You are on page 1of 26

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/220674721

A simulated model for a once-through boiler by parameter adjustment based


on genetic algorithms

Article in Simulation Modelling Practice and Theory · August 2007


DOI: 10.1016/j.simpat.2007.06.004 · Source: DBLP

CITATIONS
READS
78
1,306

3 authors:

Ali Chaibakhsh Ali Ghaffari


University of Guilan Khaje Nasir Toosi University of Technology
68 PUBLICATIONS 351 CITATIONS 253 PUBLICATIONS 3,023 CITATIONS

S.A.A. Moosavian
K. N. Toosi University of Technology
259 PUBLICATIONS 2,648 CITATIONS

Some of the authors of this publication are also working on these related projects:

Soft (Continuum) Robotics View project

Power-Water Management View project

All content following this page was uploaded by Ali Chaibakhsh on 04 July 2018.

The user has requested enhancement of the downloaded file.


https://www.sciencedirect.com/science/article/pii/S1569190X07000731

doi:10.1016/j.simpat.2007.06.004

A simulated model for a once-through boiler by


parameter adjustment based on genetic algorithms
Ali Chaibakhsh *, Ali Ghaffari, S. Ali A. Moosavian
Department of Mechanical Engineering, K.N. Toosi University of Technology, P.O. Box 19395-1999, Tehran, Iran
Received 25 November 2006; received in revised form 7 May 2007; accepted 13 June 2007
Available online 17 August 2007

Abstract

In this paper based on the physical rules, thermodynamics principles and energy mass balance, the simulation models
are developed and applied to electrical power generating plants in order to characterize the essential dynamic behavior of
the boiler subsystems and to use the corresponding models for the power plant processes. These models are developed for
a sub-critical once through Benson type boiler based on the experimental data obtained from a complete set of field exper-
iments. An optimization approach based on genetic algorithm (GA) is executed to estimate the model parameters and fit
the models response on the real system dynamics. Comparison between the responses of the corresponding models with
the response of the real plants validates the accuracy and performance of modeling approach. A similar comparison
between the responses of these models with linear parametric models shows the effectiveness and feasibility of the
developed model in term of more accurate and less deviation between the responses of the models and the
corresponding subsystems.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Power plant; Once-through boiler; Mathematical model; Genetic algorithm

1. Introduction

Facing enormous demand for electricity and growing need for more and safer power generating has moti-
vated investigation into dynamic analysis of power plants to design more reliable control systems. Better sys-
tem performance means increase in power generation efficiency and also decrease in the maintenance costs.
To design suitable controllers, adequate information about the system dynamics is required.
The boiler–turbine modeling has a wide application in power plant control and process study. The dynam-
ics of most power plants are highly nonlinear with numerous uncertainties. Thus, no mathematical model can
exactly describe such a complicated physical process, and there will always be modeling errors due to un-
mod- eled dynamics and parametric uncertainties. Besides, detailed modeling of plants dynamics is often not
effi- cient for control synthesis. The plant model should describe the plant dynamics with sufficient accuracy
and not describe the microscopic details occurring within individual plant components [1].

*
Corresponding author. Tel.: +98 21 886 748 41; fax: +98 21 886 747 48.
E-mail address: chaibakhsh@dena.kntu.ac.ir (A. Chaibakhsh).

©
103 https://www.sciencedirect.com/science/article/pii/S1569190X07000731
0

Nomenclature

C specific heat (kJ/kg K)


h specific enthalpy (kJ/kg)
¯h absolute enthalpy (kJ/kmol)
H calorific value (LHV)
(kJ/kg) k index of expansion
m mass (kg)
m_ mass flow (kg/s)
m¯ molecular weight (kg)
M accumulative mass (kg)
P pressure (MPa)
q mass flow rate
Q heat transferred (MJ)
R ratio
t time (s)
T temperature (°C)
u specific internal energy (kJ/kg)
v specific volume (m3/kg)
V volume (m3)
W power (MW)
z space

Greek symbols
a steam quality
g efficiency
q specific density (kg/m3)
s time constant (s)
n heat absorbing coefficient

Subscripts
a metal
act actual
air air
f feedwater
fur furnace
fuel fuel
in input
out output
p constant pressure
prod product
reac reactants
s steam
spray spray
v constant volume
w water
0 standard condition
mp metal part
sp steam part
There are complicated models based on finite element approximations to partial differential equations.
These models are in the form of large simulation codes for plant design, simulators and commissioning.
However, they are not normally used in control design approach because of their complexity [2]. The
analytical plant model can be formulated based on the fundamental laws of physics such as mass conservation,
momentum, and energy semi-empirical laws for heat transfer and thermodynamics state conversion. To build
such analytical models, it is necessary to define their parameters with respect to boundaries, inputs, and
outputs. Generally, the devel- oped models need to be tuned by performing tests to validate for steady-state
and transient responses [2–5].
In addition, a mathematical plant model can be developed based on the measured data obtained from real
performance of the plant. The gathered information from experiments can be used to develop test-data-based
models by using system identification techniques. In this case, it is possible to obtain black box models of rea-
sonable complexity that describe the system well in specific operating conditions [2]. The four main steps to
determine a test-data-based model from input–output data includes; collection of data from experiments,
choosing a model structure, estimation of model parameters and model validation [6].
Soft computing methods can be used to optimize model parameters over a full range of input–output data.
In recent years, genetic algorithm (GA) is widely used as an optimization method for training and adaptation
of system parameters [7–10].
Genetic algorithms have many advantages over the conventional optimization methods. It does not require
a complete system model and can be employed to globally search for the optimal solution [11]. When the
iden- tified model is nonlinear in the parameters, using conventional methods like standard least squares
technique will not provide superior results. In these cases, genetic algorithm methodologies are investigated as
potential solutions to obtain good estimation of the model parameters [12–14].
In this paper, the mathematical models with unknown parameters for subsystems of a once through Benson
type boiler are first developed based on the thermodynamics principles and energy balance. Then the related
parameters are either determined from constructional data such as fuel and water steam specification, or by
applying genetic algorithm (GA) techniques on the experimental data. Finally, the responses of the corre-
sponding models are compared with the responses of the real plant in order to validate the accuracy and per-
formance of the models.
In the next section, a brief description of the plant is presented. It consists of a general view of the power
plant and its subsystems. Inputs and outputs to the subsystems are specified in this section. It follows by the
analytical model development in Section 3. The training procedure of the proposed models based on the exper-
imental data is presented in Section 4. The next section presents the simulation results of this paper by com-
paring the proposed model responses with the actual plant responses. It also shows the comparison between
the proposed modeling approach and the linear parametric modeling responses. The last section is the conclu-
sion and some suggestions.

2. System description

A 440 MW unit is considered for identification and modeling approach. The steam generator of a sub-crit-
ical once through Benson boiler with a steam mass rate of 1408 ton per hour is shown in Fig. 1. A single fur-
nace with 14 bottom burners serves to heat all sections of the boiler.
The hot water is converted to steam in evaporator, and is superheated by passing through superheaters. The
evaporator lies in the area of the higher flue gas temperature and is supplied as a spiral belt. The evaporator
outlet temperature is about 365 °C. The superheater consists of four sections built in the boiler second pass.
The heating surface of superheater #1 is located in the boiler walls. The steam leaves superheater #1 outlet
header, and goes toward superheater #2. The outlet steam temperature from these sections is not constant
but at normal condition, is about 407 °C. These sections have an emergency temperature control that works
just when the boiler is overhauling.
The steam leaves superheater #2 towards superheater #3. For rapid steam temperature control, four spray
attemperators are located between the two sections. In the de-superheater mixing path, nozzles inject atomized
feedwater into the steam path, thereby, reducing the steam temperature. The spray flow is taken from feed
water pumps and should not normally exceed approximately 6% of the boiler load at the time. The mean water
spray flow rate from each nozzle is 10 ton per hour.
Fig. 1. Boiler cross section.

In this case, the outlet temperature should be kept constant at 470 °C. The steam for final superheating
stage goes to superheater #4. The temperature at outlet header of this part should be constant at 535 °C.
At full load condition, the outlet superheated steam pressure 18.1 MPa at the boiler outlet header.
The superheated steam from main steam header is fed towards the high pressure (HP) turbine, and
from high pressure turbine is discharged into the cold reheated header. The steam temperature in cold
reheat line is 351 °C. The reheater consists of two sections. There are four attemperators between these
sections. The reheater outlet temperature is controlled in the same way to the superheater. The outlet
reheated steam tem- perature after RH-A should be constant at 452 °C and for RH-B should be constant
at 535 °C. Also, at the full load condition, the outlet reheated steam pressure is 4.35 MPa.
The reheated steam is used to feed the intermediate pressure (IP) turbine. Exhaust steam from IP turbine is
fed into the low pressure turbine. Also the extraction steam from IP turbine is fed into the feedwater turbo
pump and feedwater storage tank (de-aerator). Four low pressure heaters are fed by the extraction steam from
LP turbine. The low quality low pressure steam from all sections is discharged into the main condenser. This
condenser is an open loop condenser using the seawater to cool up the hot water and steam.
The condensate water is pumped into the feedwater storage tank via a train of LP heaters. A feedwater
turbo pump with 13.2 MW power (or two electrical pumps with 9 MW power when the turbo pump fails)
develops compressed feedwater at 28 MPa pressure for boiler. Then, the feedwater is fed into the high
pressure heaters and its temperature is increased about 80 °C. The hot feedwater inlet into the economizer
header and the power generation cycle is repeated.
3. Analytical model development

The boiler system can be decomposed into smaller components that can be analyzed and modeled sepa-
rately. The behavior of the subsystems can be captured in terms of the balance equations and constructive
equations. Variables that account for storage of mass, energy and momentum have to be introduced as well
as parameters [15,16].

3.1. Boiler model

The steam generating unit of Fig. 1 consists of economizer, evaporator, superheaters and reheaters. In the
boiler sections, the heat released by fuel combustion is transferred to the working fluid in the boiler. Based on
this, each section of the boiler can be considered as a thermal system as shown in Fig. 2. According to the
global mass and energy balances we have,
d
½qs · V s þ qw · V w ] ¼ m_ f — m_ s ð1Þ
dt
d
½qs · us · V s þ qw · uw · V w þ ma · C p · T a ] ¼ Q þ m_ f hf — m_ s hs ð2Þ
dt
The right hand side of Eq. (2) represents the energy flow to the system from fuel and feed water and the energy
flow from the system via the steam. Since the internal energy is u = h— p/q, the global energy balance can be
written as
d
½qs · hs · V s þ qw · hw · V w — pV þ ma · C p · T a ] ¼ Q þ m_ f hf — m_ s hs ð3Þ
dt
Multiplying Eq. (1) by hw and subtracting the result from Eq. (3) we have,
dqs dhs dhw d dp
ðhs — hw ÞV s þ qs · V s þ qw · V w þ ½ma · C p · T a ] — V
dt dt dt dt dt
¼ Q þ m_ f ðhw — hf Þ — m_ s ðhs — hw Þ ð4Þ
It is shown that for the drum boiler, the changes in energy content of the water and metal masses are the phys-
ical phenomena that dominate the dynamics of the boiler [2]. The specific density of steam is not changing so
fast, particularly in steady-state condition, to have a great effect on the model dynamics. This fact is true espe-
cially in the evaporator part where there are water and steam phases. So, these terms can be neglected with
respect to other terms. Also, the last term at the left hand side of Eq. (4), V, is often neglected in modeling [17]. For
simplicity, it can be taken m_ in m_ out¼. This means that the storage of mass in the control volume is
neglected. It will be shown that the effect of steam-water fraction should be considered for this part. There- fore
Eq. (4) is rewritten as follows:
dhw d
qw · V w þ ½ma · C p · T a ] ¼ Q þ m_ in ðhw — hs Þ ð5Þ
dt dt
This equation can be used for all subsections of the boiler as a general governing equation.

Fig. 2. Boiler subsystem.


3.1.1. Superheater model
For modeling the superheater parts, it should be noted that only the steam phase is presented in these
subsystems. Also, in once-through boilers, the pressure change is only a function of the feedwater flow rate.
Let;
. Σ¼ v

Cp o h ð6Þ
oT p ; C
v
o
¼. Σ T
where ou
dh oh dT
¼
dt ð7Þ
dt
oT
Then by some additional simplifications for the superheater parts, we have;
m

zhfflfflffl·fflfflfflVfflfflfflfflffl fflfflffl fflfflfflsfflfflfflfflfflfflffl} | dT out


s þd q½ a pdþq a· V
fflfflmfflfflfflfflfflfflffl·fflfflCfflfflfflfflfflfflffl·fflfflTfflfflfflffl{] · C p ¼ Q þ Cp m_in ðTin — out Þ
s ð8Þ
s
dt dt s d
T
t
The simulation results of models with detailed representation of the temperature distribution in the metal
show that the metal temperature at steady-state condition is close to the steam temperature. Also, it is shown
that by removing the first two terms on the left side of Eq. (8), there is an off-set between the model response
and the actual plant response at the steady-state condition. In order to make the model response close to the
response of the real plant, these terms should be taking into account. Noting that the specific density is of
steam approximately is constant. Therefore, we can express the second term on the left side of Eq. (8) as a
function of mass flow rate.
d
½ma · C p · T a ] ¼ f ðm_ in Þ ð9Þ
dt
It is assumed that the left side of Eq. (9) is a linear function of inlet steam flow rate. This approximation
is good enough to fit the model response with the real experimental data.
f ðm_ in Þ ¼ k a · m_ in ð10Þ
The same approximation can be considered for the evaporator part.
Therefore, Eq. (8) is captured as follows:
dT out
¼ 1 1 1
dt q s Q þ m_ in ðT in — T out þ k a =C p Þ þ k 0 ð11Þ
| · V Cp qs · V
fflffl{zfflffl
s s
} K
2

This equation is rewritten as follows:


dT out Q
¼ K . þ m_ ðT — T þ B Þ þ B Σ ð12Þ
2
dt C
p in in out 1 2
where
K ¼1 B; ¼ ka ;B ¼kq·V ð13Þ
2 1 2 0 s s
qs · V s Cp
The heat supplied to the surfaces, Q can be derived from the heat transfer equations. Both convection and
radiation are taken into account in the heat flow rate from combustion gas to the surfaces. To develop suitable
models for the combustion gas, it is required to know the constructional information of the system (the ther-
mal capacity of combustion gas, the dynamic of heat transfer and etc.) [18].
As shown in Fig. 1, the superheater sections are located in the second pass and this causes these sections not
to be directly exposed to the flame. So, the radiative heat transfer plays a secondary role and the convection is
the dominating terms in the expression. It is more convenient to find a relation between the heat flow and the
fuel consumption instead of combustion gas parameters. The released heat from fuel combustion can be
derived as a function of fuel flow or fuel–air ratio.
The heat flow can be captured by using calorific value/lower heating value (H) of the fuel as follows:
Q ¼ H · m_ fuel ð14Þ
In this case, by considering K1 = H/Cp, the superheater model is derived as;
dT out
¼ K 2 ðK 1 m_ fuel þ m_ in ðT in — T out þ B1 Þ þ B2 Þ ð15Þ
dt
The proposed model for the superheater section is presented in Fig. 3.

3.1.2. Evaporator model


The evaporator section is the most sophisticated part of the model. The distribution of steam and water in
the evaporator depends on the heat supplied to the surfaces. The dynamics variable in the heated surfaces is
shown by Astrom and Bell for risers of a drum type boiler [2]. In the same way, these relations can be used to
obtain the dynamics of steam generator.
The specific internal energy for the steam and water mixture is as follows:
h ¼ a · hs þ ð1 — aÞ· hw ð16Þ
where a is the quality of the mixture. The mass and energy balances for the heated surface are:
oq oqh 1 oqh Q
A
oq
þ ¼ þ ¼ ð17Þ
0;
ot oz ot A oz V
In the steady-state condition, it is shown that the quality of the mixture can be obtained by the following
equation:
QAz
a¼ ð18Þ
qhcV
In this approach, it is focused on the steam quality at the evaporator outlet. The simulation results show that
the steam quality can be expressed as a linear function of the fuel consumption;
a ¼ f ðm_ fuel Þ ¼ K 3 · m_ fuel ð19Þ
The steam quality in evaporator section has a direct effect on the temperature of down-stream sections. In
once-through boilers, the outflow from the evaporator passes through a separator to extract the steam phase
from the mixture. Therefore, the steam flow rate will depend on the quality of the mixture at the evaporator
outlet.

Fig. 3. Temperature model for superheater sections.


103 https://www.sciencedirect.com/science/article/pii/S1569190X07000731
6
In order to ensure complete combustion and also reduction of heat losses, controllers of the burners adjust
the fuel/air ration and keep the excess air within a suitable range. Thus, if the heating systems work properly,
the effect of water in the steam generator can be neglected (a = 1).
It should be noted that according to Eq. (19), the quality of the steam is a function of the fuel flow rate.
The temperature of each subsystem of a boiler is affected by both fuel flow rate and feedwater flow rate.
For an ideal situation, when the combustion is perfectly done and the feedwater flow rate is adequately
adjusted, the outlet of the evaporator is uniquely steam (a = 1). However, when the fuel flow rate and the
feedwater flow rate either increase or decrease, the model should be able to represent their variations on
the response of the boiler.
The evaporator model can be proposed with the respect to Eqs. (15) and (19). The temperature model
for the evaporator section is presented in Fig. 4. It is mentioned that this structure is valid for reheater
sections.

3.1.3. Combustion model


The combustion model can be developed base on the chemical reaction [19]. However, such a model is not
directly used in all of the proposed models but it would be useful for designing the fuel/air control system [20].
In dealing with chemical reacting systems, the concept of absolute enthalpies is very important. The absolute
enthalpy is the sum of enthalpies that takes into account the energy associated with the chemical bonds [19].
D¯hðT Þ ¼ ¯hðT Þ — ¯h0 ðT ref Þ ð20Þ
A balanced relationship for any fuel air system (which can be lean, rich or stoichiometric mixture) is written
as,

C H þ cða b b
a b þ O2þ 3:76N 2Þ ! aCO 2 þ H O2 þ dN þ þf þ gCO þ · · · ð21Þ
Þð 2 eO NO
4 2 2 x

The definition of the enthalpy of reaction or the enthalpy of combustion (heating value), DhR is
DhR ¼ hprod — hreac ð22Þ

Fig. 4. Temperature model for the evaporator section.


1037

This value can be adjusted for, per mass of fuel basis. So,
DhR ðkJ=kgfuel Þ ¼ DhR =m¯ fuel ð23Þ
Also, it is in turn converted to a per-unit-mass-of-mixture basis as,

Dh ðkJ=kg . . R 1 R
ðkJ=kg mfuel ¼. ΣDh ð24Þ
R Þ¼ Þ¼
Σ Σ
mfuel
mix Dh
m
R
mfuel Dh
fuel
þ air A=F þ
mix
m 1
The combustion heat is calculated by assuming that all of the water in the products has converted to gas it is
called the lower heating value (LHV) of fuel.
The furnace pressure Pfur which corresponds to combustion air pressure can be taken apart by Eq. (25).
Dh ¼ Du þ vDP fur ð25Þ
The kinetic energy of air flow has not a considerable role and can be neglected. The important feature is the
furnace pressure control which affords additional energy savings. In the best condition, the air pressure should
be close to the product pressure. Operating under negative pressure or at high positive pressure is lead to a
significant fuel wasting and damaging the boiler.
The amount of absorbed heat in a boiler from burners depends on the type of the boiler and its subsystems.
Besides, the thermal efficiencies of these sections are different. The effect of these aspects can be introduced
by a coefficient n as follows:
Q ¼ K 3 .m_ . ΣΣ·
ðA=F Þ com þ—1 ð
fuel · K2 K 1 P air — P Þ þ B1 ð26Þ
ðA=F Þsto þ 1 por

where B1, K1, K2 = DhR and K3 = f are the parameters of model. The proposed model for the combustion sys-
tem is presented in Fig. 5.
It should be noted that for modeling the boiler subsystems, the delay time is needed to be considered. This
delay time is defined as the time required for burning the fuel in the furnace, transferring the released heat and
absorbing the transferred heat by fluid in the boiler subsystems. The delay time is an important factor that
affects the dynamics of boiler. The delay times of boiler subsystems are presented in Table 1.

3.1.4. Flow–pressure model


Next, the outlet pressure of the boiler can be calculated based on the fluid dynamics principles. Naturally,
the steam pressure will drop by passing through the boiler sections such that in the HP section of this boiler,
the pressure loss is about 6 MPa. The main pressure drop is observed where vaporization takes place. The
evaluation of water pressure drop during transition boiling is very difficult and proposed models for this case
have complexity and need information about the constructional property of tubes such as inner diameter,
length, kind of surface and etc. [21]. In the steam side, the pressure loss due to change in flow velocity is

Fig. 5. Combustion system model.


Table 1
Delay Times for boiler subsystems
Delay (min)
Outlet pressure 0.6
Re-heater 1.2
Fuel firing 0.2
Economizer 2.5
Pre-heater 4.0
HP-heater 3.0
Evaporator 0.8
SH #12 0.9
SH #4 0.9
SH #3 1.0

prevailing. So, by neglecting the other effects, the pressure loss model can be captured based on the steam
areas effect. The mass of accumulative volume is
dM
¼ m_ in — m_ out ð27Þ
dt
The total accumulative action of any control volume in boiler consists of accumulative action of steam, hot
liquid and metal parts. Therefore, differential equation for the total accumulative action can be written as
follows:
oM oMsp oM wp oM mp
oP þ þ
¼ oP oP ð28Þ
oP
Noting that the effects of second and third terms on the right hand side of Eq. (28), comparing with the first
term, are negligible. We have,
oM oMsp
oP
¼ ð29Þ
oP
In this sliding pressure Benson type boiler, the pressure differences are the driving forces for mass flow through
the components. The change of steam mass due to the pressure change is given by,
dM d 1
¼V · Σ Σ ð30Þ
dP dP vðP Þ
By considering the value of polytrophic exponent is nearly 1.0 (p Æ v = p0 Æ v0), If
dM
> 0 for dP > 0 ð31Þ
dP
Then, from Eqs. (27)–(30), the outlet pressure is obtained:
dP P0
¼
dt s · ðm_ in — m_ out Þ ð32Þ

mv
This equation can be used as a general relation for pressure drop between turbine stages.
A model for the mass flow responding to steam pressure changes is proposed by Borsi [18]. The swing of
main steam flow strictly relies on the change of steam pressure.
dm_ out dP
¼ m_
dt 2ðP out0 in0 ð33Þ
out0
—P Þ dt

Generally, in power plants, the turbine inlet flow is controlled by a governor control valve to response to the
grid frequency. So, when this valve is acting, there is an interaction between steam pressure and flow shown
by Eq. (34). When the control valve opening is completed, the pressure fluctuation is removed and the swing
of steam flow tends to zero. In this unit the load frequency has to be kept constant at 50 Hz. So, the steam flow
is limited by governor valve. The control valve acting impresses on outlet pressure and flow. In Fig. 6, the
flow–pressure model is presented (the transfer functions Gmp and Gpm are derived from Eqs. (32) and (33),
respectively).
https://www.sciencedirect.com/science/article/pii/S1569190X07000731 1039

Fig. 6. Mass flow–pressure model.

3.1.5. Attemperator model


The superheater and reheater temperatures must be kept constant at specific temperature. The spray attem-
perators is implemented between these sections to control outlet temperature. Furthermore, de-superheating
spray is used to achieve mixing between the superheated steam at the outlet of the preceding component (e.g.,
the primary superheater). The water spray is modulated by suitable valves. Because the attemperator has a
relatively small volume, the mass storage inside that is negligible. Therefore, the steady-state mass and energy
balances yield (see Fig. 7)
m_ in þ m_ spray ¼ m_ out ð34Þ
m_ in hin þ m_ spray hspray ¼ m_ out hout ð35Þ
During a normal operation, steam flow m_ in in the secondary superheater is imposed (over a wide band) by
the load controller, the specific enthalpy hin is determined by upstream superheater and hspray is nearly constant.
The inlet temperature of the second superheater, Tout is governed by the following equation [15]:

T_ 1 ð¯¯hin — m¯ in ð¯¯hin —
¼ hout Þ _ hspray Þ ð36Þ
m_
out T spray
C m¯m_
out out
DhC ¼ C m¯ þ —
in
p p out m¯out p out

This equation yields an accurate attemperator temperature model. However, a simple model can be used
in- stead of this equation based on thermal balance formula as follows:
m_ in T in þ m_ spray T
T ¼ ð37Þ
spray
out
m_ in þ out
m_
This completes the dynamics modeling of different boiler subsystems.
We precede this section with modeling the steam turbine.

Fig. 7. De-superheater spray.


3.2. Turbine model

Many linear and nonlinear models are proposed for condensing steam turbine [22,23].
Assuming that the steam expansion in turbine is an adiabatic and isentropic process, a simple model can be
estimated for turbine. For an ideal gas we have,

T out .
¼ k— 1
T in out k
ð38Þ
Σ P in
P
The generated power in turbine is obtained as [24];
. Σðk— 1 — 1!
Wt¼ C
in p ðT þ ÞP P
out k
ð39Þ
in
gthm_ 273Þ
in

The output pressure can be evaluated from the transfer function presented by Eq. (40).

GPO a
¼ e—ss ð40Þ
I bs þ 1
The HP-Turbine model is presented in Fig. 8.
The parameters of this model are calculated using the collection of input–output data. The IP-Turbine and
LP-Turbine have more complexity. The steam pressure and mass flow consistently drop and the effect of
steam condensation is obvious between each extraction stage. Eqs. (32), (39) and (40) should be combined to
derive a model for these sections [25]. It is mentioned that the presented models for turbines are accurate when
it has no frequency control. In this test case, the Unit Load Control is not responding to network frequency
changes. A superior model is presented for load frequency control by Andersson [22].
The proposed model for the once-through boiler and turbines is presented in Fig. 9.
In the next section, the related parameters of the developed analytical models are trained by genetic algo-
rithm optimization techniques using the experimental data from power plant.

4. Parameter adjustment by genetic algorithm

In this section the parameters of the developed models are evaluated and adjusted by using genetic algo-
rithm (GA) training approach.

Fig. 8. HP-Turbine model (s = 0.45 (s), a = 0.2714, b = 0.401 and k = 1.352).


https://www.sciencedirect.com/science/article/pii/S1569190X07000731 1041

Fig. 9. Overall once-through boiler and turbine model.


Genetic algorithm is specifically useful for those systems working under different operating conditions with
different behavior. The GA determines the multiple next searching points using the evaluation values of multi-
ple current searching points. The next searching points are determined by using the fitness values of the
current searching points which are widely spread throughout the searching space. In conventional optimization
meth- ods, like gradient methods, in some cases the gradient information causes searching being trapped at a
local minima point. The GA has the mutation operator to escape from these possible trapping points. These
advan- tages cause that GA is preferred to other methods.
For GA training, the error E is given by the mean value of squared difference between the target output
y*
and actual output y as follows:
P
E¼X . m Σ2
1P p¼1 y p — y ðpÞ ð41Þ

where P is the number of entries used for training process.


The flow chart of the training procedure is illustrated in Fig. 10.
The training procedure of superheater model parameters is illustrated in Fig. 11. As it is shown, the training
procedure of the proposed models consists of three phases. Noting that the response of the model is most sen-
sitive to the variation of parameter K1 on Eq. (15), in the first stage, this parameter is adjusted. In the second
stage, the additive parameters B1 and B2 in Eq. (15) which are used to remove the offset of the model response
are adjusted. At the third stage, the parameter K2 in Eq. (15) is tuned. Note that, this parameter depends on
the structural specification of the system.

Fig. 10. Flow chart of the training procedure.


1043
https://www.sciencedirect.com/science/article/pii/S1569190X07000731

Fig. 11. Training procedure for superheater models.

In order to increase the converging speed of model parameters, it is necessary to initialize the parameters
for the normal operation of the boiler. Thus, the initial parameters of models are determined based on the
steady-state conditions of the plant during the period of operation.
At the steady-state condition from Eq. (15) we have,
K 1 m_ fuel þ B1 m_ in þ B2 þ m_ in ðT in — T out Þ¼0 ð42Þ
where m_ in , m_ fuel and Tin are the values of inlet steam mass rate, inlet fuel mass rate, inlet and outlet steam tem-
perature at the steady-state, respectively.
Fig. 12. The actual power output at five different set points between 195 and 440 MW.

The parameters K1, B1 and B2 are to be calculated. For the range of operation between 195 and 440 MW
and for three middle set points at 60%, 75% and 90% of load, the response of the actual plant, based on the
experimental data, is shown in Fig. 12. Therefore, Eq. (42) is solved for these unknown parameters at three
steady-state conditions. By solving these three set of linear equations, we obtain the initial values of param-
eters used in the model training process.
The next step for developing models, is gathering appropriate data for different subsystems of the boiler. In
order to have persistent excitation data, the control systems are switched to manual control. The experimental
data are recorded for 24 h. These data include the transient and steady-state conditions. Although the param-
eters are adjusted based on a limited volume of the collected data, but the selected data are well spread and the
developed models are exploited in a wide range of load variations. The models are trained with the data when
the load is ramped down from 100% to 45%, and they are checked with different set of data when the load is
ramped up from 45% to 100%.
The developed models for the boiler subsystems are continuous time invariant systems while the recorded
data are discrete time signals with five seconds sampling time. For adjusting the model parameters, the trans-
fer function should be converted to discrete systems. However, using GA for adjustment the parameters has
also the advantage that such conversion is not necessary. The parameters adjustment is executed through
a set of 2000 points of boiler data. The models training process is performed by using Matlab Genetic
Algorithm Toolbox and is simulated by Matlab Simulink. Some adjusted parameters of the developed models
are presented in Table 2. The optimization parameters for the proposed models in this paper are presented in
Table 3.

5. Simulation results

In this section, the trained models are simulated by using Matlab Simulink. In order to validate the
accuracy and performance of the modeling approach, a comparison between the responses of the proposed
models and the responses of the real plants is performed. The responses of different subsystems for steady-
state and tran- sient conditions are shown in Figs. 13–19.
In addition, we define the error as the difference between the response of the actual subsystem and the
response of the model. Then, we evaluate the quantities such as norm errorkek , upper bound of its absolute
e , lower bound of its absolute value (r(e)), error mean value (M(e)) and error covariance (C(e)) for
value (r¯ð ÞÞ
both models and for each subsystem as illustrated in Table 4.
The responses of superheater parts are shown in Figs. 13–15. Simulation results show that the responses of
the superheater models are very close to the response of the real system over a given operating range.
https://www.sciencedirect.com/science/article/pii/S1569190X07000731 104
5
Table 2
Parameters for temperature model
B1 B2 K1 K2 K3
Superheater 12-1 —0.8667 3.743 e–1 0.8654 1.012 e–2 –
Superheater 12-2 —0.8741 3.746 e–1 0.8591 1.012 e–2 –
Superheater 12-3 —0.8522 3.742 e–1 0.8604 1.012 e–2 –
Superheater 12-4 —0.8608 3.745 e–1 0.8638 1.012 e–2 –
Superheater 31 —0.325 2.331 e–1 0.5512 1.034 e–2 –
Superheater 32 —0.332 2.450 e–1 0.5427 1.034 e–2 –
Superheater 33 —0.345 2.421 e–1 0.5483 1.034 e–2 –
Superheater 34 —0.329 2.331 e–1 0.5532 1.034 e–2 –
Superheater 41 —0.165 1.353 e–1 0.3073 1.075 e–2 –
Superheater 42 —0.113 1.373 e–1 0.3057 1.075 e–2 –
Superheater 43 —0.114 1.371 e–1 0.3086 1.075 e–2 –
Superheater 44 —0.102 1.343 e–1 0.3085 1.075 e–2 –
Reheater a-1 0.181 —2.01 0.8051 1.062 e–2 8.978 e–6
Reheater a-2 0.182 —2.01 0.8060 1.062 e–2 8.978 e–6
Reheater a-3 0.182 —2.01 0.8060 1.062 e–2 8.978 e–6
Reheater a-4 0.181 —2.01 0.8051 1.062 e–2 8.978 e–6
Reheater b-1 0.007 1.100 e–1 0.4112 1.071 e–2 5.198 e–6
Reheater b-2 0.017 1.100 e–1 0.4120 1.071 e–2 5.198 e–6
Reheater b-3 0.014 1.100 e–1 0.4120 1.071 e–2 5.198 e–6
Reheater b-4 0.013 1.100 e–1 0.4112 1.071 e–2 5.198 e–6
Evaporator —0.022 2.330 e1 0.9000 1.000 e–2 9.451 e–6
Economizer 0.047 2.67 0.3000 1.000 e–2 –
Combustion 0.168 – 3.8 e–3 9.167 e–4 1.305 e–2

Table 3
Optimization parameters for GA
Population size 50–150
Crossover rate 0.7
Mutation rate 0.1
Generations 100–300
Migration Forward, fraction: 0.2, int.: 25
Selecting Stochastic uniform
Reproduction Elite count: 2

Fig. 13. Response of the superheater #4.


Fig. 14. Response of the superheater #3.

Fig. 15. Response of the superheater #1 and #2.

Noting that in a power plant there are some cases where extra combustion air is required for a fast response
to load variation, it causes that the flame height be increased and the evaporator temperature be decreased.
This phenomenon is considered in the proposed model for evaporator by using combustion system model.
The evaporator model response is shown in Fig. 16.
Noting that the economizer is located at the end part of the boiler, and its temperature depends on the
released heat transferring by the exhaust smoke. Therefore, to have more reliable model for economizer, it
is proposed to replace this signal by the fuel flow signal. The response of the economizer as shown in
Fig. 17 indicates the response of the developed model is very close to the response of the real system such that
the maximum difference between the response of the actual system and the proposed model is much less than
0.2%.
The pressure of the steam at the outlet header of boiler is shown in Fig. 18. Simulation results show that the
proposed model have a good agreement with real system response.
Fig. 16. Response of the evaporator.

Fig. 17. Response of the economizer.

The load response over an operation range between 80% and 100% of nominal load is also shown in Fig. 19
to illustrate the behavior of the turbine-generator system. The results show the accurate response of the tur-
bine to the main steam pressure changes.
In a sliding pressure boiler, the phenomenon like shrink and swell effect which occurs immediately when
the boiler turns off or when the boiler is in the start up phase (under 35% load). In these cases, the dynamics of
once-through boiler is the same as a drum boiler. The start-up control valves are activated to cope with these
conditions. So, it is more suitable to develop a different model for these conditions.
In order to illustrate the advantages of the presented modeling approach, its responses based on the exper-
imental data, are also compared with the response of linear parametric models. The parametric models of
superheaters 3 and 4 executed based on the output error (OE) method (Fig. 20) [15].
Figs. 21 and 22 show that the responses of presented models in this paper are significantly closer to the
actual system responses.
Fig. 18. Response of the boiler outlet pressure.

Fig. 19. Response of the turbine-generator.

Table 4
Comparison of modeling error
kek r¯ðeÞ r(e) M(e) C(e)
Superheater #4 61.9912 2.5348 3.9034 e–3 —0.3746 0.0619
Superheater #3 49.3874 2.4774 8.7138 e–4 0.3153 0.0339
Superheater #1& 2 44.6287 2.1961 1.5771 e–6 —0.0032 0.1048
Evaporator 112.9793 2.5844 1.6260 e–5 0.1890 0.6361
Economizer 34.3713 0.6254 3.6294 e–5 0.1341 0.0443
Pressure 96.1058 3.7451 4.5965 e–5 —0.1559 0.4618
Reheater a 48.9887 0.9864 6.4280 e–5 —0.1492 0.1010
Reheater b 92.9899 2.8173 1.6691 e–4 0.4582 0.2851
Power 39.0707 1.1846 3.1572 e–5 0.1029 0.5203
HP-heater 85.3634 1.1911 3.7309 e–5 —0.0108 0.3834
1050 https://www.sciencedirect.com/science/article/pii/S1569190X07000731

Fig. 20. Parametric temperature model for the superheater.

Fig. 21. Response of the superheater #4.

Fig. 22. Response of the superheater #3.


Table 5
Comparison of modeling methods
Presented method model Parametric model
Accuracy Good accuracy due to model adjusting based
Good accuracy but it is needed to have good filtering on training data
on experimental Data
and choose a proper order for model
Reliability Validated for a wide range of operation Validated for normal operating conditions and has more uncertainty for
abnormal conditions
Model Low order High order
order

Table 5 illustrates a comparison between these two modeling approaches.

6. Conclusions

In this paper, based on the physical and thermo-dynamical laws, different parametric models are developed
for boiler subsystem and for the turbine in a steam power generating plant. The genetic algorithm is executed
as an optimization method to adjust the model parameters based on the experimental data. The comparison
between the responses of the corresponding models with the responses of the plant subsystems validates their
accuracy in the steady-state and transient conditions. These models can be used for design a proper temper-
ature control system for superheaters sections.
Simulation results show the effectiveness and feasibility of the developed model in term of more accurate
and less deviation in the corresponding subsystems.
Comparison between the responses of the developed models with that of a model obtained based on the
output error (OE) method indicates the accuracy and fitness of the proposed modeling approach.
These models can be improved for the abnormal conditions such as start-up and shutdown modes. It
should be considered in the model by taking into account the heat exchange and combustion process in the
furnace. For these conditions, it is more suitable to develop a different model for these conditions.
The presented models for turbines are accurate when it has no frequency control. In the future works, the
responses to gird frequency can be considered in the turbine model. This can be used to design unit load con-
trol and synchronization systems with respect to network frequency.
The further model improvements will make them proper for using in simulators and emergency control
system.

References

[1] C.K. Weng, A. Ray, X. Dai, Modeling of power plant dynamics and uncertainties for robust control synthesis, Application of
Mathematical Modeling 20 (1996) 501–512.
[2] K.J. Astrom, R.D. Bell, Drum-boiler dynamics, Journal of Automatica 36 (2000) 363–378.
[3] S. Lu, B.W. Hogg, Dynamic nonlinear modeling of power plant by physical principles and neural networks, Journal of Electrical
Power and Energy Systems 22 (2000) 67–78.
[4] F.P. De Mello, Boiler models for system dynamic performance studies, IEEE Transaction on Power Systems 6 (1991) 66–74.
[5] L. Changliang, L. Jizhen, N. Yuguang, L. Weiping, Nonlinear boiler model of 300 MW power unit for system dynamic performance
studies, International Symposium on Industrial Electronics 2 (2001) 1296–1300.
[6] L. Ljung, System Identification – Theory for the User, 2nd ed., Prentice Hall, Upper Saddle River, NJ, 1999.
[7] Y.L. Karnavas, D.P. Papadopoulos, AGC for autonomous power system using combined intelligent techniques, International
Journal of Electric Power Systems Research 62 (2002) 225–239.
[8] S.V. Wong, A.M.S. Hamouda, Optimization of fuzzy rules design using genetic algorithm, Journal of Advances in Engineering
Software 31 (1999) 251–262.
[9] H. Ishibuchi, K. Nozaki, N. Yamamoto, H. Tanaka, Selecting fuzzy if then rules for classification problems using genetic algorithms,
IEEE Transaction on Fuzzy Systems 3 (1995) 260–265.
[10] R. Dimeo, K.Y. Lee, Boiler–turbine control system design using a genetic algorithm, IEEE Transaction on Energy Conversion 10
(1995) 752–759.
[11] R. Horst, P. Pardalos, N. Thoai, Introduction to Global Optimization, 2nd ed., Kluwer Academic Publishers, Dordrech, 2000.
[12] P.J. Fleming, R.C. Purshouse, Evolutionary algorithms in control systems engineering: a survey, Control Engineering Practice 10
(2002) 1223–1241.
[13] G.J. Gray, D.J. Murray-Smith, Y. Li, K.C. Sharman, T. Weinbrunner, Nonlinear model structure identification using genetic
programming, Control Engineering Practice 6 (1998) 1341–1352.
[14] K.R. Vazguez, C.M. Fonseca, P.J. Fleming, Multi-objective genetic programming: a nonlinear system identification application,
Genetic Programming Conference 1 (1997) 207–212.
[15] C. Maffezzoni, Boiler–turbine dynamics in power plant control, Control Engineering Practice 5 (1997) 301–312.
[16] K.J. Astrom, K. Eklund, A simplified non-linear model for a drum boiler–turbine unit, International Journal of Control 16 (1972)
145–169.
[17] M. Denn, Process Modeling, Wiley, New York, 1987.
[18] L. Borsi, Extended linear mathematical model of a power station unit with a once-through boiler, Siemens Forschungs und
Entwicklingsberichte 3 (1974) 274–280.
[19] S.R. Turns, An Introduction to Combustion: Concepts and Applications, 2nd ed., McGraw-Hill Science, 2000.
[20] P. Profos, Dynamics of pressure and combustion control in steam generators, Sulzer Technical Review 37 (1955) 1–15.
[21] M.N. Dumont, G. Heyen, Mathematical modeling and design of an advanced once-through heat recovery steam generator, Journal of
Computers and Chemical Engineering 28 (2004) 651–660.
[22] G. Andersson, Dynamics and Control of Electric Power Systems, Lecture Notes in Electrical Engineering, Power Systems
Laboratory, ETH Zurich, 2003, pp. 57–90.
[23] D. Zˇ ivkoviæ, Nonlinear Model of the Condensing Steam Turbine, vol. 1, FACTA Universities Series, Mechanical Engineering, 2000,
pp. 871– 878.
[24] J.A. Rovank, R. Corlis, Dynamic matrix based control of fossil power plant, IEEE Transaction on Energy Conversion 6 (1991) 320–
326.
[25] A. Chaibakhsh, Neuro-Fuzzy Control in a Power Plant Subsystem, M.Sc. Thesis, Department of Mechanical Engineering, K.N.
Toosi University of Technology, 2004.

View publication stats

You might also like