Drum-Boiler Control Performance
Drum-Boiler Control Performance
Drum-Boiler Control Performance
Ahmed Elguindy
December 11, 2013
2nd Examiner:
Prof. Dr.-Ing. Bernd Orlik
Universität Bremen
Acknowledgment
It gives me great pleasure in expressing my sincere gratitude to everyone who
have supported and contributed into making this thesis possible.
I would like first to acknowledge my direct supervisor Dipl.-Ing. Simon Rünzi
for his enthusiasm, inspiration and huge efforts to explain things clearly and
simply. His in-depth knowledge regarding the CHP plant in Munich, related to
his PhD research, was quite helpful and beneficial for my work. Furthermore I
would like to thank my examiner Prof. Dr.-Ing Kai Michels for offering me the
project which have evolved over the course of time into an interesting thesis
topic. I wish also to address their constructive criticism following initial review
of the thesis.
My appreciation for SWM Services GmbH, specially Mr. Julian Niedermeier
for his willingness to perform experiments on the plant, its priceless valuable
information contributed significantly to improve my understanding of the real
process.
I wish to acknowledge the scholarship support provided by the Katholischer
Akademischer Ausländer-Dienst (KAAD). In particular I am very grateful to
Dr. Christina Pfestroff as I do believe that my master studies in Germany
wouldn’t have been possible without her guidance when applying for the schol-
arship. I thank as well Prof. Dr.-Ing Rainer Laur, Mr. Hans Landsberg, Mr.
Raphael Nabholz and Mrs. Claudia Dillmann for their continuous follow-up
and assistance.
Lastly and most importantly, I dedicate this thesis to my parents who raised,
supported, taught and loved me throughout my entire life.
Abstract
This thesis presents the development of an observer-based state-feedback con-
troller designed using LQ and pole placement methods to optimize pressure
and water level control performance of a drum-boiler unit that belongs to a
450 MW CHP plant in Germany. The Åström-Bell nonlinear model is initially
built within MATLAB/Simulink environment, later enlarged to include the
process PID-controllers and control valves regulating mass flow rates before
being validated against data measurements with very rich excitation. The con-
cluded simulation results adopting the newly proposed control strategy shows
that the suggested multivariable control technique outperforms the existing
PID-controller in many aspects improving the control performance significantly
and yielding much tighter reference value tracking during load changes.
Contents
1. Introduction 6
2. Process modelling 8
2.1. Combined cycle process overview . . . . . . . . . . . . . . . . . 8
2.1.1. Gas turbine . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.2. Heat recovery steam generator . . . . . . . . . . . . . . 9
2.1.3. Steam turbine . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.4. Surface condenser . . . . . . . . . . . . . . . . . . . . . 10
2.2. Steam generation process description . . . . . . . . . . . . . . . 11
2.2.1. Drum-boiler mass and energy balance . . . . . . . . . . 11
2.2.2. Drum-boiler nonlinear state equations . . . . . . . . . . 14
2.2.3. Mass flow control valve . . . . . . . . . . . . . . . . . . 15
2.2.4. Process PID-controller . . . . . . . . . . . . . . . . . . . 16
2.3. MATALB/Simulink model . . . . . . . . . . . . . . . . . . . . . 18
2.3.1. Drum-boiler model . . . . . . . . . . . . . . . . . . . . . 19
2.3.2. Control valve and actuator model . . . . . . . . . . . . . 21
2.3.3. Process PID-controller model . . . . . . . . . . . . . . . 22
4
Contents
4. Process optimization 38
4.1. Concept of state-feedback control . . . . . . . . . . . . . . . . . 38
4.1.1. Controllability and observability . . . . . . . . . . . . . 38
4.1.2. Observer-based control . . . . . . . . . . . . . . . . . . . 39
4.1.3. PI-based state-feedback control . . . . . . . . . . . . . . 40
4.2. Controller design methods . . . . . . . . . . . . . . . . . . . . . 41
4.2.1. Pole placement method . . . . . . . . . . . . . . . . . . 41
4.2.2. Linear-Quadratic method . . . . . . . . . . . . . . . . . 42
4.3. Observer-based state-feedback controller design . . . . . . . . . 44
4.3.1. Riccati controller . . . . . . . . . . . . . . . . . . . . . . 44
4.3.2. Luenberger observer . . . . . . . . . . . . . . . . . . . . 45
4.4. Simulation results . . . . . . . . . . . . . . . . . . . . . . . . . 45
A. Appendix 53
A.1. Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
A.2. MATLAB Control System Toolbox . . . . . . . . . . . . . . . . 54
A.2.1. Linear analysis functions . . . . . . . . . . . . . . . . . . 54
A.2.2. Controller design functions . . . . . . . . . . . . . . . . 54
A.3. MATLAB script . . . . . . . . . . . . . . . . . . . . . . . . . . 55
A.3.1. Drum-boiler model . . . . . . . . . . . . . . . . . . . . . 55
A.3.2. Controller design . . . . . . . . . . . . . . . . . . . . . . 57
A.4. Heat engines . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
A.4.1. Brayton cycle . . . . . . . . . . . . . . . . . . . . . . . . 58
A.4.2. Rankine cycle . . . . . . . . . . . . . . . . . . . . . . . . 59
A.5. Non-minimum phase systems . . . . . . . . . . . . . . . . . . . 60
A.6. Integral anti-windup control . . . . . . . . . . . . . . . . . . . . 61
A.7. Drum-boiler state equations coefficients . . . . . . . . . . . . . 63
A.8. Operator interface . . . . . . . . . . . . . . . . . . . . . . . . . 64
B. List of Figures 66
C. List of Tables 68
D. Bibliography 69
5
1. Introduction
1. Introduction
Energy market deregulation and integration of renewable energy resources
into the electrical grid have led to dramatic changes in the power industry which
escalated rapidly new challenges that have to be met by conventional power
plants. Such evolution caused a noticeable process modification regarding how
power plants operate, as they should become more flexible to fulfill their load
requirements which are more frequent nowadays. The process controllers have
to be designed in a way which can simultaneously fulfill the load demand as
soon as possible while at the same time bearing in mind safety and life span of
the plant crucial elements.
One common challenge is control of steam drum-boiler units handling supply
of the steam turbine continuously with steam at high pressure and temperature.
The controller should maintain drum pressure and water level within acceptable
ranges for all operating conditions. If the level exceeds upper limits, water
would be carried over to the superheater or the turbine leading to outage
in either of the turbine or the boiler. Surpassing lower limits would cause
overheating of the water wall tube resulting in serious tube rupture and severe
damage.
Drum level control in particular is quite tough due to the process physical
phenomena known as shrink/swell of steam bubbles under the water level which
causes the system to react with an initial inverse response known as a non-
minimum phase behaviour.
Classical control design methods using 2-element or 3-element PID-controllers
can behave fairly well to compensate such effect. However as the process is
quite complicated, dealing with several input variables to regulate each process
variable separately might end up with bad parameter tuning and poor level
performance observed during load changes, eventually leading the boiler unit
to trip or even worse cause emergency shutdown of the power plant. It is stated
that about 30% of the emergency shutdowns in French pressurized water reac-
tors (PWR) plants were caused by poor level control of a steam drum-boiler
unit [21].
An ongoing research project is taking place at the moment in collabora-
tion with Stadtwerke München GmbH - Munich City Utilities (SWM) in re-
gards with the process PID-controllers of the low pressure drum-boiler unit,
6
1. Introduction
located within the combined cycle plant GuD 2, short for Gas-und-Dampf-
Kombikraftwerk at Heizkraftwerk Süd (HKW) - combined heat and power
(CHP) facility. The main objective is drum level and pressure closed loop
performance optimization which have been reported to behave very poorly un-
der huge load changes taking place frequently following energy deregulation in
Germany.
The thesis is presented as follows, initially the complete process is briefly
introduced before being simplified to highlight the significant elements domi-
nating the steam generation process which are mainly focused on during mod-
elling procedure. Derivation of the differential equations is carried out for each
established featured element to develop a mathematical model capable of cap-
turing most of the system nonlinearities and later on suitable for model-based
control.
The model parameterized and implemented within MATLAB/Simulink en-
vironment will be subjected to a detailed analysis by examining stability, simu-
lating the model open loop response and validating the closed loop against data
measurements from the plant. The investigation concluded results will offer a
good insight into the system inner dynamics and shall inspect the model abil-
ity to catch the plant dynamical behaviour for a wide spectrum of operating
conditions.
In the end, the proposed control strategy is addressed. First, state-feedback
control concept and the numerous methods which applies it shall be briefly
described to illustrate their applicability and major difference between them.
The most convenient and suitable approach shall be employed to compute the
state-feedback and observer gain matrices. Finally, simulation results of the
process utilizing the newly designed observer-based state-feedback controller is
presented for various sequences to ensure stability of the optimized closed loop.
7
2. Process modelling
2. Process modelling
HKW Süd plant is classified as a combined cycle cogeneration plant, it can
handle concurrent production of electrical power and useful heat utilizing a
class of sustainable integrated technologies progressively being used.
Cogeneration plants reduce thermal and mechanical losses, harmful carbon
dioxide (CO2 ) emissions and more importantly increases the overall plant effi-
ciency to approximately 81% in comparison to stand alone plants which don’t
exceed 45%. The German government is planning to double its share of CHP
plants from approximately 12% to 25% by 2020, as part of the Integrated
Energy and Climate Protection Program (IECPP) [8].
GuD 2 at HKW Süd manages electrical power generation by combining both
Brayton and Rankine thermodynamic theoretical cycles (A.4) [12] [20] using
gas and steam turbines. Exhaust gas emitted from the gas turbine can be
reused as the heat source for steam production required to operate the steam
turbine, therefore more useful energy can be extracted, supplying additional
electricity to the grid.
Further energy can by even withdrawn from the low pressure steam leaving
the turbine when condensed using a heat exchanger where the low temperature
steam released can be utilized for district heating or water desalination.
In this chapter, the overall combined cycle process is being narrowed down
to draw the focus on one particular key element within the plant. The process
is further simplified in order to spotlight primarily our aim interest which is
the steam production using the low pressure drum-boiler unit along side with
its process PID-controllers.
8
2.1. Combined cycle process overview
Cooling water
Condenser
Steam turbine
Superheated steam
Feedwater pump
Fresh air
Electricl power
Gas turbine
Exhaust heat
9
2.1. Combined cycle process overview
10
2.2. Steam generation process description
1 3
Outflow - qs
Feedwater Tank
2
4 5
Inflow - qf
Downcomer-riser loop
Exhaust heat - Q
Figure 2.3.: Schematic diagram of the low pressure steam generation process
11
2.2. Steam generation process description
cially the two phase flow modelling attempt is usually quite complicated requir-
ing typically usage of partial differential equations. In literature there exists
a lot of research papers that were devoted into developing relatively simple
physical models [2] [7] [13] [14].
In particular the well developed Åström - Bell model3 is being considered.
The majority of the system attitude can be captured through a 4th order non-
linear model by means of defining mass flow and energy balance with the help
a physical mechanism introduced under the following elementary assumptions.
Steam demand
(to downstream) Sat. steam
Drum Internal
Upper void Separation
(saturated steam) Device Mixture
from riser
Steam rises
Steam-water
Condensation
Sat.
steam
Steam-water
mixture
Water boils
and flows
upward
Downcomer
Riser
Heat from
hot medium
Most of the system parts will be under thermal equilibrium due to their
direct contact with saturated liquid/vapour mixture. The energy stored in the
mixture is either absorbed or released quickly following drum pressure changes,
meaning that various metal parts of the system would adapt their temperatures
in the same manner.
This agrees with experimental observation which have proven that the differ-
ence between both temperatures is very small, thus a detailed representation
3
Part of an ongoing research project which started back in the early seventies
12
2.2. Steam generation process description
The empirical equation (2.3) resulted from various attempts to fit with the
experimental data. It defines mass balance of the steam bubbles under the
water level in terms of condensation flow qcd and steam flow through the liquid
surface qsd driven by density difference of the mixture and momentum of the
flow qr entering through the riser tubes. It can capture most of the process
dynamics by proper parameterizations of residence time of steam inside the
◦ and empirical
drum Td , the bubbles steam volume at hypothetical situation4 Vsd
coefficient β correspondingly.
d
(ρs Vsd ) = αr qr − qcd − qsd
dt (2.3)
ρs ◦
qsd = (Vsd − Vsd ) + αr qdc + αr β(qdc − qr )
Td
4
Theoretical state that assumes no condensation of steam inside the drum
13
2.2. Steam generation process description
Equations (2.4), (2.5) rearrange the drum mass and energy balance, equation
(2.6) combines the mass and energy balance of the downcomer-riser closed loop
in a single equation and equation (2.7) considers only the mass balance of steam
bubbles under water level. The interesting feature of this model is that the
states can be grouped in the form ((P, Vwt ), αr , Vsd ), where each term can be
computed separately in a nested manner treating the system as 2nd , 3rd or 4th
order according to modelling requirements.
14
2.2. Steam generation process description
p1 p2
H100
H0
t1
Q W
Figure 2.5.: Flow through control valve for liquid service [22]
The dynamics related to regulation of steam flow rate are quite complicated
where additional considerations have to be taken care of when compared to
feedwater mainly due to the difference in properties between both. One good
approximation to describe the flow rate through a control valve meeting prac-
tical needs can be achieved using equation (2.9) where P is the drum pressure,
the head loss coefficient m and the compressibility factor Z are taken into
account to distinguish between saturated and superheated steam.
√
Kvf · ρw · ΔP
qf = xf · (2.8)
3600
Kvs · Z · m
qs = x s · (2.9)
3600
Clearly the valve position value would vary according to the type of valve
being used. The inherent flow characteristic depicted in figure (2.6) highlight
the comparison between the commonly used control valve demonstrating that
mass flow rate for the same opening position and pressure drop across it is
obviously altered according to the category it belongs to.
Examining halfway opened linear, butterfly and relief valves correspondingly,
15
2.2. Steam generation process description
16
2.2. Steam generation process description
17
2.3. MATALB/Simulink model
18
2.3. MATALB/Simulink model
States
1
1 State variables (dx/dt)
s
Heat flow rate (z)
2 Model inputs Pressure (bar) 1
Mass flow rates (u) Pressure (bar)
Level (mm) 2
States (x) Level (mm)
Drum-Boiler m-file
19
2.3. MATALB/Simulink model
hypothetical volume Vsd◦ were quite hard to obtain, therefore were either kept
The amount of mass flow rates qf and qs at a given pressure P are first
specified in order to compute initial values. This allows computation of the
necessary heat flow rate Q̇ that preservers energy balance. A primary simula-
tion can run once these values are assigned as the model drives by itself the
variables αr and Vsd to steady state by solving equations (2.12). Finally, the
total volume Vwt is the amount required to keep the water at the relative zero
level.
2ρw Adc (ρw − ρs )g ᾱv Vr
Q̇ = qs hs − qf hf w = αr hc
K (2.12)
ρw ρs ρ w − ρs
ᾱv = 1− ln 1 + αr
ρw − ρs (ρw − ρs )αr ρs
6 ◦
K and β were kept the same, whereas Vsd is chosen as a rule of thumb
7
Changed within the range [ 2 sec - 6 sec ] until the model closed loop behaviour matched
the plant real measurements
8
Total mass including the evaporator mt = 98888 kg
20
2.3. MATALB/Simulink model
1 Pdrop (bar)
Figure 2.8.: Simulink model of the control valve combined with its actuator
21
2.3. MATALB/Simulink model
1
1 Ki Kp 1
s
Error Controller
Integrator gain Integrator Proportional gain Saturation output
Kb
Back-calculation
Td N coefficient
1
s
Filter
PID-Controller parameters
Controller Kp Ti Td Tf
Level 0.05 300 100 50
Feedwater valve 2.3 25 - -
Pressure10 1.8 12 - -
22
3. Process analysis and validation
ẋ = f (xR , u0 ) = 0 (3.1)
23
3.1. Theoretical overview
3.1.2. Linearization
For simplicity the intended stability analysis shall be performed by lineariza-
tion of the nonlinear model at typical operating points of the drum whose state
algebraic equations can be summarized into the generalized description shown
in equation (3.2).
ẋ = f (x, u)
(3.2)
y = g(x, u)
The resulting linearized model can be described in state-space form (3.3)
where A, B, G, C and D are the system, input, disturbance, output and
feedforward matrices respectively. This will come in handy when attempting
to optimize the controller since algorithm execution of the proposed strategy
requires these matrices. In addition they would reduce the nonlinear state
equations complexity offering a rather simplified overview of the states and
inputs dominating the process outputs. The open loop response of both lin-
ear and nonlinear models should be compared to inspect if both still match,
therefore answering the crucial question concerned with the linearized model
reliability during design of the optimal controller.
ẋ = Ax + Bu + Gz
(3.3)
y = Cx + Du
The matrices are computed with the help of Taylor series approximation
neglecting quadratic and higher order terms (3.4). The method intuitive basis
is that a smooth curve differs very little from its tangent line as long as the
variable doesn’t wander from the point of tangency.
∂fi ∂fi
aij = , bij =
∂xj xR ,u0 ∂uj xR ,u0
(3.4)
∂gi ∂gi
cij = , d =
∂xj ∂uj
ij
xR ,u0 xR ,u0
24
3.1. Theoretical overview
plane (A.5) [9] [23]. Alternatively poles determine stability as they are directly
associated with the system eigenvalues. That’s why inspection of the system
poles and zeros is quite efficient while analysing and predicting the system
response.
If the eigenvalues are located on the left-hand side (LHS) of the complex
plane the states will converge to zero stabilizing over the course of time. How-
ever if located on the right-hand side (RHS) the states will keep growing due
to the exponential product as depicted throughout equation (3.5) where ci
are constants coefficients related to the solution of the homogenous differential
equation describing dynamics of the system.
n
y(t) = ci eλi t (3.5)
i=1
We still need to define the relationship between the zeros, poles and eigen-
values, in addition understand how it differs when comparing Multiple-input
Multiple-Output (MIMO) systems to Single-Input Single-Output (SISO).
SISO systems
Commonly input-output (I/O) behavior is presented using transfer functions
(3.6) where zeros zi and poles pi are simply the roots of the numerator N (s)
and dominator D(s) respectively. The transfer function dominator is exactly
equivalent to the characteristic polynomial evaluated by solving equation (3.7),
that why all poles corresponds to the system eigenvalues λ.
MIMO systems
Zeros in multivariable systems do play an additional role besides affecting
system shape and performance since it might gravely influence the ability to
fully control the system [5] [6]. They are redefined with the help of Rosenbrock
matrix which benefits from the state-space description distinguishing between
transfer and decoupling zeros. The complete set consisting of both known as
invariant zeros1 is examined by computing the rank of the matrix (3.8). Not
1
Only under the assumption that feedforward matrix D = 0
25
3.2. Stability analysis
26
3.2. Stability analysis
States Inputs
Operation 3 3
P (bar) 2
Vwt (m ) αr (%) Vsd (m ) Q̇(MW) qf ( kg
s ) qs ( kg
s )
Low 5.5 21.501 0.0098 1.378 13.8473 6 6
Medium 5.5 20.391 0.0138 1.067 20.771 9 9
High 5.5 19.736 0.0178 0.756 27.6947 12 12
Table 3.1.: Drum-boiler operating points for low, medium and high load
27
3.2. Stability analysis
when inspected in the complex plane, since the conversion shall take place
instantaneously.
Since no compensation of eigenvalues have occurred, all invariant zeros are
classified as transfer zeros. In addition one can stay assured that the system is
completely controllable because any eigenvalue can be influenced by affecting its
corresponding pole. The transfer zeros which are located on the right hand side
(RHS) of the complex plane have been anticipated earlier from experimental
observation and physical understanding. They are directly correlated with the
shrink and swell physical phenomena leading the system to react in a non-
minimum phase behaviour.
In particular zeros related to the transfer behavior from steam flow rate to
water level are very close to the origin when compared with zeros linked to
feedwater and heat flow rates transfer functions respectively as seen in figure
(3.1). Therefore we should be definitely expecting a significant difference in
regards with amplitude of the water level initial inverse response when stimu-
lated by the input variables. This shall verified in the next section concerned
mainly with the open loop response to a step input.
Heat flow rate Q (MW)
1
3 2 1 1 23 1 2 3
0
1
0.28 0.21 0.14 0.07 0 0.07
1
0.28 0.21 0.14 0.07 0 0.07
1
0.28 0.21 0.14 0.07 0 0.07
Real axis
Figure 3.1.: Pole-zero plot of the linearized models at low (1), medium (2) and
high (3) load
28
3.3. Open loop step response
4
All simulations were conducted in Simulink using a fixed step size of 1 s
5
In [2] the open loop response considers only the drum-boiler unit
29
3.3. Open loop step response
30
3.3. Open loop step response
20.7
9
0 20.6
8.8
20.5
Figure 3.2.: Open loop response for a step change equivalent to decrease of
20 MW of the gas turbine electrical output power
4.5 8 0.013
0 100 200 0 100 200 0 100 200
9 20.4
50
8.8 20.2
0
8.6 20
0 100 200 0 100 200 0 100 200
Figure 3.3.: Open loop response for a step change equivalent to 10% opening
of butterfly valve position
31
3.4. Validation
4.P 8.8
4.4 8.9
Figure 3.4.: Open loop response for a step change equivalent to 10% closing of
feedwater control valve position
3.4. Validation
The system closed loop response will be validated and examined against data6
from the real plant for different scenarios to experiment its ability to capture
the real process dynamics at various operating conditions. The complete model
with the PID-controllers is shown in figure (3.5).
3.4.1. Assumptions
The heat flow rate required as an input of the model cannot be measured in
reality, yet can be predicted from the gas turbine electrical output power which
1 MW
changes as ramp function with a slope of 12 s
. The transfer function relating
st
both is assumed to be 1 order lag element whose time constant was identified
τ = 280 s assuming that the supplied heat behaviour is directly associated with
the evaporator temperature.
The feedwater valve position is always kept half-way opened in the plant
without considering the amount of feedwater which flows through it. Therefore
the pressure drop across the valve should increase or decrease accordingly to
preserve such condition which is achieved using he feedwater pump controller.
6
The measurements of the plant are being filtered and sampled at a rate of 1 Hz
32
3.4. Validation
33
3.4. Validation
Steam to 0.18182
feedwater
tank 5s+1
P_drum (bar)
qs (kg/s) qs (Kg/s)
1
5.5 PI(s) xs (%)
s P (Bar)
Pressure
Dead Zone Actuator xs Electrical Motor Butterfly valve
Reference value qs PI control
-0.04 to 0.04 10% by 18sec
Kp = 4 Ti = 1/15
1e+006
Q (W)
280s+1
Feedwater Heat flow rate (Watt)
Low pass Low pass
pump Saturation
Gain watt to MW
1
15s+1
Level (mm)
Pdrop (bar)
qf (kg/s) qf (Kg/s)
1
0 PID(s) PI(s) xf (%)
s
Level
Actuator xf Electrical Motor Feedwater valve
Reference value Level PID Controller qf PI Controller
10 % by 3 sec
Kp = 0.05 Ti = 1/300 Kp = 2 Ti = 1/25 Drum-boiler model
Td =50 Tv = 100
34
3.4. Validation
20 8
5.6
7
5.5
5.4 15 6
5.3 5
5.2 10 4
0 2000 4000 0 2000 4000 0 2000 4000
0 60 90
50
40 80
100
0 2000 4000 0 2000 4000 0 2000 4000
Figure 3.6.: Comparison between model (dashed) and plant data (solid line)
for a decrease of the gas turbine electrical output power equivalent
to 20 MW
23 6.4
4.3
6
0 2400 5400 0 2400 5400 0 2400 5400
Figure 3.7.: Comparison between model (dashed) and plant data (solid line) for
an increase of the gas turbine electrical output power equivalent
to 10 MW
35
3.5. Concluding remarks
36
3.5. Concluding remarks
be increased so that the simulation results match the plant measurements. The
drum pressure drops more when the gain isn’t changed and as a consequence
it causes higher overshoot of the level due to steam bubbles swelling.
The reasons behind such modification which as seen in the previous simu-
lation results improved the model overall performance could be summarized
throughout the following
50
10
0
6
50
100 2
0 1000 2000 3000 4000 0 1000 2000 3000 4000
Figure 3.8.: Comparison between model and plant data (solid line) for a de-
crease of the gas turbine electrical output power equivalent to
20 MW when Kp = 1.8 (dotted dashed) and Kp = 5 (dashed)
7
System identification shows that a 5th order system matches better the plant measurements
when compared to the current 4th order model
37
4. Process optimization
4. Process optimization
The concluded results brought to our attention during system modelling
and analysis suggests that an optimization of the process is achievable using
a multivariable control technique. The strategy would account for synergy
between feedwater and steam flow rates instead of just decoupling the MIMO
system into several coupled SISO systems regulated by their own noninteracting
control loops.
A state-feedback controller is suggested in order to consider the internal
variables of the system instead of the process outputs, therefore accounting for
additional aspects which were discarded using the classical control methodol-
ogy. The inner dynamics of the drum-boiler unit correspond to the developed
nonlinear model state variables which were defined in section (2.2.2).
The control concept shall be addressed presenting the available control meth-
ods and algorithms applying the approach while highlighting advantages and
disadvantages for each.
u = −F x (4.1)
38
4.1. Concept of state-feedback control
B C
Luenberger
Observer L
B C
39
4.1. Concept of state-feedback control
- -I B C
-F
State Controller
The previous control law have to be slightly modified considering error signals
as additional states h. The newly computed state-feedback matrix K would
consist of 3 parameters F , I and P affecting the states, integrated error signals
and reference tracker correspondingly.
The tunable parameters are assigned upon computation of the feedback ma-
trix K. By default I is uniquely defined as it corresponds to the left-hand
side of the matrix. However F and P can be freely selected due to the un-
derdeterministic nature of equation (4.2). Ignoring P element would lead to a
simple I-controller, while setting F to zero isn’t relevant when attempting to
design a state controller, besides, this would normally introduce an unsolvable
equation1 (4.3).
x
u = [P C − F , −I] = −Kx (4.2)
h
P = −KC −1 (4.3)
1
A solution might exist under very special certain conditions
40
4.2. Controller design methods
Ackermann’s formula
The algorithm is regarded as a standard design procedure which computes
the unique state-feedback f row vector or observer gain l column vector using
the formula (4.4) [18] where n is the system order, p are the coefficients of
characteristic polynomial calculated while defining positions of the eigenvalues
for the closed loop and finally t is the last column/row obtained from the
computed controllability/observability matrix inverse.
Unfortunately such method cannot be applied for MIMO systems as the al-
gorithm requires inversion of controllability/observability matrix which is only
attainable with square matrices with full rank.
41
4.2. Controller design methods
Riccati controller
The Riccati controller allows trade off between regulation performance and
control effort compared with the pole placement method. It’s regarded as
a robust controller since it attains infinite marginal gain and offers a phase
margin δ ≥ 60◦ [15] which is aligned with practical guidelines for control system
design. The resulting optimal feedback gain should drive the closed-loop system
without external input from any initial state to the zero state minimizing the
cost function described by equation (4.6).
∞
J= (xT (t)Qx(t) + uT (t)Ru(t))dt (4.6)
0
2
Function that summarizes dynamics of the system
42
4.2. Controller design methods
The control law (4.1) tries to minimize the quality function (4.6) whose
optimal feedback matrix F requires the symmetrical positive definite matrix
P resulting from solution of the Matrix Algebraic Riccati Equation (MARE)
(4.7) and hence the reason why a LQ regulator is named after Riccati.
F = R−1 B T P
(4.7)
AT P + P A − P BR−1 B T P + Q = 0
Kalman filter
The Kalman filter optimizes the estimation of the system states using a series
of process measurements. It takes into account the input w and measurement
v noises assumed to be random unbiased white noise. It constructs an optimal
state estimator that minimizes the cost function (4.9) where E() calculates the
expected value based on the assumed random noise.
J = E(eT W e) (4.9)
The observer gain L computed by solving the modified MARE (4.10) at-
tempts to minimize the difference between estimated and real states considering
noise influence on the process. The weighting matrices Q and R aren’t con-
sidered as punishing factors anymore as they define intensities of the expected
process noise. Choice of their values usually starts with identity matrices as
initial guess then adjusted repeatedly until a decent estimation is achieved.
L = P C T R−1
(4.10)
AP + P T A − P C T R−1 CP + Q = 0
43
4.3. Observer-based state-feedback controller design
3
P should be a 2-by-2 identity matrix in the real process, since the proportional gain of
the PID-controller reacts on the sum of all actions and not the error signal
44
4.4. Simulation results
45
4.4. Simulation results
States estimation
The observer pole placement was successful as shown in figure (4.3) since the
estimated states noise is almost negligible therefore they can be efficiently fed
back to the state-feedback controller matrix.
46
4.4. Simulation results
22
20
0.015
0.01
0 1000 2000 3000 4000 5000 6000 7000
Steam volume under water level Vsd (m3)
1.7
0.95
0.2
0 1000 2000 3000 4000 5000 6000 7000
Figure 4.3.: Estimated states using the observer for perturbations in gas tur-
bine electrical output power
47
4.4. Simulation results
5.8 10
5.6 8
5.4
6
5.2
4
5
0 2000 4000 6000 0 2000 4000 6000
Level l (mm) Flow rate qf (Kg/s)
200
100 15
0 10
100 5
200
0
0 2000 4000 6000 0 2000 4000 6000
Figure 4.4.: Comparison between state observer (dashed) and plant (solid line)
for perturbations in gas turbine electrical output power
5.6
8
5.5
5.4 7
5.3
6
0 1000 2000 3000 0 1000 2000 3000
Level (mm) Flow rate qf (Kg/s)
60
12
40
20 10
0 8
20
6
40
0 1000 2000 3000 0 1000 2000 3000
Figure 4.5.: Comparison between state observer (dashed) and plant (solid line)
for a decrease of the gas turbine electrical output power equivalent
to 10 MW
48
4.4. Simulation results
8
5.6
5.4 6
5.2 4
0 1500 3000 4500 0 1500 3000 4500
Level (mm) Flow rate qf (Kg/s)
100 15
50
10
0
50
5
100
0 1500 3000 4500 0 1500 3000 4500
Figure 4.6.: Comparison between state observer (dashed) and plant (solid line)
for a decrease of the gas turbine electrical output power equivalent
to 20 MW
15
100
0 10
100 5
200
0
0 2000 4000 6000 0 2000 4000 6000
49
4.4. Simulation results
5.45
5
5.4
5.35 0
0 500 1000 1500 0 500 1000 1500
Level (mm) Flow rate qf (Kg/s)
12
0 10
8
50
6
4
100
2
0 500 1000 1500 0 500 1000 1500
Figure 4.8.: Model closed loop response using the PI-based state-feedback con-
troller for high (dashed), medium (solid line) and low load (dotted
dashed)
Figure 4.9.: States and input variables behaviour using the PI-based state-
feedback controller at medium load
50
Set value tracking Flow rates
set value
4.4. Simulation results
Drum-boiler
model
Observer-based
State-feedback controller
-F
Figure 4.10.: Block diagram of the proposed multivariable feedback control strategy
51
5. Conclusion and future work
52
A. Appendix
A. Appendix
A.1. Nomenclature
53
A.2. MATLAB Control System Toolbox
54
A.3. MATLAB script
%% Temperature
Tfw = 104; %Feedwater (C)
T Sat = XSteam('Tsat p',P); %Saturation (C)
dT Sat dP = IAPWS IF97('dTsatdpsat p',P*0.1) * 1e−6; %(K/Pa)
%% Density
rhoV = XSteam ('rhoV p',P); %Steam (Kg/m3)
rhoL = XSteam ('rhoL p',P); %Water (Kg/m3)
%Partial derivative with pressure
drhoL dP = (2*P*0.0148 − 3.7836) * 1e−5; %Water (Kg/J)
drhoV dP = (2*P*0.0010 + 0.4450) * 1e−5; %Steam (Kg/J)
%% Specific Enthalpy
hfW = XSteam('hL T',Tfw) *1e3; %Feedwater (J/Kg)
55
A.3. MATLAB script
%% Coefficients Values
Eta = (Alpha*(rhoL−rhoV))/rhoV;
AlphaV = (rhoL / (rhoL−rhoV)) * (1 − ((rhoV/((rhoL−rhoV)*Alpha)) ...
* log(1+(((rhoL−rhoV)*Alpha)/rhoV))));
dAlphaV dP = (1/((rhoL−rhoV)ˆ2))*(rhoL*drhoV dP − ...
rhoV*drhoL dP)*(1 + (rhoL/(rhoV *(1+Eta))) − ...
(((rhoV+rhoL)*log(1+Eta))/(rhoV*Eta)));
dAlphaV dAlpha = (rhoL/(rhoV*Eta))*(((1/Eta)*log(1+Eta)) − ...
(1/(1+Eta)));
qdc = sqrt((2*rhoL*Adc*(rhoL−rhoV)*9.81*AlphaV*Vr)/K);
Vwd = Vwt − Vdc − (1−AlphaV)*Vr;
%% State variables
dP dt = (1/(e11*e22 − e12*e21))*(e11*Q + qf*(hfW*e11 − e21) + ...
qs*(e21 − hV*e11));
dVwt dt = (1/(e11*e22 − e12*e21))*(qf*(e22 − hfW*e12) + ...
qs*(hV*e12 − e22) − e12*Q);
dAlpha dt = (1/e33)*(Q − Alpha*hC*qdc − e32* dP dt);
dVsd dt = (1/e44)*(((rhoV/Td)*(Vsd0−Vsd)) + ((hfW − hL)*qf/hC) − ...
e42* dP dt − e43*dAlpha dt);
Level = (Vwd+Vsd)/Ad −0.875;
%% Model outputs
y = [ dP dt dVwt dt dAlpha dt dVsd dt Level ];
end
56
A.3. MATLAB script
57
A.4. Heat engines
Heat addition
Heat rejection
Fluid Exhaust
heat Open system
58
A.4. Heat engines
Process 1-2 Fresh air is being supplied for an open cycle, as for a closed
one it’s drawn back from the turbine to a compressor increasing its pres-
sure in an adiabatic compression process.
Process 2-3 The compressed air is mixed with fuel or natural gas before
being burned inside the combustion chamber at constant pressure.
Process 3-4 The heated pressurized air is supplied to the heat engine,
where it’s allowed to expand through the turbine driving its blades, in
an adiabatic expansion process.
Process 4-1 Finally, heat rejection to the surrounding atmosphere takes
place at constant pressure.
P T 3
q in
t.
ns
co
2 3 q in
=
p
s 4
s=
=
co
ns
co
t.
ns
t.
1 4 2
q out ns
t.
1 = co q out
p
Figure A.2.: T-S and P-V diagram of a typical ideal Brayton cycle
59
A.5. Non-minimum phase systems
Process 3-4 Vapour leaving the turbine at low pressure and temperature
is condensed, converting it into wet saturated water.
Process 4-1 Saturated water is pumped back, feeding the boiler at high
pressure, where the cycle is repeated.
60
A.6. Integral anti-windup control
Since the output signal has an initial value y(0) = 0, final value y(∞) = 1
and its area under the curve evaluated by the integral in equation (A.3) is equal
to 0, then this implies that the output signal y(t) must take negative values
over time.
Y (s) (s − z0 )
Gcl (s) = = (A.1)
R(s) s + p0
(s − z0 )
Y (s) = Gcl (s)R(s) = (A.2)
s(2s + p0 − z0 )
∞
(z0 − z0 )
y(t)e−z0 t dt = Y (s)|s=z0 = Y (z0 ) = =0 (A.3)
0 s(2s + z0 − z0 )
Figure (A.4) illustrates the step and bode responses of a simple 2nd order
minimum and non-minimum phase plant. The zero of the first plant lies in the
LHS leading to the normal expected step response, however, as it’s is shifted to
the RHS for the second plant, there exists an undershoot in the initial response
and delay in overall response, caused by the phase shift difference as seen in
the bode plot.
1
Magnitude (dB)
0.8 20
0.6 40
0.4 180
135
Phase (deg)
0.2 90
45
0 0
45
0.2 90
2 1 0 1 2
0 5 10 15 20 10 10 10 10 10
Figure A.4.: Step response and bode plot for a minimum (solid line) and a
non-minimum (dashed) system
61
A.6. Integral anti-windup control
4
0.2
3
0
2
0.2
1
0 0.4
0 10 20 30 40 50 0 10 20 30 40 50
Figure A.5.: Motor velocity and position behaviours with (dashed) and without
(solid line) anti-windup
62
A.7. Drum-boiler state equations coefficients
∂ρw ∂ρs
e11 = Vwt + Vst
∂p ∂p
e12 = ρw − ρs
∂ρw ∂hw ∂ρs ∂hs ∂tsat
e21 = Vwt hw + ρw + Vst hs + ρs − Vt + mt Cp
∂p ∂p ∂p ∂p ∂p
e22 = ρw hw − ρs hs
∂hw ∂ρw ∂hs ∂ρs
e31 = ρw − α r hc (1 − ᾱv )Vr + ρs + (1 − αr )hc ᾱv Vr +
∂p ∂p ∂p ∂p
∂ ᾱv ∂tsat
(ρs + (ρw − ρs )αr ))hc Vr − Vr + mr Cp
∂p ∂p
∂ ᾱv
e33 = ((1 − αr )ρs + αr ρw )hc Vr
∂p
∂ρs ∂ρs ∂ρw ∂ ᾱv
e41 = Vsd + αr (1 + β)Vr ᾱv + (1 − ᾱv ) + (ρs − ρw ) +
∂p ∂p ∂p ∂p
1 ∂hs ∂hw ∂tsat
ρs Vsd + ρw Vwd − Vsd − Vwd + md Cp
hc ∂p ∂p ∂p
∂ ᾱv
e43 = αr (1 + β)(ρw + ρs )V r
∂p
2ρw Adc (ρw − ρs )g ᾱv Vr
qdc =
K
h w − hf w 1 ∂hw ∂hs ∂tsat dP
qct = qf + ρw Vwt + ρs Vst − Vt + mt Cp
hc hc ∂p ∂p ∂p dt
∂ρs ∂ρw ∂ ᾱv dP ∂ α¯v dαr
qr = qdc − Vr ᾱv + (1 − ᾱv ) + (ρw − ρs ) + (ρw − ρs )Vr
∂p ∂p ∂p dt ∂αr dt
ρw ρs ρ w − ρs
ᾱv = 1− ln 1 + αr
ρ w − ρs (ρw − ρs )αr ρs
(ρw − ρs )
ζ = αr
ρs
∂ ᾱv ρw ln(1 + ζ) 1
= −
∂αr ρs ζ ζ 1+ζ
∂ ᾱv 1 ∂ρs ∂ρw ρw ρs + ρw
= ρ w − ρ s 1 + − ln(1 + ζ)
∂p (ρw − ρs )2 ∂p ∂p ρs (1 + ζ) ζρs
63
A.8. Operator interface
Figure A.6.: Screenshot of the drum-boiler unit in the real process DCS
64
A.8. Operator interface
Figure A.7.: Screenshot of the low pressure steam distribution network in the
real process DCS
65
B. List of Figures
B. List of Figures
3.1. Pole-zero plot of the linearized models at low (1), medium (2)
and high (3) load . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.2. Open loop response for a step change equivalent to decrease of
20 MW of the gas turbine electrical output power . . . . . . . . 31
3.3. Open loop response for a step change equivalent to 10% opening
of butterfly valve position . . . . . . . . . . . . . . . . . . . . . 31
3.4. Open loop response for a step change equivalent to 10% closing
of feedwater control valve position . . . . . . . . . . . . . . . . 32
3.5. Simulink validation model . . . . . . . . . . . . . . . . . . . . . 34
3.6. Comparison between model (dashed) and plant data (solid line)
for a decrease of the gas turbine electrical output power equiva-
lent to 20 MW . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.7. Comparison between model (dashed) and plant data (solid line)
for an increase of the gas turbine electrical output power equiv-
alent to 10 MW . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.8. Comparison between model and plant data (solid line) for a de-
crease of the gas turbine electrical output power equivalent to
20 MW when Kp = 1.8 (dotted dashed) and Kp = 5 (dashed) . 37
66
B. List of Figures
4.3. Estimated states using the observer for perturbations in gas tur-
bine electrical output power . . . . . . . . . . . . . . . . . . . . 47
4.4. Comparison between state observer (dashed) and plant (solid
line) for perturbations in gas turbine electrical output power . 48
4.5. Comparison between state observer (dashed) and plant (solid
line) for a decrease of the gas turbine electrical output power
equivalent to 10 MW . . . . . . . . . . . . . . . . . . . . . . . . 48
4.6. Comparison between state observer (dashed) and plant (solid
line) for a decrease of the gas turbine electrical output power
equivalent to 20 MW . . . . . . . . . . . . . . . . . . . . . . . . 49
4.7. Comparison between PI-based state-feedback controller (dashed)
and plant (solid line) for perturbations in gas turbine electrical
output power . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.8. Model closed loop response using the PI-based state-feedback
controller for high (dashed), medium (solid line) and low load
(dotted dashed) . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.9. States and input variables behaviour using the PI-based state-
feedback controller at medium load . . . . . . . . . . . . . . . . 50
4.10. Block diagram of the proposed multivariable feedback control
strategy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
67
C. List of Tables
C. List of Tables
3.1. Drum-boiler operating points for low, medium and high load . 27
68
D. Bibliography
D. Bibliography
[1] Asea Brown Boveri: Drum Level Control Systems in the Process Industries,
1997. ABB Download Center.
[2] Åström, Karl Johan und Rodney D. Bell: Drum Boiler Dynamics. Auto-
matica, 36:363–378, März 2000.
[3] Åström, K.J. und T. Hägglund: PID Controllers - Theory Design and
Tuning. International Society of America, 1995, ISBN 9781556175169.
[7] Flynn, M.E. und M.J. O Malley: A drum Boiler Model for Long Term
Power System Dynamic Simulation. IEEE Transaction Power System,
14(1):209–217, 1999.
[11] Kautsky, J., N.K. Nichols und P. Van Dooren: Robust Pole Assignment
in Linear State Feedback. International Journal of Control, 41:1129–1155,
1985.
69
D. Bibliography
70