{λ = 0.8, ny = 3} ⇒ ρ1,2,3 = 0.5, 0.8, 0.9{λ = 0.8, ny = 8} ⇒ ρ1,2,3 = 0.87± j0.1, 0.91

(49)

Clearly however PP-PFC can deliver the desired poles precisely. The closed-loop simulations are given in figure 8 and demonstrate that although the

18

Page 19: Pole-placement Predictive Functional Control for over ...

5 10 15 20 25 300

0.5

1

Samples

PPPFC RESPONSES with various ρ

ρ=0.3ρ =0.4ρ=0.6ρ=0.8

5 10 15 20 25 300

0.5

1

Samples

Figure 7: Simulation of model H3 with PP-PFC using different target poles.

responses are similar, PP-PFC has given behaviour much closer to the targetdynamic. Clearly nothing is for free and the speed up is paid for by arelatively aggressive input.

5.2. Dealing with with parameter uncertainty

This section demonstrates that the PPPFC algorithm is robust to someparameter uncertainty - a detailed sensitivity analysis forms future work.Take system H3 as the system model Gm and the real process Gp as follows:

Gp =0.12z−1 + 0.37z−2

(1− 0.4z−1)(1− 0.92z−1)(50)

The closed loop behaviour for two choices of λ are given in figures 9 and 10;a disturbance d is also added at sample 40. The figures show clearly that thelarge difference between the model output ym and actual process output ypdoes not affect the offset free tracking property. These figures demonstrateclearly the robust behaviour of the algorithm and the continuing efficacy ofλ as a tuning parameter.

5.3. Non-minimum phase system

The proposed PP-PFC technique works for non-minimum phase systems(here H4(z) given in (51)) whereas a CPFC approach requires ny ≫ 1 and λis less effective. Figure 11 shows the closed-loop PP-PFC behaviour matches

19

Page 20: Pole-placement Predictive Functional Control for over ...

10 20 30 40 50 60 70 800

0.5

1

1.5

Samples

PFC RESPONSES with ρ = 0.8

y (CPFC)y (PPPFC)

10 20 30 40 50 60 70 80

0

5

10

Samples

u (CPFC)u (PPPFC)

Figure 8: Simulation of model H2 with a target pole of λ = 0.3 using CPFC and PPPFC.

0 10 20 30 40 50 600

0.5

1

1.5

Samples

PFC RESPONSES with ρ = 0.5

ym

yp

udr

Figure 9: Simulation of model (50) with mismatching parameters and a double target poleof λ = 0.5.

20

Page 21: Pole-placement Predictive Functional Control for over ...

0 10 20 30 40 50 600

0.5

1

1.5

Samples

PFC RESPONSES with ρ = 0.9

ym

yp

udr

Figure 10: Simulation of model (50) with mismatching parameters and a double targetpole of λ = 0.9.

the target pole λ despite the request for a significant speed up. CPFC hasto use a large coincidence horizon for this case and hence can only be usedto obtain poles close to the open-loop dynamics.

H4 =−0.2z−1 + 0.6z−2

(1− 0.5z−1)(1− 0.9z−1)(51)

5.4. Constraint handling

A major reason for preferring PFC to PID, even with low-order systemswhere achievable closed-loop behaviour may be similar, is the constraint han-dling facility which is cumbersome to code and test with PID but elementarywith PFC.

The proposed control law for PP-PFC given in (40) has a simple facilityfor including constraint handling summarised as follows:

1. Test whether the proposed input satisfies input absolute and rate con-straints. If not, modify u(k) to ensure both using saturation.

2. To ensure satisfaction of output/state constraints one must form theimplied predictions of (15) over a sensible but large horizon and modifyu(k) as required to ensure satisfaction. This reduces to a simple forloop which ensures, for example, that maxi yp(k + i|k) is within limits;as yp has a linear dependence on u(k) this is easy to do.

21

Page 22: Pole-placement Predictive Functional Control for over ...

0 5 10 15 20 25 30 35

−0.2

0

0.2

0.4

0.6

0.8

1

Samples

PFC RESPONSES with ρ = 0.4

ym

yp

udr

Figure 11: Simulation of non-minimum phase system H4 with a target pole of λ = 0.4.

Figure 12 shows the PP-PFC algorithm successfully integrates input con-straints of −2 < u(k) < 2 and continues to deliver good performance.

Remark 6. It is non-trivial to guarantee the satisfaction of state constraintsin general, especially in the presence of disturbances and parameter uncer-tainty and algorithms for achieving this have substantial coding requirementsand complexity and thus are inconsistent with the context of a PFC approach.Nevertheless, the PFC approach guarantees recursive feasibility of state con-straints for the nominal case which is still a strong result.

6. Summary of coding requirements and applications on hardware

For completeness this section looks at how the PPPFC algorithm wouldbe coded and demonstrates efficacy on real hardware.

6.1. Coding requirements

The coding complexity of the proposed algorithm is simple compared tocommon alternatives, and can be implemented in just a few (5 to 10) linesof code.OFFLINE COMPUTATIONS

22

Page 23: Pole-placement Predictive Functional Control for over ...

10 20 30 40 500

0.5

1

1.5

Samples

PFC RESPONSES with ρ = 0.8

y (CPFC)y (PPPFC)

10 20 30 40 50−2

0

2

Samples

u (CPFC)u (PPPFC)

Figure 12: Simulation of system H3 with input constraints and a target pole of λ = 0.8.

1. Find partial fraction expansion of (14); this is trivial using the cover-uprule1.

2. Choose desired values for ρi and use (45) to determine βi.

3. Determine parameters γi as in (39).

ONLINE COMPUTATIONS (each sample)

1. Update model states: y(i)m (k + 1) = aiy

(i)m (k) + biu(k).

2. Update d(k) estimate using: d(k) = yp(k)−∑

i y(i)m (k).

3. Calculate u(k) using (40).

4. Ensure u(k) satisfies any relevant constraints before implementing.

6.2. Application on hardware

This section illustrate the efficacy of the proposed method on a hot airblower pilot plant (AMIRA LTR-701) (see figure 13). The air is sucked inthrough a radial fan and flows through a throttle valve for heating. The air isheated and then flows through a tube. The input signal is the heating powerand the output signal is the temperature at the end of the tube. Modelparameters were identified at a sampling time of 2s using a least squares

1We exclude discussion of systems with repeated poles for brevity. In practice, followingan identification, real systems can always be modelled with slightly different poles anyway.

23

Page 24: Pole-placement Predictive Functional Control for over ...

Figure 13: Hot air blower ARIMA LTR-701.

3

4

5

6

0 500 1000 1500 2000 2500

y p (proposed)

y ref

t [sec]

Figure 14: Real-time PFC control of ARIMA LTR-701.

method and measured data filtered by a low-pass filter with time constant10s. The identified model parameters between the measured voltages is:

ym(z) =

[0.23z−1

1− 0.71z−1+

0.019z−1

1− 0.986z−1

]u(z) (52)

The proposed PFC is designed uses poles ρ1,2 = 0.98, 0.9. For the real-time control (figure 14), the reference is changed stepwise from 2.9V to 6Vand a disturbance was introduced around 1300s by increasing the flow ratesubstantially. The proposed PFC control law has been effective at obtainingthe desired response characteristics and has robust behaviour.

7. Conclusions and future work.

Although PFC is successful in many scenarios, the underlying requirementof making a higher-order system take on output predictions close to a first-

24

Page 25: Pole-placement Predictive Functional Control for over ...

order system can give rise to inconsistencies where there are significant lagor non-minimum phase characteristics and thus tuning by trial and error isdifficult and indeed the achievable results are limited.

This paper has proposed a novel PFC law which exploits the individualdynamics making up the system model and thus is able to use a coincidencehorizon of just one. The proposed PP-PFC approach is effective and simpleto tune on high order systems with simple poles, even though this is incontradiction to the normal guidance of large coincidence horizons beingessential. This result is significant because of two main factors:

1. The use of a coincidence horizon of one means that the coding andprediction is trivial; implementation requires just 5 to 10 simple in-structions.

2. Control tuning for conventional PFC requires trial and error and oftencannot deliver the desired dynamics due to the limited flexibility of theapproach. Conversely, the proposed pole placement PFC algorithm hasmore flexibility with no overall increase in coding complexity, combinedwith a more intuitive design procedure, as no trial and error is required.

3. It is notable that the proposed approach is able to deal with non-minimum phase characteristics much better than CPFC and thus hasfar more flexibility in tuning the resultant closed loop.

The proposed approach retains the important attributes of robustness touncertainty and in particular systematic constraint handling and thus is moreflexible than the obvious alternatives of PID or pole-placement for whichsystematic constraint handling is difficult. Efficacy has been demonstratedon numerous examples and laboratory hardware.

Future work will consider the extension of this concept to systems withcomplex poles, systems with dead-time, to open-loop unstable systems andto cases where the targets are ramps. Moreover there is a need to considerthe extent to which one can exploit the additional flexibility available withinthe choice of the other closed-loop poles; as these positions come from simplelinear combinations of existing poles it is expected that a common sensedefault will be possible.

Changenet, C., Charver, J.N., Gehin, D., Sicard, F. and Charmel, B., Predic-tive Functional Control of an expansion valve for controlling the evaporatorsuperheat, IMechE Journal of Systems and Control Engineering, 222 (6),pp571–582, 2008.

25

Page 26: Pole-placement Predictive Functional Control for over ...

Fallasohi, H., Ligeret, C. and Lin-shi, X., Predictive Functional Control ofan expansion valve for minimizing the superheat of an evaporator, Inter-national Journal of Refrigeration, 33(2), pp409–418, 2010.

Haber, R., Bars, R. and Schmitz, U., Predictive Control in Process En-gineering: From the Basics to the Applications, Chapter 11: PredictiveFunctional Control, Wiley-VCH, Weinheim, Germany, 2011.

Kerrigan, E.C., Robust Constraint Satisfaction: Invariant Sets and Predic-tive Control. PhD thesis, University of Cambridge, UK, 2000.

Khadir, M.T., and Ringwood, J.V., Extension of first order predictive func-tional controllers to handle higher order internal models, Int. J. Appl.Math. Comput. Sci., 18(2), pp229-239, 2008.

Kouvaritakis, B., Rossiter, J.A., and Chang, A.O.T., Stable generalized pre-dictive control: an algorithm with guaranteed stability, Proc IEE, 139(4),pp349–362, 1992.

Kvasnica, M. (2009). Real-time model predictive control via multi-parametricprogramming: theory and tools. VDM-Verlag.

Maciejowski, J., Predictive control with constraints, Prentice Hall, 2001.

Mayne, D.Q., Seron, M.M. and Rakovic, S.V., Robust model predictive con-trol of constrained linear systems with bounded disturbances, Automatica,41 (2), pp219-224, 2005.

Richalet, J., Rault, A., Testud, J.L. and Papon, J., Model predictive heuristiccontrol: applications to industrial processes, Automatica, 14(5), pp413–428, 1978.

Richalet J. and ODonovan, D., Predictive functional control Principles andindustrial applications, Springer-Verlag, 2009, London, England.

Richalet, J., Lavielle, G. and Mallet, J., Commande Predictive, Eyrolles,2004.

Rossiter, J.A., Notes on multi-step ahead prediction based on the principleof concatenation, Proceedings IMechE., 207, pp261–263, 1993.

26

Page 27: Pole-placement Predictive Functional Control for over ...

Rossiter, J.A., Model predictive control: a practical approach, CRC press,2003.

Rossiter, J.A., Richalet, J. and Haber, R., The effect of coincidence horizonon predictive functional control, Special issue on Process Control: CurrentTrends and Future Challenges, Processes, 2015, Vol. 3, Issue 1, pp25–45,doi: 10.3390/ pr3010025

Scokaert, P.O. and Rawlings, J.B., Constrained linear quadratic regulation.Automatic Control, IEEE Trans. on AC, 43(8), 1163–1169, 1998.

27