You are on page 1of 153

Nuclear

Reactor
Dynamics and
Stability

Henryk Anglart

Nuclear Reactor Dynamics and


Stability

2011 Henryk Anglart


All rights reserved

Preface

his textbook is intended to be an introduction to nuclear reactor kinetics,


dynamics and stability for students of energy
engineering and applied sciences as well as for
I C O N K E Y
professionals working in the nuclear field. The
 Note Corner
basic aspects of transient behavior of nuclear
 Examples
reactors are presented with focus on how to solve
 Computer Program
practical problems.
 More Reading

The textbook is organized into four chapters and


each chapter is divided into several sections. Parts in the book of special interest are
designed with icons, as indicated in the table to the left. Note Corner contains
additional information, not directly related to the topics covered by the book. All
examples are marked with a pen icon. Special icons are also used to mark sections with
computer programs and suggested more reading.
Chapter one contains introductory information, such as classification of dynamic
systems and description of basic approaches to analyze such systems. Chapter two and
three are devoted to nuclear reactor kinetics and dynamics using the point-reactor
approximation. Chapters four is dealing with dynamics and stability of two-phase
flows in heated channels.

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

Table of Contents
Preface
1

Introduction ................................................................................................7
1.1

Purpose of the Book ....................................................................................... 7

1.2

Schematics of Basic Nuclear Systems ........................................................... 8

1.3

Historical Review ......................................................................................... 11

1.3.1
1.3.2

Reactor Transient Tests ...................................................................................... 11


Reactor Instability ............................................................................................... 11

1.4

Future Development .................................................................................... 12

1.5

Rudiments and Fundamental Approaches ................................................ 13

1.5.1
1.5.2
1.5.3
1.5.4
1.5.5
1.5.6
1.5.7

Types of Dynamical System ............................................................................... 13


Basic Terms Used in Dynamics .......................................................................... 14
Linear Stability Analysis .................................................................................... 16
Bifurcations......................................................................................................... 18
Time Domain Approach ..................................................................................... 19
Laplace Transform Approach ............................................................................. 24
Frequency Domain Approach ............................................................................. 26

Nuclear Reactor Kinetics .........................................................................33


2.1
2.1.1
2.1.2
2.1.3
2.1.4
2.1.5

2.2
2.2.1
2.2.2

2.3
2.3.1
2.3.2

2.4
2.4.1
2.4.2

2.5
2.5.1
2.5.2
2.5.3

Reactor Kinetics Models.............................................................................. 33


Delayed Neutrons ............................................................................................... 33
Derivation of Point Kinetics Equations .............................................................. 34
Equations for Six-Group Point Kinetics Model .................................................. 39
Equations for One-Group Point Kinetics Model ................................................. 40
Average Neutron Generation Time and Lifetime ............................................... 41

Normalized Point Kinetics Equations ........................................................ 41


Equilibrium Point of a Nuclear Reactor.............................................................. 42
Normalized Equations of Point Kinetics Model ................................................. 43

Solutions with Constant Reactivity ............................................................ 45


Solutions of Six-Group Point Kinetics Equations ............................................... 45
Solutions of One-Group Point Kinetics Equations ............................................. 48

Point Kinetics Model with Time-Dependent Reactivity ........................... 53


Small-Perturbation Approximation ..................................................................... 53
Numerical Solutions of Point Kinetics Equations ............................................... 55

Approximate Point Kinetics Models .......................................................... 59


The Prompt Jump Approximation ...................................................................... 59
The Prompt Kinetics Approximation .................................................................. 61
The Constant Delayed Neutron Source Approximation ..................................... 61
4

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

Nuclear Reactor Dynamics...................................................................... 67


3.1

Reactivity Feedbacks .................................................................................. 67

3.1.1
3.1.2
3.1.3
3.1.4
3.1.5

3.2

Influence of Fuel and Moderator Temperature on Reactor Operations ...............67


Doppler Effect .....................................................................................................69
Reactivity and Reactivity Coefficients ................................................................70
Fuel Coefficient of Reactivity..............................................................................72
Moderator Coefficient of Reactivity ....................................................................72

Point Dynamics Model of PWR ................................................................. 73

3.2.1
3.2.2
3.2.3
3.2.4
3.2.5

3.3

Derivation of Point Dynamics Equations ............................................................73


Nuclear Reactor at Equilibrium ...........................................................................75
Normalized Point Reactor Dynamics Model .......................................................76
Reactor Dynamics in Presence of Small Perturbations ........................................79
Frequency Domain Analysis of Point Dynamics Model .....................................81

Point Dynamics Model of BWR ................................................................. 84

3.3.1
3.3.2

Formulation of BWR Dynamics Model...............................................................84


Nuclear Reactor at Equilibrium ...........................................................................85

3.4

Nonlinear Effects ......................................................................................... 89

3.5

Nuclear Reactor Stability ........................................................................... 91

3.5.1
3.5.2
3.5.3
3.5.4

Open Loop Transfer Functions ............................................................................92


Closed Loop Transfer Functions..........................................................................92
Stability of Zero-Power Reactor ..........................................................................93
Stability of Closed Loop Systems ........................................................................94

Dynamics of Boiling Systems .................................................................. 99


4.1

Analysis of Two-Phase Flow Transients.................................................... 99

4.2

Flow Instabilities in Heated Channels ..................................................... 109

4.2.1
4.2.2
4.2.3
4.2.4
4.2.5

4.3
4.3.1
4.3.2
4.3.3
4.3.4

Classification of Instabilities .............................................................................109


Density Wave Oscillations ................................................................................111
Methods of Analysis of Two-Phase Flow Instabilities ......................................113
Frequency Domain Methodology ......................................................................113
Time Domain Approach ....................................................................................117

Instabilities in Heated Loops .................................................................... 119


Parallel Channel Instability................................................................................119
Multiple Parallel Channels ................................................................................121
Boiling Loop Stability .......................................................................................121
Heated Wall Dynamics ......................................................................................121

APPENDIX A LAPLACE TRANSFORMATION

125

APPENDIX B NYQUIST STABILITY CRITERION

131

APPENDIX C SELECTED STEAM-WATER DATA

135

APPENDIX D BOILING CHANNEL STABILITY MODEL

137

Chapter

1
1 Introduction

he dynamics and stability of engineering equipment influence their


economical, safe and reliable operation and for that reasons they are major
areas of concern. Dynamics of many systems is their inherent property, and
the systems are just designed to behave like that. There are however systems,
which are designed for stationary operations, but still, they will behave in a transient,
dynamical manner. To such systems belong nuclear reactors, which preferably should
operate at steady-state conditions during their whole lifetime. However, nuclear
reactors are subject to various disturbances, which cause their transient behavior. For
that reason it is necessary to investigate and understand their dynamical behavior.
Stability is another important property of engineering systems. Stability can be
expressed using the everyday experience, which suggests that stable systems are
inherently able to return to the initial stationary condition when subject to any external
perturbation. On the contrary, unstable systems will transit to some other, completely
new state, which significantly deviates from the initial one. Obviously, for technical
equipment which is typically designed and optimized to operate at given desirable
conditions, this type of behavior should be avoided. This is in particular true for
nuclear reactors. If their operational conditions significantly deviate from the nominal
ones, and the conditions undergo periodic oscillations, some important safety limits
can be compromised. Eventually, such unsteady, unstable behavior if not properly
mitigated could lead to reactor failures.

1.1 Purpose of the Book


This book is intended to give an introduction to analysis and modeling of the transient
behavior of nuclear reactors. Any transient behavior can be characterized by some
measure of the time scale, which usually is a period of time during which a given
transient parameter significantly changes (for example it changes e-folds), or undergoes
a repeatable oscillation. In this book the focus will be on such transients, which have
time scales much smaller than the reactor fuel cycle, however, much longer than the
time scales of the reactor noise. It means that changes of reactor operational
conditions due to fuel depletion or even due to poisoning will not be included in the
present book. Nor transients resulting from the turbulent fluctuations of inlet velocity
and temperature of coolant will be included.
The focus of this book is to elucidate the phenomena that govern the transient
behavior of nuclear reactors which are subject to external perturbations. The reactor
power response is investigated following sudden changes of such parameters as
coolant temperature and/or mass flow rate at the inlet to the reactor core, or sudden
7

C H A P T E R

I N T R O D U C T I O N

change of reactivity due to movement of control rods. For that purpose simplified,
practical computational models are developed.
To make the development of computational models as simple as possible, the Scilab
(www.scilab.org) computational environment is used. The Scilab program is freely
available and can be installed on various platforms. The package contains basic tools
required for model development such as: time domain solvers, available through
command ode(), frequency domain and transfer function tools applicable for linear
systems defined by command syslin(), and integrated plotting functions. The
programs developed in this book are freely available and can be downloaded from
www.reactor.sci.kth.se/downloads .

1.2 Schematics of Basic Nuclear Systems


Several different nuclear power systems have been developed in the past decades and
their detailed description can be found elsewhere[1-1]. For the purpose of this book only
two most wide-spread systems will be described and referenced: nuclear power plants
with Pressurized Water Reactors (PWR) and with Boiling Water Reactors (BWR).
A schematic of a PWR plant is shown in FIGURE 1-1. It contains two circulation
loops: the primary loop that carries coolant through the reactor core and the steam
generators, which are usually operating in a once-through manner. The primary system
is kept under high pressure (typically 15.5 MPa) to avoid boiling of coolant in the
reactor core. The secondary loop is used to produce steam in steam generators. The
steam flows through steam lines to turbines, where it expands and performes the shaft
work. After condensation in the condenser, it returns as feedwater to steam generators.

FIGURE 1-1. Schematic of a nuclear power plant with Pressrized Water Reactor.

A schematic of BWR plant is shown in FIGURE 1-2. This type of plant contains the
recirculation loop which is forcing coolant through the reactor core, and the main
loop, which is circulating steam from the steam dome in the reactor pressure vessel to
the turbines. After expansion in the turbine and performing the shaft work, steam is
condensed in the condenser and then returnes as feedwater back to the reactor
pressure vessel. Since steam is generated in the reactor pressure vessel, there is no need
for steam generators. Thanks to this feature the plant design is significantly simplified.
8

C H A P T E R

I N T R O D U C T I O N

However, the presence of the vapor phase in the reactor core influences the dynamic
behavior of the system.

FIGURE 1-2. Schematic of a nuclear power plant with Boiling Water Reactor.

To be operated, nuclear power plants require sophisticated control systems. A


simplified schematics of such systems used in PWRs and BWRs are shown in
FIGURE 1-4 and FIGURE 1-4, respectively. Typically the system contains several
dedicated controllers, such as the reactor power controller, the turbine controller, the
reactor pressure controller and the water level controller. Each of the controllers
collects plant signals as schematically shown in the figures.
Steam
Pressurizer

14

15 16

13

Pressurizer
pressure controller

Steam
generator

Steam-generator
level controller

7
Hot leg

Reactor

17
Cold leg

Feedwater

9
10

Pressurizer
level controller

Reactor power
controller

11

12

Preheater
1-17 Signals

FIGURE 1-3. Schematic of PWR control system. Sygnals: 1 neutron flux, 2 pressure in steam line, 3
control rod position, 4 steam pressure in pressurizer, 5 valve position, 6 heating power, 7 water
level in pressurizer, 8 hot leg temperature, 9 cold leg temperature, 10 valve position, 11 base
signal, 12 correction signal, 13 feedwater inlet mass flux, 14 water level in steam generator, 15
steam mass flow rate, 16 base signal, 17 valve position.

C H A P T E R

I N T R O D U C T I O N

FIGURE 1-4. Schematic of BWR control system.

The operation point of a nuclear reactor is often described in terms of the reactor
power and the core flow. A typical BWR power/core-flow map is shown in FIGURE
1-5. The map shows paths that are followed during the reactor start-up and shutdown.
It also contains exclusion areas which should be avoided during the reactor normal
operation. One such area indicated in the figure is the stability exclusion area. It is
characterized by relatively high reactor power and low core flow rate. In this region the
core is susceptible to instable behavior and thus the region should be avoided.

FIGURE 1-5. BWR power/core-flow map.

10

C H A P T E R

I N T R O D U C T I O N

1.3 Historical Review


1.3.1

Reactor Transient Tests

Prediction of transients in Nuclear Power Plants (NPP) is important from the safety
point of view. Such predictions give important information about plant behavior
during transients, and in particular, they can reveal potential risks for failures that could
lead to core damages.
Various transient tests have been performed in NPPs to learn about their behavior.
The purpose of such tests is twofold: (a) they give a direct insight into plant
performance, since major parameters of interests are measured and recorded; (b) they
provide a valuable database for validation of computational tools.
In April 1977 such transient test was performed in Peach Bottom 2 power plant with
General-Electric-designed BWR/4 reactor. Three turbine trip tests were performed at
different power levels. The tests were concerned with transients following a sudden
closure of the turbine stop valve. As a result of the valve closure, pressure waves were
generated in the steam lines and propagated with relatively little attenuation into the
reactor core. The induced core pressure oscillations resulted in significant changes of
void fraction in the core, which, in turn, caused oscillations of reactor power. The data
obtained in the Peach Bottom-2 tests were subsequently used in an International
Benchmark program, where several computational codes were used by various
organizations to predict the measurements[1-5].
A corresponding test for PWR was performed at Three-Mile-Island-1, with B&Wdesigned reactor. The analyzed transient was a main steam line break, which may occur
as a consequence of the rupture of one steam line upstream of the main steam
isolation valves. This event is characterized by significant space-time effects in the core
caused by asymmetric cooling and an assumed stuck-out control rod after the reactor
trip. The measured data were used for International Benchmark of various
computational codes[1-6].
1.3.2

Reactor Instability

Interest into nuclear reactor instability started growing after first recorded incidents
with Boiling Water Reactors (BWR). In 1978 the TVO-I reactor, which is ASEAATOMs constructed BWR, underwent self-sustained oscillations leading to the
reactor scram[1-3]. The oscillations had so-called out-of-phase mode, in which the mean
reactor power remained approximately constant, whereas one half radially increased
the power, and the other one decreased. Similar type of oscillations was observed in
the Caorso plant in 1984 and it was the first event of this type that was widely
described in the literature. However, only after the LaSalle event of March 1988[1-4] the
BWR dynamics and stability issues attracted the attention of authorities and wider
public. Shortly after this event (in 1990), an international workshop on BWR stability
held in Long Island, USA, gathered more than 100 participants, to share their
experience with modeling and analyzing the BWR dynamics and instabilities.
Until now several dozen instability events occurred worldwide: most of them taking
place in the United States, Europe and Japan. It has been observed that BWRs are
susceptible to the following types of instabilities:

11

C H A P T E R

I N T R O D U C T I O N

caused by out-of-tune controllers. This is a


malfunction of reactor control hardware and can be relatively easy removed by
adjusting the controller gains.

Channel thermal-hydraulic instability,

Coupled neutronic-thermal-hydraulic instability,

Control system instability

in which a boiling channel can


oscillate even without neutronic feedback, if the local pressure drops become
out-of-phase with the inlet flow perturbations due to the density-wave effects.
also called reactivity
instability, in which the density wave with void fraction variation is affecting
the reactivity and power feedback.

While the first and the second type of instability can be easily removed or occurs only
locally in blocked channels, the reactivity instabilities are major concern in BWRs.
Several dedicated stability test have been performed in nuclear power plants. In the
period from 1989 to 1997 such measurements were performed in Forsmark 1 and 2
reactors. The measured data were subject to an International Benchmark program
which started in 1999[1-7].
On October 26 1989 instability incident occurred in Ringhals-1 reactor, ASEAATOM-designed BWR. The incident occurred during reactor start-up, following the
planned summer outage. At power 75% and 3720 kg/s core flow the power started to
oscillate in a limit cycle at 0.5 Hz frequency. The amplitude of power oscillations
reached 16%. After few minutes the oscillations were stopped by partial scram
initiated by the operator. During measurements performed 6 hours prior to the
incident, the local power range monitor signals were recorded as being out of phase.
The largest noted phase difference was 130 degrees between the two halves of the
core. The measurements revealed a local decay ratio of above 0.9. Following the
incident, series of stability measurements were performed at Ringhals-1 reactor. In
total 41 state points were measured during 4 consecutive fuel cycles (from 14th to 17th).
The measured data were subsequently used to conduct an International Benchmark
program[1-8].

1.4 Future Development


The stability and dynamics of current nuclear reactors is quite well understood and
widely analyzed, providing basis for safe and economic operation of existing nuclear
power plants. However, the prediction tools require further development to assure
high accuracy and reliability. In particular, the coupling between neutronics and
thermal-hydraulics requires full three-dimensional treatment. Also, thermal-hydraulic
modules will need to resolve temperature and density distributions on scales which are
smaller then a single nuclear fuel pin.
The future reactor design development will focuse on new systems that are currently
under consideration, namely Generation III+ and IV reactors. The characteristic
feature of the former is the increasing presence of passive systems. Such systems, while
desirable from the safety point of view, may introduce undesirable, difficult to model,
instability features. The latter (re)introduce new type of coolants, such as supercritical
water, liquid metals (such as lead-bismuth and sodium) and various gases (e.g. Helium).
Some of the coolants (in particular supercritical water) undergo significant property
12

C H A P T E R

I N T R O D U C T I O N

variation, similar to those during evaporation, which may lead to unstable behavior of
the reactor cores. To analyze dynamics and stability of such system, a new generation
of simulation tools will be required, combining Computational Fluid Dynamics (CFD)
codes with fully three-dimensional treatment of neutronics.

1.5 Rudiments and Fundamental Approaches


The kinetics, dynamics and instabilities of nuclear reactors are investigated using three
fundamental approaches: the time domain approach, the Laplace transform approach
and the frequency domain approach. Each of these approaches serves different
purposes, as described in the following sections. The choice of the approach depends
to a large extend on the type of the dynamical system under consideration. Even
though a thorough classification is not available, some basic features of dynamical
systems can be identified, as shortly described below.
1.5.1

Types of Dynamical System

The dynamical systems can be quite loosely classified into several types:
Type 0: Static Systems
This type is given only for reference. As the name suggests, such systems are stationary
and do not change with time. Even though there are no systems in the nature that are
strictly static, this assumption is very often made to make the analysis simple.
Type 1: Solvable Deterministic Systems
Such systems are characterized by existence of solutions in analytical, closed forms.
The behavior of such systems can be strictly predicted at any time instant for given
initial conditions. Such systems typically result from linearization and simplifications of
certain existing real systems (such as a pendulum), serving as their mathematical
models.
Type 2: Non-solvable Deterministic Systems
Systems of Type 2 are usually described by non-linear differential equations that have
no analytical solutions in closed forms. The solution can be obtained by the numerical
integration only. The response of such systems can be predicted with required
accuracy, provided that a proper numerical approach is applied. An example of such
system is the van der Pol oscillator.
Type 3: Deterministic Chaotic Systems
Such systems are characterized by solutions that are very sensitive to the initial
conditions and which undergo sudden qualitative changes. Such systems are difficult to
model and the results of predictions are usually evaluated in terms of mean values and
variances. Many systems in the nature belong to this category, including the solar
system. Complex technical systems, such as nuclear reactors can exhibit chaotic
behavior as well. The best know example of the chaotic system is the Lorenz model of
natural convection in the atmosphere.
The Lorenz equations have the following form,
13

C H A P T E R

(1.1)

I N T R O D U C T I O N

dx
= 10 x + 10 y
dt
dy
= 28 x y xz
dt
dz
8
= z + xy
dt
3

Solution of the equations is shown in the xyz coordinates in FIGURE 1-6.

FIGURE 1-6. Solution of Lorenz equations in state coordinates xyz.

Type 4: Stochastic Systems


Stochastic systems have solutions that can be evaluated in terms of mean values and
variances only. Typically the equations describing the stochastic systems contain
coeeficients that can get random values.
1.5.2

Basic Terms Used in Dynamics

There are several terms that are used to describe the dynamical systems (and in
particular nonlinear dynamical systems) that are introduced and explained in this
section. The first important term is the dynamical system itself: in general it is a selfcontained entity that exhibits some temporal behavior. Such system described with
equations will constitute a model of the dynamical system.
is a variable or a list of variables (that is a vector) which
are needed to uniquely determine the system current state. For the Lorenz system
given by Eq. (1.1) the state is described by three variables: x, y and z. All possible states
of a system are called a phase space. That is for the Lorenz system the phase space
will be a three-dimensional space xyz.

State of a dynamical system

Dynamics or equations of motion for a dynamical system describe a relation between


the present and the future state of the system. For continuous dynamical systems,
equations of motion are differential equations which can be either linear or nonlinear
in nature, determining a linear or a nonlinear dynamical system, respectively.

14

C H A P T E R

I N T R O D U C T I O N

Not always the future state of the dynamical system is uniquely determined by the
present state. There are two important examples of such a situation, which apply to
nuclear reactors:

when the equations of motion explicitely depend on time (such systems are
called nonautonomuous systems). This can occur in a nuclear reactor with
insertion of a time-dependent reactivity,

when the system is stochastic.

A solution of the the equations of motion starting from a given state point (that state
point determines the initial conditions) is called an orbit or a trajectory. For
continuous systems described with differential equations it is a curve in the phase
space.
In some cases the orbits have special properties: if they start from a given point in the
space phase, they come arbitrary close and arbitrary often to this or another point in
the space phase. Sets of such points are called the non-wandering sets. There are four
types of such sets:
1.

Fixed points,

which correspond to stationary solutions of the equations of

motion.
2.

Limit cycles, which correspond to periodic solutions.

3.

Quasiperiodic orbits, which correspond to periodic solutions with at least


two incommensurable frequencies (i.e., the ratio of the frequencies is an
irrational number).

4.

Chaotic orbits, which correspond to bound non-periodic solutions.

The first three types of orbits may occur in linear systems, whereas the fourth appears
only in nonlinear systems.
A non-wandering set may be either stable or unstable. Changing a certain parameter
(this is a so-called controlling parameter) of the system can change the stability of the
non-wandering set. This is accompanied by the change of the number of nonwandering sets due to bifurcation. Change of stability and bifurcation always coincide.
There are two types of stability: a weaker (Lyapunov) and a stronger (asymptotic) one.
The Lyapunov stability (also called marginal stability) occurs when every orbit
starting in a neighborhood of a non-wandering set remains in its neighborhood. The
asymptotic stability possesses all properties of the Lyapunov stability and in addition
all orbits in the neighborhood approach the non-wandering set asymptotically. The
stability of non-wandering sets is discussed in more detail in Section 1.5.3.
Non-wandering sets with asymptotic stability are called attractors. The basin of
attraction is the set of all initial states approaching the attractor in the long time limit.

15

C H A P T E R

1.5.3

I N T R O D U C T I O N

Linear Stability Analysis

Very often the system under consideration is described with a set of non-linear
differential equations. To enable the analysis of the system with some of the well
established approaches (such as the Laplace transformation approach), the equations
have to be linearized.
Consider a system of non-linear ordinary differential equations,
(1.2)

dx
= F (x; u) , x(0) = 0 ,
dt

where x is the vector of unknown functions, u is the vector of forcing functions, and
F represents the right-hand-sides in the system of equations. The fixed point of the
system is represented by vector x0, which satisfies the equation,
(1.3)

F( x 0 ; u 0 ) = 0 .

The forcing functions u0 are either equal to zero or to given constant values. The
behavior of a dynamical system around any point (and in particular around the nonwandering sets) can be investigated using the linear stability analysis, also known as
the first method of Lyapunov. For that purpose the function F is Taylor-expanded
around the point x0 as follows,
(1.4)

F(x 0 + x ) = F (x 0 ) +

F
x + O x 2 .
x x 0

( )

Here

(1.5)

F1
x
1
F
F 2
= x
x 1
F
N
x1

F1
x2
F2
x2
FN
x2

F1
xN

F2
xN

FN
xN

is the Jacobian matrix.


Taking x which is close to x0, that is x = x0 + x and using Eq. (1.4) yields,
dx d (x 0 + x ) dx
F
=
=
F(x 0 ; u 0 ) +
dt
dt
dt
x

x
x0

Thus, the system behavior around the non-wandering set, and in particular, the
stability of the non-wandering set can be evaluated from the linearized differential
equation,

16

C H A P T E R

(1.6)

I N T R O D U C T I O N

dx
= Ax , with x(0) = 0 ,
dt

where
(1.7)

A=

F
x

.
x0

Assuming that the solutions of Eq. (1.6) have a form x(t ) = e e st , the equation yields,
(1.8)

A e = se .

This equation has a non-trivial solution for such values of s, which satisfy the following
characteristic polynomial,

(1.9)

A sI = 0 .

Here I is the unit matrix with dimensions NN. Assuming that Eq. (1.9) has N singlevalued roots, for fixed points, the solution of Eq. (1.6) has the following general form,
N

(1.10)

x(t ) = Ak e k e s t ,
k

k =1

where Ak are constant coefficients, ek are eigenvectors and sk are eigenvalues of the
Jacobian matrix evaluated at the fixed point. If the coefficients of the Jacobian matrix
are real, eigenvalues are real or complex conjugate numbers.
The solution given by Eq. (1.10) is valid only when x is small, it can be thus termed
as the small-perturbation approximation. This is a very general method and usually
gives some interesting conclusions concerning the behavior of dynamical systems. In
particular, when real parts of all eigenvalues of the Jacobian matrix are negative, the
fixed point is asymptoticallystable, as can be deduced from Eq. (1.10). Once analyzing
the stability of any system, it is thus important to determine its fixed points (or in
general, non-wandering sets) and the real parts of the eigenvalues of the Jacobian
matrix.
Analytical calculation of the roots of the characteristic function is tedious for
polynomials with the order greater than 2 and impossible for polynomials with the
order grater than 5. In such cases the Routh and Hurwitz theorem can be used to
evaluate the stability. This theorem states that the real parts of all roots of a polynomial
are negative if and only if:
(1.11)

1 > 0, 2 > 0, ... N > 0 ,

where,

17

C H A P T E R

(1.12)

I N T R O D U C T I O N

b1
b3
k = b5
M
b2 k 1

1
b2
b4
M

0
b1
b3
M

0
1
b2
M

b1
M

1
M

b2 k 2

b2 k 3

b2 k 4

b2 k 5

b2 k 6

L 0
L 0
L 0 ,
O M
L bk

and the characteristic equation is as follows,


(1.13)
1.5.4

A sI = s N + b1 s N 1 + L + bN 1 s + bN = 0 .
Bifurcations

A bifurcation manifests itself with a change of the number of attractors in a nonlinear


dynamical system. In a bifurcation point, at least one eigenvalue of the Jacobian gets a
zero real part. There are three generic types of so-called co-dimension-one bifurcations
(here the term co-dimension refers to the number of control parameters for which fine
tuning is necessary to get such a bifurcation). Two of them, which are relevant for
continuous dynamical systems, are shortly described below.
occurs, when a single real eigenvalue crosses the boundary of
stability (that is the imaginary axis). This case is depicted in FIGURE 1-7. In Hopf
bifurcation, a conjugated complex pair croses the boundary of stability, as shown in
FIGURE 1-8.

Stationary bifurcation

=Im(s)
Stability
boundary
=Re(s)

FIGURE 1-7. Occurrence of the stationary bifurcation in a continuous dynamical system when the real
eigenvalue crosses the stability boundary (shown with an arrow).

=Im(s)

Stability
boundary

=Re(s)

FIGURE 1-8. Occurrence of the Hopf bifurcation in a continuous dynamical system when the
conjugated complex pair eigenvalues cross the stability boundary (shown with two arrows).

18

C H A P T E R

I N T R O D U C T I O N

Crossing the boundary of stability indicates the bifurcation point, but it does not
indicate how many solutions bifurcate or disappear at that point. Further analysis is
necessary to determine the character of solutions after bifurcation points. This analysis
is the subject of the bifurcation theory and is not covered in this book.
EXAMPLE 1-1. Consider a dynamical system that is described by the following
set of two differential equations:

dx
= f ( x, y; ) = a + b x 2
dt
dy
= g ( x, y; ) = y
dt

Here x and y are the state variables and is the control parameter. There are two fixed points as follows:

(x, y )1 = ( a b 0) , or assuming a = b = 1, (x, y )1 = ( 0)


(x, y )2 = ( 0)
(x, y )2 = ( a b 0)
The Jacobian at both fixed points are as follows:
2
2 x 0
J ( x, y ) 1 =
=

0 1 1 0

0 ,

2
2 x 0
J ( x, y ) 2 =
=

0 1 2 0

The characteristic polynomials for the two fixed points are:


2 s

1 s

= 2 s (1 + s ) = s 2 + 1 2 s 2 and

2 s

1 s

= 2 + s (1 + s ) = s 2 + 1 + 2 s + 2

If = -1, the first fixed point has the characteristic polynomial s2 s 2 = 0 with roots s1 = -1 and s2 = 2.
That indicates an existence of a saddle point. At the same time the second fixed point has the
characteristic polynomial s2 + 3s + 2 = 0 with roots s1 = -2 and s2 = -1, which indicates an existence of an
attractor at that point. Similar analysis indicates that for = 0 there is a double fixed point at the origin
with eigenvalues s1 = 0 and s2 = -1, and for > 0 there are no real fixed points.


1.5.5

MORE READING: An excellent introduction to the bifurcation theory can be


found in an article by John David Crawford: Introduction to
bifurcation theory, Review of Modern Physics, Vol. 63, No. 4, October
1991. The article can be found at
http://prola.aps.org/thumbnail/RMP/v63/i4/p991_1?start=0.
Another good site to visit can be found at
http://www.egwald.ca/nonlineardynamics/bifurcations.php

Time Domain Approach

The time domain approach is based on the time integration of the model differential
equations. Taking a simple case of the first-order system described by the following
differential equation,
(1.14)

dx
+ g (t ) x = u (t ) ,
dt

19

C H A P T E R

I N T R O D U C T I O N

with the initial condition,


(1.15)

x(0 ) = x0 ,

the solution can be obtained in a closed form after integration as follows,


(1.16)

x(t ) = x0 e

0t g (t ' )dt '

+e

0t g (t ' )dt '

u(t ')e
0

g (t '' )dt '' dt ' .

In particular, taking constant function g(t) = 0=const and the step-function on the
right-hand side: u(t) = 0 for t<0 and u(t) = u0 for t0, yields:
(1.17)

x(t ) =

u0

(1 e ) + x e
0 t

0 t

u
+ x0 0 e 0 t ,
0
0
u0

If the forcing function has the form u(t) = A sin(t), the solution is as follows,
(1.18)

A
( sin t cos t ) .
x(t ) = 2
+ x0 e 0 t + 2
2
0 + 2 0
0 +

The first term on the right-hand-side of Eq. (1.18) decays with time, whereas the
second term represents the persistent response of the system to the sinusoidal forcing
function. As can be seen, the response signal will have a changed amplitude and phase
as compared to the forcing function, but will have the same frequency of oscillations.
The amplification factor and the phase shift can be obtained by transforming the
term as follows,

A
(0 sin t cos t ) = 2A0 2 sin t cos t =
2
+
0 +
0

2
0

A 0
(sin t + tan cos t ) = 2 A2 0
(cos sin t + sin cos t ) =
2
+
0 + cos

2
0

A 0
sin (t + )
+ 2 cos

2
0

Here,

(1.19)

= arctan

is the phase shift, and

(1.20)

A 0
A
= 2 0 2
2
2
0 + cos 0 +

1 + tan 2 =

A
2
0

+ 2
20

A 0
2
0 + 2

1+

2
=
02

C H A P T E R

I N T R O D U C T I O N

is the amplitude of the response signal. Thus, the system response can be written as,
A

A
x(t ) = 2
+ x0 e 0 t +
sin (t + ) .
2
2
0 + 2
0 +

(1.21)

For constant and positive 0, the first term on the right-hand-side of Eq. (1.21) decays
with time, whereas the second term represents the persistent response of the system.
The system amplification is usually given in decibels, using the amplification at zero
frequency as a reference,
L = 20 lg

(1.22)

response signal amplitude


.
response signal amplitude at = 0

Using the above definition, the amplification of the first-order system is,
L = 20 lg

(1.23)

1
1+ 2

where

(1.24)

FIGURE 1-9 shows the phase shift and amplification characteristics of the first order
system.
0

-10

-5

-20

-10

-30

-15

-40

-20

L [dB]

theta [deg]

-50

-25

-60

-30

-70

-35

-80

-40

-90
0.01

0.1

10

-45
0.01

100

beta

0.1

10

100

beta

(a)

(b)

FIGURE 1-9. Characteristics of the first order system: (a) phase shift, (b) amplification.

Similarly, for the second-order system describing the forced oscillations with viscous
dissipation the governing equation is as follows,

21

C H A P T E R

(1.25)

I N T R O D U C T I O N

d 2x
dx
+ c + kx = a sin t .
2
dt
dt

This equation can be transformed to the following form,


(1.26)

d 2x
dx
+ 2
+ 02 x = A sin t ,
2
dt
dt

where,
c
,
2m

(1.27)

(1.28)

0 =

k
.
m

The general solution of the equation in case of small damping ( < 0 ) and initial
conditions x(0 ) = x0 and x& (0 ) = x& 0 is as follows,

(1.29)

x& + x0

x(t ) = e t x0 cos t + 0
sin t

t
Ae
1

sin cos t + ( sin + 0 cos )sin t +

02 2 + 4 2 2

2
0

2 2

sin (t )
2

+ 4

The first two terms result from the perturbation introduced by the initial conditions
and they will decay with time, whereas the last term represents the system response
when time t goes to infinity. It can be seen that the system response signal will have the
same frequency as the forcing function, but the amplitude will be amplified with
factor,
(1.30)

2
0

+ 4 2 2

and the phase will be shifted with


(1.31)

2
2
2
0

= arctan

Using the definition of the system amplification given by Eq. (1.22) and the frequency
ratio defined in Eq. (1.24), the amplification and the phase shift are given as,

22

C H A P T E R

(1.32)

L = 20 lg

(1.33)

= arctan

(1 ) + d
2

I N T R O D U C T I O N

d
,
2
1

where
d=

(1.34)

is a dimensionless damping factor. The amplitude and phase shift characteristics are
shown in FIGURE 1-10.
40

d=0.1

d=0.1

-20

d=1

d=1

20

d=5

d=10

-40

0
L [dB]

theta [deg]

-60
-80
-100

-20
-40

-120

-60
-140

-80

-160
-180
0.01

0.1

10

-100
0.01

100

beta

0.1

10

100

beta

(a)

(b)

FIGURE 1-10. Characteristics of the second order system: (a) phase shift, (b) amplification.

An important parameter used in the evaluation of BWR stability is the decay ratio,
which is defined as the ratio of two consecutive amplitudes in a given signal, as shown
in FIGURE 1-11. The decay ratio can be calculated from the analytical solution as a
ratio of the system response at time t0 + T to the value at t0, where T is the period of
oscillations.
(1.35)

e ( t 0 +T )
DR = t 0 = e T .
e

The period of oscillations is obtained as


(1.36)

T=

thus, the decay ratio is as follows,

23

C H A P T E R

(1.37)

DR = e

I N T R O D U C T I O N

A1

DR=A2/ A1
A2

t0

t0+T

FIGURE 1-11. Definition of the decay ratio.


1.5.6

Laplace Transform Approach

Linear systems described with a set of linear ordinary differential equations with
constant coefficients can be investigated using the Laplace transform approach. The
definition of the Laplace transformation and a table with selected functions and their
images are given in APPENDIX A.
One of the most useful properties of the Laplace transformation is that it can be used
to derive the transfer function for dynamic systems. If any system is subject to a
forcing function u(t) and the response of the system is described with a function x(t),
then the system transfer function G(s) is defined as,
(1.38)

G (s ) =

x (s )
,
u (s )

where x (s ) and u (s ) are images (Laplace transformations) of functions x(t) and u(t),
respectively. Fundamental properties of system transfer functions are given in
APPENDIX A.
Transfer functions are very useful, since they a characterizing the system which they
are derived for. In particular, transfer functions can be used to investigate the signal
amplification, phase shift and system stability, once moving from the Laplace domain
to the frequency domain, as described in the next section.
In some cases transfer functions can be used directly for evaluation of the system
stability. This is the case when the transfer function can be expressed as a polynomial
quotient,
(1.39)

G (s ) =

N (s )
,
D (s )

24

C H A P T E R

I N T R O D U C T I O N

of two polynomials N(s) and D(s). The roots of the nominator polynomial N(s) are
called the roots of the transfer function, whereas the roots of the denominator
polynomial D(s) are called the poles of the transfer function. Thus, the transfer
function can be written as,
(1.40)

G (s ) = C

(s z1 )(s z2 )...(s zM ) .
(s p1 )(s p2 )...(s pN )

Transfer function G(s) describes a stable systems when all poles p1, p2, , pN have
negative real parts.
If all coefficients in polynomials that determine a transfer function are real, then the
poles of the transfer function will be either real or conjugate complex numbers. Their
possible locations for a stable system, using the s-plane, are shown in FIGURE 1-12.
Im(s)
s1

s3
3

s2

Re(s)
-

FIGURE 1-12. Locations of poles of transfer function on s-plane.

The location of poles on the s-plane suggests that the system response will be
described by the following function,
x(t ) = A1e 3t + Aet sin (t + ) .

Since 3<, the first term will decay with time faster than the second term. Thus, the
second term will dominate the system response. If both and are known, the decay
ratio for the system response signal can be found as,
DR

Ae (t 0 + T ) sin[ (t0 + T ) + ] T
= e = e 2 / .
t 0
Ae sin [t0 + ]

The method of Laplace transformation can be successfully applied to determine


dynamic system stability providing that all poles of the transfer function can be
determined. However, in many cases finding poles of the transfer function can be very
difficult or even impossible. This can occur when the transfer function is no longer
expressed in terms of a rational polynomial but instead it contains complex
transcendental functions. In such cases the properties of the dynamic system can be
investigated in the frequency domain, as described in the next section.
25

C H A P T E R

1.5.7

I N T R O D U C T I O N

Frequency Domain Approach

Similarly as the Laplace transformation approach, the frequency-domain method can


be used for linear systems only. If the system under consideration is non-linear, it
should be first linearized around a certain operational point.
A frequency response of any system is the system behavior subject to a sinusoidal
forcing function. Such forcing function can be expressed as,
(1.41)

u (t ) = u0 sin t ,

where u0 is the amplitude of the input signal (forcing function) and is the signal
frequency expressed in radians per second.
When a linear system is subject to a sinusoidal signal at input, the system response will
also be a signal of the same shape and frequency. However, the output signal will have
different amplitude and phase. Thus, the output signal can be written as,
(1.42)

x(t ) = x0 sin (t + ) .

Examples of input (forcing) and output signals for a linear system are shown in
FIGURE 1-13.

u0
x0

Tx
T
Output signal

Input signal

FIGURE 1-13. A linear system response to a sinusoidal forcing function at steady-state. The system
introduces a lag to the signal and reduces its amplitude.

The signal amplification factor has been previously defined as,


(1.43)

M=

x0
,
u0

and the phase shift is defined as,


(1.44)

Tx
2 , radians.
T
26

C H A P T E R

I N T R O D U C T I O N

It can be shown that the frequency response of a linear system can be obtained by
substituting s with j in the transfer function, and that:

the amplification factor is given as M = G ( j ) ,

the phase shift is given as = arg G ( j ) .

In addition, it can be shown that the amplification at steady-state is obtained as,


(1.45)

M 0 = lim G ( j ) = G (0 ) .
0

The system amplification can be expressed in decibels, as defined in Eq. (1.22). In


terms of the transfer function G(s), the amplification is as follows,
(1.46)

L = 20 lg

G ( j )
, decibels (dB).
G (0 )

The amplification factor and the phase shift can be plotted as a function of frequency
. These plots are so-called frequency characteristics or Bode characteristics.
The frequency approach can give an answer whether the system is stable or not
without specifically finding the poles of the transfer function. For that purpose the
Nyquist plot is used, in which the imaginary part of G(j) is plotted against the real
part of G(j). The principles of the Nyquist plot and the proof of the Nyquist stability
criterion is given in APPENDIX B.
For the first-order system described by the following differential equation,
(1.47)

dx
+ 0 x = u (t ) ,
dt

x ( 0) = 0

the transfer function is obtained after Laplace transformation of the equation,


(1.48)

sx ( s ) + x(0) + 0 x ( s ) = u ( s ) ,

which gives,
(1.49)

G (s) =

x ( s )
1
=
.
u ( s ) s + 0

The real and the imaginary parts of G(j) are readily obtained as,
G ( j ) =

1
( j + 0 )

=
= 2 0 2j 2
.
j + 0 ( j + 0 )( j + 0 ) 0 +
0 + 2

thus,

27

C H A P T E R

I N T R O D U C T I O N

Re[G ( j )] =

0
2
0

+ 2

Im[G ( j )] =

2
0

+ 2

The module and the argument of G(j) are now found as,
2

(1.50)



1
=
G ( j ) = Re G + Im G = 2 0 2 + 2
2
02 + 2
0 + 0 +
2

From Eq. (1.46) the system amplification is found now as,

(1.51)

L = 20 lg

G ( j )
= 20 lg
G (0 )

1
j + 0
1

= 20 lg

1
1+ 2

0
where = /0. Similarly, the phase shift is found as,
(1.52)

= arg G ( j ) = arctan

Im G
= arctan ( ) .
Re G

As can be seen, identical expressions have been obtained for L and as previously
derived in the time domain analysis.
The Bode plots for the first order system have been already shown in FIGURE 1-9.
The Nyquist plot can be obtained in an analytical form by expressing the imaginary
part of G(j) with its real part. This can be achieved as follows,
Im[G ( j )] =

2
0

= Re[G ( j )]
+ 0
0
2
0

Thus,

Im[G ( j )]
0
Re[G ( j )]

The imaginary part can be now expressed as,

Im G
Im G
0

1
Re G
Re G
Im G =
=
,
2
2
Im G
Im G 0
2
0 +
0
1+

Re G
Re G

or,
28

C H A P T E R

Re 2 G + Im 2 G =

Re G

I N T R O D U C T I O N

This equation represents a half-circle on the G-plane, which can be readily seen from
the following form of the equation,
2

1
1
Re G
+ Im 2 G =
.
20

20

The Nyquist plot for the first order system is shown in FIGURE 1-14.

ImG

1/20

1/0
=0

ReG

1/20

FIGURE 1-14. Nyquist plot for the first-order system.

The Bode and Nyquist characteristics can be plotted with dedicated Scilab functions
nyquist() and bode(), as shown in the Example below.
EXAMPLE 1-2. Perform the Nyquist and Bode plots for first-order system using
Scilab functions.
SOLUTION: Assume first order system with 0 = 1, that is the transfer function
is G(s) = 1/(s+1). To plot the Nyquist and Bode plots, the transfer function is first
defined as:
s = poly(0,s); G = 1/(s+1); h = syslin(c,G);
The plots are then generated as:
nyquist(h,0.001,100,First order system);
bode(h,0.001,100,First order system);
where frequency interval was chosen between 0.001 and 100 rad/sec. The obtained plots are shown in
the figures below.

29

C H A P T E R

I N T R O D U C T I O N

Nyquist plot
0.1

100

0.0

0.001

3.943

-0.1
0.025
Im(h(2i*pi*f))

0.849
-0.2

0.052

-0.3

0.439

-0.4

0.088

0.155

-0.5

-0.6
-0.2

0.0
0.2
First order system

0.4

0.6

0.8

1.0

1.2

Re(h(2i*pi*f))

FIGURE 1-15. Nyquist plot of the first order system with 0 = 1 obtained with the Scilab program.
Magnitude
0
-10

db

-20
-30
-40
-50
-60

-3

10

-2

10

-1

10

10

10

10

Hz
Phase
0
-10

degrees

-20
-30
-40
-50
-60
-70
-80
-90

-3

10

-2

10

-1

10

First order system

10

10

10

Hz

FIGURE 1-16. Bode plot of the first order system with 0 = 1 obtained with the Scilab program.
N O M E N C L A T U R E

A
A
d
DR
e
F
G
I
j
L
M
s
t
T

amplitude
matrix
dimensionless damping factor
decay ratio
eigenvector
Jacobian matrix
transfer function
unit matrix
imaginary unit, j = 1
system normalized amplification
amplification factor
Laplace transform parameter
time
period of oscillations
30

C H A P T E R

u
x

I N T R O D U C T I O N

input (forcing) signal


output (response) signal

Greek

frequency ratio
phase shift
frequency of the forcing function
system eigen frequency

Subscript
0
steady-state conditions

R E F E R E N C E S

[1-1]

Anglart, H., Applied Reactor Technology, Compendium, KTH, 2009.

[1-2]

Gialdi, E. et al., Core stability in operating BWR: Operational experience, Proc. SMORN-IV,
Dijon, France, Pergamon Press, Oxford, 1984.

[1-3]

Valtonen, K., RAMONA-3B and TRAB assessment using oscillation data from TVO-I, Proc.
Int. Workshop on BWR Stability, OECD-NEA CSNI Report 178, pp. 205-231, 1990.

[1-4]

NRC Bulletin 88-07 Supplement 1, Power Oscillations in Boiling Water Reactors, 1988.

[1-5]

NEA/NSC/DOC(2001)1, Boiling Water Reactor Turbine Trip (TT) Benchmark, Volume I:


Final Specification, February 2001.

[1-6]

NEA/NSC/DOC(99)8, Pressurized Water Reactor Main Steam Line Break (MSLB)


Benchmark, April 1999.

[1-7]

NEA/NSC/DOC(99)9, Forsmark 1 & 2 BWR Stability Benchmark, May 1999.

[1-8]

NEA/NSC/DOC(96)22, Ringhals 1 Stability Benchmark Final Report, 1996.

E X E R C I S E S

EXERCISE 1-1: A dynamical system is described by the following differential equation: &x& + 02 x = F (t ) .
Find the system response if F (t ) = const = F0 and at t = 0, x& = x = 0.
EXERCISE 1-2: A dynamical system is described by the following differential equation: &x& + 02 x = F (t ) .
Find the system response if F (t ) = at and at t = 0, x& = x = 0.
EXERCISE 1-3: A dynamical system is described by the following differential equation: &x& + 02 x = F (t ) .
Find the system response if F (t ) = F0e at and at t = 0, x& = x = 0.
EXERCISE 1-4: Perform the Nyquist plot for the second order system &x& + x& + 3x = 0 .
EXERCISE 1-5: Represent the second order system &x& + x& + 3x = 0 as a set of two differential equations.
Find the fixed point of the system and analyze its stability using the first method of Lyapunov.

31

C H A P T E R

I N T R O D U C T I O N

32

Chapter

2
2 Nuclear Reactor Kinetics

he power generated in a nuclear reactor depends on several parameters such


as the mass of the fissile material, the macroscopic fission cross section and
the level of the neutron flux. Nuclear reactor kinetics is dealing with transient
neutron flux changes resulting from a departure from the critical state. Such
situations arise during startup and shut-down of a reactor, or due to accidental
disturbances in the reactor steady-state operation. In this Chapter the point kinetics
equations are derived and solved. Various methods of solutions are applied and
compared.

2.1 Reactor Kinetics Models


2.1.1

Delayed Neutrons

The emission of delayed neutrons has significant consequences on the transient


behavior of reactors. Even though the fraction of delayed neutrons is small, it may play
a dominant role in the over-all reactor behavior.
Nearly all of the neutrons produced due to a fission process are emitted without a
noticeable delay after the fission. These neutrons are termed as prompt neutrons. The
prompt neutrons are emitted by the direct fission products, immediately after the
fission process.
Nuclei of some of the fission products may beta-decay into daughter nuclei which
then immediately emit a neutron. Out of roughly 500 possible fission products, there
are about 40 that possess this special property and they are called precursors of
delayed neutrons (in short: precursors). The process of emitting delayed neutrons is
schematically shown in FIGURE 2-1 .
X

It would be a tremendous task to track separately all 40 precursors in analyzing reactor


kinetics phenomena. In fact, the yield fractions and decay constants of all precursors
are not even known exactly. Therefore, it is customary to represent all precursors by
six groups (or families) of precursors. The yield fractions and decay constants of such
groups are obtained experimentally by exposing a sample of fissionable material to a
very short neutron pulse and then measuring the time behavior of the source of the
delayed neutrons. The results obtained for U-233, U-235 and Pu-239 are shown in
TABLE 2-1 .
X

33

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

fission

delayed
neutron

product

prompt
neutron

beta-decay
precursor
(A,Z-1)

emitter
(A,Z)

final
nucleus
(A-1,Z)

FIGURE 2-1. Schematic of emission of delayed neutrons.


TABLE 2-1. Characteristics of delayed fission neutrons in thermal fission.
Approximate half-life [s]
55
23
6.2
2.3
0.61
0.23
Total delayed
Total fission neutrons
Fraction delayed

Number of delayed neutrons per 100 fissions


U-233
U-235
Pu-239
0.053
0.060
0.024
0.197
0.364
0.176
0.175
0.349
0.136
0.212
0.628
0.207
0.047
0.179
0.065
0.016
0.070
0.022
0.70
1.650
0.630
249
242
293
0.0026
0.0065
0.0020

Energy
(MeV)
0.25
0.46
0.41
0.45
0.41
-

As in steady-state analyses of nuclear reactors, the complete neutron balance equations


for the time dependent neutron flux depend on the neutron location in the threedimensional space, the neutron direction vector and the neutron energy. For reactor
kinetics applications, however, it is neither feasible nor necessary to use the same level
of complexity. The equations can be condensed into purely time-dependent ones, socalled point kinetics equations. The name point kinetics merely indicates that all space
and angle dependencies in equations are neglected and only the time dependence is
left.
2.1.2

Derivation of Point Kinetics Equations

In the non-critical system the production and loss of neutrons are not in balance, and
in the absence of an independent neutron source, the system can not be in the steadystate condition. For the present purpose, the principle of neutron conservation can be
expressed in a general form as follows.
(2.1)

N
1
= (1 )F M + i TiCi + S ,
v t
i =1

(2.2)

Ci
= i FD iCi , i = 1, ..., N .
t
34

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

Equation (2.1) is a time and space-dependent neutron balance equation, where is


the neutron flux, Ci is the concentration of precursors in i-th group, S is the external
N

neutron source, = i is the total yield of the delayed neutrons and i is the
1

decay constant of i-th group of precursors. The time and space dependent
concentrations of precursors are given by Eqs. (2.2). The equations are given in general
forms, where operators F, FD, M and Ti represent the production of fission neutrons,
the production of the precursors of delayed neutrons, the migration and loss of
neutrons and the conversion from i-th group precursors to neutrons, respectively.
These operators can take various forms, depending on the particular model equations
that are employed. Typical choices include transport equations, diffusion
approximation equation or multi-group diffusion approximation equations. As an
example, for one-group diffusion approximation, the operators are as follows,
(2.3)

M = D(r ) + a (r ) ,

F = FD = f

Ti = 1.

The neutron flux can be factored as follows,


(2.4)

(r, , E , t ) = s (r, , E , t ) n(t ) ,

where s is the shape function and n(t) is the amplitude function. The amplitude
function is a function of time only, whereas the shape function is a function of the
same independent variables as the neutron flux, that is: r - the neutron position vector,
- the neutron direction vector, E - the neutron energy and t - time.
The amplitude function will have a simple physical interpretation if it is defined as,
(2.5)

1
n(t ) = (r, , E ) (r, , E , t )drddE .
v

Such definition suggests that the amplitude function is a mean weighted number of
neutrons in a nuclear reactor at time t. Here (r, , E ) is an arbitrary weighting
function. If (r, , E ) = 1 then n(t) is the mean number of all neutrons in the reactor
at time t. Using Eq. (2.4) in (2.5) yields,
1
n(t ) = (r, , E )S (r, , E , t ) n(t )drddE =
v
,
1
n(t ) (r, , E )S (r, , E , t )drddE
v

thus,
(2.6)

1
v (r, , E ) (r, , E , t )drddE = 1 .
S

Equation (2.6) is called the normalization condition of the shape function. Multiplying
the neutron balance equation (2.1) by the weighting function yields,
35

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

N
1

(
)

n
=
1

+
iTi Ci +S

S
S
S
t v

i =1

or,
N
1
dn
1

S
+ n S = (1 )nFS nMS + iTi Ci + S .
v
dt
t v

i =1

The last equation can be integrated over all neutron positions, direction angles and
energies. Using a shortcut notation dx = drddE , the integration yields,
N
dn 1
S dx = (1 )n FS dx n MS dx + i Ti Ci dx + Sdx .
dt v
i =1

Further, the equation is divided by a normalization factor F, that will be determined


later,
1
1
1 1
dn

= n (1 ) FS dx MS dx +
S dx
F
F

F v
dt
.
N
1
1
i Ti Ci dx + Sdx

F
F
i =1

Defining,
(t ) =

1 1
S dx ,
F v

yields,

(2.7)

dn n
1
1

= (1 ) FS dx MS dx +
dt
F
F

.
N
1
1
i
Ti Ci dx +
Sdx

F
F
i =1

In the similar manner, Eq. (2.2) is multiplied by

(2.8)

d 1
1
1

Ti Ci dx = i
Ti FDS dx i
Ti Ci dx .

dt F
F
F

Introducing the following definitions,


(2.9)

1
Ti and integrated to yield,
F

C i (t ) =

1
Ti C i dx
F

36

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

1
Sdx ,
F

(2.10)

S (t ) =

(2.11)

i (t ) = i

(2.12)

(t ) =

1
Ti FDS dx ,
F

1
FS dx ,
F

Eqs. (2.7) and (2.8) become,


(2.13)

dn n 1
N
= (F M )S dx + i Ci + S ,
dt F
i=1

(2.14)

dC i i
=
n i C i , i = 1, ... , N .
dt

The normalization factor F can be chosen in such a way that the coefficients in Eqs.
(2.13) and (2.14) will have a simple physical interpretation. If this factor is defined as,
(2.15)

F = FS dx ,

it can be interpreted as the mean weighted number of neutrons created in a reactor


due to fission per unit time at time t. In addition,

(2.16)

(F M )S dx (F M )dx
1
(F M )S dx =
=
.

F
F
dx
F
dx

As can be seen, the above term is equal to the reactivity, since it is a ratio of the net
neutron production to the total neutron production.
Finally, the point kinetics equations are as follows,
(2.17)

N
dn
=
n + i C i + S ,
dt

i =1

(2.18)

dC i i
=
n i C i , i = 1, ... , N .
dt

Equations (2.17) and (2.18) have been derived from the general neutron balance
equations without making any approximations. For this reason the equations are
termed as the exact point kinetics equations, where exact means that the time,
space, energy and direction dependent neutronics is taken into account and lumped
into the point kinetics equation. In particular, Eqs. (2.17) and (2.18) are equivalent to
initial full equations, if the effective kinetics parameters are calculated as given by Eqs.
(2.9) through (2.12).
37

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

To calculate the effective kinetics parameters it is necessary to chose proper shape and
weighting functions. From the mathematical point of view any well defined function
could be used as the weighting function. However, in practical applications the adjoint
neutron flux is chosen as the weighting function, that is,
(2.19)

( x ) = 0* ( x ) ,

where 0* ( x ) is a solution of the adjoint neutron balance equation (see Note Corner
below).
In addition, the shape function is taken equal to the neutron distribution function in
the critical reactor,
(2.20)

S ( x, t ) = 0 ( x ) ,

and the neutron creation operators will take the values for the critical reactor as well,
(2.21)

F = F0 ,

FD = FD 0 ,

where subscript 0 refers to the critical state of the reactor.


NOTE CORNER: Adjoint neutron flux has a simple physical interpretation.
Using the one-group diffusion approximation it can be easily shown that the
adjoint neutron flux at a given point r0 is proportional to the reactivity change
caused by an introduction or a removal of one neutron per second at that point.
Adjoint neutron flux is obtained from a solution of the adjoint neutron balance
equation. In such equation all operators are replaced with their adjoint forms. The
adjoint operator is defined by the scalar product of two functions and , each
out of the respective functional space: ( , ) = dx , where x represents in a short form all

independent variables of functions and , and the integration is performed over the whole region
where the functions are determined. For a given linear operator L, its adjoint operator L* is defined by
the following equality: ( , L ) = ( , L ) . If in addition L = L*, then the operator is self-adjoint. It can
be easily shown that if L is a constant, then L = L*. Also it can be shown that the Laplace operator is
self-adjoint; that is = * . For the first order derivatives, the following will be valid, however: (d/dt)* = d/dt.

If the assumptions (2.19) through (2.21) are adopted, the kinetics parameters are as
follows:
1

(2.22)

(2.23)

(x ) (x )dx
= v
,
(
)
(
)

x
F

x
dx

*
0

*
0

i = i

0 0

(x )T F (x )dx ,
(x )F (x )dx
*
0

*
0

D0 0

0 0

(2.24)

= i

i =1

38

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

As an example, the neutron balance equations with the one-group diffusion


approximation can be written as follows,
N
1
= (1 ) f + D 2 a + i Ci + S ,
v t
i =1

Ci
= i f i Ci , i = 1, ..., N ,
t

and since in steady-state, the time derivatives and the external neutron sources are
equal to zero, the equations yield,

f 0 + D 20 a0 = 0 .
Since all operators in the above equation are self-adjoint, the adjoint balance equation
is identical,

f 0* + D 20* a0* = 0 .
Further, since the boundary conditions for both equations are identical (zero flux at
the extrapolated boundary), both solutions are identical, with possibly a constant
multiplier c as follows,
(2.25)

0* = c0 .

For the critical reactor and the one-group diffusion approximation the operators are as
follows,
(2.26)

F0 = FD 0 = f ,

Ti = 1, i = 1, ... , N ,

thus,
1

(2.27)

(2.28)

v (x ) (x )dx v c (x )dx
1
=
=
=
(x )F (x )dx c (x )dx v
*
0

*
0

i = i

2
0

0 0

2
0

(x )T F (x )dx = c
(x )F (x )dx
c
*
0

*
0

D0 0

0 0

2
0

dx

2
0

dx

= i .

The reactivity could be found from Eq. (2.16), however, this would require the
knowledge of the neutron flux distribution. As an approximation, the perturbation
theory can be applied.
2.1.3

Equations for Six-Group Point Kinetics Model

Equations (2.17) and (2.18) are the fundamental equations of reactor kinetics for the
point-reactor model. In the following the overbar indicating the effective values of the
kinetics parameters will be dropped, unless it introduces confusion. It should be
X

39

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

remembered, however, that they represent some mean weighted quantities, expressed
by Eqs. (2.9) through (2.12). The six-group point kinetics model is summarized in
TABLE 2-2.
.

TABLE 2-2. Point reactor kinetics model.


Reactor kinetics equation
Delayed-neutron precursor equations: i =
1,,6

6
dn
=
n + i C i + S
dt

i =1
dC i i
=
n i C i , i = 1, ... ,6
dt

The values of decay constants and yields of delayed-neutron precursors are given in
TABLE 2-3 .
X

TABLE 2-3. Decay constants and yields of delayed-neutron precursors in thermal fission of uranium235.
Decay constants and yields of delayed-neutron precursors in thermal
fission of uranium-235
t1/2, [s]
, [s-1]

55.7
22.7
6.22
2.30
0.61
0.23

0.0124
0.0305
0.111
0.301
1.1
3.0

0.000215
0.00142
0.00127
0.00257
0.00075
0.00027
0.0065

Total

2.1.4

0.0173
0.0466
0.0114
0.0085
0.0007
0.0001
0.084

Equations for One-Group Point Kinetics Model

Some properties of the point kinetics model can be investigated using a one-group
approximation. The total yield of the one group of delayed neutrons is obtained as a
sum of yields in all groups. For uranium-235 this value is shown in TABLE 2-3 and is
equal to = 0.0065. The decay constant can be obtained from a proper averaging, e.g.,
X

(2.29)

= i =
i =1 i

i =1 i
6

Using data from TABLE 2-3, the equivalent decay constant for one-group assumption
for uranium-235 is obtained as = 0.0065 0.084 0.08 s-1.
The equations in one group approximation of point kinetics equations are as follows,
(2.30)

dn
=
n + C + S ,
dt

(2.31)

dC
= n C .
dt

40

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

Here C represents the concentration of precursors of all groups of delayed neutrons.


In the next section special cases of the point kinetics model will be considered and
their solutions will be found.
2.1.5

Average Neutron Generation Time and Lifetime

The average neutron generation time has been derived for the one-group
diffusion approximation in Eq. (2.27) and can be written in various forms as follows,
(2.32)

1
l l
1
= = =
.
v f
k k k v a

Here l is the average neutron lifetime and k is the effective multiplication factor. The
name generation time has been chosen since represents the average time between
two birth events in successive neutron generations. Firstly, 1 f is the mean free path
for fission, that is, it is the average distance a neutron travels from its birth to a fission
event. Then, (1 f ) v = t f is the average time between the birth of a neutron and a
fission event it may cause. Since -neutrons is released per fission, the averaged time
between the birth of a neutron and the birth of a new generation is as follows,
(2.33)

t f

1
=.
v f

In a similar manner, the average traveling distance of a neutron between the birth and
the death (absorption or leakage) is 1 ( a + DBg2 ) and the average neutron lifetime
can be obtained as,
(2.34)

l=

1
1
1
1
=
,
2
v a + DB g v a 1 + L2 B g2

since the one-group diffusion length is given as L = D a .


Eq. (2.34) yields,
(2.35)

l=

k
1
k
=
= k .
2 2
v a k 1 + L Bg v a k

Equation (2.35) indicates that the lifetime and the generation time are equal for a
critical reactor. For subcritical reactor (k < 1) the neutron lifetime is shorter than the
generation time and as a consequence, the neutron population will decrease. For
supercritical reactor the lifetime will be longer and the neutron population will
increase, whereas for a critical reactor the population remains constant.
X

2.2 Normalized Point Kinetics Equations


The point-reactor model equations constitute a system of ordinary linear differential
equations with variable coefficients, since the reactivity and other kinetics
41

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

parameters are in general - functions of time. For an arbitrary function (t ) it is thus


in general - not possible to obtain a rigorous analytical solution of the system of
equations. In such case it is necessary to employ approximate numerical solutions.
In the following the equilibrium of an operation reactor is investigated as well as
normalized equations of the point kinetics model are derived. Such equations are very
convenient since all initial conditions are suitably reduced to zero values.
It is convenient to represent the point reactor equations in a normalized excess form,
using some well defined reference condition as a reference. A reasonable choice is to
find the equilibrium (steady-state) parameters of the reactor and use them to derive the
normalized excess equations for point reactor kinetics. For that purpose it is necessary
to find the equilibrium point (that is the fixed point) of the model.
2.2.1

Equilibrium Point of a Nuclear Reactor

Equilibrium points of equations in TABLE 2-2 corresponding to the reactor steadystate condition can be obtained by equating to zero the time derivatives. The following
system of equations is obtained,
X

(2.36)

(2.37)

n + i Ci + S = 0 ,

i =1

n i Ci = 0 .

Substituting Eq. (2.37) to (2.36) yields,


X

(2.38)

n + i n + S = 0 .

i =1

Since
6

(2.39)

= i

i =1

Equation (2.38) becomes,


(2.40)

n + n + S n + S = 0 .

Equation (2.40) yields the equilibrium neutron density as,


(2.41)

ne =

Since ne cannot be negative, because this is the averaged weighted number of neutrons
in the reactor, the reactivity must be negative if there are neutron sources in the
reactor. It means that the equilibrium in the reactor with external neutron sources is
42

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

possible when the reactor is subcritical. In such situation the steady-state power of the
nuclear reactor is determined by the neutron source efficiency, the reactivity and the
mean neutron generation time.
When S = 0 (reactor has no external neutron sources), Eq. (2.41) yields:

n e = 0 .

(2.42)

This equation is satisfied in three different cases:


a)

= 0, ne 0 ,

which defines the critical state of reactor

b)

0, ne = 0 ,

reactor shut down condition

c)

= 0, ne = 0 ,

non-physical (critical reactor cannot have zero power)

The first equilibrium (or fixed) point can be chosen as the reference case for the
normalization of the point kinetics equations.
2.2.2

Normalized Equations of Point Kinetics Model

The power of an operating reactor can change in a very wide range, up to 1010 times.
Due to that it is useful to express the reactor power in terms of the steady-state power,
which must be larger than zero for a critical reactor.
Assuming that the mean average neutron density is ne, Eq. (2.37) yields,
X

(2.43)

Cie =

i ne
i

Introducing normalized excess variables,


(2.44)

x=

n ne
,
ne

(2.45)

yi =

Ci Cie
Cie

one gets
(2.46)

n = ne x + n e ,

(2.47)

Ci = Cie yi + Cie .

Substituting the normalized variables into the six-group point kinetics equations yields,
X

(2.48)

dx

1 6
S
= x + i yi + + + x ,
dt

i =1
ne

43

C H A P T E R

(2.49)

N U C L E A R

dyi
= i x i yi
dt

R E A C T O R

i = 1,...,6

K I N E T I C S

with the following initial conditions:


(2.50)

x ( 0) = 0 ,

(2.51)

yi (0) = 0 i = 1,...,6 .

Equation (2.48) and (2.49) together with the initial conditions given by Eqs. (2.50) and
(2.51) constitute the normalized point kinetics equations. They describe the
deviation of the reactor power from the initial, steady-state value. The model can be
easily implemented using the Scilab environment. A script containing the point kinetics
model is shown below.
X

COMPUTER PROGRAM: Point Reactor Kinetics Model

function [dy]= PKModel(t,y,yield,dconst,LAMBDA,ne)


// Point kinetics model
// dy - returned right-hand-side values
// t - time
// y - vector of unknown functions
// yield yields of the precursors
// dconst decay constants of precursors
// LAMBDA generation time
// ne neutron density at equilibrium
//
// Set parameters
//
NPre = length(yield);
// Get number of precursors
neq = 1 + NPre;
// Total number of equations
// Time functions
rho = InReactivity(t);
// Reactivity as a function of time
S
= InSource(t);
// Neutron source as a function of time
//
// Calculate Right-Hand-Sides of differential equations
//
dy = zeros(neq,1);
// Initialize RHS
dy(1) = (-sum(yield)*y(1) + yield*y(2:$) + rho + rho*y(1))/LAMBDA + S/ne;
for i = 1:NPre
dy(i+1) = dconst(i)*(y(1) - y(i+1));
end
endfunction

The point kinetics model PKModel() can be solved with the Scilab build-in solver for
the ordinary differential equations. For that purpose a dedicated solver script
PKODESol() has been implemented in Scilab. The model requires a specification of
two additional functions: InReactivity(t) and InSource(t) which describe the
time variation of the reactivity and the external sources, respectively. The usage of the
point kinetics model implemented in Scilab is elucidated in the example below.

44

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

EXAMPLE 2-1. A critical reactor operated during a long period of time with
constant power. At time t = 0 the reactivity was suddenly made 0.0022 positive.
Assuming = 10 3 s and one group of delayed neutrons with = 0.08 s 1 and
= 6.5 103 , predict the relative power change due to the reactivity insertion.
Make a plot power versus time in time interval from 0 to 5 s.
SOLUTION: Time dependent reactivity is given with the following function:

function [rho] = InReactivity(t)


if t < 0
rho = 0.;
else
rho = 0.0022;
end
endfunction

The vector of time instances at which the solution is found is given as:
->t=linspace(0,5,100);

Finally, the model is invoked as follows:


->PKODESol(t,6.5e-3,0.08,1.e-3,1);

The calculated time variation of the relative power as a function of time is shown in FIGURE 2-2.

1.9
1.8
1.7

Power

1.6
1.5
1.4
1.3
1.2
1.1
1.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Time, s

FIGURE 2-2. Relative reactor power change after sudden insertion of 0.0022 reactivity at time t = 0.

2.3 Solutions with Constant Reactivity


If the reactivity is constant, the model contains a set of linear ordinary differential
equations with constant coefficient and can be solved analytically. This condition is
valid in the case of the step change of reactivity from 0 to some finite (positive or
negative) value 0. Two types of solutions will be obtained and compared: using the
six-group and the one-group models.
2.3.1

Solutions of Six-Group Point Kinetics Equations

Point-kinetics equations can be easily solved for a step-change of the reactivity, when
the reactor operated at steady state before the step change. For t > 0 the reactivity is
then constant and Eqs. (2.48) with (2.49) are linear differential equations with constant

45

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

coefficients, which can be solved using, e.g., the Laplace transformation approach (see
APPENDIX A),

sx ( s ) =

(2.53)

syi ( s ) = i x ( s ) i y i ( s ) .

x ( s ) +

1 6
1 S ( s ) 0
x ( s ) ,
+
i yi ( s) + 0 +

ne
i =1
s

(2.52)

Here x ( s ) , yi ( s ) and S ( s ) are Laplace transforms of functions x(t), xi(t) and S(t),
respectively, e.g.,

(2.54)

x ( s ) = L{x (t )} = e st x (t )dt .
0

The system of algebraic equations formed by Eqs. (2.52) and (2.53) has to be solved to
find x ( s ) and yi ( s ) . From Eq. (2.53) one gets,
(2.55)

yi ( s ) =

i
s + i

x ( s ) .

Combining Eq. (2.55) with (2.52) yields,


(2.56)

1 6 i i 0
0 1 S ( s)

s
+

x
(
s
)
=
+
,

s
ne
i =1 s + i

or
(2.57)


1 6 i 0
0 1 S ( s )

.
s
1
+

x
(
s
)
=
+

s
ne
i =1 s + i

Finally,
(2.58)

x ( s ) =

1
1 S ( s )
+
,
sM ( s ) ne M ( s )

where
(2.59)

1 6
M ( s ) = s1 + i 0 .
i =1 s + i

To find x(t) it is necessary to apply the inverse Laplace transformation to Eq. (2.58)
and for that purpose the zero values of M(s) function have to be found. It can be
shown that all roots of function M(s) are real and their location can be conveniently
presented using graphical techniques.

46

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

Equation (2.59) can be written as:


(2.60)

1 6 i
M ( s ) = 0 s1 +
i =1 s + i

F ( s ) = 0 .

As can be seen, finding roots of M(s) is equivalent to finding crossing points of


function F(s) with 0 . FIGURE 2-3 demonstrates this approach.
X

(s )

s7

s6

s5

s4

s3

s2

s1

FIGURE 2-3. A graphical demonstration of the roots of M(s) function.

Equation (2.60) is called the inhour equation (which comes from inverse hour, when it
was used as a unit of reactivity that corresponded to e-fold neutron density change
during one hour) and is a 7th order algebraic equation with 7 roots. As shown in
FIGURE 2-3 , the first root is positive for positive reactivity. The root will change sign
and become negative when reactivity is negative. All remaining roots are always
negative.
X

When zero values sk of function M(s) are known, it is possible to find x(t) in a general
form using the following formula, which is known from the theory of the Laplace
transformation,
(2.61)

x(t ) =

1 0 1
1
L1
=
+
e sk t ,

sM ( s ) M (0) k = 0 sk M ( sk )

or shorter,
7

(2.62)

x(t ) = A + Ak e s k t ,
k =1

where
(2.63)

A=

1
, Ak = 0
, k = 1,...,7 .
M ( 0)
sk M ( sk )
47

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

When 0 > 0 then s1 > 0, sk < 0, k = 2, , 7. In such case the last terms in Eq. (2.62)
approach zero when time increases to infinity. In most cases the terms can be
neglected after relatively short period of time following the step change of reactivity
and only two first terms are significant. That is,
(2.64)

x(t ) = A + A1e s1t , s1 > 0.

When 0 < 0 then sk < 0, k = 1, , 7 and the last terms in Eq. (2.62) decrease
relatively fast compared to the second term, which also decreases with time. It means
that after short period of time Eq. (2.64) is approximately valid, however, now s1 < 0.
The reactor period, Tp, or e-folding time, is defined as the time required for the
neutron density to change by a factor e, that is,
(2.65)

n ( t ) = n0 e

t / Tp

Since roots sk have a unit of reciprocal time, Eq. (2.62) shows that for positive s1 (and
thus positive reactivity) 1/s1 is the reactor period after the laps of sufficient time to
permit the contribution of other terms to damp out. Consequently Tp = 1/s1 is called
the stable reactor period. The quantities 1/s2, 1/s3, , 1/s7 are sometimes referred to
as transient reactor periods but they are negative and have no physical significance.
If reactivity 0 is negative, all roots will be negative and s1 will determine the slowest
rate of change of the neutron density and thus will ultimately yield a stable (negative)
reactor period equal to 1/s1.
2.3.2

Solutions of One-Group Point Kinetics Equations

The general solution found in the previous section will be, for the sake of simplicity,
analyzed for a case when only one group of delayed neutron is assumed. Equations
(2.52) and (2.53) become,
X

sx ( s ) =

(2.67)

sy (s ) = x (s ) y (s ) .

x ( s ) +

y ( s)

(2.66)

0 1
s

S ( s ) 0
+
x ( s ) ,

ne

The averaged decay constant in Eqs. (2.66) and (2.67) is defined as,
(2.68)

i =1 i
6

and
6

(2.69)

= i .
i =1

Finding y (s ) from Eq. (2.67) and substituting into Eq. (2.66) yields,
48

C H A P T E R

0 1
x ( s ) =
(2.70)

N U C L E A R

R E A C T O R

K I N E T I C S

S ( s )
ne

s
=
0

s+

s+

(s + )

0 S0
(s + ) 0 + S 0
+ =


ne s (s s1 )(s s 2 ) ne
s s 2 + s 0 + 0

It has been assumed in Eq. (2.70) that the source term is constant and equal to S0,
whereas s1 and s2 are roots of the polynomial in the denominator and are as follows,
2

(2.71)

s1, 2

0 + 0 + + 4 0

=
.
2

The inverse Laplace transformation of Eq. (2.70) yields the required solution in the
time domain. It should be noted that Eq. (2.70) can be written as,
(2.72)

x ( s ) =

(s + ) 0 + S0 = A L( s) .
s (s s1 )(s s2 ) ne
sM ( s )

Using the formula given in APPENDIX A (see (A.6)), the inverse Laplace transform
is found as,
(2.73)

L( s)
L1 A
=
sM ( s )

L(0)
L( s1 ) s1t
L ( s 2 ) s2t
A
e +
e .
+

(
)
M
0
s
M
(
s
)
s
M
(
s
)

1
1
2
2

Finally,
(2.74)

S
s + s1t
s + s2t
x(t ) = 0 + 0
+ 1
e + 2
e ,
s2 (s2 s1 )
ne s1s2 s1 (s1 s2 )

and the neutron flux is obtained as,


n(t ) = ne x(t ) + ne =
(2.75)


S
s + s1t
s + s 2t .
ne 1 + 0 + 0
+ 1
e + 2
e
s2 (s2 s1 )
ne s1s2 s1 (s1 s2 )

EXAMPLE 2-2. The reactivity in a steady-state thermal reactor with no external


neutron sources, in which the neutron generation time is 10-3 s, is suddenly made
0.0022 positive. Assuming one group of delayed neutrons, determine the
subsequent change of neutron flux with time: = 0.08 s 1 , = 6.5 103 .
SOLUTION: The roots s1 and s2 are found as s1 = 0.03982 and s2 = -4.4198,
and the equation for the neutron flux becomes,

49

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

n(t ) = n e (1.4844e 0.03982 t 0.4844e 4.4198t ) .

EXAMPLE 2-3. A critical reactor operated during a long period of time and the
mean weighted number of neutrons was equal to ne = 106. Suddenly a source of
neutrons with mean weighted yield equal to S0 = 106 s-1 was introduced into the
reactor. Find the number of neutrons as a function of time, n(t). In calculations
assume one group of delayed neutrons with = 0.1 s 1 , = 6.4 10 3 and the
mean generation time = 10 3 s .
SOLUTION: The point-kinetics equations with one group delayed neutrons are as


follows,

dx
S

= x+ y+ 0 + + 0 x,
dt

ne
dy
= x y , with initial conditions, x(0) = y(0) = 0. Since the reactor was initially critical, the
dt

reactivity 0

= 0 . The source term is the following function of time,

0
S (t ) =
S 0

t<0
t0

Substituting the reactivity and the source term into the point-kinetics equations, and performing
the Laplace transform yields,
sx ( s ) =

x ( s ) +

y ( s )

S0 1 ,

ne s

sy ( s ) = x ( s ) y ( s ) .

Combining the above two equations yields,


x ( s ) =

S0

ne

s+

s2 s + +

To facilitate the inverse Laplace transform, the right-hand-side of the above equation will be
expressed in terms of basic functions as follows,
s+
a b
c
= + +

s s2

s+ +
s s + +

2
a s s + + + b s + + + c s 2 b + + s a + + b + (a + c )s

s2 s + +
s2 s + +

The constants a, b and c are found from the following equations,


a+c = 0,

a + + b = 1 ,

b + = .

50

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

And the constants are readily obtained as,


b=

a=

c=

Finally, the expression for x (s ) becomes,

1
1
S0

x ( s ) =
+

2
2
2
ne
s+ +
s + s

+
+

Applying the inverse Laplace transformation yields,

+
t

S0

x (t ) =
+
t
e
2
2

ne

+
+
+

and the number of neutrons is obtained as,

+
t


.
n (t ) = n e + S 0
t+
1 e
2

+

+

FIGURE 2-4

shows the variation of the excess normalized number of neutrons with time.

0.35

0.30

0.25

0.20

0.15

0.10

0.05

0.00
0

10

t, s

FIGURE 2-4. Excess normalized number of neutrons in a critical reactor as a function of time after
sudden insertion of a constant neutron source S = 106 s-1.
As can be seen the number of neutrons increases very fast during the first 0.5s of the transient and
then almost linearly, for time t larger than 1s. This is because the exponential term in the solution
is very small and the term linearly proportional to time prevails.

51

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

EXAMPLE 2-4: An under-critical reactor operated during a long period of time


with the reactivity equal to 0 = 9 . A source of neutrons with a mean weighted
yield equal to S0 = 106 s-1 was present in the reactor. Suddenly the source of
neutrons was removed from the reactor. Find the number of neutrons as a
function of time, n(t). In calculations assume one group of delayed neutrons with
= 0.1 s 1 , = 6.5 10 3 and the mean generation time = 10 3 s .

SOLUTION: The equilibrium number of neutrons before the removal of the source can be
calculated as,
ne =

S0

106 103
1.709 104
9 6.5 10 3

The solution of the point-kinetics model with one group of delayed neutrons have been found as,

S
s + s1t
s + s2t ,
n(t ) = ne x(t ) + ne = ne 1 + 0 + 0
+ 1
e + 2
e

n
s
s
s
(
s

s
)
s
1 1
2
2 (s2 s1 )
e 1 2

where the roots are found as,


2

s1, 2

0 + 0 + + 4 0

=
s1 = 0.09 , s2 = 65.01 .
2

Finally, taking S0 = 0, the solution becomes,

n(t ) = 1.709 10 4 0.1 e 0.09t + 0.9 e 65.01t .


X

FIGURE 2-5

shows the number of neutrons as a function of time.

18000

16000

14000

Neoutrons

12000

10000

8000

6000

4000

2000

0.05

0.1

0.15

0.2

0.25
Time, s

0.3

0.35

0.4

0.45

0.5

FIGURE 2-5. Number of neutrons in a reactor as a function of time after sudden removal of a constant
neutron source S0 = 106 s-1.
As can be seen the number of neutrons rapidly drops from initial 17090 just below 2000 and it
continues to drop, however with much smaller paste. FIGURE 2-6 shows the same function for a
longer time interval.
XX

52

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

18000

16000

14000

Neutrons

12000

10000

8000

6000

4000

2000

10

15

20

25
Time, s

30

35

40

45

50

FIGURE 2-6. Number of neutrons in a reactor as a function of time after sudden removal of a constant
neutron source S0 = 106 s-1 in a time interval from 0 to 50 s.
The rapid drop of the number neutrons during the first 0.05 s (prompt jump) is now hardly visible.
What can be seen is the slow drop of the number of neutrons from less than 2000 to 0 during
approximately 50 s.

2.4 Point Kinetics Model with Time-Dependent


Reactivity
In the previous Section the point kinetics equations were considered to be linear
equations with constant coefficients due to a special value of the reactivity: it has been
assumed that the reactivity is a step-function of time. In general case, when the
reactivity is an arbitrary function of time (and even of power, but this will be
considered in the next Chapter), the term (t )n (t ) in Eq. (2.30) will cause that the the
methods described in the previous Section can not be applied (in particular, the
Laplace transformation method will not work). To solve this type of equations other
methods should be used, such as the small-perturbation approximation method, or the
equations should be solved numerically. The former approach will be valid for small
perturbations only, but it will in general lead to an analytical solution that can give a
valuable insight into the behavior of the system under consideration. The latter
method will provide a solution which is valid for any perturbation (both small and
large), but the interpretation of the results is somewhat more complicated and requires
a generation of plots and diagrams which then can be analyzed.
2.4.1

Small-Perturbation Approximation

The exact point kinetics equations can be turned into a linear system of ordinary
differential equations with constant coefficients using a general approach presented in
the Introduction. In this section the approximate equations will be derived using the
perturbation method.
To get rid of the terms with the time-dependent coefficients, it is necessary to consider
the perturbed equations by assuming that all parameters deviate from the equilibrium
values by small amounts. That is, the following applies,

53

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

x(t ) = x0 + x(t )
(2.76)

yi (t ) = yi 0 + yi (t ) .

(t ) = 0 + (t )
Substituting Eq. (2.76) into (2.48) yields,

(2.77)

d [x0 + x(t )]
1 6
= [x0 + x(t )] + i [ yi 0 + yi (t )] +
dt

i =1
.
[0 + (t )] + S + [0 + (t )] [x + x(t )]
0

ne

The terms in Eq. (2.77) can be rearranged as follows,

(2.78)

dx0 dx(t )
1 6
1 6
+
= x0 x(t ) + i yi 0 + iyi (t ) +
dt
dt

i =1
i =1
.
0 (t ) S 0
(t )

(t )
+
+ +
x0 +
x0 + 0 x(t ) +
x(t )

ne

1
4243
small

The last term is a product of two small quantities which is negligibly small and will be
neglected. Since the terms with 0 index satisfy the initial equation, the above
equation can be written as,
(2.79)

dx(t )
1 6

(t ) (t )

= x(t ) + iyi (t ) +
+
x0 + 0 x(t ) .

i =1

dt

In a similar manner, the equation for the precursors of delayed neutrons is


transformed as,
(2.80)

d [ yi 0 + yi (t )]
= i [x0 + x(t )] i [ yi 0 + yi (t )] i = 1,...,6 ,
dt

or,
(2.81)

dyi 0 dyi (t )
+
= i x0 + ix(t ) i yi 0 iyi (t ) i = 1,...,6 .
dt
dt

Again, 0 values satisfy the initial equation, which leads to,


(2.82)

dyi (t )
= ix(t ) iyi (t ) i = 1,...,6 .
dt

Equations (2.79) and (2.82) are point-kinetics equations perturbed around an


equilibrium operational point defined by values x0 and 0 . It should be noted that the
equations constitute a system of linear ordinary differential equations with constant
coefficients.
54

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

Assuming that x and yi correspond to small perturbations designed earlier by x and yi,
and that thanks to a proper normalization both x0 and 0 are equal to zero, the
equations can be written in the standard form as,
(2.83)

dx(t )

1 6
(t )
,
= x(t ) + i yi (t ) +
dt

i =1

(2.84)

dyi (t )
= i x(t ) i yi (t ) i = 1,...,6 ,
dt

with initial conditions,


(2.85)

x ( 0) = 0 ,

(2.86)

yi (0) = 0, i = 1,...6 .

Applying the Laplace transformation, the system of equations formed by Eqs. (2.83)
through (2.86) becomes,

sx ( s ) =

(2.88)

sy i ( s ) = i x ( s ) i yi ( s ) i = 1,...,6 .

x ( s ) +

1 6
( s)
,
i y i ( s) +

i =1

(2.87)

Equations (2.87 ) and (2.88) give the following transfer function of the reactor:
))

(2.89)

G (s) =

x ( s )
1
=
6
( s )
i
s +
i =1 s + i

which is often referred to as the zero-power reactor transfer function. More detailed
analysis of this transfer function will be performed in the next Chapter.
2.4.2

Numerical Solutions of Point Kinetics Equations

The most general solution of the point kinetics equations can be obtained using the
numerical integration. In this section two cases will be considered: the first case, when
the reactivity is a linear function of time, and the second case, when the reactivity is a
sine function of time.
The linear (ramp) and sinusoidal changes of reactivity have important practical
applications. The ramp change of reactivity takes place when control rods are moved
with constant velocity under normal reactor operation. The reactivity can be given by
the following function,
(2.90)

(t ) = a t ,

55

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

where a is the rate of increase of reactivity. Typical changes of n(t) for this type of
reactivity increase are shown in FIGURE 2-7,

a3 > a2

ln n

a 2 > a1

a1

t
FIGURE 2-7. Time-variation of n(t) for the ramp change of reactivity.

The curves in FIGURE 2-7 suggest the following form of the solution,
(2.91)

ln n(t ) B + at 2 ,

or

n(t ) Ae at ,
2

(2.92)

where A, B and a are proper constants. The above qualitative analysis indicates that
2

when the reactivity linearly increases, the power increases as e t , whereas, as it was
shown earlier, with step-change of reactivity, the power change is proportional to e t .
FIGURE 2-8 shows a comparison of the relative power change in a reactor following
the step and the ramp (linear) change of reactivity. In both cases the same reactivity
equal to 0.0022 is introduced. For the step-insertion case, all the reactivity is inserted at
time t = 0. For the ramp insertion case the reactivity is linearly increased from 0 to
0.0022 during time interval equal to t = 6.6 s, and then it is kept constant. For a
comparison, FIGURE 2-9 shows the power increase when the ramp insertion of the
reactivity is performed 100-times faster, and the final reactivity 0.0022 is achieved after
t = 0.066 s. In this case the two solutions almost coincide, as it could be expected. It
can be concluded from the figures that the step-change of reactivity is a good
approximation for the case with the ramp insertion of reactivity, provided that the
insertion rate is very fast. If the insertion rate is slow, the two solutions will differ and
the step-change solution predicts higher power values during the first seconds of the
transient.

56

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

1.4
Ramp insertion
Step insertion

1.2

Relative Power

1.0

0.8

0.6

0.4

0.2

0.0
0

10

Time, s

FIGURE 2-8. Relative reactor power increase after ramp ( (t ) = 3.25 103 t ) and step ( 0 = 0.0022 )
insertion of reactivity.

1.4
Ramp insertion
Step insertion

1.2

Relative Power

1.0

0.8

0.6

0.4

0.2

0.0
0

10

Time, s

FIGURE 2-9. Relative reactor power increase after ramp ( (t ) = 0.325t ) and step ( 0 = 0.0022 )
insertion of reactivity.

The sinusoidal change of reactivity is used for determination of the reactor transfer
function using the so called pile oscillator. In this case the reactivity is the following
function of time,
(2.93)

(t ) = 0 sin t ,

where 0 and are the amplitude and the frequency of the reactivity oscillations,
respectively. Typical reactor power oscillations caused by harmonic reactivity
oscillations are shown in FIGURE 2-10. The figure shows a comparison of solutions
obtained with the exact one-group model and with the small-perturbationapproximation model. As can be seen the approximate solution can be considered
correct only during the first period of time after the insertion of the reactivity (first 0.1
0.2 s), as it could be expected.

57

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

0.7
Exact solution
0.6

Small perturb. approx.

0.5
0.4

Power

0.3
0.2
0.1
0.0
-0.1
-0.2
-0.3
0

10

15

20

25

30

35

40

45

50

Time, s

FIGURE 2-10. Power oscillation in a reactor subject to sine reactivity oscillation with 0 = 0.0022 and
= 3.14 rad/s. Comparison of the exact solution with the small-perturbation-approximation solution.

The small-perturbation approximation predicts power oscillations that have constant


amplitude, constant mean value and constant frequency, equal to the frequency of
oscillations of the pile oscillator. The solution obtained with the exact model indicates,
however, that both the amplitude and the mean value of oscillations increase with
time.
The influence of the reactivity amplitude on power oscillations is shown in FIGURE
2-11. It can be clearly seen, that the reactor power oscillations have a divergent
character. The divergence rate increases with increasing amplitude of the reactivity
oscillations.

12
Rho0 = 0.0022
Rho0 = 0.0065

10

Power

-2
0

10

15

20

25

30

35

40

45

50

Time, s

FIGURE 2-11. Power oscillation in a reactor subject to sine reactivity oscillation with frequency
= 3.14 rad/s and two values of reactivity amplitude: 0 = 0.0022 and 0 = 0.0065 .

The figures indicate that the small perturbation approximation gives an oscillatory
solution which agrees with the exact solution only for small time intervals right after
the initiation of the perturbation. As a result the two solutions diverge from each other
quite rapidly, already during the first cycle. Moreover, the small perturbation
58

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

approximation is not able to capture the long-term behavior of the reactor, since it
predicts persistent, constant-amplitude power oscillations, whereas the exact solution is
diverging with time.

2.5 Approximate Point Kinetics Models


Even though a solution of the exact point kinetics equations can be obtained relatively
easy using computers, in certain situations it is convenient and also enough to predict
the kinetics only approximately. Three such approximations are discussed in the
following sections.
2.5.1

The Prompt Jump Approximation

In the Prompt Jump Approximation (PJA) the rapid power change due to prompt
neutrons is neglected, corresponding to taking = 0 in the point kinetics equations.
In the full form, the equations are as follows,
(2.94)

dx

1 6
S
= x + i yi + + + x ,
dt

i =1
ne

(2.95)

dyi
= i x i yi
dt

i = 1,...,6

Multiplying both sides of the first equation with yields,

6
dx
S
= x + i yi + + + x
dt
ne
i =1

Since = 0 , then
(2.96)

x(t ) =

1 6

i yi +


i =1

The PJA consists of 6 ordinary differential equations with constant coefficients given
by Eq. (2.95) and one algebraic equation describing the reactor excess power in terms
of the concentration of the precursors of delayed neutrons.
Assuming one group of precursors, the equations are as follows,
(2.97)

dy
= x y ,
dt

(2.98)

x(t ) =

y +
.

The two equations can be combined into a single solvable ordinary differential
equation with time-dependent coefficients,

59

C H A P T E R

(2.99)

N U C L E A R

dy
a (t ) y = a (t ) ,
dt

R E A C T O R

a (t ) =

K I N E T I C S

, y ( 0) = 0

Equation (2.99) has a closed-form solution as follows,


t

(2.100)

y (t ) = e 0

a (t ' )dt '

a(t ')e

a (t '' )dt ''

dt ' .

Substituting the solution to Eq. (2.98) yields,


(2.101)

x(t ) =

e 0

a (t ' )dt '

a(t ')e

a (t '' )dt ''

dt ' +

In particular, when

a(t ) =

0
= const ,
0

the solution is as follows,

(2.102)

0 t
e 0 1 + 0 .
x(t ) =
0
0

The solutions given by Eq. (2.102) for various values of

= 0.08 s

r = 0 and for

are shown in FIGURE 2-12.


2.5

2
r=0.5

x(t) [-]

1.5
r=0.4

r=0.3

0.5

r=0.2
r=0.1
r=0.0

0
0

t [s]

FIGURE 2-12. Solutions obtained from PJA for different values of r = 0 and using one group of
precursors with = 0.08 s 1 .

60

C H A P T E R

2.5.2

N U C L E A R

R E A C T O R

K I N E T I C S

The Prompt Kinetics Approximation

The Prompt Kinetics Approximation (PKA) is somehow opposite to the PJA, since it
neglects the influence of the delayed neutrons on the reactor kinetics, and considers
only the prompt neutrons. The balance equation in the PKA is as follows,
(2.103)

dx S
= + + x,
dt ne

x(0) = 0 .

If there are no external neutron sources, the equation becomes,


(2.104)

dx
= ( x + 1) .
dt

This equation can be integrated as follows,


x

dx

(t )

(x + 1) =
0

dt ,

and the following exact solution, valid for any reactivity function (t), is obtained,
(2.105)

x(t ) = e

(t )

dt

1.

If, in particular, the reactivity is constant and equal to 0, the excess normalized reactor
power will be given as,
0

x(t ) = e 1 .
2.5.3

The Constant Delayed Neutron Source Approximation

The Constant Delayed neutron Source (CDS) is based on the assumption that the
production of the delayed neutrons is constant and equal to the production at the
beginning of the transient. In this approximation, the balance equations are as follows,
(2.106)

1 6
S
dx
= x + i yi + + + x ,
dt

i =1
ne

(2.107)

0 = i x(0 ) i yi

i = 1,...,6

Since x(0) = 0, then yi = 0 for i = 1,,6. As a result, the first equation becomes,
(2.108)

dx

S
= x+ + + x.
dt

ne

The above equation can be solved using the initial condition x(0) = 0.
This equation can be solved analytically. For that purpose it is written in a more
compact form as,

61

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

dx

S
and b(t ) = + .
+ a (t ) x = b(t ) , where: a (t ) =
dt

ne

For the initial condition x(0) = 0, the general solution is as follows,


(2.109)

x(t ) = e

0t a (t ' )dt ' t

b(t ')e
0

a (t '' )dt ''dt ' .

In particular, assuming S = 0 and = 0 = const , the solution is given as,


(2.110)

x(t ) =

0
0

1 e .

The solutions are shown for various positive values of r = 0 in FIGURE 2-13
and for negative values in FIGURE 2-14. As can be seen, the solutions represent with
good accuracy the details of the prompt jump, whereas for increasing time, the
solutions are converging to constant values. The latter is due to the neglect of the
delayed neutrons.
The solution given with Eq. (2.110) demonstrates the influence of time-variation of
delayed neutrons on reactor kinetics. If the concentration of precursors were not
changing with time, insertion of any positive reactivity smaller than would result with
a limited increase of the neutron flux. However, if the inserted reactivity is larger than
, the neutron flux will exponentially increase with time to infinity. For 0> the reactor
becomes prompt critical, that is the prompt neutrons are enough to sustain the chain
reaction.
1.2
r=0.5

x(t) [-]

0.8
r=0.4

0.6
r=0.3

0.4
r=0.2

0.2

r=0.1

0
0

t [s]

FIGURE 2-13. Solutions obtained from CDS for different positive values of r = 0 and = 0.001 s .

62

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

0
-0.05
r=-0.1

x(t) [-]

-0.1
-0.15

r=-0.2

-0.2
r=-0.3

-0.25
r=-0.4

-0.3
r=-0.5

-0.35
0

t [s]

FIGURE 2-14. Solutions obtained from CDS for different negative values of r = 0 and = 0.001 s .
N O M E N C L A T U R E

B
C
D
E
F
k
l
L
M
n
R
s
S
t
T
v
x
z

buckling
concentration of delayed neutron precursors
diffusion coefficient
neutron energy
neutron production operator
effective multiplication factor
average neutron lifetime
diffusion length, L = D a
migration and neutron removal operator
neutron density; amplitude function
reactor core radius
Laplace transform parameter
external neutron source
time
delayed neutrons to fission neutrons operator
mean neutron velocity
non-dimensional point kinetics variable
coordinate

fractional yield of the delayed neutrons


neutron flux
decay constant
total neutron yield per fission
reactivity
average neutron generation time
weighting function
frequency

63

C H A P T E R

a
f

N U C L E A R

R E A C T O R

K I N E T I C S

macroscopic cross section for neutron absorption


macroscopic cross section for fission
neutron direction angle

Subscript
0
critical or steady-state reactor conditions
e
equilibrium
i
pertinent to i-th group

infinite medium
Superscripts
*
adjoint

R E F E R E N C E S

[2-1]

Ott, K.O. and Neuhold, R.J., Introductory Nuclear Reactor Dynamics, ANS public., La Grange Park,
Illinois, USA, 1985.

[2-2]

Glasstone, S. and Sesonske, A., Nuclear Reactor Engineering, 3rd Ed., Van
Nostrand Reinhold Company, 1981.

E X E R C I S E S

EXERCISE 2-1: Based on the point kinetics model implemented in Scilab, write your own program to
solve the point kinetics equations with both 1 and 6 delayed-neutron precursors. Assume reactivity and
external sources as given functions of time. Assume , i and i to be known, user-specified
constants. Use two methods: (a) direct time integration of differential equations and (b) Laplacetransformed solution of small perturbation approximation equations.
EXERCISE 2-2: Apply models from EXERCISE 2-1 to solve EXAMPLE 2-1 . Compare your solution
with the solution given in EXAMPLE 2-1 .
XX

EXERCISE 2-3: Apply models from EXERCISE 2-1 to solve EXAMPLE 2-1 but assume that the
reactivity has been made 0.0022 negative. All other data are the same.
X

EXERCISE 2-4: Apply models from EXERCISE 2-1 to solve EXAMPLE 2-1 . Compare two solutions:
with one group and six groups of precursors. Use data from TABLE 2-3 .
X

EXERCISE 2-5: Apply models from EXERCISE 2-1 to find the neutron density as a function of time
n(t) in a reactor where the reactivity was step-changed from 0 = 10 to 1 = 9.5 . Weighted mean
source intensity is S = 105 s-1. Assume one group of delayed neutrons with = 0.1 s 1 , = 7 10 3 .
Assume = 10 3 s .
EXERCISE 2-6: Apply models from EXERCISE 2-1 to solve EXAMPLE 2-2. Compare the numerical
solution with the analytical solution obtained in EXAMPLE 2-2 .
X

EXERCISE 2-7: Apply models from EXERCISE 2-1 to solve EXAMPLE 2-3 . Compare the numerical
solution with the analytical solution obtained in EXAMPLE 2-3 .
X

64

C H A P T E R

N U C L E A R

R E A C T O R

K I N E T I C S

EXERCISE 2-8: A constant neutron source was suddenly introduced into a critical reactor. Some time
after the source introduction, it was observed that the neutron flux in the reactor increases linearly with
the rate 108 m-2 s-1 /s. Determine the mean weighted neutron source S , assuming one group of the
delayed neutrons with the following parameters: = 0.1 s-1, = 7.5 10-3, = 10-4 s.
EXERCISE 2-9: During the reactor start-up the reactivity was step-changed from 0 = 15 to
1 = 10 . The mean weighted neutron source was S = 106 s-1 and the generation time was equal to
= 10-4 s. Assuming one-group of delayed neutrons with = 0.1 s-1 and = 7.0 10-3, derive the
expression for the neutron density as a function of time.
EXERCISE 2-10: Determine the reactivity value 1 knowing that the ratio of the mean values of
neutron fluxes in a sub-critical reactor with 0 = 10 and 1 is equal to 1.3.
EXERCISE 2-11: A reactor oscillator (that is a rod made of strong neutron absorber moving
harmonically in the reactor core) was used to determine the transfer function of a reactor operating at low
power. Determine the reactor power response to the sine change of the reactivity (t ) = 0 sin t .
Assume = 7.5 10-3, = 0.1 s-1, = 10-3 s, 0 = 0.05 , f = 0.01 Hz.

65

Chapter

3
3 Nuclear Reactor Dynamics

uclear reactor dynamics is concerned with reactor power transients in which


the reactivity feedback caused by the reactivity change due to the power
change plays an important roll in over-all reactor behavior. In this Chapter
the reactor dynamics equations will be derived and several example
solutions of reactor dynamics problems will be shown.

3.1 Reactivity Feedbacks


When power changes in a nuclear reactor are large enough to influence the value of
the reactivity, the transient behavior of the reactor is termed as the nuclear reactor
dynamics. As can be expected, the influence of power on the reactivity has to be
quantified in order to properly describe the dynamic behavior of the nuclear reactor.
3.1.1

Influence of Fuel and Moderator Temperature on Reactor Operations

During operation of a nuclear reactor the energy released due to nuclear fissions is
transferred to the coolant. The resulting temperature distribution in the fuel and
coolant (in BWRs even the void fraction distribution) is a subject of the thermalhydraulic analysis of the nuclear reactor. The temperature distribution, which in a
general case is a function of both the time and the location, is influencing the values of
microscopic cross-sections for various nuclear reactions caused by neutrons. As a
result the reactivity will depend on the temperature changes.
From a practical point of view it is important to know what the influence of
temperature on reactivity is, and how it will influence the operation of the nuclear
reactor. In general the following two cases can be considered:

Reactivity increases with the temperature: in this case the increasing reactivity
will cause the increase of the reactor power, which, in turn, will cause the
increase of temperature, etc. It means that in this case the reactor will be
inherently unstable.

Reactivity decreases with temperature: in this case the decreasing reactivity will
cause the decrease of the reactor power which will be followed by the decrease
of the temperature, and so on. Clearly the reactor will be inherently stable in
such a case.

The conclusion is that reactors should be constructed in such a way which assures the
decreasing reactivity in function of increasing reactor power. This can be achieved by
using proper materials and a proper nuclear reactor configuration.
67

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

The reactivity changes with temperature because the reactivity depends on


macroscopic cross sections, which themselves involve the atomic number densities of
materials in the core,
(3.1)

(r, t ) = N (r, t ) (r, t ) .

The atomic density N(r,t) depends on the reactor power level because:
a) material densities depend on temperature T,
b) the concentration of certain nuclei is constantly changing due to neutron
interactions (poisons and fuel burnup).
The microscopic cross section is given in Eq. (3.1) as an explicit function of the spatial
location r and the time t. This must be done since the cross sections that appear in
one-speed diffusion model are actually averaged over energy spectrum of neutrons
that appear in the reactor core. Since this spectrum is itself dependent on the
temperature distribution in the core and hence the reactor power level, this
dependence must be taken into account in Eq. (3.1).
The reactivity variation with the temperature is the principal feedback mechanisms
determining the inherent stability of a nuclear reactor with respect to short-term
fluctuations in the power level. In principle, the reactivity feedback could be evaluated
by solving heat transfer equations, both in fuel and coolant regions and predicting the
temperature distribution in the reactor core for a given reactor power. However, such
approach would lead to a very complex system of partial nonlinear differential
equations. Typical simplification is based on the so-called lumped-parameter model
in which the major parts of the reactor core are represented by a single averaged value
of temperature, such as an average fuel temperature, moderator temperature and
coolant temperature.
The subject of the reactor dynamics analysis is to accommodate the core average
temperatures such as TF (fuel) and TM (moderator) in suitable models of reactivity
feedback. To this end one can write the reactivity change as a sum of two
contributions,
(3.2)

(t ) = ext (t ) + f ( P ) .

The reactivity in Eq. (3.2) is measured with respect to the equilibrium power level P0
(for which the reactivity is just equal to zero) and is a superposition of the externally
controlled reactivity insertion (for example, such as by the control rod movement) and
internal reactivity change due to inherent feedback mechanisms, indicated here as a
function of the power level.
For steady-state power level P0, Eq. (3.2) becomes,
(3.3)

0 = 0 + f ( P0 ) ,

68

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

which merely states that to sustain the criticality of the system one has to supply
external reactivity 0 to counteract the negative feedback reactivity f ( P0 ) . The
incremental reactivities appearing in Eq. (3.2) are then defined as,
(3.4)

ext (t ) = ext (t ) 0 ,

(3.5)

f ( P ) = f ( P ) f ( P0 ) .

The feedback mechanism in reactivity is schematically shown in FIGURE 3-1 .


XX

ext
+

Neutron
kinetics

Incremental power

ext

f
Feedback
mechanisms

FIGURE 3-1. Reactivity feedback mechanisms.

As can be seen, the output signal from the system, which in this case is the incremental
power, affects the input signal, which is the reactivity. For stability analysis, it is
interesting to investigate the feedback signal characteristics (gain and phase shift), since
this will determine the over-all stability of the closed-loop system.
3.1.2

Doppler Effect

With increasing material temperature, the nuclei will move with increasing speed. Since
the nuclei move in a chaotic manner, their relative velocity against a mono-energetic
neutron flux will no longer be constant and will have some distribution. This is
equivalent to a situation in which neutrons would have a certain energy distribution
when approaching stationary nuclei. This is the so-called Doppler effect, in analogy to
a similar phenomenon known in acoustics and optics.
Due to the Doppler effect, the number of nuclear reactions caused by mono-energetic
neutrons will depend on the temperature of the material. Without going into details it
can be mentioned that with the increasing material temperature the microscopic
absorption cross section will increase (e.g. of U-238) resulting in decreasing reactivity.
Also the microscopic fission cross-section will decrease with the temperature leading
to additional reduction of the reactivity.
One can estimate the influence of the Doppler effect on reactivity using the expression
for the resonance escape probability as,
(3.6)

N I
p = exp F ,
s

where for the metallic uranium and the uranium dioxide fuel at 300K temperature, the
effective resonance integral I is given as,

69

C H A P T E R

(3.7)

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

A
[b]; IUO2 = 4.45 + 84.5 A [b] .
M
M

IU = 2.95 + 81.5

The temperature dependence of the integral I is described by the following correlation


obtained from experimental data,
(3.8)

I (T ) = I (300 K ) 1 +

)]

T 300 ,

where, for 238UO2:


(3.9)

A
.
M

= 6.1 10 3 + 4.7 10 3

A and M in Eqs. (3.7) and (3.9) are area (in m2) and mass (in kg), respectively, of a fuel
rod.
3.1.3

Reactivity and Reactivity Coefficients

When the effective multiplication factor k remains constant from one generation of
neutrons to another, it is possible to determine the number of neutrons at any
particular generation by knowing the number of neutrons at zero generation, N0,
and the value of k. Thus, after n generations the number of neutrons is equal to
N = N0k n .
In particular, if in the preceding generation there are N 0 neutrons, then there are
N 0 k neutrons in the present generation. The change of the number of neutrons
expressed as a fraction of the present number of neutrons is referred to as reactivity
and is expressed as,
(3.10)

N 0k N 0 k 1
=
.
N 0k
k

As can be seen, reactivity is a dimensionless number; however, since its value is often
rather small, artificial units are defined. From the definition given by Eq. (3.10), the
value of reactivity is in units of k / k . Alternative units are % k / k and pcm
(percent milli-rho). The conversions between these units are as follows,
(3.11)

1 k k = 100% k k = 10 5 pcm ,

(3.12)

1% k k = 0.01 k k = 10 3 pcm ,

(3.13)

1pcm = 10 5 k k = 10 3 % k k .

EXAMPLE 3-1. Calculate the reactivity in a reactor core when k is equal to 1.002
and 0.998.
SOLUTION: The reactivity is as follows: for k= 1.002, = (1.002-1/1.002 =
0.001996 k/k = 0.1996 % k/k = 199.6 pcm. For keff = 0.998: = (0.9981/0.998 = 0.002 k/k = 0.2 % k/k = 200 pcm.

70

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

Other units often used in reactor analysis are dollars ($) and cents. By definition, 1$ is
reactivity which is equivalent to the effective delayed neutron fraction, , and, as can
be expected, one cent (1c) reactivity is equal to one-hundredth of a dollar.
As already mentioned, the dependence of the reactivity on temperature has an
important influence on the reactor dynamics and stability. In particular, a reactor will
be stable when the reactivity is a decreasing function of temperature. Needless to say
that the dependence of the reactivity on various parameters should be known.
However, it is not possible (nor necessary for most practical purposes) to estimate
such functions with all details. Instead a linearized form of the function is applied to
evaluate the reactivity change due to various parameter changes, that is,

(3.14)

(TC , TF , TM ,...)

TM + ...
TC +
TF +
TC
TF
TM
,

= TCTC + TF TF + TM TM + ...

where TC , TF , TM are the coolant, fuel and moderator temperature coefficient of


reactivity, respectively. These coefficients play important role in safe operation of
nuclear reactors.
A single temperature coefficient of reactivity can be defined as a partial derivative of
the core reactivity with respect to temperature,
(3.15)

= T( j ) .
T

T
j
j
j

Here j indicates that separate temperatures in the reactor (j = C for coolant, j = F for
fuel, etc.) are taken into account.
The two dominant temperature effects in most reactors are the change in resonance
absorption (Doppler effect) due to fuel temperature change and the change in the
neutron energy spectrum due to changing moderator or coolant density (due to
temperature, pressure or void fraction changes).
Noting that,
(3.16)

1 k 1 k
=

,
T k 2 T k T

one obtains,
(3.17)

T =

1 k 1 k
+
= TF + TM .
k TF k TM

Now changes in fuel temperature TF do not affect the shape of the thermal neutron
energy spectrum. It should be mentioned, however, that fuel and moderator
temperature effects can not always be separated. In LWRs a change in moderator
temperature will significantly change the moderator density, thereby influencing
71

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

slowing down and hence resonance absorption. In spite of such interference, it is


customary to analyze both coefficients separately.
3.1.4 Fuel Coefficient of Reactivity
The fuel temperature coefficient has an

important influence on reactor safety in case


of a large positive reactivity insertion. A negative fuel temperature coefficient is
generally considered more important than a negative moderator temperature
coefficient. The reason is that the negative fuel coefficient starts adding negative
reactivity immediately, whereas the moderator temperature cannot turn the power rise
for several seconds.
This coefficient is also called the

prompt reactivity coefficient

or the

fuel Doppler

reactivity coefficient. Its value can be readily obtained from Eq. (3.6) as,

(3.18)

TF =

1 dk
1 dp
1 dI
=
= ln p
.
k dTF p dTF
I dTF

Using Eq. (3.8) in (3.18) yields,


X

p(300 K ) 2 TF

(3.19)

TF = ln

3.1.5

Moderator Coefficient of Reactivity

When the moderator is at the same time used as a coolant (this is the case in LWRs),
the moderator coefficient of reactivity will in principle reflect the influence of the
coolant density changes on the reactivity.
The dominant reactivity effect in water-moderated reactors arises from changes in
moderator density, due to the thermal expansion of the coolant fluid or due to the
void formation. The principal effect is usually the loss of moderation that accompanies
a decrease in moderator density and causes a corresponding increase in resonance
absorption. It can be calculated as follows,
(3.20)

TM =

1 dk
1 dp
1 d
NF
=
=
exp
k dTM
p dTM
p dTM N M s

1 dN
I = ln M .
p dTM

Since dNM/dTM is negative and may be quite large, particularly if the coolant
temperature is close to the saturation temperature, the reactivity coefficient can also be
large.
Typical values of reactivity coefficients are given in TABLE 3-1 .
X

TABLE 3-1. Reactivity coefficients[3-1].


Type of coefficient
Fuel Doppler (pcm/K)
Coolant void (pcm/%void)

BWR

PWR

HTGR

LMFBR

-4 to -1

-4 to -1

-7

-0.6 to -2.5

-200 to -100

-12 to +20

72

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

Moderator (pcm/K)

-50 to -8

-50 to -8

+1.0

Expansion (pcm/K)

~0

~0

~0

-0.92

3.2 Point Dynamics Model of PWR


A simple point dynamics model of PWR should take into account the influence of the
fuel and the moderator temperature on the reactivity. Since the moderator - which is at
the same time the reactor coolant - is not boiling, the moderator reactivity feedback
will result from the density changes caused by temperature changes of single-phase
coolant. Thus, the simplest point dynamics model of PWR should take into
consideration the time variations of the fuel and coolant temperature.
3.2.1

Derivation of Point Dynamics Equations

The equations of point-reactor dynamics model consist of the point-reactor kinetics


equations complemented with a system of equations to describe the non-nuclear
(thermal-hydraulic) reactor parameters. In general, the equations are as follows,
(3.21)

6
dn
=
n + i Ci ,
dt
i =1

(3.22)

dCi i
= n i Ci , i = 1, , 6,

dt

(3.23)

= 0 + C (t ) + f (u1 , u2 ,..., u K ) ,

(3.24)

duk
= f k (n, u1 , u2 ,..., u K ; w1 ,..., wm ) .
dt

The two first equations are the point kinetics equations, which have already been
discussed in the previous Chapter. It should be noted that the source term S is
neglected in the point reactor dynamics equations since it is usually small compared to
other terms. The third equation expresses the dependence of the reactivity on various
parameters: 0 is the initial reactor reactivity with withdrawn control rods and in
cold condition, C (t ) is the reactivity change due to the control system (control
rods, fluid absorbents, etc.) and f is the reactivity change due to non-nuclear
parameter (e.g. temperature) changes.
Since non-nuclear parameters uk, k=1, , K do not change much in respect to the
steady-state (equilibrium) values, Eq. (3.23) can be written as,
K

(3.25)

= 0 + C (t ) + k uk .
k =1

There are various possible formulations of Eq. (3.24) leading to a variety of point
dynamics models. In the simplest case of a point dynamics model suitable for PWRs,

73

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

the influence of the coolant and the fuel temperature is taken into account, and the
corresponding equations are as follows,
(3.26)

dTC
= aC n C TC ,
dt

(3.27)

dTF
= a F n F TF ,
dt

(3.28)

= 0 + C (t ) + TC TC + TF TF .

The above model may be improved by a more accurate treatment of the heat transfer
between coolant and fuel using the Newtons equation of cooling. The heat balance
for fuel can be written as,
(3.29)

dTF
= a F n h (TF TC ) ,
dt

mF c pF

and the corresponding heat balance for coolant is as follows,


(3.30)

dTC
= h (TF TC ) WC c pC (TCex TCin ) ,
dt

mC c pC

where mF and mC are the mass of fuel and coolant in the core, respectively, cpF and cpC
are specific heat of fuel and coolant, h is the heat transfer coefficient between fuel and
coolant, WC is the coolant mass flow rate [kg/s] and TCex and TCin are the coolant exit
and inlet temperature, respectively. Assuming further that,
(3.31)

TC =

1
(TCex + TCin ) ,
2

the model is closed, that is, the number of unknowns is equal to the number of
equations. The complete system of equations for the point dynamics model is
summarized in TABLE 3-2.
X

TABLE 3-2. Point reactor dynamics model.


Reactor kinetics equation

6
dn
=
n + i C i
dt
i =1

Delayed-neutron precursor equations: i =


1,,6

dC i i
=
n i C i
dt

Fuel heat balance

mF c pF

74

dTF
= a F n h (TF TC )
dt

C H A P T E R

R E A C T O R

Coolant heat balance

mC c pC
Reactivity changes

D Y N A M I C S

A N D

S T A B I L I T Y

dTC
= h(TF TC ) 2WC c pC (TC TCin )
dt

= 0 + C (t ) + TC TC + TF TF

The point dynamics model is schematically illustrated in FIGURE 3-2.

C (t )

Tcin(t), Wc(t)

Tcex

TF
TC

FIGURE 3-2. Schematics of the point reactor dynamics model with reactivity, coolant inlet temperature
and coolant mass flow rate as external forcing functions.

To solve the point dynamics equations it is necessary to specify the initial conditions.
For that purpose, the reactor equilibrium condition will be determined.
3.2.2

Nuclear Reactor at Equilibrium

Equilibrium state of a nuclear reactor is described by Eqs. (3.21) through (3.24) in


which the time derivatives are equal to zero and the reactivity is constant. In particular,
using the model given in TABLE 3-2, the equilibrium is described by the following
system of algebraic equations,
X

(3.32)

(3.33)

n + i Ci = 0 ,

i =1

n i Ci = 0 ,

(3.34)

a F n h (TF TC ) = 0 ,

(3.35)

h (TF TC ) 2WC c pC (TC TCin ) = 0 ,

(3.36)

= 0 + TC TC + TF TF .

Two cases have to be considered. If n = 0 then Ci = 0, TC = TF = TCi and this type of


equilibrium corresponds to a shut-down reactor. When n 0 then = 0 and the
following expressions are obtained (index e means the equilibrium value),
75

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

i
ne ,
i

(3.37)

Ci ,e =

(3.38)

a F n e h (TF ,e TC ,e ) = 0 ,

(3.39)

h (TF ,e TC ,e ) 2WC ,e c pC (TC ,e TCin ,e ) = 0 ,

(3.40)

= 0 + TC TC ,e + TF TF ,e .

Solution of the system of equations yields,

0 + ( TC + TF )TCin,e

(3.41)

ne =

(3.42)

1
1
TF ,e = TCin ,e +
+ a F ne ,
2W c

C ,e pC h

(3.43)

TC ,e = TCin ,e +

1
F
(
TC + TF ) + T
aF
h
2WC ,e c pC

Cie =

i
ne ,
i

a F ne
.
2WC ,e c pC

The above equations determine steady-state operation conditions of a nuclear reactor.


3.2.3

Normalized Point Reactor Dynamics Model

Introducing normalized variables,


n ne
,
ne

(3.44)

x=

(3.45)

yi =

Ci Cie
, i = 1,...,6 ,
Cie

(3.46)

zF =

TF TF ,e
,
TF ,e

(3.47)

zC =

TC TC ,e
,
TC ,e

and substituting them into equations in TABLE 3-2 the following


equations of point reactor dynamics are obtained,
X

(3.48)

dx

1 6

= x + i yi + + x ,
dt

i =1

76

normalized

C H A P T E R

R E A C T O R

D Y N A M I C S

(3.49)

dyi
= i x i yi
dt

(3.50)

hTC ,e
dz F
a F ne
h
=
x
zF +
zC ,
dt
mF c pF TF ,e
mF c pF
mF c pF TF ,e

(3.51)

(3.52)

A N D

S T A B I L I T Y

i = 1,...,6 ,

2c W + h
hTF ,e
2WC ,eTCin,e
dzC
=
z F pC C ,e
zC +
u
dt
mC c pC TC ,e
mC c pC
mC TC ,e
2WC ,e (TC ,e TCin,e )
2WC ,e
2WC ,eTCin ,e
w
w zC +
uw
mC TC ,e
mC
mC TC ,e

= C (t ) + TC TC ,e zC + TF TF ,e z F ,

with zero initial conditions.


Equation (3.51) contains two forcing functions defined as,
)

(3.53)

u=

TCin TCin ,e
,
TCin ,e

(3.54)

w=

WC WC ,e
.
WC ,e

The forcing functions are the dimensionless coolant inlet temperature, u(t), and the
dimensionless coolant flow rate, w(t).
The derived model is represented by a set of non-linear ordinary differential equations
with time-dependent coefficients and one additional algebraic equation describing the
total reactivity. Such model can be solved using a numerical approach only. An
example of the numerical implementation of the model using the the Scilab
environment is presented below.
COMPUTER PROGRAM: Point Dynamics Model


//
//
//
//
//
//
//
//
//
//
//
//
//
//

function [dy]=
PDModel(t,y,yield,dconst,LAMBDA,ne,cpF,cpC,mF,mC,WCe,aF,h,alph
aF,alphaC,TCine,TCe,TFe)
// Point Dynamics model
// dy - returned RHS values
// t - time
y - vector of unknown functions
yield - yields of the precursors of delayed neutrons
dconst - decay constants of the precursors of delayed neutrons
LAMBDA - generation time
ne - neutron density at equilibrium
cpF - fuel specific heat
cpC - coolant specific heat
mF - mass of fuel
mC - mass of coolant
WCe - coolant flow at equilibrium
aF - power-to-neutron-density factor
h
- fuel-coolant heat flux coefficient
alphaF - fuel (Doppler) reactivity coefficient
alphaC - moderator/coolant reactivity coefficient
77

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

// TCine - coolant inlet temperature at equilibrium


// TCe - coolant temperature at equilibrium
// TFe - fuel temperature at equilibrium
//
// Set parameters
//
NPre = length(yield);
neq = 3 + NPre;
// Total number of equations
rho = InReactivity(t);
u
= InCoolTemp(t);
w
= InCoolFlow(t);
//
// Calculate Right-Hand-Sides of differential equations
//
dy = zeros(neq,1); // Initialize RHS
rho
= rho + alphaC*TCe*y(3) + alphaF*TFe*y(2);
dy(1) = (-sum(yield)*y(1) + yield*y(4:$) + rho + rho*y(1))/LAMBDA;
dy(2) = (aF*ne*y(1)/TFe - h*y(2) + h*TCe*y(3)/TFe)/(mF*cpF);
dy(3) = (h*TFe*y(2)/TCe - (2*cpC*WCe+h)*y(3))/(mC*cpC) ...
+ 2*WCe*(TCine*u - (TCe-TCine)*w)/(mC*TCe) ...
- 2*WCe*w*(y(3) - TCine*u/TCe)/mC;
for i = 1:NPre
dy(i+3) = dconst(i)*(y(1) - y(i+3));
end
endfunction

The above Scilab function can be used with any ordinary differential equation solver to
find the solution. For that purpose a dedicated solver script is provided: PDODESol().
In addition three forcing functions have to be supplied: InReactivity(t) which
gives the external reactivity as a function of time, InCoolTemp(t) which gives the
inlet coolant temperature as a function of time and InCoolFlow(t) which provides
the coolant mass flow rate as a function of time. The usage of the point dynamics
model implemented in Scilab is elucidated in the example below.
EXAMPLE 3-2. Predict the neutron density change in a reactor following a positive step change
of reactivity with 0.0022 at time t = 1 s. Use the following reactor data: = 0.001s , = 0.1

s-1, = 0.0075, cpF = 200 J kg-1 K-1, cpC = 4000 J kg-1 K-1, mF = 40000 kg, mC
= 7000 kg, WC,e = 8000 kg s-1, TCin,e = 550 K, aF = 7x106 J m3 s-1, ne = 200 m-3,
TF = -10-5 K-1, TC = -5x10-5 K-1, h = 4x106 J K-1 s-1. Plot the calculated neutron
density as a function of time in the time interval from 0 to 5 s.

SOLUTION: The reactivity time change is given in the InReactivity() function


as follows:
function [rho] = InReactivity(t)
if t < 1
rho = 0.;
else
rho = 0.0022;
end
endfunction

The other forcing functions have default zero value and do not need to be updated.
The input parameters are specified at the Scilab prompt as follows:
->yield=7.5e-3; dconst=0.1; LAMBDA=1.e-3; ne=200; cpF=200; cpC=4000; mF=40000;
mC=7000; WCe=8000; aF=7e6; h=4e6; alphaF=-1e-5; alphaC=-5e-5; TCine=550;
Two input parameters are still missing: the equilibrium fuel and coolant temperatures. These parameters
can be found from Eqs. (3.42) and (3.43), and are specified in Scilab as follows:
->TFe=TCine+aF*ne*(1/(2*WCe*cpC)+1/h);

78

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

->TCe=TCine+aF*ne/(2*WCe*cpC);
The vector of time instances at which the solution will be found is defined as:
->t=linspace(0,5,100);
Finally, the model solver is invoked as follows:
->PDODESol(t,yield,dconst,LAMBDA,ne,cpF,...
cpC,mF,mC,WCe,aF,h,alphaF,alphaC,TCine,TCe,TFe,blue);
The calculated neutron density change with time is shown in FIGURE 3-3.

280
270
260

ne, 1/m^3

250
240
230
220
210
200
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Time, s

FIGURE 3-3. Neutron density change in a reactor following a sudden insertion of positive reactivity
0.0022 at time t = 1 s.
3.2.4

Reactor Dynamics in Presence of Small Perturbations

The point dynamics model can be written in a short form as,


(3.55)

dx
= F(x; u) , x(0) = 0 ,
dt

where x is the vector of unknown functions, which for one-group approximation is as


follows,
(3.56)

x = [x, y, zC , z F ] ,
T

u is the vector of the forcing functions,


(3.57)

u = [C , u, w]

and F represents the right-hand-sides in the point dynamics model.


The fixed (or equilibrium) point of the reactor is represented by vectors x0 and u0,
which satisfy the equation,
(3.58)

F( x 0 ; u 0 ) = 0 .
79

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

The present model has been formulated in terms of the excess normalized variables,
using the equilibrium point as the reference. Due to that the fixed point of the model
is given as,
(3.59)

x 0 = [0 0 0 0] , and u 0 = [0 0 0] .
T

One of the important questions concerning the operation of nuclear reactors is their
behavior in the vicinity of the fixed point, since typical power reactors stay in this
condition during the most part of their lifetime. In particular, one should investigate
whether the reactor is stable around the fixed point.
The behavior of a dynamical system around any point (and in particular around the
fixed point) can be investigated using the first method of Lyapunov. For the present
model the linearization can be easily achieved by dropping the product terms
x, u w, w zC and the equations will be transformed into a set of linear equations
with constant coefficients,
(3.60)

dx

1 6

= x + i yi + ,
dt

i =1

(3.61)

dyi
= i x i yi
dt

(3.62)

hTC , e
dz F
aF ne
h
=
x
zF +
zC ,
dt
mF c pF TF , e
mF c pF
mF c pF TF , e

i = 1,...,6 ,

2c W + h
hTF , e
2WC , eTCin , e
dzC
=
z F pC C , e
zC +
u
dt
mC c pCTC , e
mC c pC
mCTC , e

(3.63)

2WC , e (TC , e TCin , e )


w
mCTC , e

(3.64)

= C (t ) + TC TC ,e zC + TF TF ,e z F .

The Jacobian matrix for the one-group approximation is as follows,

(3.65)

F aF ne
=
x mF c pF TF , e

0
0

hTC , e
.
mF c pF TF , e
2c W + h
pC C , e

mC c pC

0
h

mF c pF
hTF , e
mC c pCTC , e

The eigenvalues of the Jacobian are as follows,

80

C H A P T E R

(3.66)

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

s1 = 0 ,
s2 =

(3.67)

(c

pC

2c pC mC

h
2c pF mF

WC ,e
1

mC 2c pC mC c pF mF

h mC + c pF h mF + 2c pC c pF mFWC ,e ) 8c 2pC c pF hmC mFWC ,e

(3.68)

(c

s3 =

(3.69)

h
2c pC mC

h
2c pF mF

WC ,e
1
+

mC 2c pC mC c pF mF

h mC + c pF h mF + 2c pC c pF mFWC ,e ) 8c 2pC c pF h mC mFWC ,e


2

pC

s4 =

The obtained expressions indicate that the Lyapunov exponents are always real and
negatives for the second, the third and the fourth eigenvalue, whereas it is always zero
for the first one.
3.2.5

Frequency Domain Analysis of Point Dynamics Model

The linearized point-dynamics model can also be investigated in the frequency domain.
Equations (3.60) through (3.64) can be Laplace transformed and the following transfer
functions can be derived,
x ( s )
,
C ( s )

(3.70)

H ( s) =

(3.71)

L( s ) =

(3.72)

M ( s) =

x ( s )
,
w ( s )

(3.73)

N ( s) =

z F ( s )
.
u ( s )

x ( s )
,
u ( s )

Transfer functions (3.70) through (3.73) describe the nuclear reactor response to
various forcing functions.
Transfer function H(s) describes the nuclear reactor power response to a change of the
external reactivity. The block diagram of a reactor undergoing this type of forcing
function is shown in FIGURE 3-4.
X

81

C H A P T E R

C ( s )

R E A C T O R

( s )

D Y N A M I C S

GR ( s ) =

A N D

S T A B I L I T Y

x ( s )

x ( s )
( s )

T ( s )
F ( s )

CL ( s )

H F ( s) =

H C ( s) =

F ( s )
x ( s )

CL ( s )
x ( s )

FIGURE 3-4. Block diagram of a reactor with a perturbation of the external reactivity.

The transfer functions GR(s), HF(s) and HC(s) describe the reactor response, the fuel
feedback and the coolant feedback, respectively. The transfer functions can be derived
from Eqs. (3.60) through (3.64) in the following way.
Performing the Laplace transformation of the point dynamics equations yields:

sx ( s ) =

(3.75)

sy i ( s ) = i x ( s ) i y i ( s ) i = 1,...,6 ,

(3.76)

sz F ( s ) =

hTC ,e
a F ne
h
x ( s )
z F ( s ) +
zC ( s ) ,
mF c pF TF ,e
mF c pF
mF c pF TF ,e

szC ( s ) =

2c W + h
hTF ,e
z F ( s ) pC C ,e
zC ( s ) +
mC c pC TC ,e
mC c pC

x ( s ) +

1 6
( s )
,
i y i ( s ) +

i =1

(3.74)

(3.77)

2WC ,eTCin,e
2WC ,e (TC ,e TCin ,e )
u ( s )
w ( s )
mC TC ,e
mC TC ,e

(3.78)

( s ) = C ( s ) + TC TC ,e zC ( s ) + TF TF ,e z F ( s ) .

The transfer function GR(s) can be readily obtained from Eqs. (3.74) and (3.75) as
follows,

82

C H A P T E R

(3.79)

R E A C T O R

GR ( s ) =

1 6
s 1 + i
i =1 s + i

D Y N A M I C S

A N D

S T A B I L I T Y

Laplace transforms zF ( s ) and zC ( s ) can be expressed in terms of x ( s ) solving Eqs.


(3.76) and (3.77) and noting that u ( s ) = w ( s ) = 0 . Equation (3.77) yields,
(3.80)

zC ( s ) =

hTF ,e
zF ( s ) .
TC ,e (mC c pC s + 2c pCWC ,e + h )

Substituting Eq. (3.80) into (3.76) yields,


(3.81)

zF ( s ) =

aF ne

h2

TF ,e mF c pF s + h

+
+
m
c
s
2
c
W
h
C pC
pC C ,e

x ( s ) .

Since F ( s ) = TF TF ,e zF ( s ) , the following expression for HF(s) is obtained:


(3.82)

H F ( s) =

TF TF ,e z F ( s)
x ( s )

TF aF ne

h2
mF c pF s + h

m
c
s
+
2
c
W
+
h
C
pC
pC
C
,
e

In a similar manner the coolant feedback transfer function can be obtained as,

H C (s) =

C
T C ,e C

T z ( s )
x ( s )

(3.83)

TC h

C
T C ,e
F
T F ,e

T
T

TF (mC c pC s + 2c pCWC ,e + h )

s+

hTF ,e
mC c pC TC ,e
TF TF ,e zF ( s )
=
2c pCWC ,e + h
x( s)
.

mC c pC

H F ( s)

The block diagram shown in FIGURE 3-4 can be replaced with a simplified one
shown in FIGURE 3-5.
X

C ( s )

H ( s) =

x ( s )
C ( s )

x ( s )

FIGURE 3-5. Simplified block-diagram of a reactor subject to a perturbation of external reactivity.


83

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

The transfer function of the simplified block-diagram H(s) can be obtained as,
(3.84)

H ( s) =

x ( s )
GR ( s )
=
.
C ( s ) 1 [H F ( s ) + H C ( s )]GR ( s )

3.3 Point Dynamics Model of BWR


In BWR the major moderator feedback is caused by the change of the density due to
boiling and the change of vapor content in the moderator. However, under low-void
conditions, the feedback will also be caused by the density change due to temperature
variations in the single-phase regions of the moderator/coolant.
3.3.1

Formulation of BWR Dynamics Model

A simple model of BWR dynamics, which contains the essential properties of BWR
behavior has been formulated as follows[3-2],
(3.85)

dx(t ) (t )
(t )
=
x(t ) + y (t ) +
,
dt

(3.86)

dy (t )
= x(t ) y (t ) ,
dt

(3.87)

dz (t )
= a1 x(t ) a2 z (t ) ,
dt

(3.88)

d (t )
= t (t ) ,
dt

(3.89)

d t (t )
= k z (t ) a3 t (t ) a4 (t ) .
dt

The first two equations of the model describe the neutronic behavior of the reactor,
where x(t) is the excess neutron density normalized to the steady-state neutron density,
and y(t) is the excess delayed neutron precursor concentration, also normalized to the
initial steady-state precursor concentration. The third equation represents the time
behavior of the excess average fuel temperature, z, and the last two equations describe
the time dependence of the excess void reactivity feedback. The total reactivity change
is given as,
(3.90)

(t ) = (t ) + TF z (t ) .

As can be seen, the total reactivity feedback consists of the void and the fuel
temperature reactivity feedback.
The model given by Eqs. (3.85) through (3.90) has been applied to an analysis of the
stability test 7N at the Vermont Yankee reactor[3-4]. The recommended[3-2] constants in
the model are given in the Table below.

84

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

TABLE 3-3. Parameters in the BWR dynamics model valid for Vermont Yankee test 7N.

Parameter

Value

Units

a1

25.04

K s-1

a2

0.23

s-1

a3

2.25

s-1

a4

6.82

s-2

k0

-3.70 x 10-3

K-1 s-2

TF

-2.52 x 10-5

K-1

0.0056

4.00 x 10-5

0.08

s-1

The parameter k, which is proportional to the void reactivity coefficient and the fuel
heat transfer coefficient, determines the gain of the feedback and, thus, defines the
linear stability of the reactor point model. The value k0 given in the Table is the critical
value above which the model becomes unstable.
3.3.2

Nuclear Reactor at Equilibrium

The full power (fixed) operating point is obtained by setting time derivatives in Eqs.
(3.85) through (3.89) equal to zero. That is,

x(t ) + y (t ) ,

(3.91)

0=

(3.92)

0=

(3.93)

0 = a1 x(t ) a2 z (t ) ,

(3.94)

0 = t (t ) ,

(3.95)

0 = k z (t ) a3 t (t ) a4 (t ) .

x(t ) y (t ) ,

In addition, the total reactivity must be equal to zero,


(3.96)

0 = (t ) + TF z (t ) .

The full power point corresponds then to the following state variable vector,
85

C H A P T E R

(3.97)

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

x 0 = [x, y, z , , t ]0 = [0,0,0,0,0] .
T

The Jacobian matrix of the system is as follows,

(3.98)




F
=
x a1
0

a2

0
0

0
k

0
a4

0
.
0
1

a 3

The eigenvalues of the matrix are as follows,


(3.99)

1 = 0 , 2 = a 2 , 3, 4 =

a 3 a 32 4a 4
2

5 =

Substituting the parameters from

TABLE 3-3, the following eigenvalues are obtained,


(3.100)

1 = 0 , 2 = 0.23 , 3, 4 = 1.125 m 2.357 j , 5 = 140.08 .

As can be seen only the first eigenvalue has a non-negative real part. Using the constants given in

TABLE 3-3 and applying an external reactivity perturbation:


(3.101)

ex (t ) = 0.0022 cos(10t ) , for 0 < t < 0.1 s, otherwise ex (t ) = 0

a limit cycle solution is obtained, as shown in FIGURE 3-6.

12
Power
Fuel temperature

Norm. Excess Variable

10

-2
0

20

40

60

80

100
Time, s

86

120

140

160

180

200

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

FIGURE 3-6. Reactor normalized excess power and fuel temperature as a function of time after stepinsertion of reactivity; k = -3.7e-3.

0.8

0.6

Power

0.4

0.2

0.0

-0.2

-0.4
0

10

12

Fuel temperature

FIGURE 3-7. Phase-space representation of limit cycle after step-insertion of reactivity; k = -3.7e-3.

To investigate the behavior of the model, the coefficient k is first changed to -3.510-3
and then to -3.910-3. The model response is shown in FIGURE 3-8 through
FIGURE 3-11.

8
Power
7

Fuel temperature

Norm. Excess Variable

6
5
4
3
2
1
0
-1
0

20

40

60

80

100

120

140

160

180

200

Time, s

FIGURE 3-8. Development of power and fuel temperature as a function of time with k=-3.510-3.
87

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

0.7
0.6
0.5
0.4

Power

0.3
0.2
0.1
0.0
-0.1
-0.2
-0.3
0

Fuel temperature

FIGURE 3-9. Phase-space representation of the model response with k=-3.510-3.

16
Power
14

Fuel temperature

Norm. Excess Variable

12
10
8
6
4
2
0
-2
0

20

40

60

80

100

120

140

160

180

200

Time, s

FIGURE 3-10. Development of power and fuel temperature as a function of time with k=-3.910-3.

88

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

1.5

1.0

Power

0.5

0.0

-0.5

-1.0
-2

10

12

14

16

Fuel temperature

FIGURE 3-11. Phase-space representation of the model response with k=-3.910-3.

In the case of k = -3.510-3 the amplitude of oscillations is decreasing and the system
is returning to the equilibrium point with zero excess power and fuel temperature.
When k is changed to -3.910-3, the oscillation amplitude is growing and the system
undergoes the limit cycle oscillations.

3.4 Nonlinear Effects


For some special cases it is possible to compare the linear and non-linear solutions and
evaluate the reactor stability using analytical methods in both cases. As an example,
consider a simplified version of the point-dynamics model with one-loop feedback.
The delayed neutrons will be taken into account indirectly by assuming
+ . It will be assumed that T(t) is an over-all average reactor
temperature and the rate of change of T(t) in the reactor is proportional to the
deviation of the reactor power P(t) from its equilibrium, steady state value P0. Thus, the
system can be described as follows,
(3.102)

+ T (t )
dP(t ) (t )
=
P (t ) = 0
P (t ) ,
dt

(3.103)

MC

dT (t )
= P (t ) P0 ,
dt

with initial conditions:


P(t=0) = P0, and T(t=0) = T0.

89

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

Here MC is the system total mass multiplied by the average heat capacity of fuel and
coolant and the system reactivity (t ) has only two parts: constant term 0 and the
temperature feedback term T (t ) . Introducing non-dimensional variables
(t ) = [T (t ) T0 ] / T0 , (t ) = P(t ) / P0 and = t / t0 into Eqs. (3.102) and (3.103)
yields,
(3.104)

d ( ) + ( )
=
( ), (0) = 1 ,
d

(3.105)

d ( ) 1
= ( ( ) 1), (0) = 0 .
d

Here new constants have been introduced to reflect the variable change.
Equations (3.104) and (3.105) can be solved analytically. Dividing both equations with
each other yields,
(3.106)

d ( )
+ ( ) ( )
=
,
d ( )

( ) 1

or
(3.107)

( ) 1
+ ( )
d ( ) =
d ( ) .

( )

Both sides of Eq. (3.107) can be integrated as follows,

( ) 1
d ( )
d ( ) = d ( )
= ( ) ln ( ) + C1 ,
( )
( )

(3.108)

(3.109)

+ ( )

d ( ) =

1
2
( ) + ( ) + C2 .

Thus, the solution of Eq. (3.107) reads as follows,


(3.110)


+C.

1
2

( ) ln ( ) = ( ) + ( ) 2

The constant C can be found from the initial conditions:


(3.111)


+ C C = 1.

1
2

(0) ln (0) = (0) + (0) 2

Thus the final solution of the nonlinear problem is as follows,


(3.112)

2
= ( ) ln ( ) 1 .
( ) + ( )
2


90

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

If now (t ) is plotted as a function of (t ) for different values of , closed orbits,


looking like distorted circles, will be obtained, as shown in FIGURE 3-12 . A simple
analysis shows that as increases, both (t ) and (t ) remain finite for all values of
time t. Both these quantities oscillate but remain limited and do not diverge as t .
X

rel. temp. difference


0.40
k = 1.0e-3
0.35

k = 4.0e-3
k = 6.0e-3

0.30

0.25

(t )

0.20

0.15

0.10

0.05

rel. power

0.00
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

5.5

(t )
FIGURE 3-12. Limit cycle oscillations of reactor power and temperature.

Equations (3.104) and (3.105) can be linearized and solved using e.g. the Laplace
transform approach. Thus, a direct comparison of the linear and nonlinear model is
possible.

3.5 Nuclear Reactor Stability


The stability of nuclear reactors can be analyzed by using two main approaches:

the time-domain approach, when the reactor dynamics equations are solved in
time domain and the stability property of the reactor is evaluated based on the
calculated power oscillations,

the frequency-domain approach, in which the reactor-dynamics equations are


linearized around a fixed point and the stability properties of the system are
investigated based on the location of poles of the transfer function.

The first approach is quite straightforward, providing that a proper numerical model
for the reactor dynamics is available. The advantage of this approach is that all
nonlinear effects are included, and the predictions are valid even far from the fixed
(unperturbed steady-state) point. The disadvantage is usually the high computational
cost.

91

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

The second approach is valid only for small perturbations, thus it can be used for the
prediction of the instability threshold, but not to predict the nonlinear behavior of the
system far from the equilibrium point. The main advantage of the frequency domain
approach is that it is fast running and computationally inexpensive. The frequency
domain approach is outlined in the following sections.
3.5.1

Open Loop Transfer Functions

The Laplace transform representation is particularly useful in the development of the


important property of a system called the transfer function (see Appendix A). In a
general sense, the transfer function is a mathematical expression which describes the
effect of a physical system on the signal transferred through it. For a system shown in
FIG. 2-4 , the transfer function G(s) is defined as a ratio of the Laplace transform of
the output signal, Y(s) to the Laplace transform of the input signal, U(s); G(s) =
Y(s)/U(s).
X

U(s)

G(s)

Y(s)

FIGURE 3-13. Open loop system.


3.5.2

Closed Loop Transfer Functions

The system shown in FIGURE 3-13 represents a so-called open-loop system, that is, a
system without feedback. A system with feedback, also called as a closed loop system,
is shown in FIGURE 3-14 .
X

U(s)

G(s)

Y(s)

H(s)

FIGURE 3-14. Closed loop system with feedback.

G(s) represents the system (forward) transfer function and H(s) is the feedback transfer
function.
Performing a summation of signals at the input to the system yields,
(3.113)

[U ( s ) + H ( s ) Y ( s )]G( s ) = Y ( s ) .

Thus, the transfer function for the system with feedback (closed loop transfer
function) becomes,
(3.114)

GT ( s )

Y ( s)
G( s)
=
.
U ( s) 1 H ( s) G( s)

92

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

Both the forward and multiple feedback reactor transfer functions have been derived
in previous sections.
3.5.3

Stability of Zero-Power Reactor

Reactor kinetics equations represent a typical example of an open-loop system. The


input signal is then the reactivity and the output signal is the neutron flux. The transfer
function, referred usually as a zero-power reactor transfer function, is defined as,
(3.115)

G( s) =

x ( s )
.
( s )

Using the linearized reactor point kinetics equations derived in the previous Chapter
(with dropped non-linear terms, which holds for small perturbations and < 0.1 ),
the transfer function of the zero power reactor is obtained as,
(3.116)

G( s) =

x ( s ) 1
1
=
.
( s )
1 6 i

s 1 +
s

+
1
i
=
i

The gain G ( j ) and the phase shift angle arg (G ( j ) ) of the transfer function are
shown in FIGURE 3-15 and FIGURE 3-16, respectively.
X

20
=103s
4
=10 s
6
=10 s

15

10

Gain, decibeles

10

15

20

25
2
10

10

10
Frequency, 1/s

10

10

FIGURE 3-15. Gain plot for various neutron generation time values.

The gain plot (often called the Bode diagram) indicates that an open-loop reactor
system tends to be unstable as frequency becomes small, since the gain becomes
infinite when frequency approaches zero. Thus, a reactor without feedback is
intrinsically unstable.

93

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

10

=103s
4
=10 s
6
=10 s

20

Phase angle, degrees

30

40

50

60

70

80

90
2
10

10

10
Frequency, 1/s

10

10

FIGURE 3-16. Phase plot for various neutron generation time values.
3.5.4

Stability of Closed Loop Systems

Reactor dynamics equations represent a typical example of a closed loop system. This
is due to the presence of multiple feedbacks which exist in this case, as discussed in
previous sections.
It is interesting to consider a special case of the open loop in which,
(3.117)

U ( s) G( s) H ( s) = U ( s) .

In this case the system will be self-excited if the feedback loop is closed. There will be
no need for any external input signal since the feedback signal is in-phase with the
external input. Equation (3.117) can be written as,
(3.118)

[1 H ( s ) G( s )]U ( s ) = 0 .

Since the input perturbation is arbitrary, the condition for the instability is,
(3.119)

1 H ( s) G( s) = 0 .

Equation (3.119) is called the characteristic equation of the system and is the same as
the denominator of the closed loop transfer function, given by Eq. (3.114).
One can observe that the roots of the characteristic function will be poles of the closed
loop transfer function. If the closed loop transfer function has the form,
(3.120)

F ( s) =

1
,
sa

the original f(t) of the transfer function F(s) is,


(3.121)

f (t ) = L1 {F ( s )} = e at .

Transfer function F(s) has a pole at s = a, which corresponds to the zero value of its
denominator s a. If now Re(a) > 0 the function f(t) will diverge with time which
94

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

indicates unstable system. In other words, roots of the characteristic equation may
imply exponential divergence in the time domain. The limiting case when Re(a) = 0 is
the marginal stability case in which a perturbation causes the system to oscillate
sinusoidally, but not diverge with time. Naturally system is stable when Re(a) < 0, since
then any perturbation will damp out with time.
One can formulate the following stability criterion for a closed loop system:
The necessary and sufficient condition for the closed loop system to be stable to small perturbations is
that all the roots of the characteristic equation have negative real parts.
If one can factor the closed loop transfer function using partial fraction expansion, the
roots can be easily determined. This is the case when the closed loop transfer function
has a form of a rational polynomial. In the case of complicated transcendental
algebraic equations direct evaluation of roots is not trivial, however, and an approach
using the Nyquist Criterion (see Appendix B) has proven to be an efficient way of
investigation of the system stability.
N O M E N C L A T U R E

C
D
E
G
h
H
j
k
l
L
n
s
S
t
T
W
x
y
z

concentration of delayed neutron precursors


diffusion coefficient
neutron energy
reactor forward transfer function
heat transfer coefficient
reactor feedback transfer function
imaginary unit 1
effective multiplication factor
average neutron lifetime
diffusion length, L = D a
neutron density; amplitude function
Laplace transform parameter
external neutron source
time
temperature
mass flow rate
excess normalized reactor power
excess normalized concentration of precursors
excess normalized reactor temperature

Greek

TM
TF

moderator reactivity coefficient


fuel (Doppler) reactivity coefficient
fractional yield of the delayed neutrons
neutron flux
decay constant
total neutron yield per fission
reactivity
95

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

average neutron generation time


macroscopic cross section for neutron absorption
macroscopic cross section for fission

Subscript
0
critical or steady-state reactor conditions
e
equilibrium
i
pertinent to i-th group of delayed neutrons
F
fuel
M
moderator

infinite medium
R E F E R E N C E S

[3-1]

Duderstadt, J.J. and Hamilton, L.J., Nuclear Reactor Analysis, John Wiley & Sons, Inc., 1980.

[3-2]

March-Leuba, J., Cacuci, D.G. and Perez, R.B., Nonlinear Dynamics and Stability of Boiling
Water Reactors: Part 1 Qualitative Analysis, Nucl. Sci. Eng., Vol. 93, pp. 111-123, 1986.

[3-3]

Ott, K.O. and Neuhold, R.J., Introductory Nuclear Reactor Dynamics, ANS public., La Grange Park,
Illinois, USA, 1985.

[3-4]

Sandoz, S.A. and Chen, S.F., Trans. Am. Nucl. Soc., Vol. 45, p. 727, 1983.

E X E R C I S E S

EXERCISE 3-1: Derive transfer functions of a detailed and a simplified block-diagram of a nuclear
reactor undergoing perturbation of the inlet temperature of coolant.
Hint:
The detailed diagram in this case is as shown in figure below. As can be seen the perturbation of
the inlet coolant temperature u(s) is transformed in the block with transfer function Ku(s) and next
is summed with temperature feedback effects and enters the reactor with transfer function GR(s) as
the total reactivity change ( s) . All transfer functions shown in figure can be derived from (3.74)
through (3.78) assuming C ( s ) = w(s) = 0.

u ( s )

G R (s )

Ku(s)

x ( s )

H F (s )

H C (s )
FIGURE 3-17. Figure for EXERCISE 3-1 . Block diagram of a reactor with a perturbation of the inlet
temperature of coolant.
X

96

C H A P T E R

R E A C T O R

D Y N A M I C S

A N D

S T A B I L I T Y

EXERCISE 3-2: Derive transfer functions of a detailed and a simplified block-diagram of a nuclear
reactor undergoing perturbation of the inlet mass flow rate of coolant.
Hint:
The detailed diagram in this case is as shown in figure below. As can be seen the perturbation of
the inlet coolant flow w(s) is transformed in the block with transfer function Kw(s) and next is
summed with temperature feedback effects and enters the reactor with transfer function GRw(s) as
the total reactivity change ( s ) . All transfer functions shown in figure can be derived from (3.74)
through (3.78) assuming C ( s ) = u(s) = 0.
X

w ( s )

G R (s )

Kw(s)

x ( s )

H F (s )

H C (s )
FIGURE 3-18. Figure for EXERCISE 3-2 . Block diagram of a reactor with a perturbation of the inlet
mass flow rate of coolant.
X

EXERCISE 3-3: Extend the program from EXERCISE 2-1 for modeling of the reactor dynamics
problems. Assume that the reactivity, the coolant mass flow rate and the inlet coolant temperature are
given functions of time. All other parameters are user-defined constants.
EXERCISE 3-4: Using the transfer functions H(s), L(s) and M(s) derived in this Chapter and in
EXERCISE 3-1 and EXERCISE 3-2 , evaluate the stability of the following reactor: = 0.001s , =
0.1 s-1, = 0.0075, cpF = 200 J kg-1 K-1, cpC = 4000 J kg-1 K-1, mF = 40000 kg, mC = 7000 kg, WC,e = 8000
kg s-1, TCin,e = 550 K, aF = 7x106 J m3 s-1, ne = 200 m-3, TF = -10-5 K-1, TC = -5x10-5 K-1, h = 4x106 J K-1
s-1. Perform the Nyquist plot and use the Nyquist criterion to evaluate the reactor stability. Find roots of
the characteristic equation and evaluate the reactor stability based on roots values. Compare the two
methods.
X

EXERCISE 3-5: Using the time-domain model of the reactor point dynamics developed in EXERCISE
3-3, determine the time response of the reactor caused by a step change of reactivity. Use the same
reactor data as in EXERCISE 3-4 . The reactivity change is described as C (t ) = 0.05 (t ) , where (t ) is
the step function. Plot the reactor power as a function of time.
X

EXERCISE 3-6: Using the time-domain model of the reactor point dynamics developed in EXERCISE
3-3 , determine the response of a reactor caused by a step change of the inlet coolant temperature. Use
the same reactor data as in EXERCISE 3-4 . The inlet temperature change is described as
TCin (t ) = 55 (t ) K, where (t ) is the step function. Plot the reactor power as a function of time.
X

EXERCISE 3-7: Using the time-domain model of the reactor point dynamics developed in EXERCISE
3-3, determine the response of the reactor caused by a step change of the inlet coolant mass flow rate.
Use the same reactor data as in EXERCISE 3-4 . The inlet coolant mass flow rate change is described as
WC (t ) = 0.1W C ,e (t ) kg s-1, where (t ) is the step function. Plot the reactor power as a function of time.
X

97

C H A P T E R

R E A C T O R

D Y N A M I C S

98

A N D

S T A B I L I T Y

Chapter

4
4 Dynamics of Boiling
Systems

ynamics of boiling systems has important implications on over-all dynamics


of nuclear power plants. For that reason it is necessary to investigate the
dynamic characteristic of channels with phase change. The first section of
this Chapter is dealing with analytical solutions of transient two-phase flows
in boiling channels subject to sudden perturbations of inlet or heating conditions. The
channel response is analyzed in time domain. The second section is devoted to a study
of boiling channel stability. The governing two-phase flow equations are linearized and
the channel stability is investigated in the frequency domain. In the third and last
section of this Chapter a similar stability analyses are performed for boiling loops.

4.1 Analysis of Two-Phase Flow Transients


The analysis of transients in boiling loops and systems involves the modeling of several
coupled phenomena such as: single-phase and two-phase flows in heated channels and
volumes, boiling heat transfer, phase separation and mixing, transient heat conduction
in solid structures, effects of void fraction on power generation (e.g. due to neutronic
feedback in BWRs), and others. In this section the basic equations that describe all
these processes will be given.
The simplest description of dynamics of a boiling channel is usually based on a onedimensional drift-flux or homogeneous two-phase flow model. The following
conservation equations are used in the one-dimensional drift-flux model.
Mass conservation equation for the liquid phase,
(4.1)

[ l (1 )A] + ( l jl A) = A .
t
z

Mass conservation equation for the vapor phase,


(4.2)

[ v A] + ( v j v A) = A .
t
z

Energy conservation equation for the liquid phase,


(4.3)

[( l il p )(1 )A] + ( l il jl A) = qlPH .


t
z

99

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

Energy conservation equation for the vapor phase,

[( viv p )A] + ( viv jv A) = qvPH .


t
z

(4.4)

Momentum conservation equation for the two-phase mixture,


G 1 G 2 A p

+
+
+
t A z M z

(4.5)

2 4C f , lo
G2
2
(
)

z
+ m g sin = 0
i i lo,i
i
lo D
h

2l

Here the Dirac delta function ( z z i ) has been used to indicate positions of local
pressure losses. Equation (4.5 ) employs the static and dynamic mixture densities,
defined as,
X

m = k k = (1 ) l + v ,

(4.6)

(4.7)

x2
= k
k kk

x2
(1 x )2 .
=
+

v l (1 )

The closure relationship for void fraction in the drift-flux model is as follows,

(4.8)

jv
.
C0 j + U vj

Here,
jv =

(4.9)

Gx

and
(4.10)

jl =

G (1 x )

are superficial velocities of vapor and liquid phase, respectively, Uvj is the drift velocity
and C0 is the void concentration parameter. The channel cross section area A in Eqs.
(4.1) through (4.5 ) can be, in general, a given function of the axial distance z.
X

The transient two-phase flow model described by Eqs. (4.1) through (4.10) is quite
complex and includes five conservation equations. Solution of such a system is quite a
demanding task, and for some purposes a simple model can be used, where only three
conservation equations are formulated and solved.
X

100

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

For the homogeneous equilibrium model (HEM) the conservation equations are as
follows.
Mass conservation equation for the mixture:
(4.11)

m G
+
= 0.
t
z

Energy conservation equation for the mixture:


(4.12)

[( miM p )A] (Gim A)


+
= qPH .
z
t

Momentum conservation equation for the mixture:

(4.13)

G 1 G 2 A p

+
+
+
t A z m z
n
2 4C f , lo
G2
2

(
z

z
)
+ m g sin = 0

i lo , i
i
lo D
i =1
h

2l

Closure relationships for mixture thermodynamic enthalpy, im, and mixture center-ofmass enthalpy, iM, are as follows:
(4.14)

im = i f (1 x ) + i g x ,

(4.15)

iM =

f i f (1 ) + g i g
m

Mixture densities are given by Eqs. (4.6) and (4.7 ), whereas the remaining notation is
conventional.
X

A similar three-equation model can be obtained by assuming slip between phases and
using, e.g., the drift-flux model equations as the starting point. In such model the
original mass and energy conservation equations are replaced by modified continuity
equation and the void propagation equation.
Assuming the thermodynamic equilibrium between liquid and vapor phases
( v = g , l = f , etc., where subscript g is used to indicate the saturated vapor and
subscript f the saturated liquid phase) and combining mass conservation equations
(4.1 ) and (4.2) yields the following volumetric continuity equation:
XX

(4.16)

1 d f d g
j
= v fg
+
dp dp
z
f
g

p j f d f
j d g

+ g
t dp dp
g

Combining energy conservation equations (4.3 ) and (4.4) yields,


X

101

p
.
z

C H A P T E R

=
(4.17)

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

q PH
+
Ai fg

di f
di g p
di fg
di f
1
g
G
+x
1 f (1 )

i fg
dp
dp t
dp
dp

p

z

In addition, from mass conservation equations and using Eq. (4.8) one can derive the
void propagation equation,
X

C K

d g p j g d g p
+ CK
=

,
t
z g g dp t g dp z
z

(4.18)

where,
C K = C 0 j + U gj .

(4.19)

EXAMPLE 4-1: Find the two-phase mixture density as a function of the axial
distance and time for a boiling channel with a step change in inlet velocity. Assume
uniformly heated channel with saturated inlet conditions, homogeneous flow
model and constant system pressure.
With this assumptions Eq. (4.16) becomes,
X

j
= v fg = .
z

(4.20)

From Eq. (4.11) one gets,


X

m G m ( m j ) m

j
+
=
+
=
+ m
+ j m = 0,
t
z
t
z
t
z
z

(4.21)
or,

j
+ j m = m
= m .
t
z
z

(4.22)

Equation (4.20) can be integrated as,


X

j ( z , t ) = z + f (t ) ,

(4.23)

where f(t) is a function of time only. Its value can be found substituting z = 0 in Eq. (4.23),
X

(4.24)

j(0, t ) = f (t ) = jin (t ) ,

and Eq. (4.23) becomes,


X

(4.25)

j( z, t ) = z + jin (t ) .

Assuming that the inlet velocity is given as,


(4.26)

j
jin (t ) = 1
j2

for t 0 .
for t > 0

102

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

and taking into account the equation for a trajectory of a particle moving with velocity j,
(4.27)

dz
= j = j in (t ) + z ,
dt

one gets,
(4.28)

j1 (t t0 ) j2 j1 t j 2
+
e
e

z (t , t 0 ) =
j 2 (t t 0 )
e
1

if

t0 0

if

t0 > 0

for t > 0, where t0 is the particle entrance time (z(t0) = 0).


Solving Eq. (4.22) along particle trajectories (i.e., along the characteristic dz/dt = j) and combining with
Eq. (4.27) yields,
X

(4.29)

f
z j
j j1 t

+ 2 2
e
j1
j1
j1
m ( z, t ) =
f
z
1 + j

for 0 t

if

t>

1
z

ln 1 +

j2

z
1

ln1 +

j2

From Eq. (4.25) it is obvious that the step change of inlet velocity is immediately transferred along the
whole channel. That is at time t1 = 0 the superficial velocity at the channel exit is equal to j(L,0) = j1 +
L, whereas at any time t2 > 0, this velocity will be j(L,t2) = j2 + L. This is obviously a non-physical
behavior caused by the fact that pressure changes are neglected.
X

An instant change of the superficial velocity in the whole channel does not mean an instant change of the
mixture density, void fraction and enthalpy. All these properties will change gradually as the particles will
move along the channel.
Assuming that a particle enters the channel at time t0 = 0, it will arrive at a position z1 in the channel at
time t = 1 ln1 + z1 . From Eq. (4.29) results that the mixture density at that location will start
1

j
X

changing between time t = 0 and time t = t1 and then it will remain constant.
Having expression given by Eq. (4.29) for the mixture density, one can derive similar time-and-location
dependent expressions for the void fraction and the thermodynamic mixture enthalpy, using Eqs. (4.17)
and (4.14), respectively. FIGURE 4-1 shows the time evolution of the mixture density in a boiling
channel with 70 bar pressure, pipe diameter equal to 10 mm, heat flux equal to 6105 W m-2, velocity of
the saturated water at the inlet equal to 2 m s-1. At time t = 0 the inlet velocity is reduced to 1.5 m s-1. The
evolution of the mixture density is shown at different location from the inlet to the channel.
X

As can be seen, at z = 0.5 m the density of the mixture drops from the initial value of 363 kg m-3 to a new
value of 310 kg m-3 within 0.22 s, whereas at z = 3.5 m density drop is from 90 to 70 kg m-3 within 0.6 s.

103

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

400
z = 0.5 m
z = 1.0 m

350

z = 1.5 m
z = 2.0 m

Density, kg/m^3

300

z = 2.5 m
z = 3.0 m

250

z = 3.5 m
200
150
100
50
-0.2

0.0

0.2

0.4

0.6

0.8

1.0

Time, s

FIGURE 4-1. Mixture density evolution with time at different locations in a boiling channel: pressure 70
bar, pipe diameter 10 mm, heat flux 6105 W m-2, inlet velocity 2 m s-1, reduced to 1.5 m s-1 at t = 0.

800
t = 0.0 s
700

t = 0.1 s
t = 0.2 s

Density, kg/m^3

600

t = 0.5 s
t = 1.0 s

500
400
300
200
100
0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Distance, m

FIGURE 4-2. Mixture density along axial distance at different time instances in a boiling channel:
pressure 70 bar, pipe diameter 10 mm, heat flux 6105 W m-2, inlet velocity 2 m s-1, reduced to 1.5 m s-1
at t = 0.
The mass flux of the mixture can be calculated as G (z , t ) =
given by Eqs. (4.25) and (4.29 ), the result is as shown in
X

104

j (z, t ) m (z, t ) . Substituting functions

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

1500
z = 0.5 m
1450

z = 1.0 m
z = 1.5 m

Mass flux, kg/m^2.s

1400

z = 2.0 m
z = 2.5 m

1350

z = 3.0 m
z = 3.5 m

1300
1250
1200
1150
1100
-0.2

0.0

0.2

0.4

0.6

0.8

1.0

Time, s

FIGURE 4-3. Mixture mass flux evolution with time at different locations in a boiling channel: pressure
70 bar, pipe diameter 10 mm, heat flux 6105 W m-2, inlet velocity 2 m s-1, reduced to 1.5 m s-1 at t = 0.

1500
t = -0.0 s
1450

t = +0.0 s
t = 0.2 s

Mass flux, kg/m^2.s

1400

t = 0.5 s
t = 1.0 s

1350
1300
1250
1200
1150
1100
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Distance, m

FIGURE 4-4. Mixture mass flux along axial distance at different time instances in a boiling channel:
pressure 70 bar, pipe diameter 10 mm, heat flux 6105 W m-2, inlet velocity 2 m s-1, reduced to 1.5 m s-1
at t = 0.
It is interesting to note that the mass flux change consists of two parts: the first drop is caused by the
sudden change of the inlet velocity. This drop takes place in the whole channel at the time of the inlet
velocity change (t = 0). The second drop is caused by the change of the mixture density, which is
transported in the boiling channel with the flow velocity.

105

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

EXAMPLE 4-2: Perform similar analysis as in EXAMPLE 4-1 but using the driftflux model. All other assumptions are the same.
X

The governing equations are obtained from Eqs. (4.16) and (4.18) and, after taking
into account the assumptions, will be as follows:
X

j
= v fg = ,
z

(4.30)
and

C
+ CK
=
K = g C0
v

t
z g
z
fg

(4.31)

The boundary/initial conditions are as follows,


(4.32)

(0, t ) = 0 ,

(4.33)

j
jin (t ) = 1
j2

for t 0 .
for t > 0

Integration of Eq. (4.30 ) yields,


X

(4.34)

j( z, t ) = jin (t ) + z .

Assuming that,
dz
= C K = C K ,in + C0z ,
dt

(4.35)

EE

Equation (4.31) can be re-written as,


v

dz D
+

= g C0
v

t dt z
Dt
fg

(4.36)

where D/Dt is the total derivative.


Equation (4.35) can be integrated and the particle trajectories can be obtained. If a particle enters the
channel at time t0 < 0, then Eq. (4.35) becomes,
X

(4.37)

dz
= C0 j1 + U gj + C0z ,
dt

or,
(4.38)

dz
dz
= dt
= dt .
C0 j1 + U gj + C0z
C
j
+
U
+
C

z
0
1
gj
0
0
t0

The integration is performed from the channel inlet (z = 0, t = t0) to a certain position (z,t<0) and it
yields,
z

(4.39)

C j + U gj + C0z
1
= C0(t t0 ) ,
ln (C0 j1 + U gj + C0z ) = t t0 ln 0 1

C0
C
j
+
U
0 1
gj
0

106

C H A P T E R

(4.40)

z ( t ; t0 ) =

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

C0 j1 + U gj C 0 ( t t 0 )
e
1 .
C0

Equation (4.40) describes the particle path from the inlet at time t = t0 < 0 until t = 0. Substituting t = 0
into Eq. (4.40 ) one can obtain the axial position of the particle at the time of the step change of the inlet
velocity:
X

(4.41)

z0 = z (0; t0 ) =

C0 j1 + U gj
C0

[e

C0t0

1 .

To find the particle trajectory for t > 0 one has to integrate Eq. (4.35 ) from z = z0 and t = 0 to certain
location (z,t>0). Eq. (4.35 ) now becomes,
X

dz
= C0 j2 + U gj + C0z ,
dt

(4.42)

and the solution is,


z

C 0 j 2 + U gj + C 0 z
1
ln (C 0 j 2 + U gj + C 0 z ) = t 0 ln
C j + U + C z
C0
gj
0
0
0 2
z

(4.43)

(4.44)

z (t ; t0 ) =

C0 j2 + U gj
C0

(e

C0t

= C 0 t ,

1 + z0e C0t .

Substituting z0 given by Eq. (4.41 ) into (4.44 ) yields,


X

z ( t ; t0 ) =

(4.45)

C0 j2 + U gj C0 t
C j + U gj C 0 t 0
e
e
1 + 0 1
1 e C 0 t =
.
C0
C0

C0 j1 + U gj C 0 (t t 0 ) j2 j1 C 0 t C0 j2 + U gj
e
e
+

C0
C0

Paths described by Eq.(4.45 ) are valid for t0 < 0. The solutions for t0 > 0 can be readily obtained from
Eq. (4.44 ) substituting z0 = 0 and t = t t0. The result is as follows,
X

(4.46)

z ( t ; t0 ) =

C0 j2 + U gj C0 (t t0 )
e
1 .
C0

The void fraction distribution can be obtained by integration of Eq. (4.36 ) and using Eqs. (4.45) with
(4.46 ). The resultant expression is as follows,
X

(4.47)

*
vg 1
for 0 < t t ( z )
v fg C0
C
j
+
U

j
0
2
gj
C
t

z+
2 1 e 0

C0 j1 + U gj

C0


( z, t ) =

vg
1
for t > t * ( z )

z
v
C
0
fg 0 1 +

C0 j2 + U gj

where,
(4.48)

t* ( z) =

1
C0z .
ln1 +
C0 C0 j2 + U gj

107

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

The mixture density can be found from Eq. (4.6 ) and using expression for void fraction given by Eq.
(4.47 ). In the same manner one can obtain the mixture enthalpy from Eq. (4.14 ) or (4.15).
X

EXAMPLE 4-3: Solve same problem as in EXAMPLE 4-1 assuming subcooled


liquid at the inlet to the channel.
X

The energy equation in the single phase flow region is,


(4.49)

Since
(4.50)

il
i qPH .
+G l =
t
z
A

G = j l , the equation becomes,


Dil il
i
qPH .

+ j l =
Dt t
A l
z

Integrating Eq. (4.50) along characteristics dz/dt = jin(t) yields,


(4.51)

il (t ) il (0) =

qPH
t
A l

Here i(0) = in is the inlet enthalpy, which is assumed to be constant. The residence time of liquid particle
in the subcooled region can be found from Eq.(4.51 ) as,
X

(4.52)

t1 =

A l
(i f iin ) = const .
qPH

Integrating Eq. (4.30 ) for z > (t), where (t) is the boiling boundary yields,
X

(4.53)

j( z, t ) = jin (t ) + [z (t )].

Again solving along characteristics dz/dt = j(z,t) yields,


(4.54)

z (t , t0 ) = [ j2 t1 + ( j2 j1 )t0 ]e

( t t1 t0 )

( t )

[ j2 ( )d ] ,

t1 + t0

when t > 0.
For a step change of the inlet velocity, Eq. (4.54 ) becomes,
X

(4.55)

j1 t0 j2 j1 (t t1 )
1

+
+ j2 t1
e
for t > t1 and t0 0
e

1
j (t t t )

z (t , t0 ) = 1 e 1 0 + ( j2 j1 )t + j1 t1
for t1 + t0 t < t1 and t0 0

j2 (t t1 t0 )
1

e
+ j2 t1
for t > t1 and t0 > 0

Solving Eq. (4.22) along trajectories z(t,t0),


X

(4.56)

m ( t , t 0 ) = f e (t t ) ,
0

and eliminating t0 from Eqs.(4.55 ) and (4.56) yields the solution for m ( z, t ) .
X

108

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

EXAMPLE 4-4: Consider uniformly heated boiling channel subject to a step


change in the wall heat flux. Other assumptions are the same as in EXAMPLE
4-1.
X

Equations (4.20 ) and (4.22 ) are still valid with,


X

(4.57

= 1
2

for t 0

for t > 0

Integrating the equation of the characteristics given by Eq. (4.27) yields,


X

(4.58)

j
jin (2t 1t0 )
e 2t + in e 2t 1
e

2
z (t , t0 ) = 1
j
in e 2 (t t0 ) 1
2

if

t0 0 ,

if

t0 > 0

for t > 0. Next, integrating Eq. (4.22 ) along the trajectories z(t,t0) and eliminating t0, the density variation
m ( z, t ) can be obtained as,
X

(4.59)

f
z
1 2t
1 + 1 + 2
e
jin
2
2
m ( z, t ) =
f
2 z

1 + j

in

for 0 t

if

t>

1
z
ln 1 + 2
jin
2
.

1
z
ln1 + 2
2
jin

4.2 Flow Instabilities in Heated Channels


Various thermal-hydraulic instabilities have been observed in two-phase flows. As
already mentioned, instabilities are undesirable since they may degrade system control
and performance, erode thermal margins and lead to mechanical damages. Nuclear
reactors have to be designed and operated in such a way that sustainable instabilities
are avoided. This requires a good understanding of the phenomena that govern
various modes of instabilities in nuclear power plants as well as knowledge of stability
margins, which are valid for particular system and operating conditions.
4.2.1

Classification of Instabilities

Two-phase instabilities can be conveniently classified into static instabilities, which


can be explained in terms of steady-state laws, and dynamic instabilities, which
require a consideration of the transient conservation equations.
Static Instabilities:
1. Excursive (i.e., Ledinegg) instabilities are non-periodic flow transients.
Instabilities of this type plagued early low-pressure fossil boilers, since flow
excursions could lead to burn-out of the boiler tubes. Ledinegg instability may
occur in heated channels with low system pressure and low inlet loss
coefficient, where pressure drop may decrease with increasing flow. If the
pump characteristics has less negative slope, excursive instability will occur.
2. Flow regime relaxation instabilities are caused by the pressure-drop
characteristics of the different flow regimes. For example, pressure drop in
109

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

slug flows is less then in bubbly flow with the same flow rates of gas and
liquid. If a system is operating in the bubbly flow regime near the flow regime
boundary, a small negative perturbation in liquid flow rate may cause a
transition to slug flow. As a result, pressure drop in the channel will reduce. If
the channel operates at a constant pressure drop condition (as in case of a
large number of parallel channels), more liquid will enter the channel to satisfy
the boundary conditions. This, in turn, may cause the system to return to
bubbly flow regime.
3. Nucleation instabilities include bumping and geysering phenomena. These
instabilities are characterized by a periodic relaxation of the metastable
condition that builds up due to insufficient nucleation sites. In particular, if the
liquid superheat builds up until the existing nucleation sites are activated, rapid
boiling and expulsion of the resultant two-phase mixture may occur.
Dynamic Instabilities:
1. Density-wave oscillations can occur in both diabatic and adiabatic two-phase
systems and in diabatic single-phase systems. Generally speaking, density wave
oscillations are caused by the lag introduced into the thermal-hydraulic system
by the finite speed of propagation of density perturbations. This type of
instability is one of the most important and of practical concern in BWRs and
will be discussed in more detail in the following part of this section.
2. Pressure-drop oscillations can occur in loops having a negative slope (similar
to the situation described for the excursive instability) and containing a
compressible volume (e.g. an accumulator). In such systems excursions may
occur periodically.
3. Flow regime excited instabilities can occur when a particular flow regime,
normally slug flow, induces a periodic disturbance in the system operating
state. If this disturbance is at a frequency that is close to the natural frequency
of the two-phase system, a resonance can occur.
4. Acoustic instabilities may occur in two-phase system having the proper
combination of geometric characteristics and sonic speed. As in single-phase
gas flows, organ-pipe-type standing waves can be set up when a pressure pulse
propagates through two-phase mixtures flowing in a conduit. On reaching an
area change or obstruction, the change in acoustic impedance causes a
pressure pulse of opposite polarity to propagate in the opposite direction. If
the excitation frequency and geometry of the conduit is such that an integral
number of one-quarter wavelengths can fit within it, the standing waves may
appear. Such acoustic-induced channel pressure drop oscillations of large
amplitude have been observed for subcooled systems operating in the
negative-slope region of the system pressure-drop versus flow curve.
5. Condensation-induced instabilities are known to lead to large water-hammertype loads; however, their nature is not fully understood. A typical example is
the so-called chugging phenomena that have been observed in the vent pipes
of steam relief valves which are submerged in a liquid pool. When the steam
first exits into the subcooled pool of liquid it is normally at a high enough
110

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

velocity to form a jet within the pool. However, as the steam flow rate drops
off, the condensation rate in the pool may be large enough to completely
collapse the steam jet, and cause a liquid slug to surge up into the discharge
line. Subsequently, the steam can heat up the interface of the liquid slug to
saturation, allowing the pressure of the discharging steam to increase such that
it blows the slug back into the liquid pool. A cyclic process can occur with
large inertial loads associated with the liquid slug motion being transmitted to
the walls of the vessel containing the pool.
4.2.2

Density Wave Oscillations

To understand the mechanisms of density wave oscillations, consider an air/water


flow channel connected to a tank filled with water, (see FIGURE 4-5, depicting a
system investigated by Svanholm and Friedly). Water is entering into the pipe through
an inlet orifice and gas is constantly introduced into the pipe and mix with water at a
certain distance downstream from the inlet orifice. To simplify the analysis, the
following assumptions are made. The density of the air is negligible in relation to that
of the water and the volume occupied by the water is negligible compared to that of
the air. Finally, slip is also neglected.
X

Liquid

pex
ptot

pin
Uin

Uex

Gas inlet
FIGURE 4-5. Gas/liquid flow channel.

The airflow determines the velocity of the two-phase mixture and the density of the
mix is proportional to that of the liquid. Assuming that the pipe has no losses except at
the inlet and exit, the hydraulic head loss at the inlet orifice is,
2

(4.60)

pin = in

U in l
.
2

The pressure drop at the exit is,


2

(4.61)

pex = ex

U ex ex
.
2

Here Uin and Uex are mean inlet and exit velocities, respectively, in and ex are the inlet
and outlet pressure loss coefficients, respectively, and ex is the outlet (two-phase)
111

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

density. The constant pressure head, provided by the water tank, is always equal to the
sum of these two pressure drops, i.e.
(4.62)

ptot = pin + p ex = l gh .

For example, opening the inlet orifice for a short time induces a density wave that
propagates through the pipe and passes the exit orifice after a time . During the
passage, the corresponding pressure drop at the exit increases temporarily. Since the
total pressure drop always remains constant, the increased pressure drop over the exit
will bring a corresponding pressure decrease at the inlet. This means that less water is
sprayed into the channel creating a sudden decrease in density, which induces the wave
that propagates through the channel. This dynamic process causes density wave
propagation in the channel.
If the amplitude of the waves is converging the system is said to be stable and if the
amplitude is diverging it is said to be unstable. These finite propagation times, , induce
time-lag effects and phase-angle shifts between the channel pressure drop and the inlet
flow, which may cause self-exciting oscillations.
In general, any increase in the frictional pressure drop in the liquid region has a
stabilizing effect, since this pressure drop is in phase with the inlet flow and acts to
damp the fluctuations. Inlet orificing can be used to stabilize an unstable flow. An
increase in the two-phase region pressure drop has a destabilizing effect, since the
pressure drop is out of phase with the inlet flow, due to the wave propagation time, .
Thus, an exit flow restriction is a strongly destabilizing factor.
For a fixed channel geometry, an increase in inlet velocity has a stabilizing effect in
terms of heat flux, because the extent of the two-phase flow region and the density
change due to boiling are significantly reduced by the increase in inlet velocity. An
increase in the system pressure has a stabilizing effect in terms of exit quality, since at
higher pressure the density change due to phase change is less significant.
Density-wave instabilities can be further classified as follows:
1. Loop instabilities
2. Parallel-channel instabilities
3. Channel-to-channel instabilities
4. Neutronically-coupled instabilities
The most important modes of density-wave instabilities are loop and parallel-channel
instabilities. The parallel-channel mode corresponds to a system of a big number of
channels connected in parallel, in which a constant pressure condition governs flow
through each of the channels. The principles of density-wave instability that are
described in this section correspond just to this mode of instability. The loop instability
is very similar; however, the boundary condition of zero pressure drop in the loop is
imposed.

112

C H A P T E R

4.2.3

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

Methods of Analysis of Two-Phase Flow Instabilities

A rigorous stability analysis of boiling systems is difficult due to the nonlinear


mathematical form of the underlying conservation equations and is only possible if
some simplifying assumptions are made. In particular, if the threshold of instability is
of interest, linearized models are often used. Such models are obtained by perturbing
the governing equations around a given steady-state operating point. The linearized
model is next Laplace-transformed and a frequency domain methodology is used to
study the system stability.
While the linear stability analysis can be used to determine the instability threshold, this
approach does not provide information concerning other characteristics of nonlinear
systems, such as the magnitude and frequency of any limit cycle oscillations. For this
purpose a non-linear stability analysis must be performed.
In general, the methods of linear analysis of two-phase flow instabilities consist of
the following steps (so-called frequency-domain methodology):
1. Linearize governing equations around steady-state operating point
2. Obtain transfer functions
3. Examine properties of roots of characteristic equations
The methods of nonlinear analysis of the phase instabilities are as follows:
1. Hopfs bifurcation method
2. Method of Lyapunov
3. Harmonic quasi-linearization (describing function methods)
4. Fractals as measure of strange attractors (chaotic vibrations)
Nonlinear methods are beyond the scope of the present course and will not be
discussed here (the interested reader may consult literature, e.g.,[4-1]). In what follows
the basic principles behind the linear methods will be discussed.
4.2.4

Frequency Domain Methodology

Frequency domain methodology has been previously used to investigate the stability of
a nuclear reactor. As a short remainder, a generic block diagram of a negative
feedback-control system is shown in FIGURE 4-6. This is a simple linear system with
a single input and single output (SI/SO).
X

In

G(s)
Out

H(s)

FIGURE 4-6. Block diagram of a feedback control system

113

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

The transfer function G(s) is called the forward loop transfer function, and H(s) the
feedback loop transfer function. The output and input are related to each other as
follows,
(4.63)

X out ( s ) = [ X in ( s ) H ( s ) X out ( s )] G ( s ) ,

from which the output signal can be obtained as


(4.64)

G (s)
X out ( s ) =
X in ( s ) .
+

1
G
(
s
)
H
(
s
)

The transfer function G/(1+GH) is called the closed loop transfer function and
1+GH=0 is the characteristic equation of the close loop. If any of the roots of 1+GH
will appear in the right-halve of the complex plane, the closed system will be unstable.
Perturbation analysis
The first order perturbation of a time dependent function f (t ) can be viewed as a first
order Taylors series expansion,
(4.65)

f f ( x (t )) f 0 =

f
x (t ) ,
x 0

where f 0 is the value of the unperturbed function. Consider a few examples:


Let f ( x) = c1 x a . The perturbed equation is,
f ( x) = ac1 x 0a 1x

Let f ( x) =

(4.66)

f
x
=a
f0
x0

dx
: then,
dt

dx d (x)
.
=
dt
dt
g2 ( x)

Let f ( x) =

G ( x, z)dz : then, based on the Leibniz rule,

g1 ( x )
g20

(4.67)

f (x) = G(x, z)dz + G0 (x, g 2 (x))g 2 G0 (x, g1 (x))g1 .


g10

Perturbations are quite similar to differential calculus. The only real difference is that
perturbations are taken around steady state point and all variables not perturbed are
denoted by a subscript zero to indicate the steady state value.

114

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

Derivation of perturbed equations


The time-dependent pressure drop in a boiling channel can be obtained from the
integration of the momentum equation (e.g. Eq. (4.5 ) or (4.13 )). The integration can
be simplified by dividing the channel into two parts: a single-phase flow part, which is
stretching from the inlet to the boiling boundary, located at distance from the inlet,
and the two-phase flow part, stretching from the boiling boundary to the outlet from
the channel, see FIGURE 4-7 .
X

z=L

Two phase

z=

Single phase

z=0

FIGURE 4-7. Single-phase and two-phase sections of a boiling channel.

Using the homogeneous equilibrium model, the pressure drop across each section is
given as,

(p )

1 H

(4.68)

= pin p( ) =

G 1 G 2 4C f ,lo G 2

G 2 ( zi ) ,

+
+
+
g
dz
+
i
0 t l z Dh 2l l i
2 l
L1

for the single phase flow part and,

(p )

2 H

(4.69)

= p() pex =

G G2 2 4Cf ,lo G2

G2(zi ) ,
2
lo,i (zi )i
t + z M +lo Dh 2l +mgdz+i
2l
L2

LH

for the two-phase flow part. Here is the position of the boiling boundary, L1 is the
non-boiling length of the channel and L2 is the boiling length of the channel.
Combining Eqs. (4.68 ) and (4.69) with equations for mass and energy conservation,
perturbing around a steady state point and then Laplace transforming yields the
X

115

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

following general expressions for the single-phase and two-phase pressure drop
perturbations,
(4.70)

(p1 )H = 1, H ( s )jin + 2, H ( s )q H + 3, H ( s )iin ,

(4.71)

(p2 )H = 1, H ( s )jin + 2, H ( s )q H + 3, H ( s )iin .

Here 1, H , 2, H and 3, H are transfer functions of the perturbations of inlet velocity,


heat volumetric source and inlet enthalpy, respectively, on the perturbation of the
single-phase pressure drop. 1, H , 2, H and 3, H are the corresponding transfer
functions for the perturbation of the two-phase pressure drop. Derivation of Eqs.
(4.70 ) and (4.71 ), as well as expressions for transfer functions can be found in
Appendix D.
X

Stability criterion
As an example, parallel channels with constant power and constant inlet enthalpy will
be considered. In such case, q H = iin = 0 and since the total pressure drop is
constant,
(4.72)

(p1 )H + (p 2 )H = 0 .

Combining Eqs. (4.70 ) and (4.71 ) with (4.72) yields


X

(4.73)

1, H

( s ) + 1, H ( s )]jin = 0 .

In the case of parallel channel instabilities density-wave oscillations are manifested by


self-sustained oscillations in the channel inlet flow rate caused by a feedback between
the pressure drop perturbations in the single and two-phase portions of the channel.
Using the technique of linear system control theory, the parallel channel stability can
be considered as a feedback system. The appropriate block diagram is shown in
FIGURE 4-8.
X

jin ,tot
jin ,ext
1, H1

(pH )2

(pH )1
1

jin ,tot
1, H

FIGURE 4-8. Block diagram for the parallel channel model.

The block diagram shown in this figure yields the following relationship between the
external perturbation jin , ext and the system response, jin , tot ,

116

C H A P T E R

jin , tot

(4.74)

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

1
jin , ext .
=
(
)

s
1 + 1, H

1, H (s )

It follows from this equation that the characteristic equation of the boiling channel is,
1, H (s )

1+

(4.75)

1, H (s )

= 0.

The necessary and sufficient condition for the boiling system under consideration to
be stable to small perturbations is that all the roots of the characteristic equation,
(4.75) have negative real parts.

In the case of complicated transcendental algebraic equations, such as those that occur
in Eq. (4.75 ), direct evaluation of roots is not trivial, and approach using the Nyquist
Criterion (see Appendix B) has proven to be an efficient way of investigating system
stability.
X

For boiling parallel channel,


(4.76)

G (s) = 1 +

1, H (s )
1, H (s )

and its poles are equivalent to the combined poles of [ 1, H (s ) + 1, H (s )] and zeros of
1, H (s ) . From the shape of transfer functions 1, H (s ) and 1, H (s ) is clear that all the

singularities of [ 1, H (s ) + 1, H (s )] are removable, whereas 1, H (s ) has no zeros


within C, thus, P = 0. Consequently, the parallel channel model derived herein is stable
if, and only if, the Nyquist plot of 1, H (s ) 1, H (s ) does not encircle the point (-1,0).
4.2.5

Time Domain Approach

In addition to the frequency-domain approach, the time-domain approach is used in


practical calculations of the nuclear reactor stability. In such calculations, the dynamic
behavior of the system caused by a certain perturbation of input parameters is
calculated as a function of time. FIGURE 4-9 shows an example of the perturbation
of an input parameter (this could be for instance the inlet subcooling) and an example
of the output parameter (this could be for instance the reactor power).
X

117

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

Perturbation

C H A P T E R

A1
Answer

A2

FIGURE 4-9. Examples of a perturbation and the resulting answer.

The answer shown in FIGURE 4-9 is typically obtained from an analysis performed
with a transient code where both reactor kinetics and reactor thermal-hydraulics
equations are solved simultaneously. If the answer has an oscillatory character then
after a certain period of time the least damped eigenfrequency will dominate. The
amplitude ratio, or the so-called decay ratio, A2/A1 is a practical measure how
effective the damping is and how far from the instability the system is. For undamped
system A2/A1 = 1. In such case the system is at the threshold of the instability. For
decay ratio A2/A1 < 0.25 the system is well damped.
X

Stability map of a boiling channel


It is instructive to represent the dynamic behavior of a boiling channel on a so-called
One such map proposed by Ishii and Zuber is shown in FIGURE
4-10 . The map shows the region where the boiling channel is unstable in function of
two non-dimensional parameters: the subcooling number and the phase-change
number. The subcooling number is defined as follows,

stability map.

(4.77)

N sub =

g )(i f iin )

g i fg

and the phase-change number as,


(4.78)

N pch =

g )q0PH LH
jin A f g i fg

Here LH is the heated length, PH is the heated perimeter, q 0 is the heat flux, A is the
channel cross-section area, jin is the inlet velocity, iin is the inlet enthalpy and ifg is the
latent heat.
118

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

Nsub
Subcooled
liquid

Two-phase
flow
xex=1

xex=0
Stable

Stability
boundary
Superheated
vapor

Unstable

Npch

Constant exit quality line

FIGURE 4-10. Stability map proposed by Ishii and Zuber.

It can be seen in FIGURE 4-10 that stability boundary curve for higher inlet
subcoolings is nearly parallel with the line of a constant exit quality xex, given by,
X

(4.79)

N sub = N pch

f g
x ex .
g

Ishii used this observation and derived a simple stability criterion for the high inlet
subcoolings as follows,
(4.80)

2( + 2 + ex ) g
xex in
.

1 + + ex f g

The boiling channel is stable as long as Eq. (4.80 ) is satisfied. Here in and ex are the
inlet and outlet loss coefficients, respectively and is the friction number given as,
X

(4.81)

C f LH
.
2 Dh

4.3 Instabilities in Heated Loops


4.3.1

Parallel Channel Instability

Parallel-channel system is often a part of a boiling loop. In particular, a core of a BWR


is a multi-channel system composed of a number of parallel boiling channels (fuel
assemblies) having common inlet and outlet plena. Clearly, all such channels must
satisfy an equal-pressure-drop boundary condition. FIGURE 4-11 shows a schematic
of a loop containing parallel boiling channels.
X

119

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

Condenser

Upper Plenum

Parallel
channels
Downcomer

Lower Plenum

Pump

FIGURE 4-11. A schematic of a loop containing parallel boiling channels.

The parallel channels may operate at different conditions. In LWR reactor core, the
operating powers of channels vary depending on the radial location of the channel in
the core. Inlet mass flow rate is typically regulated for different zones in the core. Due
to that, usually one channel will be least stable in the core. If the number of channels is
large and only one of them becomes unstable, the effect of oscillations in this single
channel will be small and will not affect the operation of the other, stable channels.
Therefore the analysis of the onset of the so-called parallel channel instabilities can be
done by considering the most unstable channel subject to a constant pressure drop
boundary conditions.
At steady-state conditions, pressure drop in each channel is given as,
(4.82)

pk = r3,k C f ,lo ,k

G2
G 2 N loc
2 L Gk2
+ r4,k L l g + r2,k k + lo2 ,i ,k i ,k k ,
Dh , k l
l i =1
2l

Here index k denotes specific channel, which may have individual geometry and local
loss coefficients and Nloc is the number of local losses, including the inlet and the outlet
pressure loss.
The constancy of the pressure drop for all channels implies that,
(4.83)

p1 = p1 = ... = p N = p .

The mass conservation equation for parallel channels is as follows,


(4.84)

k =1

Gk Ak = W ,

120

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

where N is the total number of parallel channels, W is the total mass flow rate through
the core and Ak is the cross section area of channel k. Eqs. (4.82 ) through (4.84 )
constitute a system of N+1 equations with N+1 unknowns (Gk and p). Solution of
the equation will give the individual mass flow rate in each channel.
X

4.3.2

Multiple Parallel Channels

In this situation the constant pressure drop condition is valid. Assuming a constant
power of the heater and ignoring changes in the lower plenum temperature, the
pressure drop perturbation of the channel can be obtained from Eqs. (4.70 ) and (4.71 )
as,
X

(4.85)

(p )H = (p1 )H + (p 2 )H =

f A

1, H

( s ) + 1, H ( s )]Win =

G H ( s )Win

If (p )H = 0 Eq. (4.85) will always have a trivial, steady-state solution Win . It may
also have periodic nonzero solution provided that
X

(4.86)

G H ( s ) = 1,H ( s ) + 1,H ( s ) = 0 ,

for some s = j 0 . In this case the channel is self-excited and will undergo selfsustained periodic oscillations at the angular frequency .
4.3.3

Boiling Loop Stability

The analysis of transients in boiling loops and systems involves the modeling of several
inter-related phenomena, such as: single-phase and two-phase flows in channels and
plena, boiling heat transfer, phase separation and mixing, unsteady heat conduction
and others. The equations for heat transfer and flow in different parts of the system
can be combined and solved using proper boundary conditions and using various
possible external perturbations. Additional possible models of different parts are
shortly described below.
Mixing in large plenum: boiling loop contains additional elements which need
modeling. Examples are lower and upper plenum, pump, condenser and pipelines.
Modeling of lower and upper plena requires some assumptions about mixing. Using
perfect mixing homogeneous flow model, lumped-parameter mass and energy
conservation equations become, respectively,
(4.87)

(4.88)

4.3.4

Vp

Vp

d m , p
dt

Ni

Nj

k =1

j =1

= Win,k Wex , j ,

d ( m, p im, p )
dt

Ni

Nj

k =1

j =1

= Win ,k iin ,k Wex , j iex , j + V p

dp
.
dt

Heated Wall Dynamics

So far the thermal energy added to a boiling channel was considered in the form of a
given heat flux q, at the channel wall. In reality the heater power rather than the wall
heat flux should be used as a given controlled parameter. Any changes in the power
121

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

generated in the heater will be transmitted to the coolant with a delay depending on
the heater geometry and material properties. Two distinct models can be considered:
One dimensional heat conduction model: in this case the lateral heat transfer (in fuel
rods it will be heat transfer through pellet, gas gap and cylindrical cladding.) can be
described by one-dimensional heat conduction equation,
(4.89)

T
q
= a h 2T +
,
c p
t

with wall heat flux given as,


(4.90)

q =

T
n

.
wall

Here T(r,t) is the temperature of the heater, q is the volumetric heat generation rate
and ah is the thermal diffusivity of the heater.
Lumped-parameter model: for simplicity, the distributed parameter model can be
replaced with a lumped-parameter model as follows,
(4.91)

H c p ,H AH

dTw
= qAH qPH ,
dt

with wall heat flux given as,


(4.92)

q = h (Tw Tl ) ,

where h is the effective heat transfer coefficient, Tl is the coolant bulk temperature, AH
the heater cross-section area.
Pressure boundary conditions: the boundary condition for a boiling loop is obtained
by summing-up pressure drops over the closed flow around the loop,
(4.93)

ploop p pump = 0 ,

where p pump is the pump head (for a natural circulation loop this part is equal to
zero) and ploop is pressure drop in the rest of the loop.
N O M E N C L A T U R E

A
cp
C0
Cf
D
g
G

cross-section area
specific heat
drift-flux distribution parameter
Fanning friction coefficient
diameter
gravity constant
mass flux
122

C H A P T E R

h
i
j
L
p
P
q
q
q
Q
r
t
T
U
V
W
x
z

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

heat transfer coefficient


specific enthalpy
superficial velocity
length
pressure
perimeter
linear heat
heat flux
volumetric heat source
heat
radius
time
temperature
velocity
volume
mass flow rate
quality
axial coordinate

Greek

void fraction
heat conductivity
molecular kinematic viscosity
molecular dynamic viscosity
density
evaporation rate

Subscript
f
fluid; saturated liquid phase
g
gas phase; saturated vapor
h
hydraulic
H
heated
k
pertinent to phase k
l
liquid phase
m
mixture
v
vapor
w
wall
R E F E R E N C E S

[4-1]

Lahey R.T. Jr. Ed. Boiling Heat Transfer, Modern Developments and Advances, ISBN 0-444-89499-3,
1992.

E X E R C I S E S

EXERCISE 4-1: In a BWR fuel assembly operating with the total power 1.5 MW and having constant
inlet mass flow rate 2.4 kg/s the inlet coolant velocity was suddenly reduced with 10%. Plot the change of
void fraction, mixture density, mixture enthalpy and mixture velocity at 3.3 m distance from the inlet to
the assembly. Use homogeneous model and assumptions adopted in EXAMPLE 4-1 . Use steam-water
property at 70 bar pressure. The assembly has total length L = 3.65 m, cross-section area A = 0.0016 m2,
heated perimeter PH = 0.7m, hydraulic diameter Dh = 0.01 m. Assume saturated conditions at the inlet
and uniform distribution of heat flux along the assembly.
X

123

C H A P T E R

D Y N A M I C S

O F

B O I L I N G

S Y S T E M S

EXERCISE 4-2: Plot the stability map for a boiling channel with length LH = 3.6 m, hydraulic diameter
D = 10 mm and pressure p = 70 bar. The inlet loss coefficient is equal to 4 and the outlet is equal to 0.
Use the Blasius formula for the Fanning friction factor Cf assuming Re = 105. What is the maximum exit
quality to keep the channel stable?

124

Appendix

A
Appendix A - Laplace
Transformation
Many analysis techniques center on the use of transformed variables to facilitate
mathematical treatment of the problem. In the analysis of continuous time dynamical
systems, the use of the Laplace transform predominates.
The Laplace transform are used to convert time domain relationships to a set of
equations expressed in terms of the Laplace operator 's'. Thereafter, the solution of the
original problem is affected by simple algebraic manipulations in the 's' or Laplace
domain rather than the time domain. The one-sided Laplace Transform of a time
function f(t) is defined as:

F ( s ) = L{ f (t )} = e

b
st

f (t )dt = lim e st f (t )dt

a 0
b a

(0 < a < b ) .

(A-1)

The transform associates a unique result or image function F(s) of the complex
variable s = a+jb with every single-valued object or original function f(t) (t real) such
that the improper integral (A-1) exists. F(s) is called the (one-sided) Laplace transform
of f(t).
It is customary to design the image function with the capital letter, corresponding to
the original function, and replace t argument with s. That is, U(s) is a Laplace transform
of a function u(t). Another notation, used throughout this book is to use a hat symbol
for the image function, that is u (s ) is an image of u(t).
Basic properties of the Laplace transform are as follows:
The Laplace transform is linear:

L{ f (t )1 + f 2 (t )} = L{ f (t )1 } + L{ f 2 (t )} = F1 ( s ) + F2 ( s )

(A-2)

Laplace transforms of derivatives are given by the following expression:


df
L = L{ f } = sF ( s ) f (0)
dt

(A-3)

In general:
125

A P P E N D I X

L A P L A C E

T R A N S F O R M

d n f
L n = L{f n }= s n F ( s ) s n 1 f (0) ... f n 1 (0)
dt

(A-4)

Laplace transforms of selected functions are given below.


Step function
Consider a function f(t) = Ku(t), where u(t) is a step-function defined as,
1
u (t ) =
0

for t > 0
for t 0

Applying the Laplace transformation yields,

K
L{Ku(t )} = Ku(t )e dt = K e dt = e st
s
0
0
st

t =

st

=
t =0

K
(0 1) = K .
s
s

Impulse (Dirac delta) function


Strictly speaking Dirac delta function is not a function, but a so-called distribution and
symbolically described as , , However, here the function form will be used, in
which the delta function is defined as follows,

(t ) = 0

for t 0, (0) = +,

(t )dt = 1

From the definition of the Laplace transformation one obtains,

L{ (t )} = 1 .
Exponential function
f (t ) = e at ,
L{e

at

}= e

at

e dt = e
st

( s + a )t

1 ( s +a )t
dt =
e
s+a

=
0

1
(0 1) = 1
s+a
s+a

To return to the time-domain from the Laplace domain, inverse Laplace transform is
used. Again this is analogous to the application of ant-algorithms and as in the use of
logarithms, tables of Laplace Transform pairs help to simplify the task, see Table A-1.
Table A-1. The Laplace transform pairs.
f(t)

F(s)

x(t)+y(t)

X(s)+Y(s)

126

A P P E N D I X

N Y Q U I S T S

S T A B I L I T Y

A f(t)

A F(s)

sF(s)-f(0)

fn

s n F ( s ) s n1 f (0) ... f

1/s

1/s2

e at

( 0)

1
s+a

sin t , cos t

s +
tn ,

n 1

C R I T E R I O N

t n e at

n!
s

n +1

s
s +2
2

n!

(s a )n+1

Of particular interest for nuclear-reactor kinetics are the following properties of the
Laplace transformation.
If L(s) and K(s) are polynomials with non-zero, single-valued roots and the degree of
the polynomial L(s) is lower than the degree of polynomial K(s) = n, then:
L ( s ) n L ( s k ) sk t
L1
e
=
K ( s ) k =1 K ( s k )

(A-5)

L( s ) L(0 ) n L( sk ) sk t
L1
+
e
=
sK ( s ) K (0 ) k =1 sk K ( sk )

(A-6)

The following example illustrates the use of the Laplace transform in the system
analysis, showing how it is used to solve linear ordinary differential equation (ODE).
Consider the first order process, described by the following differential equation:

dY (t )
+ Y (t ) = K U (t )
dt

(A-7)

Here Y(t) is the output variable and U(t) is the forcing input. The time-domain
solution of this ODE will be found using the Laplace transform approach. Let define
new variables, which are deviations from steady state solutions,
y ( t ) = Y ( t ) Y0

u (t ) = U (t ) U 0

(A-8)

127

A P P E N D I X

L A P L A C E

T R A N S F O R M

Here Y0 is the steady-state value that Y(t) will attain given a steady input U0. Assuming
that initially the system is at the steady-state condition, the following is valid:
y ( 0) = u ( 0 ) = 0

(A-9)

Substituting ((A-8) into (A-7) yields,

dy (t )
+ y (t ) = K u (t )
dt

(A-10)

Application of the Laplace transform gives,

sY ( s ) y (0) + Y ( s ) = KU ( s )
or, since y(0) = 0,

sY ( s ) + Y ( s ) = KU ( s )

(A-11)

Equation (A-11) can be transformed as follows,


G( s) =

Y ( s)
K
=
U ( s ) 1 + s

(A-12)

G(s) is called the transfer function of the considered process and describes the
relationship between input U(s) and the output Y(s). The time-domain solution of ((A11) depends on the shape of the input function. Assuming a unit step change in u(t),
i.e. u(t) = 1, t > 0, ((A-12) yields,

Y ( s) =

1
K
1
1
1
=K
= K

(1 + s ) s
s (1 + s )
s 1 + s

(A-13)

Using Table A-1, the time solution to ((A-13) is found as,


y (t ) = K (1 e t / ) Y (t ) = K (1 e t / ) + Y0

(A-14)

Transfer functions play a central role in the analysis of dynamic systems behavior since
they fully describe the relation between input-output pairs. In the previous example,
the transfer function G(s) encapsulates the behavior of the system given by ((A-7) in
the ratio given in ((A-12).
Transfer function possesses several interesting features:

Transfer functions are independent of the form of the input.

Transfer functions obey algebraic rules.

Commutative property

G1(s)G2(s) = G2(s)G1(s)
128

A P P E N D I X

N Y Q U I S T S

Associative property

Transfer functions are linear functions

S T A B I L I T Y

C R I T E R I O N

G1(s) + G2(s) = G2(s) + G1(s)

Transfer functions can be visualized using block diagrams, as shown below:


Block diagram

U(s)

G(s)

Transfer Function

Y(s)
Y (s)
= G (s)
U (s)

U(s)

G1(s)

Y(s)

G2(s)

Y ( s)
= G1 ( s )G 2 ( s )
U ( s)

G1(s)
Y(s)

U(s)
+
G2(s)

Y ( s)
= G1 ( s ) + G 2 ( s )
U ( s)

U(s)
+

G(s)

Y(s)

Y (s)
G (s)
=
U ( s) 1 + G ( s)

129

Appendix

B
Appendix B Nyquist
Stability Criterion
The Nyquist stability criterion is used to investigate stability of any (open or closed)
system using its frequency characteristics. Almost always the criterion is used for
closed systems, though. A useful, but not the most general formulation of the Nyquist
criterion is as follows: if the mapping of the transfer function of an open system
G(s)*H(s) on the plane GH encircles point (-1,0) then the system is unstable if it is
closed. The criterion is illustrated in Figure B-1.

Im(GH)
Plane GH

=
-1

=0
Re(GH)

A
B

Figure B-1: Nyquist plots of a stable and unstable closed system. Curve A does not
encircle point (-1,0) and the system is stable. Curve B encircles point (-1,0) and the
system is unstable.
The Nyquist criterion can be easily proved, but first some introductory on complex
variables and functions is required.
A complex variable s can be expressed as,
s=x+jy

(B-1)

here x and y are real numbers and j is the unit imaginary number: j = 1 . Another
representation of the complex variable can be obtained in the polar coordinates:

131

A P P E N D I X

N Y Q U I S T S

S T A B I L I T Y

C R I T E R I O N

s = r (cos + j sin ) = r e j .

Here r =

x 2 + y 2 = s is the

number s and

(B-2)
absolute value (norm, modulus) of the complex

= arctan ( y x ) is the argument of the complex number.

A complex function f (s ) = u (x, y ) + jv (x, y ) = f e j associates one (single-valued


function) or more (multiple-valued function) values of complex dependent variable f(s)
with each value of the complex independent variable s.
A single-valued function f(s) is called analytic (regular, holomorphic) at point s = a if
and only if f(s) is differentiable throughout the neighborhood of s = a.
A singular point or singularity of the function f(s) is any point where f(s) is not
analytic. The singularity can be a removable singularity if and only if f(s) is finite
throughout a neighborhood of s = a, except possibly at s = a itself. The singularity can
be a pole of order m if and only if (s-a)mf(s) (but not (s-a)m-1f(s)) is analytic at s = a.
The points s for which f(s) = 0 are called the zeros (roots) of f(s).
Function f(s) is meromorphic throughout a certain region D if and only if its only
singularities throughout D are poles.
Nyquist criterion

Let f(s) be meromorphic throughout the bounded region inside and continuous on a
closed contour C on which f (s ) 0 . Let Z be the number of zeros and P be the
number of poles of f(s) inside C, respectively, where a zero or pole of order m is
counted m times. Then
1
f ( )
d = Z P .

2j C f ( )

(B-3)

For P = 0 one obtains the principle of the argument


N=

C
,
2

(B-5)

where C is the variation of the argument of f(s) around the contour C. This
equation means that f(s) maps a moving point s describing the contour C once into a
moving point f(s) which encircles the f plane origin N = 1, 2, times if f(s) has,
respectively, 0, 1, 2, zeros inside the contour C in the s plane. The above criterion is
used for locating zeros and poles of f(s) and is known as the Nyquist criterion.
As an example, consider a function f(s) that has two zeros (s = z1 and s = z2) and one
pole (s = p1). The function can be given as
f (s ) =

(s z1 )(s z 2 )
(s p1 )

(B-6)

132

A P P E N D I X

N Y Q U I S T S

S T A B I L I T Y

C R I T E R I O N

The location of the zeros and the pole are shown in Figure B-2.
s
Im(s)

Pole
z1

p1

z2
Zeros

Contour C
Re(s)

Figure B-2: Location of zeros and poles on s plane.


Assume that point s is moving along the contour and encircles it. The argument of the
number (s-z1) will increase by 2 . It means that with a single encirclement of the
contour C by point s, the argument of function f(s) will increase by 2 for each zero
which is inside of the contour. In a similar way it is obtained that the argument of
function f(s) decreases with 2 for each pole located inside of the contour C. As a
consequence, the mapping of f(s) on plane f will encircle the plane origin the
corresponding number of times. In the considered case, f(s) will encircle the origin only
once, since its argument will increase with (2 -1) multiplied by 2 .
The Nyquist criterion is very useful in the theory of stability. To evaluate the stability
of any system, it is necessary to find the locations of zeros of the characteristic
function of the system. If any of the zeros is located on the right-hand-side of the s
plane, then the system is unstable. The contour on the s plane is shown in Figure B-3.

133

A P P E N D I X

N Y Q U I S T S

S T A B I L I T Y

C R I T E R I O N

Im(s)
CR
C+
R

Re(s)
C-

Figure B-3: Contour in the s plane that covers the whole right-hand-side half-plane
when R .
An example of the corresponding mapping of function f(s) on the f plane is shown in
Figure B-4.
Im(f)

C-

=0

CR

Re(f)

C+

Figure B-4: Mapping of function f(s) for s encircling the right half-plane as shown in
Fig. B-3.

134

Appendix

C
Appendix C - Selected
Steam-Water Data
Steam-water properties at saturation for pressure p = 7 MPa (tsat = 285.83 C)
Phase

Density

Viscosity

Enthalpy

Conductivity

Spec. heat

[kg m-3]

[Pa s]

[J kg-1]

[W m-1 K-1]

[J kg-1 K-1]

Water

739.724

9.12488 10-5

1267660

0.571881

5402.48

Steam

36.525

1.89601 10-5

2772630

0.0629146

5356.59

Steam-water properties at saturation for pressure p = 15.5 MPa (tsat = 344.79 C)


Phase

Density

Viscosity

Enthalpy

Conductivity

Spec. heat

[kg m-3]

[Pa s]

[J kg-1]

[W m-1 K-1]

[J kg-1 K-1]

Water

594.38

6.83274 10-5

1629880

0.458470

8949.98

Steam

101.93

2.31084 10-5

2596120

0.121361

14000.6

135

Appendix

D
Appendix D Boiling
Channel Stability Model
Expressions for pressure drop perturbations in a boiling channel are derived in this
Appendix. It is assumed that the boiling channel has a uniform heat flux distribution
and a uniform shape (and thus area cross section) along its length. All flow parameters
are averaged across the flow area, thus the model is one-dimensional. The model
presented in this Appendix follows the derivation presented in [D-1].
For a given system pressure, p, and the inlet velocity of the subcooled liquid,jin,0 , the
steady state mass flux in the channel, G0, is given by,
G0 = f jin,0 .

(D-1)

Here for simplicity it is assumed that the fluid density in the single phase (subcooled)
region is constant and equal to the saturation liquid velocity f . The inlet subcooling
is isub,0 and the uniform axial heat flux is q0 . The steady state length of the singlephase (subcooled) region, 0 , can be evaluated from the energy equation for the
homogenous flow model:
(m iM ) (G im ) qPH p
+
=
+ .
t
z
A
t

Here m is the two-phase mixture density, PH is the heated perimeter, A is the


channel cross-section area, and the mixture enthalpy im and iM are defined as follows,
im = i f (1 x ) + ig x

iM =

f i f (1 ) + g ig
.
m

For steady-state both terms given below are equal to zero, thus,
p
= 0,
t

and
137

A P P E N D I X

B O I L I N G

C H A N N E L

S T A B I L I T Y

( m iM )
=0 ,
t

so the expression for 0 becomes,

0 =

G0 Aisub,0
.
q0PH

(D-2)

In the two-phase region, the volumetric flux can be obtained as,


j0 ( z ) = 0 (z 0 ) + jin,0 ,

(D-3)

where
0 =

q0PH v fg
.
A i fg

(D-4)

The homogenous density can be obtained from,


G0
.
j0 ( z )

m0 ( z) =

(D-5)

For transient analysis Equations (D-1)-(D-5) generalize to


G (t ) = f jin (t ) ,

(D-6)

j (t , z ) = (t , z )(z (t ) ) + jin (t ) ,

(D-7)

q(t , z ) PH v fg
,
A
i fg

(D-8)

(t , z ) =

m (t, z ) =

G (t , z )
.
j (t , z )

(D-9)

Using the homogeneous equilibrium model, the pressure drop across single-phase and
two-phase section is given as,

(p )

1 H

G 1 G 2 4C f ,lo G 2

G 2 ( zi )
,
= pin p( ) = +
+
+ l g dz + i
t l z
Dh 2l
2 l
iL1

for the single phase flow part and,

(p )
2

LH
G G2 2 4Cf ,lo G2

G2 (zi )
2

=
p
(
)

p
=
+
+
+
g
dz
+
(
z
)
,
ex
lo,i i i
t z M lo Dh 2l m i
H
2l
L2

138

A P P E N D I X

B O I L I N G

C H A N N E L

S T A B I L I T Y

for the two-phase flow part. Here is the position of the boiling boundary, L1 is the
non-boiling length of the channel and L2 is the boiling length of the channel.
These equations can be linearized and Laplace transform to obtain,

(p1 )H

4G0 2C f

0
= s0 f + 4C f
+ i G0 jin (s) +
+ g f (s) (D-10)

2 f Dh

Dh iL1

LH

(p2 )H = (s + 0 ) m 0 ( z ) +

G0 j ( s, z ) +
Dh

4C f

4C f 2

~
0 j0 ( z ) +
j0 ( z ) + g m ( s, z ) + G0( s, z ) dz
2 DH

4C f

G0 0 +
G0 jin ,0 + g l ( s ) +
2 Dh

i G0j ( s, zi ) +

iL2

(D-11)

j02 ( z i )
m ( s, zi )
2

The next step is to relate the variables in Equation (D-10) and (D-11) to the channel
external perturbations.
The volumetric flux perturbation can be obtained from Equation (D-7) as

( s, z ) ( s ) + j ( s ) .
j ( s, z ) = (z 0 )
0
in

(D-12)

The boiling boundary perturbation, , can be expressed in terms of i( s, 0 ) by


integrating the energy equation from 0 to (t ) and then perturbing it. This gives,
i (t , ) i (t , 0 ) =

q0PH
(t ) .
G0 A

(D-13)

Taking into account that i (t , 0 ) = i f = i0 (0 ) and Laplace-transforming yields

( s ) =

G0v fg

i ( s, 0 ) = 0 i( s, 0 ) .
0i fg
isub

(D-14)

The enthalpy perturbation can be obtained from the energy equation for the
homogenous flow model, perturbing and Laplace-transforming this equation yields,

d i( s, z )
s
P q q( s, z ) jin
.
+

i ( s, z ) = H 0
dz
jin ,0
G0 A q0
jin ,0

139

(D-15)

A P P E N D I X

B O I L I N G

C H A N N E L

S T A B I L I T Y

The specific relationship between q and q depends on the model used to


quantify the heated wall-dynamics and, in general, the following expression can be
obtained,

Z1 ( s )q ( s, z ) + Z 2 ( s )q H ( s ) = TW ( s, z ) .

(D-16)

If a lumped parameter model is used to describe the heater dynamics, the following
equation is obtained,
dTW
= qH AH qPH ,
dt

H c p , H AH

(D-17)

where H is the heater material density, c p , H is the heater specific heat and AH is the
effective heater cross-section area.
Thus,

Z1 ( s ) =

PH

H c p , H AH s

and

Z 2 ( s) =

1
.
H c p,H s

In the single-phase region, TW , is given by the Newtons law of cooling, as,

TW Tb =

q
.
h1

(D-18)

where Tb is the bulk fluid temperature.


Taking into account that
a

j (t )
h1 (t ) = h1 ,0 in ,
jin ,0

where normally a = 0.8 , yields,

q = h1 , 0 TW Tb + q0a

jin
jin ,0

(D-19)

where

Tb = i c pf .

140

A P P E N D I X

B O I L I N G

C H A N N E L

S T A B I L I T Y

Here, cpf is the fluid specific heat.


Equations (D-16) and (D-19) can be combined to obtain q ( s, z ) in the single-phase
region as a function of q H ( s ) , jin ( s ) and i( s, z ) . Substituting the result into
Equation (D-15) and rearranging, we get,

( )

j ( s )
q ( s )
d i
+ ( s )i( s, z ) = ( s ) in
+ ( s) H
dz
jin ,0
q0

(D-20)

with,

( s) =

s
jin ,0

PH h1 ,0

G0 Ac pf (1 h1 ,0 Z1 ( s ) )

(D-21)

( s) =

PH q0
a
1 ,
G0 A 1 h1 ,0 Z1 ( s )

(D-22)

( s) =

PH q0 h1 ,0 Z 2 ( s )
,
G0 A 1 h1 ,0 Z1 ( s )

(D-23)

Solving the differential Equation (D-20), z = 0 to z = 0 , with the boundary


condition i( s,0) = iin ( s ) gives

i( s, 0 ) = e ( s ) iin +
0

1
j
q ( s )
1 e ( s ) 0 ( s ) in + ( s ) H
jin , 0
q0
( s)

(D-24)

Now the expression for ( s ) becomes,

( s ) = 1 ( s )jin + 2 ( s )q H + 3 ( s )iin ,

(D-25)

where

1 ( s ) =

2 ( s) =

3 ( s) =

( s)
[1 e ( s ) ],
isub jin ,0 ( s )

(D-26)

( s)
[
1 e ( s ) ] ,
isub q0 ( s )
0
isub

(D-27)

[e

( s ) 0

],

(D-28)

141

A P P E N D I X

B O I L I N G

C H A N N E L

S T A B I L I T Y

from the expression for j in Equation (D-12) we can use an


In order to eliminate
equation that describes the heated wall dynamics combined with the Newtons law of
cooling.
1

q = C ( p )(TW Tsat )m ,

(D-29)

Here C(p) is a known function of pressure and m is a constant.


Perturbation of Eq. (D-29) yields,

TW = (TW ,0 Tsat )

m
q H ( s ) = Z 3q H ( s ) .
q0

(D-30)

Substituting this result into Equation (D-16) yields

( s) =

Z 2 ( s) PH v fg
i q = Z q .
(Z3 Z1 ( s))A fg H 4 H

(D-31)

Substituting Equations (D-25) and (D-31) into (D-12) we obtain an expression


,
without

j(s, z) = [1 01(s)]jin (s) + [(z 0 )Z4 (s) 02 (s)]qH (s)


03 (s)iin (s)

(D-32)

To get an expression for m ( s, z ) we can rewrite the mass equation as,


m

+ j m + m = 0 .
t
z

(D-33)

and if one perturbs and Laplace-transforms this equation, and uses the continuity
equation, j z = , one gets,

d
(m(s, z)) + s +m(s, z) = d m0(z) j(s, z) m0(z) (s)
dz
j0 (z)
dz
j0 (z) j0 (z)

(D-34)

Substituting Equations (D-3) and (D-5) into this equation gives

(s)

(D-35)
(m(s, z)) + s 0 +1 m(s, z) = G02 j(s, z) 3
2
dz
z 0 + jin,0 0
0 (z 0 + jin,0 0 ) (z 0 + jin,0 0 )
In order to integrate this equation, a boundary condition must be established at
z = 0 . One can integrate the energy equation for the homogenous flow model from
z = 0 to z = ,

142

A P P E N D I X

B O I L I N G

C H A N N E L

S T A B I L I T Y

m
dz+ m(t, ) j(t, ) m(t, 0 ) j(t, 0 ) .
t
0

(D-36)

After perturbing and Laplace-transforming this equation one gets,

m ( s, 0 ) =

f
jin ,0

(j

in

( s ) j ( s, 0 )

(D-37)

and this can be rewritten, using Equation (D-13), as,

m( s, 0 ) =

0 f
jin ,0

( s ) .

(D-38)

Now one can integrate Equation (D-35) from 0 to z with this boundary condition.
This yields an equation on the form y ( z ) + f ( z ) y ( z ) = g ( z ) and can be solved by
using the technique with integrating factor. Using equation (D-7) and (D-12) and
( s, z ) =
( s ) gives the result,
assuming that
s

1
jin,0 0

G

m (s, z) = s
0 0 2 0 (s) +
j0 ( z)
s 0 j0 ( z)

1
j

1 in,o 0 0 G0 j (s)
j ( z)
s j 2 ( z) in
0
0 0

(D-39)

1
1 jin,o 0 1 G0 jin,o

j ( z)
s j 2 ( z ) ( s)
0
0
0

Assuming that the only local losses are those at the inlet and the exit of the channel
and substituting Equation (D-25) into (D-10) results in the following expression for
the single-phase pressure drop perturbation, depending on perturbations in the inlet
velocity, jin , inlet enthalpy, iin and in the heater internal heat generation q H .

(p1 )H = 1, H ( s )jin + 2, H ( s )q H + 3, H ( s )iin ,

(D-40)

where
s

jin ,0 4C f

g
,

s
1, H = G0 0 + 4C f 0 + in +
+
(
)
1

jin , 0

D
2
D
j
h
h
in
,
0

143

(D-41)

A P P E N D I X

B O I L I N G

C H A N N E L

jin , 0 4C f
g
i , H = G0
+
i ( s) ,
jin, 0
2 Dh

S T A B I L I T Y

i = 2,3 .

(D-42)

With the same assumptions as for (D-40) and substituting equation (D-25), (D-32) and
(D-39) into (D-11) we get expressions for the pressure drop perturbations in the twophase region as,

(p2 )H = 1, H ( s )jin + 2, H ( s )q H + 3, H ( s )iin .

(D-43)

where,
1,H = G0 (F1 ( s ) F2 ( s ) 1 ( s ) ) ,

(D-44)

2,H = G0 (F2 ( s ) 2 ( s ) F3 ( s ) ) ,

(D-45)

3,H = G0 F2 ( s ) 3 ( s ) ,

(D-46)

s 2 ex 4C f (LH 0 ) 2s 0
0
F1 ( s) =
+
+
e (0 s ) ex 1 +
s 0
2 Dh
s 0 (s 0 )2
2

jin , 0 4C f
2 Dh
g
jin ,0

0
e (20 s ) ex 1 +
(s 0 )(s 2 0 )

1
( s )
0

1 e ex + 0 e ex 1 +
(s 0 )
s

0
+ ex 1 +
1 e (0 s ) ex
2(s 0 )

) .

(D-47)

4C f (LH 0 ) 0 (2s 0 )
s 2
0 s
F2 ( s) = 0 ex +
+
e (0 s ) ex 1 +
2
s 0
2 Dh
s 0
(s 0 )
2

jin,0 4C f
2Dh

jin , 0 4C f
2 Dh

0 s
0
g

e(20 s ) ex 1 +
es ex e 0 ex + 0 +
(s 0 )(s 20 )
jin,0 (s 0 )

g
jin , 0

1
0
ex 0 e (0 s ) ex 1
1 e (0 s ) ex
2(s 0 )
2

(D-48)

s + 2 0
0 jin ,0
s2
(LH 0 )
F3 ( s ) =
jin , 0 ex
e (0 s ) ex 1 +
2
(s 0 ) 0
(s 0 )
0
2
4C f
jin ,0
jin , 0 e (20 s ) ex 1
2

+
( LH 0 )
( L H 0 )
(s 0 )(s 2 0 )
2 Dh
s 0

144

A P P E N D I X

g
s 0

B O I L I N G

1
1

1 e (0 ex ) 1 e ( s ex )
s
0

] [

jin , 0

+ (LH 0 )
1 e ( 0 s ) ex
2 (s 0 )

ex =

C H A N N E L

S T A B I L I T Y

] +

ex

Z 4 ( s ) .

(D-49)

1 j0 ( LH )

.
0 jin ,0

(D-50)

Considering local losses at other locations for the single-phase region yields,
s

4C f jin , 0

g
.
1, H = G0 0 + 4C f 0 + i +
+

(
s
)
1

jin , 0

D
2
D
j
iL1
h
h
in , 0

(D-51)

For the two-phase region, only the ex -terms in F1, F2 and F3 are changed. For F1 we
get:

s 2 ex 4C f (LH 0 ) 2s 0
0
F1 ( s) =
+
+
e (0 s ) ex 1 +
s 0
2 Dh
s 0 (s 0 )2
2

jin , 0 4C f
2 Dh

g
jin ,0

0
e (20 s ) ex 1 +
(s 0 )(s 2 0 )

1
( s )
0

1 e ex + 0 e ex 1 +
(s 0 )
s

0
+ i 1 +
1 e (0 s ) i
i
2(s 0 )

) .

(D-52)

F2 transforms into,

0 s 2 ex 4C f (LH 0 ) 0 (2s 0 )
0 s
+
+
e (0 s ) ex 1
2
s 0
2 Dh
s 0
(s 0 )
2

F2 ( s) =

jin,0 4C f
2Dh

jin , 0 4C f
2 Dh

0 s
0
g

e(20 s ) ex 1 +
es ex e 0 ex + 0 +
(s 0 )(s 20 )
jin,0 (s 0 )

g
jin , 0

1
0
i 0 e (0 s ) i 1
1 e (0 s ) i
(
)
2
2
s

i
0

145

A P P E N D I X

B O I L I N G

C H A N N E L

S T A B I L I T Y

1
0
i 0 e (0 s ) i 1
1 e (0 s ) i
2(s 0 )
i
2

) .

(D-53)

and F3,
s + 2 0
0 jin ,0
s2
(LH 0 )
F3 ( s ) =
jin , 0 ex
e (0 s ) ex 1 +
2
(s 0 ) 0
(s 0 )
0
2
4C f
jin , 0
jin ,0 e (2 0 s ) ex 1
2

+
( L H 0 )
( L H 0 )
(s 0 )(s 2 0 )
2 Dh
s 0

g
s 0

1
1

1 e (0 ex ) 1 e ( s ex )
s
0

] [

jin ,0

+ (LH 0 )
1 e (0 s ) i
2(s 0 )

] +

) Z
i

(s)

(D-54)

where

i =

1 j0 ( zi )

.
0 jin ,0

(D-55)

Calculation of the individual heat transfer coefficients involves the following


dimensionless numbers:
The Nusselt number, Nu

Nu =

The Prandtl number, Pr

Pr =

The Reynolds number, Re

Re =

h Dh

cp

v Dh

(D-56)

(D-57)

(D-58)

The following equation is used:


Nu = (Re, Pr ) = 0.023 Re0.8 Pr 0.33 .

(D-59)

This yields:
h=

0.023 Re0.8 Pr 0.33


.
Dh

(D-60)

The friction-factor can be calculated with the Blasius correlation:

146

A P P E N D I X

Cf =

B O I L I N G

0.0791
.
Re 0.25

C H A N N E L

S T A B I L I T Y

(D-61)

R E F E R E N C E S

[D-1] Lahey R.T. Jr. Ed. Boiling Heat Transfer, Modern Developments and Advances, ISBN 0-444-89499-3,
1992

147

149

INDEX
Adjoint neutron flux ............................... 38

Non-wandering sets ............................... 15

Amplification factor ......................... 20, 26

Normalized point dynamics equations .. 76

Amplitude function ................................ 35

Normalized point kinetics equations .... 44

Asymptotic stability............................... 15

Nyquist plot............................................ 27

Attractor ................................................ 15

One group approximation ...................... 40

Basin of attraction ................................. 15

Orbit

Bifurcation ............................................. 15

Percent milli rho - pcm .......................... 70

................................................. 15

Bode characteristics ............................. 27

Phase shift ........................................20, 26

Channel thermal-hydraulic instability ... 12

Phase space........................................... 14

Chaotic orbit .......................................... 15

Phase-change number ......................... 118

Characteristic polynomial ..................... 17

Point dynamics equations ..................... 73

Control system instability ..................... 12

Point kinetics equations........................ 37

Coupled neutronic-thermal-hydraulic

Poles of the transfer function................ 25

instability ................................. 12

Precursors of delayed neutrons ............ 33

Daughter nuclei ..................................... 33

Prompt criticality ................................... 62

Decay ratio..................................... 23, 118

Prompt neutrons .................................... 33

Density wave oscillations.................... 111

Prompt reactivity coefficient ................ 72

Doppler effect ........................................ 69

Quasiperiodic orbit ................................ 15

Dynamic instabilities ........................... 109

Reactivity............................................... 37

Dynamical systems................................ 14

Reactivity instability ............................. 12

First method of Lyapunov ...................... 16

Reactor period ....................................... 48

First-order system ................................. 19

Routh and Hurwitz theorem................... 17

Fixed point ............................................. 15

Second-order system ............................. 21

Frequency characteristics .................... 27

Shape function ....................................... 35

Frequency response .............................. 26

Six-group point kinetics model.............. 40

Fuel Doppler reactivity coefficient ........ 72

Stability map ........................................ 118

Fuel temperature coefficient................. 72

Stable reactor period............................. 48

Hopf bifurcation ..................................... 18

State of a dynamical system ................. 14

Jacobian matrix ..................................... 16

Static instabilities ............................... 109

Limit cycle ............................................. 15

Stationary bifurcation............................ 18

Linear stability analysis ........................ 16

Subcooling number .............................. 118

Lyapunov stability ................................. 15

Trajectory .............................................. 15

Marginal stability ................................... 15

Transfer function ................................... 24

Neutron generation time ....................... 41

Transient reactor periods ...................... 48

Neutron lifetime ..................................... 41

Zero-power reactor transfer function ... 55

Nonautonomuous systems .................... 15

151

You might also like