Advanced Methods of PID Controller Tuning For Specified Performance
Advanced Methods of PID Controller Tuning For Specified Performance
76069
Provisional chapter
Chapter 4
Methods of
Advanced Methods of PID
PID Controller
ControllerTuning
Tuningfor
for
Performance
Specified Performance
http://dx.doi.org/10.5772/intechopen.76069
Abstract
This chapter provides a concise survey, classification and historical perspective of practice-
oriented methods for designing proportional-integral-derivative (PID) controllers and
autotuners showing the persistent demand for PID tuning algorithms that integrate per-
formance requirements into the tuning algorithm. The proposed frequency-domain PID
controller design method guarantees closed-loop performance in terms of commonly used
time-domain specifications. One of its major benefits is universal applicability for both
slow and fast-controlled plants with unknown mathematical model. Special charts called
B-parabolas were developed as a practical design tool that enables consistent and system-
atic shaping of the closed-loop step response with regard to specified performance and
dynamics of the uncertain controlled plant.
1. Introduction
How to tune a controller for any control application quickly and appropriately? This question
raised in 1942 is still up to date and constantly occupies the automation community worldwide.
The answer is very intricate; its intricacy is comparable with the open hitherto unresolved
Hilbert problems known from mathematics.
Will the PID controllers, historically the oldest but currently still the most used ones, control
industrial processes in the near and far future? Based on the increase of the number of PID
tuning methods from 258 to 408 during 2000–2005, a positive response can be assumed [23].
© 2016 The Author(s). Licensee InTech. This chapter is distributed under the terms of the Creative Commons
© 2018 The Author(s). Licensee IntechOpen. This chapter is distributed under the terms of the Creative
Attribution License (http://creativecommons.org/licenses/by/3.0), which permits unrestricted use,
Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits unrestricted use,
distribution, and eproduction in any medium, provided the original work is properly cited.
distribution, and reproduction in any medium, provided the original work is properly cited.
74 PID Control for Industrial Processes
The remarkably simple ability of the PID controller to generate a difference equation using the
present, past and future values of the control error is often projected into the philosophical
understanding [1] and forecast this controller a long-term perspective.
Beginnings of PID controllers date back to 1935 when the Taylor Instruments Companies
launched their pneumatic controller with a derivative channel [1]. Owing to rapid develop-
ments in the control theory, it was supposed that the conventional PID controllers would be
gradually replaced by advanced ones; however, this did not come to pass mainly due to the
simple PID structure and its commercial usability in practice. For 83 years, control loop
designers preferred the PID controllers for their outstanding ability to eliminate the control
error using the integrator, their ability to improve the performance using the “trend” of the
controlled variable through the derivative channel and for many other benefits. PID controllers
are important parts of distributed control systems, predictive control structures; their coeffi-
cients are often adapted by means of fuzzy and neural control and set by genetic algorithms
[20, 21, 35]. In multiloop control structures, they are able to stabilize unstable objects and
difficult-to-control systems. The 46 existing PID variants and reported 408 diverse tuning
methods are a good prerequisite for achieving a satisfactory performance in simple as well as
demanding industrial applications [23, 25].
PID controllers are widely applied in technological processes of heavy and light industries, for
example in control of tension in the roll during paper winding, boiler temperature, chemical
reactor pressure, lathe spindle position in metalworking, and so on; they can be found in
modern cars controlling combustion control or vehicle dynamics [9], valve opening and
robotic arm position. In the interconnected power system, they are used to control turbine
power and speed in both primary and secondary regulation of active power and network
frequency. Being easy to implement on both Arduino and Raspberry Pi platforms allows them
to be used in mobile “unplugged” applications as well.
Commercial applicability of PID controllers is confirmed by studies referring that more than
90% out of all installed controllers in industrial control loops are PID controllers [36]. The
alarming fact, however, is that only 20% of them are tuned correctly, and in 30% of all PID
applications, the regulation is unsuitable due to an incorrect selection of synthesis method.
Another 30% of poor performance is due to ignorance of nonlinear properties of actuators, and
the remaining 20% represent an inadequate choice of sampling period or poor signal filtering
[5]. Some controllers not only do not provide the required performance yet often even stability
of the control loop being operated only in open loop and manually switched off by the service
staff when approaching the setpoint [9]. According to other statistics, 30% of controllers
operate in manual mode and require continuous fine-tuning and supervising by the process
technologist. A 25% of PID applications use coefficients pre-set by the manufacturer with no
update of their values with respect to the particular process [47].
Therefore, a natural requirement for innovative PID tuning methods has come up to ensure the
specified performance [19, 22] in terms of the maximum overshoot and settling time not only
for processes with constant parameters but for their perturbed types. In this chapter, a novel
original robust PID tuning method is presented; hopefully, it will help reverse the above
mentioned unfavorable statistics of incorrectly tuned PID controllers.
Advanced Methods of PID Controller Tuning for Specified Performance 75
http://dx.doi.org/10.5772/intechopen.76069
Despite the fact that there are more than 11,000 PID controllers in 46 variants operating in
industrial processes [23], mostly three basic forms are used to control industrial processes: the
ideal (textbook) PID controller, the real PID controller with derivative filter, and the ideal PID
controller in series with the first-order filter given by the following transfer functions, respectively.
!
1 1 Tds 1 1
GR ð s Þ ¼ K 1 þ þ Tds , GR ðsÞ ¼ K 1 þ þ , GR ð s Þ ¼ K 1 þ þ Tds , (1)
Tis T i s 1 þ TNd s Tis Tf s þ 1
where K is the proportional gain, Ti and Td are the integral and the derivative time constants,
respectively, Tf is the filter time constant and N∈<8,16> in practical applications [37, 38]. The
PID controller design objectives are:
1. tracking of setpoint or reference variable w(t) by y(t),
2. rejection of disturbance d(t) and noise n(t) influence on the controlled variable y(t).
Time response of the controlled variable y(t) is modifiable by parameters K, Ti and Td, respec-
tively; the objective is to achieve a zero steady-state control error e(t) irrespective if caused by
changes in the reference w(t) or the disturbance d(t). This section presents practice-oriented
PID controller design methods based on various performance criteria.
Consider the control loop in Figure 1 with control action u(t) generated by the PID controller
(switch SW in position “1”).
A controller design is a two-step procedure consisting of controller structure selection (P, PI,
PD or PID) followed by tuning coefficients of the selected controller type.
An appropriate structure of the controller GR(s) is usually selected with respect to:
• zero steady-state error condition (e(∞) = 0),
Figure 1. Feedback control loop with load disturbance d(t) and measurement noise n(t).
76 PID Control for Industrial Processes
2.1.1. PID controller structure selection based on zero steady-state error condition
Consider the unity feedback control loop (Figure 1), where G(s) is the controlled system.
According to the final value theorem, the steady-state error
1 sνq
eð∞Þ ¼ lim sEðsÞ ¼ lim s W ðsÞ ¼ q!wq lim ν (2)
s!0 s!0 1 þ LðsÞ s!0 s þ K L
is zero if the integrator orders νL = νS + νR in the open-loop L(s) = G(s)GR(s) is greater than the
order q of the reference signal w(t) = wqtq, i.e.,
νL > q, (3)
where νS and νR are integrator orders of the plant and the controller, respectively, KL is the
open-loop gain and wq is a positive constant [46].
where Kc and ωc are critical gain and critical frequency of the plant, respectively, the normal-
ized time delay μj and the parameter ϑj can be used to select appropriate PID control strategy.
According to Table 1 [46], the derivative part is not used in presence of intense noise and a PID
controller is not appropriate for plants with large time delays.
Performance measures for industrial control loops can be expressed both in the time and the
frequency domains. The time-domain performance indicators allow to directly expressing the
desired process parameter, whereas the frequency-domain performance indicators can be used
as PID tuning parameters.
Advanced Methods of PID Controller Tuning for Specified Performance 77
http://dx.doi.org/10.5772/intechopen.76069
0.6 < μ1 < 1; 1.5 < ϑ1 < 2.25 I or PI I+A PI + A (PI or PID) + A + C
A: forward compensation suggested, B: forward compensation necessary, C: dead-time compensation suggested, D: dead-
time compensation necessary, E: set-point weighting necessary, F: pole-placement.
Figure 2. Performance measures: δDR, ts, ηmax and e(∞); (a) setpoint step response; (b) load disturbance step response;
(c) over-, critically- and underdamped closed-loop step-responses.
In the time-domain, satisfactory setpoint tracking (Figure 2a) and disturbance rejection
(Figure 2b) are indicated by small values of maximum overshoot and a decay ratio, respec-
tively, given as
y
max yð∞Þ Aiþ1
ηmax ¼ 100 ½%; δDR ¼ , (5)
yð∞Þ Ai
where y(∞) denotes the steady-state value of y(t) [4]. A measure of the y(t) response decay is
the ratio of two successive amplitudes Ai+1/Ai, where i = 1…N, and N is half the number of
points of intersections of y(∞) and y(t). The settling time ts is the time after which the output
y(t) remains within ε% of its final value (Figure 2b); typically ε = [1%÷5%]y(∞), δDR∈(1:4;1:2),
78 PID Control for Industrial Processes
ηmax∈(0%;50%). Figure 2c depicts underdamped (plot 1), overdamped (plot 2) and critically
damped (plot 3) closed-loop step responses.
If a designed controller GR(jω) guarantees, that |S(jω)| or |T(jω)| do not exceed prespecified
values Ms or Mt, respectively, defined by
Ms ¼ sup jSðjωÞj ¼ sup
1 ; Mt ¼ sup jT ðjωÞj ¼ sup LðjωÞ (6)
ω ω 1 þ LðjωÞ ω ω 1 þ LðjωÞ
for ω∈〈0,∞), then the Nyquist plot of the open-loop transfer function L(s) = G(s)GR(s) avoids
the respective circles MS or MT each given by its center and radius as follows:
" #
1 M2t Mt
CS ¼ ½1; j0, RS ¼ ; CT ¼ 2 ; j0 , RT ¼ : (7)
Ms Mt 1 1 M2t
By avoiding the Nyquist plot of L(s) to enter the circles corresponding to MS or MT, a safe
distance from the critical point is kept (Figure 3a). Typical |S(jω)| and |T(jω)| plots for
properly designed controller are in Figure 3b. The disturbance d(t) is sufficiently rejected if
Figure 3. (a) Definition and geometrical interpretation of ϕM and GM in the complex plane; (b) sensitivity and comple-
mentary sensitivity magnitudes |S(jω)|, |T(jω)| and performance measures Ms, Mt.
Advanced Methods of PID Controller Tuning for Specified Performance 79
http://dx.doi.org/10.5772/intechopen.76069
Ms∈(1.2;2). The reference w(t) is properly tracked by the process output y(t) if Mt∈(1.3;2.5).
With further increasing of Mt the closed-loop tends to be oscillatory.
From Figure 3a results, that by increasing open-loop phase margin ϕM the gain crossover
L(jωa*) on the unit circle M1 moves away from the critical point (1,j0); similarly by increasing
open-loop gain margin GM the phase crossover L(jωf*) moves away from (1,j0). Therefore, the
stability margins ϕM or GM given by
1
φM ¼ 180 þ arg L ω∗a ; GM ¼ (8)
L jω∗f
are frequently used performance measures, their typical values are ϕM∈(20 ;90 ), GM∈(2;5).
Relations between individual stability margins and respective magnitude peaks are given by
the following inequalities
1 1 Ms 1
φM ≥ 2 arcsin ; φM ≥ 2 arcsin ; GM ≥ ; GM ≥ 1 þ : (9)
2Ms 2Mt Ms 1 Mt
The point in which the Nyquist plot L(jω) touches the MT circle defines the closed-loop
resonance frequency ωMt.
When synthesizing a control loop, if the controller type is already known and the designer has
just to select a suitable method to appropriately adjust its coefficients, we speak about PID
controller tuning methods. Controller design is a more complex problem which includes deter-
mining controller structure and then calculating its parameters. When setting coefficients of
industrial PID controllers {K, Ti, Td}, basically the following procedures are applied:
2. Analytical methods are used to generate a control law based on the mathematical model of
the plant; the plant model is obtained from first principles or via experimental identifica-
tion. The success of these methods depends on the accuracy of the mathematical model of
the plant, and is not always achievable in practical cases (e.g., for a cement kiln).
3. Classical tuning methods use only a limited number of characteristic parameters of the
plant obtained from the step response or critical system parameters [11, 27, 31]. Their main
advantage is a simple and short calculation of controller parameters. The control objective
is to provide a satisfactory response to reference change, or disturbance rejection and often
their combination. The main drawback is that the designer cannot influence the perfor-
mance by means of the adjustable parameters of the algorithm. Also, the resulting closed-
loop response may not be satisfactory if the step response of the plant is nonmonotonic, or
80 PID Control for Industrial Processes
when the plant has nonminimum-phase dynamics or large time delay. Most of these
methods are implemented in autotuners of industrial PID controllers [1].
4. Autotuners are a modern and convenient means for adjusting coefficients of industrial
controllers [33, 34, 49]. They implement a two-step algorithm of automatic acquisition of
characteristic parameters of the controlled process followed by automatic calculation and
adjustment of the controller coefficients. After activating the autotuning function on the
industrial controller, the control-loop synthesis is performed automatically in a very short
time. The ABB, Emerson, Fischer-Rosemont, Foxboro, Honeywell, Siemens, Yokogava, or
ZPA controllers have a built-in PID autotuning function implemented on a microcomputer
[47]. In many situations, however, these methods are unreliable because of the imperfec-
tion of the plant identification algorithm and the subsequent controller design.
5. Robust PID controller tuning methods for specified performance represent a modern area of
industrial control-loop synthesis. They improve the PID tuning methods by providing
stability and required performance also for processes with variable parameters. The con-
troller tuned only by conventional method is just “intuitively” invariant against perturba-
tions of the controlled plant; robust operation of the control loop is usually possible only
for small changes of plant parameters. The major disadvantage of these methods is that the
control law is not based on the knowledge of uncertainties of the controlled object and a
further research on their possible expansion is needed. The proposed original method
which eliminates this drawback, its theoretical analysis and verification on benchmark
examples are the core of the chapter.
Tuning methods are commonly used engineering tool for the synthesis of industrial control
loops as they do not require a full knowledge of the mathematical description of the controlled
plant. This differentiates them from analytical methods which, on the contrary are based on a
precise knowledge of the mathematical model of the controlled system. In the tuning methods,
the controlled process is considered as a black-box which is to be revealed only to such
extent that the controller synthesis is successful and the control objectives are achieved.
Thus, only those characteristic parameters of the unknown plant have to be acquired via
appropriate identification that are inevitable for the PID controller design. In this way, the
PID controller coefficients can be obtained in a relatively short time. The implicit knowledge
about the controlled system and the ambient influences affect the choice of the PID coeffi-
cients calculation method . According to the way of using the identified data of the controlled
plant, the tuning methods are classified into as follows:
a. model-free PID controller tuning methods,
lag plus dead time) system models which are the most commonly used approximation of plant
dynamics in industries (thermal plants, chemical and woodworking industries). Equal 9%
shares belong to IPDT (integral plus dead time) and FOLIPDT (first order lag integral plus
dead time) models with integral behavior encountered mainly in drives and power industry in
modeling mechanical subsystems of rotating machines, valves and servo-systems and 18% of
algorithms are used to control plants with SOSPDT (second order system plus dead time)
models [26]. Controllers for other system types are tuned by methods from the 15% portion.
Due to its simplicity, the Ziegler-Nichols frequency-domain methods are still used in industrial
autotuners in the original version, although they have undergone various modifications dur-
ing the last 70 years of its existence. Due to the technological development after the industrial
revolution and major electrification, PID tuning for stability was no more sufficient because a
fast setpoint attainment could bring about important savings of time and money and
Advanced Methods of PID Controller Tuning for Specified Performance 83
http://dx.doi.org/10.5772/intechopen.76069
accelerate the entire production process. More and more demanding requirements on control
performance were formulated, and an intense demand for effective tuning methods guarantee-
ing required performance has arisen.
As a rule, application of Ziegler-Nichols methods usually leads to oscillatory closed-loop
responses; hence, many scientists have become interested in their possible improvement.
Forty-two modifications of the Ziegler-Nichols frequency method were developed in the
period from 1967 to 2010. They differ from the classical algorithm in using various other
combinations of the weights (α1, α2, α3). An overview of selected model-free methods is given
in Table 2.
Tuning rules No. 1–3 are the well-known Ziegler-Nichols frequency-domain method which
objective is a fast rejection of the disturbance d(t) and δDR = 1:4. In the complex plane interpre-
tation (Figure 6), the method corresponds to shifting the critical point C = [1/Kc + j0], into the
points CP = [0.5 + j0], CPI = [0.45 + j0.0896] and CPID = [0.6-j0.28] using respectively P, PI
and PID controllers tuned according to Table 2. Put simply, the open-loop Nyquist plot is
shaped into a sufficient distance from the limit of instability specified by the point (1.j0).
Table 2. Model-free PID controller tuning rules based on critical plant parameters.
Figure 6. Moving the critical point C = [1/Kc + j0] of the plant using P, PI and PID controllers designed by Ziegler-
Nichols frequency-domain method for critical frequency ωc of the plant.
84 PID Control for Industrial Processes
Related methods (No. 4–10) use various weighting of critical parameters thus allowing to vary
the closed-loop performance requirements (see the last column in Table 2). All presented
methods (No. 1–10) are applicable for various plant types, easy-to-use and time efficient.
2π 4M 4ðM 0:5ΔDB Þ
ωc ¼ ; Kc_IR ¼ ; Kc_HR ¼ , (10)
Tc πAc πAc
where the period Tc and the amplitude Ac of critical oscillations are read off from y(t) of the
recorded limit cycle (Figure 7b); ΔDB is the width of the hysteresis plot, the relay amplitude M
is chosen as (3÷10)% of the control u(t) limits. A typical limit cycle is depicted in Figure 7b. A
hysteresis relay is used if y(t) corrupted by a noise n(t) [47].
The advantage of these methods is their applicability for different types of systems, simplicity
and the short time needed for the controller design of the—approx. (3÷4)Tc.
In these methods, the identified characteristics of the unknown system are used to create its
typical model, and the controller design algorithm is derived for this particular model. Formu-
las for calculation of the controller coefficients include process model parameters that are
function of the identified process data. Each method works perfectly for the system whose
model has been used in the design algorithm. However, if the system is approximated differ-
ently, the achieved performance may be impaired or even insufficient. The advantage is that
control objectives can be clearly defined and expressed using analytical relationships (e.g., it is
possible to derive the relationship for maximum overshoot of the step response). A small
flexibility due to the “tailor-made” design for one type of model limits the widespread appli-
cation of these methods in autotuners of industrial controllers.
PID tuning algorithms of indirect tuning methods have two more steps compared with direct
methods, as shown in the flow chart in Figure 8. When choosing the procedure for creating a
Figure 7. Determination of critical parameters Kc and Tc of the controlled plant from the limit cycle.
Advanced Methods of PID Controller Tuning for Specified Performance 85
http://dx.doi.org/10.5772/intechopen.76069
Figure 8. Flow chart of the indirect engineering method for PID tuning.
typical model . , it is important how the implicit information about the controlled system is
considered (if we deal with a driving system, a thermal process, a mechanical or pneumatic
system, etc.). If the typical model for the given controlled system has already been selected, the
model parameters are calculated in step . and subsequently used in calculation of the PID
coefficients.
The static and dynamic properties of most technological processes can be expressed by one of
the FOLPDT, IPDT, FOLIPDT, or SOSPDT models. Model parameters are identified from the
recorded step response of the controlled system (Figure 9) and are further used in calculation
of PID controller coefficients. According to Figure 1, step response of the controlled process is
obtained by switching SW into position “2” and performing step change in u(t).
Transfer functions of the model are found from the step response parameters according to Figure 9.
Figure 9. Typical step responses of (a) FOLPDT; (b) IPDT and (c) FOLIPDT models.
86 PID Control for Industrial Processes
11. Ziegler and Nichols, 1942 [48] P 1/ϑ1 — — Quarter decay ratio
12. Ziegler and Nichols, 1942 [48] PI 0.9/ϑ1 3D1 —
13. Ziegler and Nichols, 1942 [48] PID 1.2/ϑ1 2D1 0.5D1
14. Chien et al., 1952, regulator tuning [18] PI 0.6/κ1 4D1 — ηmax = 0%, D1/T1∈(0.1;1)
15. Chien et al., 1952, regulator tuning [18] PID 0.95/ϑ1 2.38D1 0.42D1
16. Chien et al., 1952, regulator tuning [18] PI 0.77/ϑ1 2.33D1 — ηmax = 20%, D1/T1∈(0.1;1)
17. Chien et al., 1952, regulator tuning [18] PID 1.2/ϑ1 2D1 0.42D1
18. Chien et al., 1952, servo tuning [18] PI 0.35/ϑ1 1.17D1 — ηmax = 0%, D1/T1∈(0.1;1)
19. Chien et al., 1952, servo tuning [18] PID 0.6/ϑ1 D1 0.5D1
20. Chien et al., 1952, servo tuning [18] PI 0.6/ϑ1 D1 — ηmax = 20%, D1/T1∈(0.1;1)
21. Chien et al., 1952, servo tuning [18] PID 0.95/ϑ1 1.36D1 0.47D1
22. ControlSoft Inc., 2005 [23] PID 2/K1 T1 + D1 max(D1/3;T1/6) Slow loop
23. ControlSoft Inc., 2005 [23] PID 2/K1 T1 + D1 min(D1/3;T1/6) Fast loop
Table 3. PID tuning rules based on FOPDT model, ϑ1 = K1D1/T1 is the normalized process gain.
2.4.2.3. PID controller tuning formulas for IPDT and FOLIPDT models
While dynamics of slow technological processes (polymer production, heat exchange, etc.) can
be approximated by an IPDT model (11b), electromechanical subsystem of rotating machines
and servo drive objects are typical examples for using a FOLIPTD model [42] (11c) (Table 4).
The gain K in the rule No. 27 is variable with respect to the normalized time delay υ3 = D3/T3 of
the FOLIPDT model; for the corresponding pairs holds: (υ3;x3) = {(0.02;5), (0.053;4); (0.11;3);
(0.25;2.2); (0.43;1.7); (1;1.3); (4;1.1)}.
Table 4. PID tuning rules based on IPDT and FOLIPDT model parameters.
Advanced Methods of PID Controller Tuning for Specified Performance 87
http://dx.doi.org/10.5772/intechopen.76069
K4 eD4 s K6 eD6 s
GSOSPDT ðsÞ ¼ ; GSOSPDT ðsÞ ¼ , (12)
ðT 4 s þ 1ÞðT 5 s þ 1Þ T 26 s2 þ 2ξ6 T 6 s þ 1
where for SOSPDT model (12b) the relative damping ξ6∈(0;1) indicates oscillatory step
response.
If ξ4 > 1, SOSPDT model (12a) is used; its parameters are found from the nonoscillatory step
response in Figure 10a using the following relations
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 t0:33 t0:7 ðt0:33 t0:7 Þ S
T 4:5 ¼ C2 C22 4C21 ; D4 ¼ ; C1 ¼ ; C2 ¼ , (13)
2 0:516 1:067 1:529 yð∞Þ
where S = K4(T4 + T5 + D4) is the area above the step response of the process output y(t), and
y(∞) is its steady-state value.
Parameters of the SOSPDT model (12b) can be found from evaluation of 2–4 periods of step
response oscillations (Figure 10b) using following rules [39]
qffiffiffiffiffiffiffiffiffiffiffiffiffi " #
ln aiþ1 ai
1 ζ26 1 XN
Nþ1
ξ6 ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; T 6 ¼ ðtNþ1 t1 Þ; D6 ¼ ti ðtNþ1 t1 Þ : (14)
π2 þ ln 2 aiþ1 πN N i¼1 2
ai
Figure 10. Step response of (a) nonoscillatory, (b) oscillatory SOSPDT model.
88 PID Control for Industrial Processes
34. Visioli, 2001, Regulator tuning [36] 1.37ν1/K1 2.42T1ν11.18 0.60T1 — Minimum ISE
35. Visioli, 2001, Regulator tuning [36] 1.37ν1/K1 4.12T1ν10.90 0.55T1 — Minimum ISTE
36. Visioli, 2001, Regulator tuning [36] 1.70ν1/K1 4.52T1ν11.13 0.50T1 — Minimum IST2E
37. Chandrashekar et al., 2002 [15] 10.3662/K1 0.3874T1 0.0435T1 0.0134T1 ts = 0.1 T1: ν1 = 0.1
38. Chandrashekar et al., 2002 [15] 2.0217/K1 4.65T1 0.2366T1 0.0696T1 ts = 0.8 T1: ν1 = 0.5
method No. 33 in Table 5 gain margin GM = 1.96, phase margin ϕM = 44.1 and maximum peak
Ms = 1.5 of the sensitivity to disturbance d(t)).
Minimization of performance indices can be applied also for unstable FOLPDT models
K1 eD1 s
GFOLPDT_US ðsÞ ¼ (15)
T1s 1
leading to simple tuning rules for PID controller (1a) (No. 34–38 in Table 6). Tuning rules No.
37 and 38 for PID controller (1c) show that settling time ts increases with growing normalized
time delay ν1 = D1/T1 of the FOLPDT model (15).
The main benefit of these methods consists of that all tuning rules are based on a single tuning
parameter that enables to systematically affect the closed-loop performance by step response
shaping [32].
Advanced Methods of PID Controller Tuning for Specified Performance 89
http://dx.doi.org/10.5772/intechopen.76069
39. Hang and Åström, 1988, Nonmodel [13] Kc sin φM T c ð1 cos φM Þ T c ð1 cos φM Þ
π sin φM 4π sin φM
Kc cos φM
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
41. Wojsznis et al., 1999, FOPDT [45]
GM
Tc
π tg φM þ 1 þ tg 2 φM Tc
4π tg φM þ 1 þ tg 2 φM
T 1 þ0, 5D1
42. Morari and Zafiriou, 1989, FOPDT [22]
K1 ðλþD1 Þ
T 1 þ 12 D1 T 1 D1
2T 1 þD1
Table 7. PID design for specified performance based on tuning parameters ϕM, GM, Mt and λ.
valid for second-order closed-loop with relative damping ωa*∈(0.25;0.65) where ωa* is the gain
crossover frequency [14]. Relations
1:18Mt jT ð0Þj 3
ηmax ≤ 100 ½%; ts ≈ forMt ∈ ð1:3; 1:5Þ (17)
jT ð0Þj ω∗a
are general for any order of the closed-loop T(s); if the controller has the integral part then
|T(0)| = |T(ω = 0)| = 1 [14].
The engineering practice is persistently demanding for PID controller design methods that
simultaneously guarantee several performance criteria [24], especially the maximum
90 PID Control for Industrial Processes
overshoot ηmax and the settling time ts. However, we ask the question: how to suitably
transform the abovementioned engineering requirements into frequency-domain specifica-
tions applicable for PID controller coefficients tuning? The response can be found in Section 3
in which a novel original PID controller design method is presented.
The proposed original method [6] enables to guarantee required closed-loop performance for a
whole family of plants specified by the uncertainty description. The core of it is the recently
developed PID controller design method based on external harmonic excitation [7]—a two-
step PID tuning method for performance specified in terms of maximum overshoot ηmax and
settling time ts.
In the first step, the plant is identified using external harmonic excitation signal (a sinusoid)
with the frequency ωn. In the second step, two developed PID controller design approaches
can be applied:
1. the approach based on guaranteed phase margin ϕM suitable for nonintegrating systems
with/without time delay and for integrating systems as well;
2. the approach based on guaranteed gain margin GM suitable for nonintegrating systems
with unstable zero.
For the ϕM–based approach, the specified performance is achieved by means of developed
quadratic dependences ηmax = f(ϕM,ωn) and ts = f(ϕM,ωn) parameterized by ωn; the
corresponding plots are called B-parabolas. For the GM–based approach similar quadratic
dependences for both the maximum overshoot ηmax = f(GM,ωn) and the settling time ts =
f(GM,ωn) were constructed. These approaches enable to achieve fulfillment of the following
performance measures (ωc is the plant ultimate frequency):
• for plants without integration behavior: ηmax∈〈0%, 90%〉 and ts∈〈6.5/ωc, 45/ωc〉,
• for plants with integration behavior: ηmax∈〈9.5%, 90%〉 and ts∈〈11.5/ωc, 45/ωc〉,
• for plants with unstable zero: ηmax∈〈0%, 90%〉 and ts∈〈8.5/ωc, 45/ωc〉.
A setup for the proposed harmonic excitation based method [7] is in Figure 11, where G(s) is a
transfer function of the controlled plant with unknown mathematical model, GR(s) is a PID
controller transfer function, and SW is a switch.
A sinusoidal excitation signal u(t) = Unsin(ωnt) is injected into the plant G(s) when the switch is
in the position SW = 4. The plant output y(t) is sinusoidal as well with the same frequency ωn,
magnitude Yn and a phase lag φ, that is y(t) = Ynsin(ωnt + φ), where φ = argG(ωn) (Figure 12).
Advanced Methods of PID Controller Tuning for Specified Performance 91
http://dx.doi.org/10.5772/intechopen.76069
Figure 11. A setup for implementation of the external harmonic excitation based method.
Figure 12. Time responses of (a) u(t); (b) y(t), and (c) location of G(jωn) in the complex plane.
After obtaining Yn and φ from the recorded time responses u(t) and y(t), one point of the
(unknown) plant frequency characteristics related with the excitation frequency ωn can be
plotted in the complex plane (Figure 12)
Y n ðωn Þ jφðωn Þ
Gðjωn Þ ¼ jGðjωn Þjejarg Gðωn Þ ¼ e : (18)
U n ð ωn Þ
Based on identified plant parameters, PID controller can be tuned using the phase margin and/
or gain margin approaches. In the control loop in Figure 11, switch SW in ”5” and the PID
controller in manual mode. To guarantee a specified phase margin ϕM at the gain crossover
frequency ωa*, the closed-loop characteristic equation under a PID controller 1 + L(jω) =
1 + G(jω)GR(jω) = 0 can be easily broken down into the magnitude and phase conditions (ωa* =
ωn, and ϕM is the required phase margin, L(jω) is the loop transfer function)
jGðjωn ÞjjGR ðjωn Þj ¼ 1, arg Gðωn Þ þ arg GR ðωn Þ ¼ 180 þ φM : (19)
92 PID Control for Industrial Processes
To guarantee a specified gain margin GM at the phase crossover frequency ωp*, the closed-loop
characteristic equation can be expressed by the magnitude and phase conditions [6] as follows
(ωp* = ωn)
jGðjωn ÞjjGR ðjωn Þj ¼ 1=GM , arg Gðωn Þ þ arg GR ðωn Þ ¼ 180 : (20)
Graphical interpretation of (19), (20) is shown in Figure 13. Let us denote φ ¼ argGðωn Þ,
Θ ¼ argGR ðωn Þ, and consider the ideal PID controller (1a), where K is proportional gain, and
Ti, Td are the integral and the derivative time constants, respectively. Substituting for s = jωn
into (1a) we obtain
1
GR ðjωn Þ ¼ K þ jK T d ωn : (21)
T i ωn
GR ðjωn Þ ¼ jGR ðjωn ÞjejΘ ¼ jGR ðjωn Þj½ cos Θ þ j sin Θ (22)
yields a complex Eq. (23) for phase margin approach and (24) for gain margin approach
1 cos Θ sin Θ
K þ jK T d ωn ¼ þj , (23)
T i ωn jGðjωn Þj jGðjωn Þj
1 cos Θ sin Θ
K þ jK T d ωn ¼ þj : (24)
T i ωn GM jGðjωn Þj GM jGðjωn Þj
Finally, PID controller parameters are obtained from (23), (24) using the substitution |GR(jωn)|
= 1/|G(jωn)| for the phase margin approach and |GR(jωn)| = 1/[GM|G(jωn)|] for the gain
Figure 13. Graphical interpretation of (a) ϕM, ωa* and shifting G into LA at ωa* = ωn; (b) GM, ωf* and shifting G into LF at
ωf* = ωn.
Advanced Methods of PID Controller Tuning for Specified Performance 93
http://dx.doi.org/10.5772/intechopen.76069
margin approach resulting from (24). The complex equations (23), (24) are solved as a set of
two real equations (25) for the phase margin or (26) for the gain margin approaches, respec-
tively
cos Θ 1 sin Θ
K¼ , K T d ωn ¼ , (25)
jGðjωn Þj βT d ωn jGðjωn Þj
cos Θ 1 sin Θ
K¼ , K T d ωn ¼ , (26)
GM jGðjωn Þj βT d ωn GM jGðjωn Þj
where (25a, 26a) represent general rules for calculating the controller gain K. After substituting
(25a), (26a) and the ratio β = Ti/Td into (25b), (26b), after some manipulations we obtain a
quadratic equation in Td for both approaches
Hence, the PID controller parameters are calculated using the expressions (25a), (26a), Ti = βTd
and (28); Θ is obtained from (19b) using (29) for the phase margin approach and (20b) for the
gain margin approach
Θ ¼ 180 þ ϕM arg Gðωn Þ ¼ 180 þ ϕM φ, (29)
Θ ¼ 180 arg Gðωn Þ ¼ 180 φ: (30)
Using the PID controller designed for the phase margin ϕM, the identified point G of the plant
Nyquist plot G(jω) with co-ordinates (1) is moved into the point LA of the open-loop Nyquist
plot located on the unit circle M1 (Figure 13a). In this way, the gain crossover LA of the open-
loop L(jω) is specified
LA Lðjωn Þ ¼ ½jLðjωn Þj; arg Lðωn Þ ¼ 1; ϕM , (31)
for which the designed PID controller guarantees the required phase margin ϕM; so for ωn is
|L(jωn)| = 1. In case of PID controller design for gain margin GM, the identified point G of the
plant Nyquist plot G(jω) with co-ordinates (1) is moved into the open-loop frequency response
point LF lying on the negative real half-axis of the complex plane (Figure 13b). In this way, the
phase crossover LF of the open-loop L(jω) is specified
1
LF L jω∗f ωn ¼ ½jLðjωn Þj; arg Lðωn Þ ¼ ; 180 : (32)
GM
Location of the points G(jωn) and L(jωn) in the complex plane is shown in Figure 13.
94 PID Control for Industrial Processes
Table 8. PI, PD and PID controller tuning rules using the harmonic excitation method.
PI, PD and PID tuning formulas for both approaches (ϕM and GM) are summed up in Table 8.
The excitation frequency can be adjusted according to the empirical relations [6].
The argument Θ in the tuning rules in Table 8 indicates the angle to be contributed to the
identified phase φ at ωn by the controller to obtain the resulting open-loop phase (180 + ϕM)
necessary to guarantee the required phase margin ϕM (or the gain margin GM). Working range
of the PID controller argument is given by the union of PI and PD controllers phase ranges
ΘPID ∈ ΘPI ∪ΘPD ¼ 90 ; 0 ∪ 0 ; þ90 ¼ 90 ; þ90 , (34)
which is symmetric with respect to 0 and due to frequency properties of PI, PD and PID
controllers also upper- and lower-bounded. The working range (34) can be interpreted using a
pretended transparent triangular ruler turned according to Figure 14; its segments to the left
and right of the axis of symmetry represent the PD and PI working ranges, respectively.
Figure 14a shows the situation, when the identified point G is situated in the 1st quadrant of
the complex plane. In case of phase-margin approach, put this ruler on Figure 14a, the middle
of the hypotenuse on the origin of the complex plane and turn it so that its axis of symmetry
merges with the ray (0,G). Thus, the ruler determines in the complex plane the cross-hatched
area representing the full working range of the PID controller argument. The controller type is
chosen depending on the situation of the ray (0,LA) forming with the negative real half-axis the
angle ϕM: situation of the ray (0,LA) in the left-hand sector suggests a PD controller, and in the
Advanced Methods of PID Controller Tuning for Specified Performance 95
http://dx.doi.org/10.5772/intechopen.76069
Figure 14. Controller structure selection using the “triangle ruler“ rule with respect to the situation of (a) G and LA; (b) G
and LF.
right-hand sector the PI controller. Figure 14a shows the case, when the phase margin ϕM is
achievable using both PI or PID controller. According to Figure 14b, the identified point G is
placed in 2nd quadrant of the complex plane. Applying the gain-margin approach, the ruler is
to be put on Figure 14b according to the similar setup than in case of phase-margin approach.
The controller type is chosen depending on the situation of the ray (0,LF) lying on the border of
the second and third quadrants of the complex plane; in this case, PI or PID controller type has
to be chosen.
This subsection answers the following question: how to transform the practical performance
requirements in terms of maximum overshoot ηmax and settling time ts into the couple of
frequency-domain parameters (ωn,ϕM) needed for identification and PID controller coeffi-
cients tuning?
j = 1…8. Let us split (33a) into 5 equidistant sections Δωn = 0.15ωc, and generate the set of
excitation frequencies
96 PID Control for Industrial Processes
k = 1…6. Its elements divided by the plant critical frequency ωc determine the set of so-called
excitation levels
k = 1…6. Let us demonstrate the qualitative effect of ωnk and ϕMj on closed-loop step response
for the plant
1
G 3 ðsÞ ¼ (38)
ðs þ 1Þð0:5s þ 1Þð0:25s þ 1Þð0:125s þ 1Þ
under PID controllers designed for three phase margins ϕM = 40 , 60 , 80 on three excitation
levels σ1 = ωn1/ωc = 0.2; σ3 = ωn3/ωc = 0.5 and σ5 = ωn5/ωc = 0.8. Related closed-loop step
responses are shown in Figure 15.
Achieving required ts and ηmax was tested by designing PID controller for a vast set of
benchmark examples [2] at excitation frequencies and phase margins expressed by a Cartesian
product ϕMjωnk of the sets (35) and (36) for j = 1…8, k = 1…6. Resulting dependences ηmax =
f(ϕM,ωn) and ts = (ϕM,ωn) are plotted in Figure 16 [6].
Considering the frequencies ωa* = ωn are equal which results from the assumptions of the
sinusoidal excitation method, the settling time can be expressed by the relation
γπ
ts ¼ (39)
ωn
similar to (16c) [6], where γ is the curve factor of the step response; in the relation (16c) for the
2nd order closed-loop, γ is from the interval (1;4) and depends on the relative damping [14]. In
case of the proposed sinusoid excitation based method γ varies over a considerably broader
interval (0.5;16) found empirically and depends strongly on ϕM, that is γ = f(ϕM) at the given
excitation frequency ωn. To examine closed-loop settling times for plants with different
dynamics, it is advantageous to define the relative settling time [7]
Figure 15. Closed-loop step responses of G3(s) for various ϕM and ωn.
Advanced Methods of PID Controller Tuning for Specified Performance 97
http://dx.doi.org/10.5772/intechopen.76069
τ s ¼ t s ωc : (40)
Substituting for ωn = σωc into (39) and (40), we obtain a relation for the relative settling time
π
τs ¼ γ, (41)
σ
where ts is related to the critical frequency ωc. Due to introducing ωc the right-hand side in (41)
is constant for the given plant and independent of ωn. The dependency (41) obtained empiri-
cally for different excitation frequencies ωnk is depicted in Figure 16b; it is evident that at every
excitation level σk with increasing phase margin ϕM the relative settling time τs first decreases
and after achieving its minimum τs_min it increases again. The empirical dependences in
Figure 16 have been approximated by quadratic regression curves, thus they are called B-
parabolas [7].
3.4.1.1. Discussion
When choosing ϕM = 40 on the B-parabola corresponding to the excitation level σ5 = ωn5/
ωc = 0.8 (further denoted as B0.8 parabola), maximum overshoot ηmax = 40% and relative
settling time τs≈10 are expected (see Figure 16). Point corresponding to these parameters
and is located on the left (falling) portion of B0.8 yielding oscillatory step response (see
response in Figure 15c). If the phase margin to ϕM = 60 increases, the relative settling time
decreases into the point on the right (rising) portion of the B0.8 parabola; the corresponding
step response in Figure 15c is weakly aperiodic. For the phase margin ϕM = 80 , the B0.8
parabola indicates a zero maximum overshoot, the relative settling time τs = 20 corresponds to
the position on the B0.8 parabola with aperiodic step response (Figure 15c). If the maxi-
mum overshoot ηmax = 20% is acceptable, then ϕM = 53 yields the least possible relative
settling time τs = 6.5 on the given level σ5 = 0.8 (“at the bottom” of B0.8).
Figure 16. Dependences: (a) ηmax = f(ϕM,ωn); (b) τs = ωcts = f(ϕM,ωn) for ϕMjωnk, j = 1…8, k = 1…6 (relative settling time
τs = tsωc).
98 PID Control for Industrial Processes
3.4.1.2. Example 1
Using the sinusoid excitation method, design ideal PID controllers (1a) for an operating
amplifier modeled by the transfer function GA(s)
1 1
GA ðsÞ ¼ 3
¼ : (42)
ðT A s þ 1 Þ ð0; 01s þ 1Þ3
The control objective is to guarantee maximum overshoots ηmax1 = 30%, ηmax2 = 5% and a
maximum relative settling time τs = 12 in both cases.
3.4.1.3. Solution
1. Critical frequency of the plant identified by the Rotach test is ωc = 173.216[rad/s] (the
process is “fast”). The prescribed settling time is ts = τs/ωc = 12/173.216[s] = 69.3[ms].
2. For the expected performance (ηmax1;τs) = (30%;12) (Design No. 1) a satisfactory choice is
(ϕM1;ωn1) = (50 ;0.5ωc) resulting from the B0.5 parabola in Figure 16. The performance in
terms of (ηmax2;τs) = (5%;12) (Design No. 2) can be achieved by choosing (ϕM2;ωn2) =
(70 ;0.8ωc) resulting from the B0.8 parabola in Figure 16.
3. Identified points for the first and second designs are GA(j0.5ωc) = 0.43ej120 and
GA(j0.8ωc) = 0.19ej165 , respectively. According to Figure 17a, both points are located in
the quadrant II of the complex plane, on the Nyquist plot GA(jω) (continuous curve) which
verifies the identification.
Figure 17. (a) Open-loop Nyquist plots; (b) closed-loop step responses of the operational amplifier, required performance
ηmax1 = 30%, ηmax2 = 5% and τs = 12.
Advanced Methods of PID Controller Tuning for Specified Performance 99
http://dx.doi.org/10.5772/intechopen.76069
4. Using the PID controller designed for (ϕM1;ωn1) = (50 ;0.5ωc), the point GA(j0.5ωc) is
moved into the gain crossover LA1(j0.5ωc) = 1ej130 on the unit circle M1, which verifies
achieving the phase margin ϕM1 = 180 130 = 50 (dashed Nyquist plot). The point
GA(j0.8ωc) has been moved by the PID controller designed for (ϕM2;ωn2) = (80 ;0.8ωc) into
LA2(j0.8ωc) = 1ej110 yielding a phase margin ϕM2 = 180 110 = 70 (dotted Nyquist plot).
5. Achieved performance read-off from the closed-loop step response in Figure 17b (dashed
line) is ηmax1* = 29.7%, ts1* = 58.4[ms]. Performance in terms of ηmax2* = 4.89%, ts2* = 60.5
[ms] identified from the closed-loop step response in Figure 17b (dashed line) complies
with the required performance.
φ ¼ φ0 þ φD (44)
The only modification in using the PID tuning rules in Table 9 is that an increased required
phase margin is to be specified.
Model ηmax,τs ωc [rad/s] ts [s] B-par. ϕM/GM ωn/ωc G(jωn) GR(jωn) ηmax* ts* [s]
GA(s) 30%,12 173.2 0.069 Figure 16 50 0.5 0.43ej120 2.31ej10 29.7% 0.058
j165 j55
GA(s) 5%,12 173.2 0.069 Figure 16 70 0.8 0.19e 5.20e 4.89% 0.061
GB(s) 30%,12 0.352 34.1 Figure 16 55 + 45.9 0.35 1.03ej23 0.97ej56 18.6% 24.78
j13 j71
GB(s) 5%,12 0.352 34.1 Figure 16 70 + 26.2 0.2 1.09e 0.92e 0.15% 28.69
GC(s) 30%,20 0.241 83.1 Figure 21 53 + 10.1 0.35 12.7ej122 0.08ej5.8 29.6% 81.73
j129 j28
GC(s) 20%,20 0.241 83.1 Figure 21 62 + 14.5 0.5 8.10e 0.12e 19.7% 82.44
GD(s) 30%,12 0.049 245.9 Figure 25 15 dB 1.25 0.14ej204 1.47ej24 24.5% 241.9
j136 j44
GD(s) 5%,12 0.049 245.9 Figure 25 18 dB 0.65 0.38e 0.38e 4.55% 243.4
Table 9. PID controller design parameters, required and achieved performance, identified plant parameters for GA(s),
GB(s), GC(s) and GD(s).
100 PID Control for Industrial Processes
ϕ0M ¼ ϕM þ ωn D (46)
The phase delay ωnD increases with increasing frequency ωn of the sinusoidal excitation
signal. It is recommended to use the smallest possible added phase φD = ωnD to lessen the
impact of time delay on closed-loop dynamics.
3.4.2.1. Discussion
The time delay D can be easily specified during identification of the critical frequency as a time
D = TyTu, that elapses since the start of the test at time Tu until time Ty, when the system
output starts responding to the excitation signal u(t). A small added phase φD = ωnD due to
time delay can be achieved by choosing the smallest possible ωn attenuating the effect of D in
(47) and subsequently in the PID controller design. Therefore when designing a PID controller
for time delayed systems, it is recommended to choose the lowest possible excitation level
when using B-parabolas (most frequently ωn/ωc = 0.2 resp. 0.35) and corresponding couples of
B-parabolas in Figure 16. From the given couple (ηmax;ts), ϕM is specified using the chosen
couple of B-parabolas, however its increased value ϕM0 given by (46) is to be supplied in the
design algorithm thus minimizing effect of the time delay on closed-loop dynamics.
3.4.2.2. Example 2
Using the sinusoid excitation method, design ideal PID controllers (1a) for a distillation col-
umn model given by the transfer function GB(s)
KB eDB s 1:11e6, 5s
GB ðsÞ ¼ ¼ : (48)
T B s þ 1 3:25s þ 1
2. Because DB/TB = 2 > 1, the plant is a so-called “dead-time dominant system.“ Due to a large
time delay, it is necessary to choose the lowest possible excitation frequency ωn to mini-
mize the added phase ωnDB in (47). Hence, for the required performance (ηmax2;τs) =
(5%;12) (Design No. 2) we choose the B0.2 parabolas in Figure 16 at the lowest possible
level ωn/ωc = 0.2 to find (ϕM2;ωn2) = (70 ;0.2ωc). The added phase value is ωn2DB =
0.2ωcDB = 0.20.35216.5180/π = 26.2 , hence the phase supplied to the PID design algorithm
Advanced Methods of PID Controller Tuning for Specified Performance 101
http://dx.doi.org/10.5772/intechopen.76069
́
is ϕM2 = ϕM2 + ωn2DB = 70 + 26.2 = 96.2 (instead of ϕM2 = 70 for a delay-free system). The
required performance (ηmax1;τs) = (30%;12) (Design No. 1) can be achieved by choosing
(ϕM1;ωn1) = (55 ;0.35ωc) from the B0.35 parabolas in Figure 16 (i.e., ωn/ωc = 0.35). The phase
́
margin ϕM1 = 55 + 45.9 supplied into the design algorithm was increased by ωn1DB =
0.35ωcDB = 0.350.35216.5180/π = 45.9 compared with ϕM1 = 55 in case of delay-free system.
3. Figure 18a shows that identified points GB(j0.35ωc) = 1.03ej23 and GB(j0.2ωc) = 1.09ej13
are located in the quadrant I of the complex plane at the beginning of the frequency
response GB(jω) (continuous curve).
4. The point GB(j0.2ωc) (Design No. 2) was shifted by the PID controllers to the open-loop
amplitude crossover LB2(j0.2ωc) = 1ej110 (dotted Nyquist plot in Figure 18a). Note that
LB2 has the same position in the complex plane as LA2 in Figure 17a, however at a
considerably lower frequency ωn2B = 0.20.3521 = 0.07[rad/s] compared to ωn2A = 0.8173.216
= 138.6[rad/s] (ts2_B* = 28.69[s] is almost 500 times larger than ts2_A* = 0.0584[s] which
demonstrates the key role of ωn in achieving required closed-loop dynamics). The identi-
fied point GB(j0.35ωc) (Design No. 1) was moved by the designed PID controller into the
amplitude crossover LB1(j0.35ωc) = 1ej125 (dashed Nyquist plot in Figure 18a).
5. Achieved performances (ηmax1* = 18.6%, ts1* = 24.78[s], dashed line), (ηmax2* = 0.15%,
ts2* = 28.69[s], dotted line) in terms of the closed-loop step responses in Figure 18b comply
with the required performance specification.
Figure 18. (a) Open-loop Nyquist plots; (b) closed-loop step responses of the distillation column, required performance
ηmax1 = 30%, ηmax2 = 5% and τs = 12.
102 PID Control for Industrial Processes
Figure 19. B-parabolas: (a) ηmax = f(ϕM,ωn); (b) τs = ωcts = f(ϕM,ωn) for systems with integrator β = 4.
Figure 20. B-parabolas: (a) ηmax = f(ϕM,ωn); (b) τs = ωcts = f(ϕM,ωn) for systems with integrator β = 8.
ϕMjωnk of the sets (35) and (36), j = 1…8, k = 1…6 and three various ratios Ti/Td: β = 4, 8
and 12).
3.4.3.1. Discussion
Inspection of Figures 19a, 20a and 21a reveals that increasing β results in decreasing of the
maximum overshoot ηmax, narrowing of the B-parabolas of relative settling times τs = f(ϕM,ωn)
for each identification level ωn/ωc and consequently increasing the settling time.
Consider for example, the B0.95 parabolas in Figures 19b, 20b and 21b: if ϕM = 70 and β = 4 the
relative settling time is τs = 30, for β = 8 it grows up to τs = 40, and for β = 12 even to τs = 45. If a
10% maximum overshoot is acceptable for the given system with integrator, then the standard
Advanced Methods of PID Controller Tuning for Specified Performance 103
http://dx.doi.org/10.5772/intechopen.76069
Figure 21. B-parabolas: (a) ηmax = f(ϕM,ωn); (b) τs = ωcts = f(ϕM,ωn) for systems with integrator β = 12.
interaction PID controller can be used with no need to use the setpoint filter; however a larger
settling time is expected.
3.4.3.2. Example 3
Using the sinusoidal excitation method, let us design ideal PID controllers for a flow valve
modeled by the transfer function GC(s) (system with an integrator and a time delay)
KC eDC s 1:3e2:1s
G C ðsÞ ¼ ¼ : (49)
sðT C s þ 1Þ sð7:51s þ 1Þ
The control objective is to guarantee a maximum overshoot of the closed-loop step response
ηmax1 = 30%, ηmax2 = 20% and a maximum relative settling time τs = 20.
1. Critical frequency of the plant identified by the Rotach test is ωc = 0.2407[rad/s]. Then, the
required settling time is ts = τs/ωc = 20/0.2407[s] = 83.09[s].
2. For GC(s) the time delay/time constant ratio is DC/TC = 2.1/7.51 = 0.28 < 1, hence, the
influence of the time constant prevails—GC(s) is a so-called “lag-dominant system” with
integrator, therefore B-parabolas are to be chosen carefully. From one side, due to time
delay it would be desirable to choose B-parabolas from Figures 19, 20 or 21 with the lowest
identification level ωn/ωc = 0.2. However, the minima of B0.2 parabolas in Figure 19b (for β
= 4), Figure 20b (for β = 8) and Figure 21b (for β = 12) indicate that the smallest feasible
relative settling time τs = 36.5 (for β = 4), τs = 33 (for β = 8) and τs = 34 (for β = 12), which do
not satisfy the required value τs = 20.
104 PID Control for Industrial Processes
3. The first performance specification (ηmax1;τs) = (30%;20) can be provided using the B0.35
parabolas for β = 12 (Figure 21b) at the level ωn/ωc = 0.35 and for parameters
(ϕM1;ωn1) = (53 ;0.35ωc) (Design No. 1), supplying the augmented open-loop phase margin
ϕ´M1 = ϕM1 + ωn1DC = 53 + 10.1 = 63.1 into the PID controller design algorithm. The
second performance specification (ηmax2;τs) = (20%,20) can be achieved using the B0.5
parabolas in Figure 21 for β = 12 and ωn/ωc = 0.5 and parameters (ϕM2;ωn2) = (62 ;0.5ωc)
(Design No. 2). To reject the influence of DC, instead of ϕM2 = 62 the augmented open-loop
phase margin ϕ´M2 = ϕM2 + ωn1DC = 62 + 14.5 = 76.5 was supplied into the PID controller
design algorithm.
4. Identified points GC(j0.35ωc) = 12.7ej122 and GC(j0.5ωc) = 8.10ej129 are located on the
plant frequency response GC(jω) (continuous curve) in Figure 22a verifying correctness of
the identification.
5. Using the PID controller, the first identified point GC(j0.35ωc) (Design No. 1) was moved
into the gain crossover LC1(j0.35ωc) = 1ej127 located on the unit circle M1; this verifies
achieving the phase margin ϕM1 = 180 127 = 53 (dashed Nyquist plot in Figure 22a).
Figure 22. (a) Open-loop Nyquist plots; (b) closed-loop step responses of the flow valve, required performance
ηmax1 = 30%, ηmax2 = 20% and τs = 20.
Advanced Methods of PID Controller Tuning for Specified Performance 105
http://dx.doi.org/10.5772/intechopen.76069
for j = 1…8. Let us split (33b) into five equal sections and generate the set of excitation
frequencies
for k = 1…6. Its elements divided by the plant critical frequency ωc determine excitation levels
fσk ¼ ωnk =ωc g ) fσk g ¼ f0:5; 0:65; 0:8; 0:95; 1:1; 1:25g (52)
for k = 1…6. Figure 23 shows closed-loop step response shaping using different GM and ωn in
the PID tuning for the plant (53b) with parameters T3 = 0.75, α3 = 1.3, for four required gain
margins GM = 5 dB, 9 dB, 11 dB and 13 dB, and three different excitation levels σ1 = ωn1/ωc = 0.5,
σ3 = ωn3/ωc = 0.8 and σ5 = ωn5/ωc = 1.1.
α2 s þ 1 α3 s þ 1
G 2 ðsÞ ¼ , G 3 ðsÞ ¼ : (53)
ðT 2 s þ 1Þn2 ðs þ 1ÞðT 3 s þ 1Þ T 23 s þ 1 T 33 s þ 1
The proposed method has been applied for each element of the Cartesian product ωnk GMj of
the sets (51) and (50). Significant differences between dynamics of individual control loops
under designed PID controllers can be observed for the benchmark systems (53).
Consider the benchmark plants G2(s) and G3(s) with following parameters: G2.1(s): (T2,
n2,α2) = (0.75,8,0.2); G2.2(s): (1,3,0.1); G2.3(s): (0.5,5,1); G3(s): T3 = 0.5, α3 = 1.3.
Couples of examined plants [G3(s), G2.3(s)] and [G2.2(s), G2.1(s)] differ principally by the ratio α/
T, which is significant for the closed-loop performance assessment for plants with an unstable
zero (for the 1st couple [α3/T3 = 2.6, α2.3/T2.3 = 2], for the 2nd couple [α2.2/T2.2 = 0.1, α2.1/
T2.1 = 0.27]).
Figure 23. Closed-loop step responses of the plant G3(s) under PID controllers designed for various GM and ωn.
106 PID Control for Industrial Processes
According to the ratio α/T unknown plants with an unstable zero can be classified in following
two groups [7]:
1. plants with α/T < 0.3;
3.4.4.1. Example 4
Using the sinusoid excitation method, ideal PID controllers are to be designed for a heating
plant described by the transfer function GD(s) (a system with an unstable zero)
KD ðT z s þ 1Þ 0:8ð7:5s þ 1Þ
G D ðsÞ ¼ ¼ : (54)
ðT D s þ 1Þ3 ð27:5s þ 1Þ3
The control objective is to guarantee a maximum overshoot ηmax1 = 30%, ηmax2 = 5% and
maximum relative settling time τs = 12.
1. Critical frequency of the plant identified by the Rotach test is ωc = 0.0467[rad/s], the system
is ”slow“. The required settling time is ts = τs/ωc = 12/0.0488 = 245.90[s].
2. Because α/TD = 7.5/27.5 = 0.27 < 0.3, the gain margin GM and the excitation frequency ωn of
the controlled object GD(s) will be determined using B-parabolas in Figure 25. For the
required performance (ηmax1,τs) = (30%,12) the appropriate values of gain margin and
Figure 24. B-parabolas: (a) ηmax = f(GM,ωn); (b) τs = ωcts = f(GM,ωn) for nonminimum phase systems, α/T > 0.3.
Advanced Methods of PID Controller Tuning for Specified Performance 107
http://dx.doi.org/10.5772/intechopen.76069
Figure 25. B-parabolas: (a) ηmax = f(GM,ωn); (b) τs = ωcts = f(GM,ωn) for nonminimum phase systems, α/T < 0.3.
excitation frequency are (GM1,ωn1) = (15 dB,1.25ωc), that is “gray parabolas” in Figure 25.
Similarly, the performance (ηmax2,τs) = (5%,12) can be achieved by choosing
(GM2,ωn2) = (18 dB,0.65ωc) according to “violet” B-parabolas in Figure 25.
3. Examination of the Nyquist plots of the controlled object GD(jω) and the open-loops
LD1(jω), LD2(jω) in Figure 26a reveals that the first identified point GD(j1.25ωc) is located
in the quadrant III of the complex plane, and its identification is carried out under a
relatively low frequency 1.25ωc = 1.250.0467 = 0.0584[rad/s], hence no high-frequency
Figure 26. (a) Open-loop Nyquist plots; (b) closed-loop step responses of the heating system, required performance
ηmax1 = 30%, ηmax2 = 5% and τs = 12.
108 PID Control for Industrial Processes
noise corrupting the excitation and output signals u(t) and y(t), respectively, is expected
during identification. If, however, the identification at the excitation level ωn = 1.25ωc were
carried out for a “fast” object with a high value of ωc, it would be necessary to choose the
lowest possible excitation level in order to reject the identification noise. The second
identified point GD(j0.65ωc) is placed in the quadrant II of the complex plane.
4. Using the PID controller designed for (GM1,ωn1) = (15 dB,1.25ωc) the point GD(j1.25ωc) was
h i
compensated into the target point LD1 ðj1:25ωc Þ ¼ 1=10GM =20 ej180 located on the negative 1
real half-axis where the gain margin GM1 of the open-loop LD1(jω) (red Nyquist plot) is
satisfied. The achieved performance evaluated from the closed-loop step response in
Figure 26b is ηmax1* = 24.5%, ts1* = 241.88[s].
5. Using the PID controller designed for (GM2,ωn2) = (18 dB,0.65ωc), the point GD(j0.65ωc)
h i
was moved to the target point LD2 ðj0:65ωc Þ ¼ 1=10GM =20 ej180 where the gain margin GM2 of
2
the open-loop LD2(jω) (green Nyquist plot) is satisfied. The achieved performance
ηmax2* = 4.55%, ts2* = 243.42[s] evaluated from the closed-loop step response in Figure 26b
satisfies the control objective.
Time-domain performance requirements specified by the process technologist, identified plant
parameters needed for PID controller tuning (for two PID controllers of all four plants GA(s), GB(s),
GC(s) and GD(s)) along with specified and achieved performance measure values are summarized
in Table 9. The asterisk “*“ indicates closed-loop performance complying with the required one.
When identifying an uncertain plant, the sinusoidal excitation with the frequency ωn is
repeated for individual parameter changes to obtain a set of points Gi from the set of frequency
responses of the uncertain plant
i = 1,2…N. Plant parameter changes are reflected in changes of the magnitude and the phase |
Gi(jωn)| and argGi(ωn), respectively; i = 1…N; N = 2p is the number of identification experi-
ments and p is the number of varying technological quantities of the plant. The nominal model
G0(jωn) is obtained from mean values of the real and imaginary parts of Gi(jωn)
1X N
1X N
G0 ðjωn Þ ¼ a0 þ jb0 ¼ ai þ j bi , (56)
N i¼1 N i¼1
i = 1,2…N. Obviously
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
hX PN
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 N
i2 hXN i2 b0 bi
jG0 ðjωn Þj ¼ a20 þ b20 ¼ a i þ i¼1 i
b ; arg G0 ðjωn Þ ¼ arctg ¼ arctg Pi¼1
N
,
N i¼1 a0 i¼1 ai
(57)
Advanced Methods of PID Controller Tuning for Specified Performance 109
http://dx.doi.org/10.5772/intechopen.76069
where φ0(ωn) = arg{G0(jωn)}. The points Gi represent some elements of the family of plants and
can be enclosed by a circle MG centered in G0(jωn) with the radius RGRG(ωn) corresponding
to the maximum distance between Gi(jωn) and G0(jωn)
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
RG ¼ max ðai a0 Þ2 þ ðbi b0 Þ2 , (58)
i
i = 1,2…N. Actually, the control law generated by the robust controller GRrob(s) designed for
the nominal point G0(jωn) performs the mapping
of the set of identified points Gi(jωn) encircled by MG with the radius RG onto the set of points
Li(jωn) delineated by ML and calculates the radius RLRL(ωn) of the dispersion circle ML which
encloses the points Li(jωn) of the Nyquist plot so as to guarantee fulfillment of the robust
performance condition.
A robust PID controller is designed using the sinusoidal excitation method with input data for
the nominal model G0(jωn): {|G0(jωn)|; φ0 = argG0(ωn)}. Substituting them into (25a), (26a), (29)
and (30), the expressions for calculating robust PID controller parameters according to Table 8
are obtained. Obviously, the phase and gain margins ϕM and GM, respectively, are robust PID
controller tuning parameters and at the same time attractive robustness measures [6].
where ϕM and GM are the required phase and gain margins, respectively, ωn is the excitation
frequency, χL = RL+/RL and χS = RS+/RS are safety factors of radii of the dispersion circles ML
and MS, respectively, delineating prohibited areas; RG(ωn) is the radius of the dispersion circle
at the Nyquist plot of the plant at ωn, and G0(ωn) is a point at the Nyquist plot of the nominal
plant at ωn. The prohibited area MS can be defined in terms of ϕM or GM using the expressions
ϕS = arcsin(RS) or GS = 1/(1RS), respectively.
Proof:
The proof is straightforward using Figures 27 and 28. If the nominal open-loop L0(s) = G0(s)
GR(s) is stable, then according to the Nyquist stability criterion the closed-loop with the
uncertain plant will be stable if the distance between L0 and (1, j0), that is |1 + L0(jωn)| is
110 PID Control for Industrial Processes
greater than the sum of the radii RL(ωn) of the circle ML centered in L0, and RS(ωn) of the circle
MS centered in (1.0j), that is
From the principles of the sinusoidal excitation PID controller tuning method results that the
robust controller shifts the nominal point of the plant frequency response G0 to the point L0
situated on the unit circle. Thus, ωn becomes the gain crossover frequency. As the point L0 is
situated on the unit circle M1, the magnitude |L0(jωn)| equals one, that is |L0| = |G0||GR| = 1,
yielding the transformation ratio |GR| = |G0|1 between the radii RG and RL of the circles MG
and ML, respectively. The radius RL of the dispersion circle ML can be expressed as
RG
R L ¼ R G jG R j ¼ : (63)
jG0 j
Substituting (62) and (63) in (61) yields the robust performance condition
2
χL RG ðωn Þ
þ χS RS < 2 2 cos φM , (64)
jG0 ðjωn Þj
Figure 27. Dispersion circles MG and ML and the prohibited area delineated by the circle MS for the phase-margin
approach.
Advanced Methods of PID Controller Tuning for Specified Performance 111
http://dx.doi.org/10.5772/intechopen.76069
which after some manipulations is identical to the proven condition (60a). Typical values of
safety factors are χL = 1.1 and χS = 1.2. The value of ϕM chosen according to the robust
performance condition used in the tuning rules in Table 1 yields robust PID controller coeffi-
cients. The design procedure is illustrated in Section 3.5.1.3.
From the principles of the proposed PID tuning method results that the robust controller shifts
the point G0(ωn) of the plant nominal frequency response to L0 situated on the negative real
half-axis of the complex plane. From the relation |L0(jωn)| = |G0(jωn)||GR(jωn)| = 1/GM results
the ratio |GR(jωn)| = 1/[GM|G0(jωn)|] between the radii RG and RL = |GR|RG of the circles MG
and ML, respectively. The radius RL of the dispersion circle ML is calculated as follows
RG
RL ¼ RG jGR j ¼ : (66)
GM jG0 ðjωn Þj
Substituting (65) and (66) into the general robust performance condition (61) and considering
the safety factors χL and χS, the following inequality is obtained
GM 1 χL RG
> þ χS RS , (67)
GM GM jG0 ðjωn Þj
which after some manipulations is identical to the proven condition (60b). According to the
robust performance condition, the chosen value GM is substituted into (26a) and robust PID
controller parameters are obtained from Table 8. ϕS and GS are found from the B-parabolas
(Figures 16, 19–21, 24, 25) considering ηmaxN and τsN of the worst-case plant.
Figure 28. Dispersion circles MG and ML and the prohibited area delineated by the circle MS for gain-margin approach.
112 PID Control for Industrial Processes
3.5.1.3. Examples
3.5.1.3.1. Example 5
Consider the plant model GA(s) from Subsection 3.4
KA0 1
GA0 ðsÞ ¼ ¼ (68)
ðT A0 s þ 1Þ3 ð0:01s þ 1Þ3
to be the nominal model of an uncertain system where KA and TA are uncertain parameters
varying within 15% from their nominal values KA0 and TA0 (i.e., the total dispersion is κ =
30%). Let us design a robust PID controller to guarantee ηmaxN = 30% and a relative settling
time τsN = 12 for the worst-case model of GA(s).
3.5.1.3.2. Robust PID controller design for the uncertain plant GA(s)—solution and discussion
1. The measured ultimate frequency of the nominal model is ωc0 = 173.216[rad/s]. From the
robust performance condition results tsN = τsN/ωc = 12/173.216 = 69.3[ms].
2. For the required performance (ηmaxN,τsN) = (30%,12) the corresponding values of phase
margin and excitation frequency have been selected (ϕM,ωn0) = (50 ,0.5ωc0) using the pair
of ”red“ B-parabolas in Figure 16. As there are two uncertainties in GA(s) (KA and TA), the
number of identification experiments is N = 22 = 4.
3. For ωn0 = 0.5,ωc0 = 0.5173.21 = 86.61[rad/s], four points of the family of Nyquist plots
corresponding to the uncertain plant model were identified using the sinusoidal excitation:
GA1(jωn0), GA2(jωn0), GA3(jωn0) and GA4(jωn0) (blue ”x “in Figure 29a). The nominal point
GA0(jωn0) calculated from the coordinates of all identified points GAi(jωn0), i = 1, 2, 3, 4 is
Figure 29. (a) Nyquist plots for GA(s), ηmaxN = 30% and τsN = 12; (b) closed-loop step responses satisfy the required
performance ηmaxN = 30% and τsN = 12 (upper plot: worst-case plant model; lower plot: nominal plant model).
Advanced Methods of PID Controller Tuning for Specified Performance 113
http://dx.doi.org/10.5772/intechopen.76069
located on the “blue” Nyquist plot of the nominal model GA0(jωn). The dispersion circle
MG is centered in GA0(jωn0) with the radius RG = 0.199.
4. The desired robust performance ηmaxN = 30%, τsN = 12 can be achieved using ϕS = 50 at
the excitation level ωn0 = 0.5ωc0.
5. R.H.S. of the robust performance condition (60a) is δ0_RP = 89.39 , thus the robust perfor-
mance condition (60a) ϕM>δ0_RP will be satisfied if choosing for example ϕM = 90 .
6. Using the designed PID controller, the nominal point GA0(jωn0) is shifted to LA0(jωn0) =
GA0(jωn0)GR_rob(jωn0) = 1ej90 on the unit circle M1 (Figure 29a).
7. The smallest phase margin estimated from the location of the worst-case point LAN =
1.13ej114 is ϕMN = 66.2 . The achieved smallest phase margin ϕ+MN = 61 is given by the
intersection of the “red” Nyquist plot and the unit circle M1 (closest to the negative real
half-axis).
8. Radius of the prohibited area RS = sinϕS = sin(503.14/180) = 0.766/χS = 0.6383 multiplied by
the expansion coefficient χS = 1.2, as well as the χS = 1.1-times enlarged radius RL of the
dispersion circle ML guarantee that none of the open-loop Nyquist plots enters the
prohibited area delineated by the MS circle. The enlarged circles ML+ and MS+ (dotted plots
in Figure 29a) are touching, which indicates fulfillment of the robust performance condition.
9. From the closed-loop step response of the worst-case plant model (Figure 29b, red plot)
results ηmaxN_obtained = 8.2% and the relative settling time τsN_obtained = ωc0tsN_obtained =
173.2160.0671 = 11.62, which proves achievement of the specified performance. Using the
phase margin ϕM = 90 at the identification level ωn0 = 0.5ωc0 (“red” B-parabolas in
Figure 16a) corresponds to the nominal performance ηmax0 = 2% and τs0 = 10. The closed-
loop step response in Figure 29b (green plot) corresponding to the nominal model satisfies
ηmax0_obtained = 0% and τs0_obtained = 173.2160.0469 = 8.12 as expected.
3.5.1.3.3. Example 6
Consider the plant model from the Subsection 3.4
to be the nominal model of the uncertain plant GD(s) with parameters KD, TD and αD varying
within 15% from their nominal values KS0, TS0 and αD0 (the total dispersion is κ = 30%). A
robust PID controller is to be designed to guarantee specified performance in terms of a maxi-
mum overshoot ηmaxN = 5% and a relative settling time τsN = 12 for the worst-case model of GD(s).
3.5.1.3.4. Robust PID controller design for the uncertain plant GD(s)—solution and discussion
1. The measured ultimate frequency of the nominal model is ωc0 = 0.0488[rad/s]. From the
requirements on the nominal closed-loop performance results: ts = τs0/ωc = 12/0.0488 =
245.9[s].
114 PID Control for Industrial Processes
2. For the required performance (ηmaxN,τsN) = (5%,12) the corresponding values of gain
margin and excitation frequency have been selected (GM,ωn) = (18 dB,0.65ωc0) using the
pair of ”red“ B-parabolas in Figure 25. As there are three uncertain parameters in GD(s)
(KD, TD and αD), the number of identification experiments is N = 23 = 8.
3. For ωn0 = 0.65ωc0 = 0.650.04880 = 0.03172[rad/s], eight points of the family of Nyquist plots
corresponding to the uncertain plant model were identified using the sinusoidal excitation:
GD1(jωn)…GD8(jωn) (depicted by blue ”x“ in Figure 30). The nominal point GD0(jωn)
calculated from the coordinates of all identified points GDi(jωn), i = 1…8 is located on the
Nyquist plot of the nominal model GD0(jωn) (blue curve) thus proving correctness of the
identification. Radius of the dispersion circle MG centered in the nominal point GD0(jωn0)
with the radius RG = 0.164.
4. The desired robust performance ηmaxN = 30%, τsN = 12 can be achieved using ϕS = 50 at
the excitation level ωn0 = 0.5ωc0.
5. Using the designed robust PID controller, the nominal point GD0(jωn0) of the plant is
shifted to the point LD0(jωn0) = GD0(jωn0)GR_rob(jωn0) = 0.0841ej180 located on the unit
circle. The nominal open-loop Nyquist plot (green plot) crosses LD0(jωn0) (Figure 14), the
radius of the circle ML is RL = 0.0400.
6. The smallest gain margin G+MN = 18.8 dB is estimated from the position of the worst-case
point LDN(jωn0) = 0.112ej197 . The achieved smallest gain margin is given by the intersec-
tion point of the red Nyquist plot with the negative real half-axis of the complex plane
G+MN = 16.9 dB.
Figure 30. (a) Nyquist plots for GD(s), ηmaxN = 5% and τsN = 12; (b) closed-loop step responses with GD(s) for the required
performance ηmaxN = 5% and τsN = 12 (upper plot: worst-case plant model; lower plot: nominal plant model).
Advanced Methods of PID Controller Tuning for Specified Performance 115
http://dx.doi.org/10.5772/intechopen.76069
7. Both the radius of the prohibited area RS = (GS1)/GS = (1018/20–1)/1018/20 = 0.8741/χS = 0.699
multiplied by the expansion coefficient χS = 1.2, as well as the radius RL of the dispersion
circle ML enlarged χS = 1.1-times guarantee that none of the open-loop Nyquist plots enters
the prohibited area delineated by the MS circle. The enlarged circles ML+ a MS+ in Figure 30a
(dotted curves) touch, which indicates fulfillment of the robust performance condition.
8. As GM = 21.5 dB at the excitation level ωn0 = 0.65ωc0 has been considered, according to ”pink
“B-parabolas in Figure 25a nominal performance ηmax0 = 1.5% and τs0 = 21 is expected. The
nominal closed-loop step response in Figure 30b (green plot) shows the nominal perfor-
mance in terms of ηmax0_obtained = 0% and τs0_obtained = 0.0488381 = 18.59 as expected.
9. From the closed-loop step response of the worst-case plant model (Figure 30b, red plot)
results ηmaxN_obtained = 4.8% and the relative settling time τsN_obtained = ωc0tsN_obtained =
0.0488237 = 11.57 which proves achievement of the specified performance. Using the gain
margin GM = 21.5 dB at the excitation level ωn0 = 0.65ωc0 (“pink“ B-parabolas in Figure 25a)
indicates the expected nominal performance ηmax0 = 1.5% and τs0 = 21. The closed-loop
step response in Figure 30b (green plot) corresponding to the nominal model satisfies
ηmax0_obtained = 0% and τs0_obtained = 0.0488381 = 18.59 as expected.
4. Conclusion
A novel frequency-domain PID design method for performance specified in terms of maxi-
mum overshoot and settling time is presented applicable for uncertain systems with paramet-
ric uncertainties. One of the main results is developed empirical charts called B-parabolas; this
insightful graphical tool is used to transform engineering time-domain performance specifica-
tions (maximum overshoot and settling time) into frequency-domain performance measures
(phase margin and gain margin). The developed PID design method is based on shaping the
closed-loop step response using various combinations of excitation signal frequencies and
required phase and gain margins. Using B-parabolas, it is possible to shape time responses of
processes with various types of dynamics. By applying appropriate PID controller design
methods including the above presented, it is possible to achieve cost-effective control of
processes with uncertainties. The presented advanced external harmonic excitation-based
design method contributes to improve the unfavorable statistical ratio between the properly
tuned to all implemented PID controllers in industrial control loops.
Author details
References
[1] Åström KJ, Hägglund T. PID Controllers: Theory, Design and Tuning. 2nd ed. Research
Triangle Park: Instrument Society of America; 1995. ISBN 1-55617-516-7
[2] Åström KJ, Hägglund T. Benchmark Systems for PID Control. In: IFAC Workshop on
Digital Control PID'00; 5-7 April 2000; Terrassa, Spain. 2000. pp. 181-182
[5] Bialkowski WL. Dreams versus reality: A view from both sides of the gap. Pulp and Paper
Canada. 1993;94(11):19-27
[6] Bucz Š. Engineering methods of PID controller tuning for specified performance [doctoral
thesis]. Bratislava, Slovenská Republika: Slovenská Technická Univerzita v Bratislave;
2011 (in Slovak)
[7] Bucz Š, Veselý V, Kozáková A, Kozák Š. A novel PID controller design methodology for
specified performance using ultimate plant parameters. In: Proceedings of the 19th World
Congress of the International Federation of Automatic Control. IFAC; 2014. pp. 4909-4914.
ISBN 978-3-902823-62-5
[8] Coon GA. How to find controller settings from process characteristics. Control Engineer-
ing. May 1956;3(5):66-76
[10] Ford RL. The determination of the optimum process-controller settings and their confir-
mation by means of an electronic simulator. Proceedings of the IEE, Part 2. 1953;101(80):
pp. 141-155 and pp. 173-177
[11] Grabbe EM, Ramo S, Wooldrige DE. Handbook of Automation Computation and Control.
Vol. 1, 2, 3. New York; 1959–61
[12] Haalman A. Adjusting controllers for a deadtime process. Control Engineering. 1965;
(July):71-73
[13] Hang CC, Åström KJ. Practical aspects of PID auto-tuners based on relay feedback. In:
Proceedings of the IFAC Adaptive control of Chemical Processes Conference, Copenha-
gen, Denmark; 1988. pp. 153-158
[15] Chandrashekar R, Sree RP, Chidambaram M. Design of PI/PID controllers for unstable
systems with time delay by synthesis method. Indian Chemical Engineer Section A. 2002;
44(2):82-88
[16] Chau PC. Process Control—A First Course with MATLAB. 1st ed. New York: Cambridge
University Press; 2002. ISBN 978-0521002554
[17] Chen D, Seborg DE. PI/PID controller design based on direct synthesis and disturbance
rejection. Industrial and Engineering Chemistry Research. 2002;41:4807-4822
[18] Chien KL, Hrones JA, Reswick JB. On the automatic control of generalised passive sys-
tems. Transactions of the ASME. 1952;74(February):175-185
[19] Kozáková A, Veselý V, Osuský J. Decentralized digital PID design for performance. In:
12th IFAC Symposium on Large Scale Systems: Theory and Applications; 12-14 July 2010;
Lille, France. Ecole Centrale de Lille; 2010. ISBN 978-2-915-913-26-2
[20] Leva A, Maggio M. A systematic way to extend ideal pid tuning rules to the real structure.
Journal of Process Control. 2011;21:130-136
[21] Leva A, Negro S, Papadopoulos AV. PI/PID autotuning with contextual model
parametrisation. Journal of Process Control. 2010;20:452-463
[22] Morari M, Zafiriou E. Robust Processs Control. Englewood Cliffs, New Jersey, USA:
Prentice-Hall Inc.; 1989. ISBN 0137821530, 07632
[23] O’Dwyer A. Handbook of PI and PID Controllers Tuning Rules. 2nd ed. London: Imperial
College Press; 2006. ISBN 1860946224
[24] Osuský J, Veselý V, Kozáková A. Robust Decentralized Controller Design with Perfor-
mance Specification, Vol. 4(1). Kumamoto, Japan: ICIC Express Letters; 2010. pp. 71-76.
ISSN 1881-803X
[25] Padula F, Visioli A. Tuning rules for optimal PID and fractional-order PID controllers.
Journal of Process Control. 2011;21:69-81
[26] Padula F, Visioli A. On the stabilizing PID controllers for integral processes. IEEE Trans-
actions on Automatic Control. 2012;57:494-499
[27] Pettit JW, Carr DM. Self-tuning controller. US Patent No. 4669040; 1987
[29] Rotach V. Calculation of the Robust Settings of Automatic Controllers, Vol. 41(10). Mos-
cow, Russia: Thermal Engineering (Russia); 1994. pp. 764-769
[30] Suyama K. A simple design method for sampled-data PID control systems with adequate
step responses. Proceedings of the International Conference on Industrial Electronics,
Control, Instrumentation and Automation. 1992:1117-1122
118 PID Control for Industrial Processes
[31] Veronesi M, Visioli A. Performance assessment and retuning of PID controllers. Industrial
and Engineering Chemistry Research. 2009;48:2616-2623
[33] Veselý V. Easy tuning of PID controller. Journal of Electrical Engineering. 2003;54(5–6):
136-139, Bratislava, Slovak Republic. ISSN 1335-3632
[34] Vilanova R. IMC based robust PID design: Tuning guidelines and automatic tuning.
Journal of Process Control. 2008;18:61-70
[35] Visioli A. Fuzzy logic based set-point weight tuning of PID controllers. IEEE Transactions
on Systems, Man, and Cybernetics—Part A. 1999;29:587-592
[36] Visioli A. Tuning of PID controllers with fuzzy logic. IEE Proceedings—Control Theory
and Applications. 2001;148(1):180-184
[37] Visioli A. Practical PID control. Advances in industrial control. Springer-Verlag London
Limited; 2006. ISBN 1-84628-585-2
[38] Visioli A. Research trends for PID controllers. Acta Polytechnica. 2012;52(5/2012)
[41] Vítečková M, Víteček A, Smutný L. Controller tuning for controlled plants with time
delay. In: Preprints of Proceedings of PID'00: IFAC Workshop on Digital Control; 5-7 April
2000; Terrassa, Spain. 2000. pp. 83-288
[42] Wang L, Cluett WR. Tuning PID controllers for integrating processes. IEE Proceedings—
Control Theory and Applications. 1997;144(5):385-392
[43] Wang Y-G, Shao H-H. PID autotuner based on gain- and phase-margin specification.
Industrial and Engineering Chemistry Research. 1999;38:3007-3012
[45] Wojsznis WK, Blevins TL, Thiele D. Neural network assisted control loop tuner. In: Pro-
ceedings of the IEEE International Conference on Control Applications, Vol. 1; USA: 1999.
pp. 427-431
[46] Xue D, Chen Y, Atherton DP. Linear Feedback Control: Analysis and Design with
MATLAB. SIAM Press; 2007. ISBN 978-0-898716-38-2
Advanced Methods of PID Controller Tuning for Specified Performance 119
http://dx.doi.org/10.5772/intechopen.76069
[47] Yu C-C. Autotuning of PID Controllers. A Relay Feedback Approach. 2nd ed. Springer-
Verlag London Limited; 2006. ISBN 1-84628-036-2
[48] Ziegler JG, Nichols NB. Optimum settings for automatic controllers. ASME Transactions.
1942;64:759-768
[49] Zhuang M, Atherton DP. Automatic tuning of optimum PID controllers. IEE Proceedings
Part D: Control Theory and Applications. 1993;140(3):216-224. ISSN 0143-7054