0% found this document useful (0 votes)
83 views64 pages

4 - Reactor Statics

Reactor Statics .1

Uploaded by

elsayed
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
83 views64 pages

4 - Reactor Statics

Reactor Statics .1

Uploaded by

elsayed
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 64

1

CHAPTER 4
Reactor Statics
Prepared by
Dr. Benjamin Rouben, 12 & 1 Consulting, Adjunct Professor, McMaster University & University
of Ontario Institute of Technology (UOIT)
and
Dr. Eleodor Nichita, Associate Professor, UOIT
Summary:
This chapter is devoted to the calculation of the neutron flux in a nuclear reactor under special
steady-state conditions in which all parameters, including neutron flux, are constant in time.
The main calculation method explored in this chapter is the neutron-diffusion equation. Analyti-
cal solutions are derived for simple neutron-diffusion problems in one neutron energy group in
systems of simple geometry. Two-group diffusion theory and the approximate representation of
the diffusion equation using finite differences applied to a discrete spatial mesh are introduced.
The rudimentary reactor-physics design of CANDU reactors is presented. The two-step approach
to neutronics calculations is presented: multi-group lattice transport calculations, followed by
full-core, few-group diffusion calculations. Finally, the chapter covers fuel-property evolution
with fuel burnup and specific features of CANDU neutronics resulting from on-line refuelling.

Table of Contents
1 Introduction ............................................................................................................................ 3
1.1 Overview ............................................................................................................................. 3
1.2 Learning Outcomes ............................................................................................................. 3
2 Neutron Diffusion Theory ....................................................................................................... 4
2.1 Time-Independent Neutron Transport ............................................................................... 4
2.2 Fick’s Law and Time-Independent Neutron Diffusion......................................................... 7
2.3 Diffusion Boundary Condition with Vacuum at a Plane Boundary ..................................... 9
2.4 Energy Discretization: The Multi-Group Diffusion Equation............................................. 12
3 One-Group Diffusion in a Uniform Non-Multiplying Medium.............................................. 12
3.1 Plane Source...................................................................................................................... 13
3.2 Point Source ...................................................................................................................... 14
3.3 Flux Curvature in Source-Sink Problems and Neutron In-Leakage ................................... 16
4 One-Group Diffusion in a Uniform Multiplying Medium with No External Source .............. 16
4.1 Uniform Infinite Reactor in One Energy Group................................................................. 17
4.2 Uniform Finite Reactors in One Energy Group.................................................................. 18
4.3 Uniform Finite Reactors in Various Geometries ............................................................... 21
4.4 Flux Curvature and Neutron Out-Leakage ........................................................................ 30
5 Reactors in Two Neutron-Energy Groups with No External Source...................................... 31
5.1 The Neutron Cycle and the Four-Factor Formula ............................................................. 31
5.2 The Two-Energy-Group Model.......................................................................................... 34
5.3 Neutron Diffusion Equation in Two Energy Groups .......................................................... 34
5.4 Uniform Infinite Medium in Two Energy Groups.............................................................. 34
5.5 Uniform Finite Reactors in Two Energy Groups ................................................................ 36

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
2 The Essential CANDU

6 Solution for Neutron Flux in a Non-Uniform Reactor ........................................................... 38


6.1 Finite-Difference Form of the Neutron-Diffusion Equation.............................................. 38
6.2 Iterative Solution of the Neutron-Diffusion Equation....................................................... 42
7 Energy Dependence of Neutron Flux.................................................................................... 43
7.1 The Fission-Neutron Energy Range................................................................................... 43
7.2 The Thermal Energy Range ............................................................................................... 44
7.3 The Slowing-Down Energy Range ..................................................................................... 44
7.4 Neutron Flux over the Full Energy Range ......................................................................... 46
8 Basic Design of Standard CANDU Reactors........................................................................... 47
9 CANDU Reactor Physics Computational Scheme.................................................................. 50
9.1 Lattice Calculation............................................................................................................. 51
9.2 Reactivity-Device Calculation............................................................................................ 52
9.3 Full-Core Calculation ......................................................................................................... 52
10 Evolution of Lattice Properties.............................................................................................. 53
11 Summary of Relationship to Other Chapters........................................................................ 56
12 Problems ............................................................................................................................... 58
13 Appendix A: Reactor Statics .................................................................................................. 62
14 Acknowledgements............................................................................................................... 64

List of Figures
Figure 1 Negative flux curvature in homogeneous reactor .......................................................... 19
Figure 2 Infinite-slab reactor......................................................................................................... 22
Figure 3 Flux with physical and unphysical values of buckling ..................................................... 23
Figure 4 Infinite-cylinder reactor .................................................................................................. 24
Figure 5 Ordinary Bessel functions of first and second kind ........................................................ 25
Figure 6 Rectangular-parallelepiped reactor ................................................................................ 26
Figure 7 Finite-cylinder reactor..................................................................................................... 28
Figure 8 Energies of fission neutrons............................................................................................ 31
Figure 9 Sketch of cross section versus neutron energy............................................................... 32
Figure 10 Neutron cycle in thermal reactor.................................................................................. 33
Figure 11 Face view of very simple reactor model ....................................................................... 38
Figure 12 Flux in a cell and in immediate neighbours .................................................................. 39
Figure 13 Linear treatment of flux in central cell and one neighbour.......................................... 40
Figure 14 Flux in cell at reactor boundary .................................................................................... 42
Figure 15 Energy distribution of fission neutrons......................................................................... 44
Figure 16 Variation of flux with energy if absorption is smooth .................................................. 46
Figure 17 Variation of flux with energy in the presence of resonances ....................................... 46
Figure 18 Sketch of flux variation over full energy range ............................................................. 47
Figure 19 CANDU basic lattice cell ................................................................................................ 51
Figure 20 Supercell for calculating device incremental cross sections......................................... 52
Figure 21 Simple model with superimposed reactivity device..................................................... 53
Figure 22 Evolution of fuel isotopic densities............................................................................... 54
Figure 23 Infinite-lattice reactivity vs. irradiation......................................................................... 55
Figure 24 Finite-reactor reactivity vs. irradiation.......................................................................... 55
Figure 25 Averaging reactivity with daily refuelling...................................................................... 56

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 3

1 Introduction
1.1 Overview
This chapter is devoted to the calculation of the neutron flux1 in a nuclear reactor under special
steady-state conditions in which all parameters, including neutron flux, are constant in time.
The position-, energy-, and angle-dependent steady-state neutron-transport equation is derived
by writing the detailed neutron-balance equation. This is done in Chapter 3 of the book.
However, it is also done here for completeness. The steady-state diffusion equation is subse-
quently derived using a linear approximation of the angular dependence of the neutron flux.
Multi-group neutron-energy discretization is also introduced.
Analytical solutions are derived for simple neutron diffusion problems. First, the one-group
diffusion equation is solved for a uniform non-multiplying medium in simple geometries.
Subsequently, the one-group diffusion equation is solved for a uniform multiplying medium (a
homogeneous “nuclear reactor”) in simple geometries. The importance of neutron leakage and
the concepts of criticality and the neutron cycle are introduced.
Of course, real reactors are almost never homogeneous, and rarely can neutron energies be
accurately represented by a single energy group. However, two energy groups are often suffi-
cient to represent neutron diffusion in a thermal reactor because most of the fissions are
induced by thermal neutrons. This chapter therefore proceeds to introduce two-group diffusion
theory and the approximate representation of the diffusion equation using finite differences
applied to a discrete spatial mesh. The latter enables the treatment of non-uniform reactor
cores.
Following this treatment of the basic theory of neutron transport and diffusion, the rudimen-
tary reactor-physics design of CANDU reactors is presented, and the associated neutron energy
spectrum is discussed. Subsequently, more general core modelling concepts are presented,
including the two-step approach to neutronics calculations: multi-group lattice transport
calculations, followed by full-core, few-group diffusion calculations. The final section in the
chapter concentrates on fuel-property evolution with fuel burnup and specific features of
CANDU neutronics resulting from on-line refuelling.

1.2 Learning Outcomes


The goal of this chapter is for the student to understand:
 The neutron transport resulting from the detailed neutron balance;
 The neutron-diffusion equation resulting from the linear approximation of the angu-
lar dependence of the neutron flux;
 Analytical solutions of the diffusion equation in simple geometries for both multiply-
ing and non-multiplying media;

1 In this document, the term “neutron flux” is used to denote the quantity  = n, where n is neutron density and
 is neutron speed. The units of flux are 1cm-2s-1 or 1m-2s-1. Some newer texts use the term flux density or
fluence rate. All three terms refer to the same physical quantity.

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
4 The Essential CANDU

 The concepts of neutron cycle, criticality, and buckling;


 Energy discretization of the diffusion equation using neutron energy groups and spa-
tial discretization using finite differences;
 The two steps of neutronics calculations: lattice-cell transport calculations and full-
core diffusion calculations; and
 Basic design elements of CANDU reactors and their implications for the physical at-
tributes of such reactors.

2 Neutron Diffusion Theory


2.1 Time-Independent Neutron Transport
The time-independent neutron-transport equation is an equation in six variables: three vari-

ables r for the position of the neutron in the reactor, two variables ̂ for the direction of
neutron motion, and one variable E for neutron energy (we could also use neutron speed v
instead of E, because there is a one-to-one relationship between v and E).
Because we are dealing with time independence (i.e., reactor statics), the neutron-transport
equation expresses the fact that the number of neutrons at any position, moving in any direc-
tion, and with any energy, is unchanging over time. In other words, if we consider a differential
  ˆ
volume in the six-dimensional variable space, i.e., dr d̂dE about r ,  
, E , the rate of neutrons
“entering” the differential volume must be equal to the rate of neutrons “exiting” the differen-
tial volume.
To write the equation, we must therefore consider all phenomena by which neutrons enter or
exit the differential volume or are created or destroyed within the volume. The phenomena
which we must consider are:
 Neutron production by nuclear fission;
 Neutron scattering, which can change the neutron’s energy and/or its direction of
motion;
 Neutron absorption;
 Neutron spatial leakage into or out of the differential volume.
[Note: Another process for neutron production is the (n, 2n) nuclear reaction. However, this
reaction produces many fewer neutrons than fission and is neglected here.]
The rates at which neutrons enter or exit the differential volume by virtue of these various
phenomena, expressed per unit differential volume per second, are given one by one in the
following equations.
The rate of birth of neutrons by fission is given by:
 E   ˆ
 
4 E ' ˆ '

 f  r , E '   r ,  
ˆ ' dE ' ,
', E ' d  (1)

where
 
   f r , E ' is the neutron-yield cross section at position r for fissions induced by
neutrons of energy E’;

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 5

 ˆ
 
 r, 
' , E ' is the angular flux of neutrons at position, of energy E’, and moving in di-

rection ' ; and
 (E) is the fraction of all fission neutrons born with energy E. Note: In this chapter,
for simplicity, a single fission spectrum (E) is used for fissions induced in all nu-
clides, whereas in reality the fission spectrum should be taken as different for differ-
ent nuclides. This fully correct treatment is what is done in Chapter 3.
Note: In the above equation, delayed neutrons are not referred to separately. The neutron
production rate includes both prompt and delayed neutrons. This is acceptable because it can
be shown that in (true) steady state, accounting separately for delayed-neutron production
reduces in any case to the above expression.
The rate of production of neutrons from an “external” source (not related to fissions in the
reactor fuel) is given by:

S r , E , (2)
 
where S r , E  is the external source strength for neutrons of energy E at position r .
Note: We will assume that the source is isotropic.
The rate of neutron gain from neutrons entering the differential volume by scattering from
other neutron directions of motion or other neutron energies is given by:
 

   s r , E '  E , ˆ '  ˆ  r , ˆ ', E ' d ˆ ' dE ',

  (3) 
E ' '


 ˆ ' 
where  s r , E '  E ,  ˆ  is the cross section for scattering neutrons from energy E’ to
energy E and from direction ' to direction ̂ .
The rate of neutron loss from absorption and from neutrons exiting the differential volume by
scattering is given by:
  ˆ

t  r , E  r ,  ,E ,  (4)
 
where  t r , E  , the total cross section at position r for neutrons of energy E,
 

  a  r , E     s r , E  E ', 
ˆ 
ˆ ' d
ˆ ' dE ', 

where  a r , E  is the absorption cross section.
The net rate of neutron spatial leakage (i.e., diffusion) out of the differential volume is given by:
  
 ˆ ,E ,
  J r,  (5)

ˆ , E is the angular current at position r of neutrons moving in direction ̂ with
where J r ,  
energy E.

ˆ , E   r , 
Note: Because J r ,    
ˆ ,E  
ˆ , the neutron leakage [Eq. (5)] can also be written as:

ˆ    r , 
  ˆ ,E .  (6)

Taking into account all the above rates, the neutron balance can then be expressed in the time-
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
6 The Essential CANDU

independent neutron-transport equation as follows:


   E   ˆ ˆ ' dE '  r, E '  E,  ˆ  r, 
1
4
S r, E  
4 E ' ˆ ' 
 f  r , E '   r ,  
', E ' d    s  ˆ '  
ˆ ', E ' d
ˆ ' dE ' 
ˆ
E ' ' (7)
  ˆ   
 
 t  r , E   r , , E   J r ,   
ˆ , E  0.

Note again that for simplicity a single fission spectrum (E) has been used here. See Chapter 3
for the fully correct treatment. The second term of the equation can be simplified because the
fission cross section does not depend on ̂' , so that the integral over ̂' in the second term
reduces to the angle-integrated flux, and Eq. (7) becomes:
  E    ˆ  r, 
1
4
S r, E 
4 E '  ˆ

 f  r , E '    r , E '  dE '   s r , E '  E, 
ˆ '   
ˆ ', E ' d  
ˆ ' dE '
E ' '
  ˆ   
  
 t  r , E   r , , E   J r , , E  0, ˆ  (7)’
  ˆ

where  r , E '    r ,  
ˆ ' is the total (angle-integrated) flux of neutrons with energy E
' , E ' d
ˆ'
 
at position r .
The neutron-transport equation (7) or (7)’ is an exact statement of the general steady-state
neutron-balance problem. However, it can immediately be seen that it is very complex: in
addition to its dependence on six independent variables, it is an integrodifferential equation.
Because of its complexity, this equation cannot be solved analytically except for problems in the
very simplest geometries, and real problems require numerical solution by computer.
Note that when there is no external source S, the equation appears to be a linear homogeneous
equation, which does not generally have a solution (except a trivial zero solution for the flux) for
arbitrary values of the nuclear properties. In this case, a solution can be found by modifying the
nuclear properties. Mathematically, this can be done by modifying the yield cross section f by
dividing it by a quantity, keff, which can be selected to ensure a non-trivial solution. This quan-
tity keff is called the multiplication constant.
Two general categories of codes exist to solve the neutron-transport equation: deterministic
codes (which solve the equation directly by numerical means) and Monte Carlo codes (where
stochastic methods are used to model a very large number—typically millions or even hundreds
of millions—of neutron births and their travel and event histories, from which the multiplication
constant and flux and power distributions can be evaluated using appropriate statistics of these
histories). Although the application of either type of transport computer code to full-core
reactor models requires very significant computer resources and execution time to achieve a
high degree of accuracy, both methods, especially Monte Carlo codes, have seen much greater
application in whole-reactor analysis in the last decade or so. While core-wide pin-power
reconstruction and time-dependent kinetics calculations are still beyond reach, static eigenvalue
calculations and global flux and power distributions can now be carried out routinely using
Monte Carlo codes. However, detailed discussion of either type of transport code will not be
covered in the present work.
The traditional way of attacking neutronics problems in reactors has been to solve the transport
equation numerically in relatively small regions (such as a basic lattice cell or a small collection
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 7

of cells) to compute region-averaged properties, and then to use these to solve the full-core
reactor problem with a simplified version of the transport equation, the neutron-diffusion
equation. This computational scheme is discussed in greater detail in Section 9.

2.2 Fick’s Law and Time-Independent Neutron Diffusion


To derive a simplified version of the neutron-transport equation, we note that, because the
fission and total cross sections do not depend on the neutron direction of motion ̂ (i.e.,
“nuclei do not care from which direction neutrons interact with them”), and also because the
most important quantity in reactor physics is the fission rate (which determines power produc-
tion), it may be very effective to try to obtain an equation which is independent of ̂ and which

involves only the angle-integrated flux  r, E  [also called the integral flux].
To achieve this, we can attempt to remove the neutron direction of motion by simply integrating
Eq. (7) over ̂ . Let us see what this gives, term by term. The first two terms in (7) do not
contain ̂ , and because the integral of ̂ over a sphere is 4, their integrals, which we can call
T1 and T2, can be written directly as:

T1  S  r , E  (8)
 
T2    E    f  r , E '    r , E '  dE '. (9)
E'

Similarly, integrating the fourth term gives a result that depends on the integral flux only:
  ˆ ˆ    r, E    r, E  .
T4   t  r , E   r , 
ˆ

, E d t  (10)

Integrating the third term gives a more complex relation:


 ˆ  r, 
 ˆ '
T3      s r , E '  E , 
ˆ E' 
ˆ'
ˆ ', E ' d  
ˆ ' dE ' d 
ˆ.  (11)

We can, however, simplify T3 by noting that, because no absolute direction in space is “special”,
the scattering cross section s cannot depend on the absolute directions ̂ and ̂ ' , but only on
the scattering angle between the directions of the incoming and scattered neutrons, or even
more specifically on the cosine  of that angle, i.e., on   
ˆ '
ˆ . Using this fact, we can then
simplify the integral in Eq. (11) by integrating over ̂ first (actually over , because s does not
depend on the “azimuthal” angle of scattering, and integrating over that azimuthal angle simply
gives 2). The result is then a product of two integrals:
 ˆ ' 2  r, E '  E ,   d dE '
 ˆ ' d
T3     r , E '  E , 
ˆ'
 s 
E'  
  (12)
    r , E '   s  r , E '  E  dE ',
E'

where we have defined


 
 s  r , E '  E   2   s  r , E '  E ,   d  . (13)

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
8 The Essential CANDU


Equation (12) is a useful result because it depends on the integral flux  r , E' only.
For the fifth term, we get:
  
 ˆ   
ˆ d
T5     J r , E ,   ˆ   r , E ,  ˆ 
ˆ d 
 (14)
ˆ
  ˆ 
  
   J r, E ,
where
   

J r , E    J r , E, 
ˆ d
ˆ  (15)

is the current of neutrons of energy E at position r .
Unfortunately, this result is not at all of the form we would like, i.e., it is not at all expressible in
terms of the integral flux, because it is clear that integrating a function of the vector quantity

 
J r , E , ̂ over ̂ will obviously give in general a result totally unrelated to the integral flux
 r, E  !
In summary, integrating Eq. (7) over ̂ gives:
    
S  r , E     E    f  r , E '    r , E '  dE '   s  r , E '  E    r , E '  dE '
E' E'

    
 t  r , E    r , E     J  r , E   0 , (16)

but the last term on the left-hand side (the leakage term) has thwarted our efforts to achieve an
equation in the integral flux only.
However, one approximation which is often used in diffusion problems (diffusion of one mate-
rial through another) can help us here. This approximation, called Fick’s Law, applies in low-
absorption media if the angular flux varies at most linearly with angle and if the neutron source
is isotropic [already assumed in writing Eq. (7)]. Under these conditions it states that the
  
current J r , E  is in the direction in which the integral flux  r, E  decreases most rapidly, i.e.,

it is proportional to the negative gradient of  r, E  :
   
J  r , E    D  r , E    r , E  , (17)

where the quantity Dr , E  is called the diffusion coefficient and can be written as 1/(3tr),
with tr being the neutron transport cross section.
The reader is referred to Appendix A (Reactor Statics) of this chapter for the derivation of Fick’s
Law. Here, we will continue to derive the final form of the neutron-diffusion equation.
Equation (17), when substituted into Eq. (16), finally gives an equation in the integral flux only,
the time-independent neutron-diffusion equation:
    
S  r , E     E    f  r , E '    r , E '  dE '   s  r , E '  E    r , E '  dE ' 
. (18)

E' E'
   
 t  r , E    r , E     D  r , E    r , E   0

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 9

The time-independent neutron-diffusion equation is much simpler than the neutron-transport


equation and can be used to solve for the flux in specially prepared full-core reactor models.
Note that the out-leakage term (the last term on the left-hand side) has a positive sign, even
though this term nominally represents a loss of neutrons; the + sign arises as a result of the –
sign in relationship (17) between the current vector and the flux gradient.
Before the diffusion equation can be used, it is important to understand that Fick’s Law is only
an approximation; it is valid only in low-neutron-absorption media, when the integral flux does
not vary too quickly and when angular flux varies weakly with angle (at most linearly). There-
fore, Fick’s Law is not an especially good approximation in the vicinity of strong absorbers,
where the spatial flux variation is very large, or, for the same reason, near interfaces between
regions with large variations in nuclear properties or near external surfaces.

2.3 Diffusion Boundary Condition with Vacuum at a Plane Boundary


To solve the diffusion equation, which is a second-order partial differential equation, through-
out the reactor volume, we need to define boundary conditions at the surface of the reactor.
The vacuum boundary conditions in transport theory are quite clear: the angular current (or
angular flux) at the boundary must be zero for any direction pointing to the inside of the reactor
(assuming that the reactor has no re-entrant surface):
 ˆ

 r, ,E  0  (19)
 ˆ  eˆ  0, where ê is the unit outgoing normal
for r at the boundary and any ̂ such that   

at r .
This condition cannot be applied in diffusion theory, which depends on the integral flux only,
because we have lost all the directional information of the angular flux. In diffusion theory, we
need boundary conditions, at most, on the integral flux and its gradient. The conditions can be
generalized to demand that the total rate of incoming neutrons be zero. We will now proceed
to derive the boundary conditions used in diffusion theory.
We first derive a general relationship between the angular flux and the current using the
approximation that the angular flux is linear in angle. Linearity in angle means that the angular
flux can be written as a linear function of the x-, y-, and z-components of the angle ̂ :

 
ˆ ab  b  b ,
  x x y y z z
(20)

where a, bx, by, bz are constants which can be determined in terms of the integral flux and
current.
Let us first consider the integral flux     ˆ dˆ .

From Eq. (20),

ˆ  b  d
  a d x x
ˆ  b  d
y y
ˆ  b  d
z z
ˆ. (21)
   

The first integral in Eq. (21) has value 4, whereas the others have value 0 (it is clear that in
integrating a single component of the angle over all solid angles, the + and – components cancel
out). This means that  = 4a, which implies that

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
10 The Essential CANDU


a . (22)
4
Now let us consider the current:

J     ˆ  ˆ d ˆ .

(23)

From Eq. (20),



ˆ d
J  a  ˆ b  
x x
ˆ d
ˆ b  
y y
ˆ d
ˆ b  
z z
ˆ d
ˆ. (24)
   

It is clear that the first integral in Eq. (24) is zero, for the same reason that the + and – compo-
nents cancel out. Consider the other integrals in Eq. (24), for instance,   x  ˆ d
ˆ . This is a

vector integral, and only the x-component will survive because   x  y or  z d 
ˆ  0 , for the

same reason, cancellation.


ˆ , which can easily be shown to be equal to
On the other hand, the x-component gives 

 2x d 

ˆ  2 1  2 d  2  3 1  4 ].
4/3 [by symmetry, it must be equal to 
 2z d 1 3 1 3
Therefore, we can conclude that
 4
J
3

bxiˆ  by ˆj  bz kˆ ,  (25)

i.e.,
3
bx  Jx (26)
4
(and similarly for y and z).
Now, substituting Eqs. (22) and (26) into Eq. (20),

ˆ    3 J   J   J      3 J 
 
  ˆ. (27)
4 4 4 4
x x y y z z

Let us now apply Eq. (27) at a plane boundary with vacuum. Assume that the boundary plane is
perpendicular to the z-axis and the polar axis is the outward normal to the boundary, i.e., the
unit vector k̂ in the positive z-direction. The total rate of incoming neutrons is:

  
J   
ˆ 
ˆ  kd ˆ, (28)

where the integral over the “half-space” - means integrating over the entire azimuthal angle 
(i.e., over 2), but only over half the polar angle , i.e., over  from /2 to  (or over   cos 
from -1 to 0).
Using Eq. (27), the total rate of incoming neutrons is:
  3 
   
ˆ 
  ˆ 
ˆ  kd ˆ   
 ˆ ˆ
  4  4 J ˆ  ˆ ˆ  kd . (29)

ˆ  k̂     , then, in the integral of the second term, only the part involving z will
Because  z

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 11

survive (for the same reason stated before), and:



 3   J  kˆ
Total rate of incoming neutrons =
0
* 2 *  d  ˆ 0
* 2 * J  k *   d   
2
.
4 1 4 1 4 2
Applying Fick’s Law and setting the total rate of incoming neutrons to 0 gives:
 D ˆ D d 
    k  0, i.e.,   0 (30)
4 2 2 dz 4
1 d 1
  ,
 dz boundary 2D
which can be written as
1
 ,
d extr
which leads to
2
d extr  2 D  , (31)
3 tr
2
d extr  2 D  , (32)
3 tr
where tr is the transport cross section.
Actually, a more advanced analysis gives a slightly higher, more correct value for dextr:
2.1312 0.7104
d extr  2.1312 D   . (33)
3 tr  tr
Equations (30)-(33) result in a zero net incoming current at the physical boundary.
The geometric interpretation of Eq. (30) is that the relative neutron flux near the plane bound-
ary has a slope of -1/dextr, i.e., the flux would extrapolate linearly to 0 at a distance dextr beyond
the boundary. This is often stated as “the flux goes to 0 at an extrapolation distance dextr
beyond the boundary”. Such an interpretation is not literally correct: the flux cannot go to zero
in a vacuum, because there are no absorbers to remove the neutrons; the flux only appears to
be heading to the zero value at the extrapolation point. In fact, instead of using the “slope”
version (30), the boundary condition is often applied by “extending” the reactor model to the
extrapolation point and demanding a zero flux at that point.
Now, typical values for the diffusion coefficients D in CANDU reactors are  1 cm, and therefore
the extrapolation distance dextr is of the order of 2–3 cm. For systems of large dimension, e.g.,
large power reactors which are several metres in size, the extrapolation distance is sometimes
neglected if a slightly approximate answer is sufficient.
[Note: Equation (33) applies in principle to plane boundaries only. Slightly different formulas for
the extrapolation distance would apply to curved boundaries; however, the difference is small
unless the radius of curvature of the boundary is of the same order of magnitude as dextr.]

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
12 The Essential CANDU

2.4 Energy Discretization: The Multi-Group Diffusion Equation


Neutron energy E is of course a continuous variable. However, the diffusion equation (18) with
continuous E is not usually solved as is. Instead, the equation is rewritten in “multi-group” form
by subdividing the energy range into a number G of intervals, called energy “groups” and
labelled with g from 1 to G (g =1 corresponding to the interval with the highest energies and g =

G to the interval with the lowest energies). In each group g, the continuous flux  r, E  is

replaced by an average group flux,  g r  .

The nuclear cross sections in each group are assumed to have been appropriately averaged over
the corresponding energy intervals. In multi-group notation, the cross sections and variables
are written as follows:
 
S  r , E   Sg  r 
  E   g
 
 f  r , E     f , g  r 
 
 s  r , E '  E    s , g ' g  r 
 
 t  r , E   t ,g  r 
 
D  r , E   Dg  r  .

The multi-group diffusion equation then takes the form:


 G
 
Sg  r    g  f ,g '  r  g '  r  
g ' 1

G
      

g ' 1
s , g ' g  r  g '  r   t ,g  r  g  r     Dg  r  g  r   0, g  1,..., G.

3 One-Group Diffusion in a Uniform Non-Multiplying Medium


We will start the analysis of systems with simple systems and increase complexity gradually.
Assume that all neutrons are lumped into a single energy group, i.e., G = 1. There is then no
need for the subscript g. With a single energy group, the notion of scattering across groups has
no meaning, and the “incoming scattering” term in s cancels the “outgoing scattering” term
inherent in the total cross section t (another way of saying this is that the “out-scattering” and
“in-scattering” terms cancel one another if there is only one group). This means that we can
drop the second term in the diffusion equation and write a instead of t in the third term.
In this section, we will deal with neutron diffusion in non-multiplying media, i.e., in media
where the fission cross section is zero and the neutron flux is driven by an external neutron
source. This type of problem is sometimes called a “source-sink” problem. We will assume that
the medium is uniform outside the source, and also that it is infinite in size. Uniform properties
also mean that the diffusion coefficient can be taken outside the divergence operator. With
these assumptions, the diffusion equation becomes:
    
S  r    a  r   D    r   0,
   (34)
or S  r   a  r   D2  r   0.
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 13

3.1 Plane Source


We take the source as an infinite plane source, which we can place, without loss of generality, in
the y-z plane. In this case, the flux is a function of x only, (x), and the Laplacian can be written
as:
d2
2  .
dx 2
The diffusion equation outside the source, i.e., for any x different from 0, becomes:
d 2  x 
D  a  x   0. (35)
dx 2
We can simplify this equation by dividing by D. If we define
D
L2  , (36)
a
called the diffusion area (and L called the diffusion length), the equation becomes:
d 2  x  1
 2   x   0. (37)
dx 2 L
For x > 0, Eq. (37) has mathematical solutions exp(x/L) and exp(-x/L), which give a general
solution:
  x   Ae x / L  Ce  x / L .

However, the term ex/L goes to  as x   and therefore cannot be part of a physically accept-
able solution for x > 0. The solution for x > 0 must then be  x   Ce  x / L .
By left-right symmetry, we can see that the full solution for any x must be
  x   Ce  x / L , (38)

with C being a constant which we can determine from the boundary condition at x = 0.
If S is the source strength per unit area of the plane, then the number of neutrons crossing
outwards per unit area in the positive x-direction must tend to S /2 as x  0. Therefore,
S
lim  J  x   
x 0 2
 d  x   S
 lim   D  2
x 0
 dx 
 CD  x / L  S
 lim  e 
x 0
 L  2
CD S SL
  , i.e., C  .
L 2 2D
The neutron flux outside the source is then finally:

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
14 The Essential CANDU

SL  x / L
  x  e . (39)
2D
It is interesting to try to interpret the “physical” meaning of the diffusion area L2. Let us calcu-
late the mean square distance that a neutron travels in the x-direction from the source (at x =
0) to its absorption point. We can do this by averaging x2 with the absorption rate a as a
weighting function:

 x  dx
2
a

x2  0

  dx a
0 .
With the form (39) for the flux, we can evaluate the integrals and show that
 x 2  2L2 ,
i.e., we can interpret L2 as one-half the square of the average distance (in one dimension)
between the neutron’s birth point and its absorption point.

3.2 Point Source


Let us now take the source as a single point source, assumed isotropic. Without loss of general-

ity, we can place the source at the origin of co-ordinates, r  0.
To solve for the flux in this geometry, we need to write the Laplacian  2 in spherical co-
ordinates. In view of the spherical symmetry of the problem, there is no dependence on angle
(whether polar or azimuthal), the flux is a function of radial distance r only, i.e., (r), and the
Laplacian is then:
1 d  2 d  d2 2 d
  2 r
2
  . (40)
r dr  dr  dr 2 r dr
The diffusion equation outside the source, i.e., for all points except the origin, is then:
D 2  r    a  r   0,
 d 2 2 d  (41)
i.e., D  2    a  r   0,
 dr r dr 

D 2  r    a  r   0,
 d 2 2 d  (42)
i.e., D  2    a  r   0,
 dr r dr 

d 2  r  2 d  r  1
  2   r   0. (43)
dr 2 r dr L
 r 
To solve this equation, we can write  in the form  r   . Then, in terms of , Eq. (43)
r
becomes:

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 15

d  r 
dr    r  ,
r r2
d 2  r  d  r  d  r 
d 2  r  2 2  r 
2
 dr  dr2  dr2  .
dr r r r r3
.
Substituting these forms into Eq. (43) results in the simpler form:
d 2  r  1
 2   r   0. (44)
dr 2 L
Equation (44) has mathematical solutions exp(r/L) and exp(-r/L), which give a general solution:
  r   Ae r / L  C e  r / L
er/ L e r /L
  r  A C .
r r
er / L
Now, the term   as r   and therefore cannot be part of a physically acceptable
r
solution. The solution must then be
e r / L
 r  C , (45)
r
with C being a constant which remains to be determined. To determine C, we can use the
continuity condition at the origin.
If S is the source strength, then the number of neutrons crossing the surface of a small sphere
outwards must tend to S as the sphere’s radius tends to 0, and therefore:
lim 4 r 2 J  r    S
r 0

  d  
 lim  4 r 2   D  S
r 0
  dr  
   r  r/ L  r/ L  
   e e  
 lim  4 r 2  CD  L 2    S
r 0 r
    
     
S
 4 CD  S , i.e., C  .
4 D
The neutron flux outside the source is then, finally:
S e r / L
 r  . (46)
4 D r
Again, it is interesting to interpret the physical meaning of the diffusion area L2. Let us calculate
the mean square distance that a neutron travels outwards from the source (at r = 0) to its

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
16 The Essential CANDU

absorption point. We can do this by averaging r2 with the absorption rate a as a weighting
function:

 r   4 r dr
2 2
__ a

r 
2 0

.
   4 r dr
2
a
0

With the form (46) for the flux, we can evaluate the integrals and show that
 r 2   6L2 , (47)
i.e., we can interpret L2 as one-sixth of the square of the average distance outwards between
the neutron’s birth point and its absorption point.

3.3 Flux Curvature in Source-Sink Problems and Neutron In-Leakage


Let us consider the “out-leakage” term (call it Leakout) in the above examples.
     
Leakout  leakage out per differential physical volume dr    J    D r   r  . Applying
this to the flux from a plane source (Eq. 39) gives, for x > 0:
 SL  x / L   S   x/ L
Leakout   D 2  e    e  0.
 2D   2L 
We can also show that Leakout < 0 for x < 0. If we apply the formula to the flux from a point
source (Eq. 46):
 S er/L   d 2 2 d   S e r/L 
Leakout   D 2    D  2 
r dr   4 D r 
, which we can show
 4 D r   dr
S e r/ L
  0.
4 L2 r
We see that in all cases, Leakout < 0, i.e., the leakage is inwards (in-leakage) at any point outside
the source. This can actually be seen in the general case from
  
S  r    a  r   D2  r   0,
 
stated earlier as Eq. (34), which for any point outside the source gives  D 2 r    a  r   0
. In other words, we can say that any point outside an external source in a non-multiplying
medium sees a positive flux curvature and neutron in-leakage, i.e., any differential volume is a
net sink or absorber of neutrons:
Positive flux curvature  neutron in-leakage.

4 One-Group Diffusion in a Uniform Multiplying Medium with No


External Source
We will now switch to the analysis of reactor systems with no external source, meaning that all
neutrons are produced by neutron-induced nuclear fission. In this section, as in the previous
one, we will start by assuming that all neutrons are lumped into a single energy group, i.e., G =
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 17

1. There is then, again, no need for the subscript g. Also, as explained in Section 3, we can drop
the term in s and replace t by a.
The neutron-diffusion equation in this case is:
  
 f   r   a  r   D 2  r   0. (48)

A very important point to note is that without the external source, this is a linear homogeneous
equation in the flux (if the properties are truly constant and independent of the flux). This
means that if we find one solution of the equation, then any multiple is also a solution. There-
fore, the absolute value of the flux cannot possibly be deduced from the diffusion equation
(incidentally, not from the transport equation either). This is totally different from problems
with external sources, which drive the absolute value of the flux.

4.1 Uniform Infinite Reactor in One Energy Group


Let us first assume a uniform reactor, infinite in size. This assumption makes all points in space
“equivalent”, which means that the neutron flux  would also have to be constant throughout
space. Moreover, there can be no leakage from one point to another; therefore, the leakage
term can be dropped from the equation.
The equation in the infinite uniform multiplying medium is then:
 f   a  0. (49)

This equation is interesting. The only solution is a trivial solution, i.e., a null flux,  = 0, unless
 f  a . (50)

In other words, unless the composition of the medium is exactly balanced so that Eq. (50) is
satisfied, the uniform infinite reactor cannot really operate in steady state (except in a trivial
zero-flux situation). We can call Eq. (50) the criticality condition for a uniform infinite reactor.
What happens if the criticality condition in Eq. (50) is not satisfied? Then there is no non-trivial
solution, but is this all that we can say? Actually, we can ensure that there is always a non-trivial
solution if we modify Eq. (49) by “tuning” the neutron-yield cross section by dividing it by a new
“modifying factor” which we call k∞, as in:
 f
   a  0. (51)
k
A non-trivial solution of Eq. (51) can always be guaranteed by selecting the value of k∞ as:
 f
k  . (52)
a
What this means is that if the composition of the uniform infinite reactor is modified from the
original composition by dividing the neutron-yield cross section by k∞ defined as in Eq. (52), the
modified uniform infinite reactor can then be operated in steady state with a non-zero flux; this
modified reactor is now critical. What is the value of the flux? We can see from Eq. (51) that
with the modified neutron-yield cross section, the flux can have any value; in other words, the
critical uniform infinite reactor can operate at any flux (and therefore power) value! This is a
direct consequence of the (apparent) homogeneity of the equation (i.e., the equation is of the
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
18 The Essential CANDU

form F = 0, with F an arbitrary operator independent of the flux).


Comparing Eqs. (50) and (51), the criticality condition for a uniform infinite reactor can be seen
to be:
 f
k  1 (53)
a
for criticality. The deviation of k∞ from 1 tells us how far the reactor with the original composi-
tion ( f ,  a ) is from criticality. k∞ is called the infinite reactor multiplication constant and can
have the following values:
k∞ < 1: the infinite reactor is said to be subcritical;
k∞ = 1: the infinite reactor is critical;
k∞ > 1: the infinite reactor is said to be supercritical.
The physical interpretation of k∞ can be stated as follows:
 f  f  neutron production rate
k    . (54)
a  a neutron loss rate
A quantity related to k∞, but “centred” at 0 instead of at 1, is the “reactivity”, ∞, defined as:
1
  1  . (55)
k
Therefore:
k∞ < 1   < 0: the infinite reactor is subcritical
k∞ = 1   = 0: the infinite reactor is critical
k∞ > 1   > 0: the infinite reactor is supercritical.
Real-life reactors are designed to be operated at critical or very close to critical. Therefore, in
most situations, we would expect k∞ to be very close to 1 and ∞ to be very close to 0. As a
result, while ∞ (and k∞) are absolute, dimensionless numbers, a new fractional “unit” of 1
milli-k (or mk) is defined for reactivity, where:
1 mk  0.001. (56)
For example, ∞ = + 2 mk (or – 1 mk) means that ∞ = +0.002 (or – 0.001). Another reactivity
unit, “pcm”  0.01 mk, is commonly used in Europe.

4.2 Uniform Finite Reactors in One Energy Group


We will now analyze uniform reactors of finite size (at least in some dimensions). The nuclear
properties are assumed uniform throughout the reactor, but not all points in the reactor are
equivalent spatially (some are further from or closer to the reactor boundary), so that the flux is

a function of space, r  , and also the leakage term must remain in the diffusion equation:
  
 f   r   a  r   D 2  r   0,
presented earlier as Eq. (48). This can be rewritten as:
  
 D2  r    f   r   a  r  . (57)
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 19


The left-hand side of Eq. (57) is the leakage out of the “point” r . If we integrate Eq. (57) over
the volume V of the reactor:
   
  D2  r  dr   f   a     r  dr . (58)
V V

The left-hand side of Eq. (58) is the total leakage out of the reactor. This of course must be
positive: neutrons can only go out of the reactor, not into it, because there are no sources of
neutrons outside which can “feed” neutrons into the reactor. Therefore, if Eq. (58) is true as is,
the right-hand side must be positive, and because the integral of the flux must also be positive,
so must the quantity ( f   a ) . We can therefore write Eq. (57) as:

    a 
2  r   f  r 
D (59)
 
i.e.,  2  r   Bg2  r  ,

where we have defined


 f   a
B g2  . (60)
D
The quantity Bg2 , as it appears in Eq. (59), is called the geometrical buckling of the reactor. Its
physical meaning can be understood by rewriting Eq. (60) as:

2  r 
Bg  
2
 . (61)
 r 
In other words, the geometrical buckling is the negative relative curvature of the neutron flux.
Because we have established that B g2 is definitely positive, this means that the flux curvature is
negative (see Figure 1), and in fact, in view of Eq. (59), the relative curvature is uniform (and
negative) in a homogeneous finite reactor. Note that this statement is strictly true for homo-
geneous reactors only. In real reactors, the relative flux curvature (and the buckling) can vary
and even change sign.

Figure 1 Negative flux curvature in homogeneous reactor


Because the neutron flux curvature/geometrical buckling must “bend” the flux to bring it to 0

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
20 The Essential CANDU

almost at the reactor boundary (actually, at an extrapolation distance dextr beyond the physical
boundary), it is clear that the flux curvature will be large for reactors of small dimensions and
small if the reactor dimensions are large. Therefore, B g2 in Eq. (59) is purely a geometrical
quantity.
Returning now to Eq. (57) and using the definition of geometrical buckling, the diffusion equa-
tion for homogeneous reactors can be rewritten as:
  
 f   r   a  r   DBg2  r   0. (62)

This equation has the same characteristics as Eq. (49). The only solution is a null flux,  r   0 ,
unless
 f  a  DBg2 . (63)

In other words, unless the composition of the reactor is exactly balanced so that Eq. (63) is
satisfied, the uniform reactor cannot really operate in steady state (except in a trivial zero-flux
situation). We can call Eq. (63) the criticality condition for a uniform finite reactor.
What happens if the criticality condition Eq. (63) is not satisfied? Then there is no non-trivial
solution. However, just as we did for the infinite medium, we can ensure that there is always a
non-trivial solution if we modify Eq. (62) by tuning the neutron-yield cross section by dividing it
by a similar parameter, called keff, as in:
 f   
  r   a  r   DBg2  r   0. (64)
keff
A non-trivial solution of Eq. (64) can always be guaranteed by selecting the value of keff as:
 f
keff  , (65)
a  DBg2
which can also be written as:
 f
 a
keff
Bg 
2
. (66)
D
This means that if the composition of the uniform reactor is modified from the original compo-
sition by dividing the neutron-yield cross section by keff defined as in Eq. (65), the modified
uniform reactor can then operate in steady state (i.e., as a critical reactor) with a non-zero flux.
Again, as in the infinite medium, the flux in the modified critical reactor can have any value, that
is, the critical uniform reactor can operate at any flux (and therefore power) value!
In summary, the criticality condition for a uniform finite reactor is:
 f
keff  1 (67)
a  DBg2
or

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 21

 f
1
 f   a a k 1
Bg 
2
  2 . (68)
D D L
a
Equation (67) is clearly a generalization of the criticality condition Eq. (53) for the infinite
medium, where the buckling was 0, i.e., B g2  0 (flat flux). The deviation of keff from 1 tells us
how far the reactor with the original composition ( f ,  a , D ) is from criticality.

Equation (68) is intriguing. The left-hand side, B g2 , is a geometrical quantity, as already noted.
The right-hand side, on the other hand, is a function of the nuclear properties only and is not a
geometrical quantity. It is nonetheless called a “buckling”, the material buckling:
k  1
Bm2  . (69)
L2
In view of this, the criticality condition for a uniform reactor in one neutron-energy group can
be expressed as:
Geometrical Buckling  Material Buckling, Bg2  Bm2 . (70)

The physical interpretation of keff can be obtained from:



 f  f   r  neutron production rate
keff      . (71)
 a  DBg  a  r   DBg   r  neutron loss rate (by absorption and leakage)
2 2

This ratio of production to loss is the same at any point in a homogeneous reactor and is of
course then also the same as the ratio of the reactor-integrated production and loss.
Incidentally, Eq. (64), a linear equation with a boundary condition (zero flux at the extrapolation
distance beyond the boundary), is mathematically an eigenvector problem, which can best be
seen by rewriting the equation in the form:

 a  DBg2   r    f   r  , (72)

where
1
 (73)
keff

is the eigenvalue of the problem.


Eigenvalues of eigenvector problems can take on only distinct, non-continuous values. This tells
us that keff cannot have just any value, but it can have only a certain number of distinct values.
Later, we will see that the “physical” keff is the largest of these distinct values.

4.3 Uniform Finite Reactors in Various Geometries


Equations (67) and (68) relate the reactor multiplication constant to the reactor properties and
to geometrical buckling, but we do not yet have a value for the buckling or for the distribution
of the neutron flux in the reactor. In this section, we will solve for the flux distribution and for
the value of the reactor multiplication constant keff for reactors of various geometries (and we
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
22 The Essential CANDU

can therefore also derive the value of the buckling). We will do this by showing how to solve
the diffusion equation in the following form [rewritten from Eq. (59)]:
 
2  r   Bg2  r   0. (74)

4.3.1 Infinite slab reactor


Let us consider a reactor in the shape of a slab of physical width a in the x-direction and infinite
in the y- and z-directions. The reactor is centred at x = 0; see Figure 2. This is a one-dimensional
d2
problem, meaning that  is a function of x only. The Laplacian reduces to , and the diffu-
dx 2
sion equation (74) is now:
d 2  x 
2
 Bg2  x   0. (75)
dx
This equation has mathematical solutions (x)  sin(Bgx) and (x)  cos(Bgx). However, by left-
right symmetry, sin(Bgx) is not a viable solution (because the flux would be negative in half the
space). Therefore, the only physical solution must have the form (x) = Acos(Bgx), where A is a
constant.

Figure 2 Infinite-slab reactor


a 
However, we must also satisfy the boundary condition for the problem, which is   ex   0 ,
 2 
where aex is the “extrapolated width” and x = aex /2 is the point at which the flux appears to go
 a 
to zero. Therefore, we must have A cos Bg ex   0 , which means that the values of Bg are
 2 
n
limited to B g  , where n = any odd integer.
a ex
While any odd integer value of n gives a mathematical solution of Eq. (75), only n = 1 is a physi-
cally acceptable solution, because higher values of n would give cosine functions which would
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 23

become negative for some values of x before returning to 0 at aex (see Figure 3). Therefore, the
final solution for the flux distribution in an infinite slab reactor is
 x 
  x   A cos   , (76)
 aex 
and the buckling is
2
 
B   .
2
g (77)
 aex 

Figure 3 Flux with physical and unphysical values of buckling


Note: The value of buckling which gives the only physical solution is the smallest of the mathe-
matically possible values, and the corresponding value of keff for the physical solution is the
largest of the mathematically possible values (i.e., the physical eigenvalue is the smallest of the
possible values).
Equation (76) gives the flux distribution in the reactor. However, the amplitude A, which gives
the absolute value of the flux, cannot be obtained from the diffusion equation, as noted at the
beginning of Section 4. To find A, an extraneous condition has to be imposed, for instance the
actual value of the flux at a point, or the total power of the reactor, or the power in some region
of the reactor.
For example, if we assume a value P for the power per unit area of the y-z plane and note that Ef
is the energy released per fission (typically 200 MeV  3.2*10-11 joules), we have:

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
24 The Essential CANDU

a /2 a /2
 x 
P  Ef f    x  dx  E f  f  A1 cos  dx
 a /2  a /2  ex 
a
a /2
a  x  a  a 
i.e., P  ex E f  f A sin    2 A ex E f  f sin  
  aex   a /2   2aex 
P
A ,
 a 
2aex E f  f sin  
 2aex 
and the absolute flux in the slab is
P  x 
  x  cos  . (78)
 a   aex 
2aex E f  f sin  
 2aex 
If the extrapolation distance is ignored, i.e., aex = a, this reduces to:
P x 
  x  cos   . (79)
2aE f  f  a 

4.3.2 Infinite cylindrical reactor


We now consider a reactor in the shape of a uniform cylinder of physical radius R and of infinite
length in the axial (z) direction, as shown in Figure 4. The problem is independent of the
azimuthal angle, and the flux is a function of the r variable only.

Figure 4 Infinite-cylinder reactor


The Laplacian can now be written as
1 d  d 
 r ,
r dr  dr 
and the diffusion equation (74) becomes:

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 25

1 d  d  r  
  Bg   r   0
2
r
r dr  dr 
(80)
d   r  1 d  r 
2
i.e., 2
  B g2  r   0.
dr r dr
This differential equation is actually well known to mathematicians: it is called Bessel’s equation
of order 0, and its mathematical solutions are the ordinary Bessel functions of the first and
second kind, J (Br) and Y (Br) respectively; see Figure 5.
0 0

Figure 5 Ordinary Bessel functions of first and second kind


From Figure 5, we can see that Y (Br) goes to - as r → 0 and is therefore not acceptable as a
0
physical solution for the neutron flux. The only solution is then:
  r   AJ 0  Bg r  . (81)

The flux must go to 0 at the extrapolated radial boundary Rex, i.e., we must have:
J 0  Bg Rex   0. (82)

Figure 5 shows that J0(r) has several zeroes, called ri: the first is at r1  2.405, and the second at
r2  5.6. However, because the neutron flux cannot have regions of negative values, the only
physically acceptable value for Bg is
2.405
Bg  . (83)
Rex
As a result, the flux in the infinite cylinder is given by:
 2.405r 
  r   AJ 0  , (84)
 Rex 
and the buckling for the infinite cylinder is

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
26 The Essential CANDU

2
 2.405 
B 
2
g  . (85)
 Rex 
If we use the reactor power P per unit axial height to determine A and neglect the extrapolation
distance, the following value is obtained:
0.738P
A , (86)
E f  f R2

and the absolute flux is


0.738P  2.405r 
 r  J0  . (87)
E f  f R2  R 

4.3.3 Parallelepiped reactor


We now look at a fully finite reactor geometry (i.e., the reactor is finite in every dimension, and
is uniform): the parallelepiped of sides a, b, c (see Figure 6). In this case, we use the Cartesian

co-ordinate system, and the flux is written as  r     x, y , z  . The diffusion equation becomes:
d 2  x, y , z  d 2  x, y , z  d 2  x, y , z 
2
 2
 2
 Bg2  x, y , z   0. (88)
dx dy dz

Figure 6 Rectangular-parallelepiped reactor


To solve this equation, we try a separable form for the neutron flux, i.e.,
  x, y , z   f  x  g  y  h  z  , (89)

where f, g, and h are functions to be determined. Substituting form (89) into Eq. (88), we get:
d 2 f  x d 2g  y d 2h  z 
2
g ( y ) h ( z )  f ( x ) 2
h ( z )  f ( x ) g ( y ) 2
 Bg2 f  x  g ( y )h( z )  0. (90)
dx dy dz
Dividing the equation by f(x)g(y)h(z):
1 d f  x 1 d g  y 1 d h z
2 2 2
    Bg2 . (91)
f  x  dx 2
g  y  dy 2
h  z  dz 2

The left-hand side of Eq. (91) is a sum of three terms which are functions only of x, y, and z

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 27

respectively. The right-hand side is a constant. The only way in which this can happen is if each
of the three terms on the left-hand side is a constant on its own, i.e., if we can write:
1 d f  x 1 d g  y 1 d h z
2 2 2
  Bx ,
2
  By ,
2
  Bz2 , (92)
f  x  dx 2
g  y  dy 2
h  z  dz 2

where B x2 , B y2 , B z2 are constants and

Bx2  By2  Bz2  Bg2 . (93)

Each of the equations in (92), e.g.,


1 d f  x d 2 f x
2
  Bx 
2
 Bx2 f  x   0, (94)
f  x  dx 2
dx 2

is exactly the same equation as for the slab reactor, with the same solution
f  x   cos  Bx x  , (95)

where

Bx  . (96)
aex
The full solution for the neutron flux in the parallelepiped reactor is therefore:
 x   y  z 
  x, y, z   A cos   cos   cos   . (97)
 aex   bex   cex 
The quantities
2 2 2
     
B    , By2    , Bz2    ,
2
x (98)
 aex   bex   cex 
are called the “partial bucklings” in the three directions, and the total buckling is
2 2 2
     
B       .
2
g (99)
 aex   bex   cex 
If we normalize the flux to the total fission power P of the reactor and neglect the extrapolation
distance, we can show that the normalization constant, A, is given by
 3P
A . (100)
8abcE f  f

4.3.4 Finite-cylinder reactor


In the case of a finite-cylinder uniform reactor of radius R and height H (see Figure 7), we use

the cylindrical co-ordinate system to write the flux as  r    r , z  . The Laplacian can be
written as:

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
28 The Essential CANDU

d2 1 d d2
2    , (101)
dr 2 r dr dz 2
and the diffusion equation becomes
d 2  r, z  1 d  r, z  d 2  r, z 
2
  2
 Bg2  r, z   0. (102)
dr r dr dz

Figure 7 Finite-cylinder reactor


To solve this equation, we again try a separable form for the neutron flux, i.e.,
  r, z   f  r  g  z  , (103)

where f and g are functions to be determined. Substituting Eq. (103) into Eq. (102), we get:
d 2 f r 1 df  r  d 2g  z
2
g  z    f  r  2
 Bg2 f  r  g  z   0. (104)
dr r dr dz
Let us divide this equation by f(r)g(z):
1 d f r 1 1 df  r  1 d g z
2 2
    Bg2 . (105)
f  r  dr 2
f  r  r dr g  z  dz 2

The left-hand side of Eq. (105) is the sum of a function of r and a function of z. The right-hand
side is a constant. The only way in which this can happen is if the parts in r and z are each
individually equal to a constant, i.e., if we can write:
1 d f r 1 1 df  r  1 d g z
2 2
   B 2
,   Bz2 , (106)
f  r  dr f  r  r dr g  z  dz
2 r 2

where Br2 ,B g2 are constants and

Br2  B g2  Bg2 . (107)

The equations in (106) have been seen before; they are the same equations as for the infinite
cylinder and the slab reactor respectively and therefore have the same solutions:
f  r   J 0  Br r  , g  z   cos  Bz z  , (108)

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 29

where
2.405 
Br  and Bz  , (109)
Rex H ex
where Rex and Hex are the extrapolated radius and the extrapolated axial dimension of the
reactor. The full solution for the neutron flux in the finite-cylinder reactor is therefore:
 2.405r   z 
  r, z   AJ 0   cos  , (110)
 Rex   H ex 
2 2
 2.405    
B 
2
r  , Bz  
2
 (111)
 Rex   H ex 
are called the radial and axial bucklings respectively, and the total buckling is
2 2
 2.405    
B 2
g    . (112)
 Rex   H ex 
If we normalize the flux to the total fission power P of the reactor and neglect the extrapolation
distance, we can show that the normalization constant is:
3.63P
A . (113)
 R 2 HE f  f

4.3.5 Spherical reactor


The last geometry we will look at is a spherical uniform reactor of radius R. We use the spheri-
cal co-ordinate system, and because there is spherical symmetry, the flux can be written as

 r    r  . The Laplacian can be written as:
1 d  2 d 
2  r , (114)
r 2 dr  dr 
and the diffusion equation becomes
1 d  2 d  r  
  Bg  r   0. (115)
2
2 r
r dr  dr 

Let us try to represent (r) in the form

 r
 r  , (116)
r
where the function   r  is to be determined. Equation (115) then reduces to:

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
30 The Essential CANDU

1  d  2  1 1 d  r    2  r
 r   2   r       Bg 0
r2  dr   r r dr   r

1 d  d  r   2  r
i.e.,     r   r    Bg 0
r2  dr  dr   r
1  d  r  d  r  d 2  r   2  r
i.e,     r   Bg  0,
r2  dr dr dr 2
 r

d 2  r 
which finally reduces to 2
 Bg2  r   0. (117)
dr
The general solution of this equation is:
  r   A sin  Bg r   C cos  Bg r  ,

which gives
sin  Bg r  cos  Bg r 
 r  A C . (118)
r r
However, the cosine term goes to  as r → 0, which is physically not acceptable, whereas the
sine term is acceptable because it has a finite limit as r → 0 (as can be verified using L’Hôpital’s
rule). Therefore, the final solution for the spherical reactor is
sin  Bg r 
 r  A ,
r
where, for the same reason as in the other geometries (to guarantee no negative flux), Bg must
take the lowest allowable value, i.e.,

Bg  , (119)
Rex
so that finally
 r 
sin  
 r  A  Rex 
. (120)
R
If we normalize the flux by imposing a value P for the total fission power of the reactor and
neglect the extrapolation distance, we find that:
P  4 AR 2 E f  f
P (121)
 A .
4R E f  f
2

4.4 Flux Curvature and Neutron Out-Leakage


Let us consider the “out-leakage” term (call it Leakout) in the above examples: Leakout  outleak-

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 31

     
age per differential physical volume dr    J    D  r    r  . Applying this to the flux in a
uniform slab reactor (Eq. 76) gives, for x > 0:

   x 
2
   x 
Leakout   D  A cos 
2
    DA   cos    0.
  aex    aex   aex 0 
We can similarly show for all other uniform reactors that Leakout > 0. This can actually be seen
  
in the general case from  f   r   a  r   D2  r   0 (presented earlier as Eq. (48)), which
  
for any point in the uniform reactor gives  D 2 r    f   a  r   DB 2 r   0 . This
means that we can say that any point in a uniform reactor sees neutron out-leakage (i.e., any
differential volume is a source of neutrons).
Note again that the above finding applies to uniform reactors, not to all reactors in general.
Real reactors may have regions which are net sinks of neutrons, where the flux curvature is
positive and where there is a net in-leakage of neutrons. The general rule that we can infer
from the above and also from what we learned in source-sink problems is that:
Positive flux curvature  Net neutron in-leakage, while
Negative flux curvature  Net neutron out-leakage.

5 Reactors in Two Neutron-Energy Groups with No External Source


5.1 The Neutron Cycle and the Four-Factor Formula
Neutrons born in fission have high energy (see the sketch of their energy distribution in Figure
8).

Figure 8 Energies of fission neutrons

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
32 The Essential CANDU

It can be seen that the peak in their energy distribution is ~0.73 MeV. A 1-MeV neutron has a
speed of ~13,800 km/s: fission neutrons are fast neutrons. There are essentially no fission
neutrons born with thermal energies ~1 eV; the reference thermal energy is 0.025 eV, speed
2,200 m/s, T = 293.6 K.

Figure 9 Sketch of cross section versus neutron energy


However, the probability of neutrons inducing fission in fissile nuclei (such as 235U) is orders of
magnitude larger for thermal neutrons (with energies of a small fraction of 1 eV) than for fast
neutrons (see Figure 9). This is why thermal reactors (such as CANDU reactors) use a neutron
moderator (such as heavy water in CANDU) to slow neutrons down. This leads to the neutron
cycle in thermal reactors: neutrons are born “fast” (i.e., with relatively high energy, typically in
the range of MeV) and are “encouraged” to slow down to the thermal range, where they induce
more fissions. Of course, fast neutrons do induce a small number of fissions; in fact, only fast
neutrons can fission non-fissile nuclides (e.g., 238U), because there is usually a minimum energy
threshold for such fission to occur. Moreover, while following the cycle, neutrons can also
escape (leak out of the reactor) or be captured in non-productive absorptions at any energy, in
particular in the “resonance range” (between ~1 eV and ~105 eV), where many strong reso-
nance-absorption peaks (mostly leading to non-productive captures) exist. To reduce resonance
capture, most power reactors are designed with the fuel lumped into (mostly) cylindrical
elements in fuel-rod assemblies (inside fuel channels in the case of CANDU) surrounded by a
moderator. Note also that some up-scattering (increase in neutron energy on collision with a
nucleus) can occur in the neutron cycle; this is much less than the rate of down-scattering
(moderation) of neutrons, and the up-scattering cross section is sometimes neglected. The
neutron cycle in a thermal reactor and with no up-scattering (i.e., scattering from a low-energy
group to a high-energy group neglected) is illustrated in Figure 10.

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 33

Figure 10 Neutron cycle in thermal reactor


We can derive an equation for the infinite-lattice multiplication constant k∞ by “following” the
various phases of the neutron cycle in the “neutron-generation” view, as follows:
 Imagine that N neutrons are born in thermal fissions in one generation (each ther-
mal fission produces about 2.5 neutrons on average).
 There will also be some neutrons born in fast fissions in the same generation. Define
 as the ratio of the total number of fission neutrons born in both fast and thermal
fissions to the number born in thermal fissions.  is called the fast-fission ratio ( >
1). Therefore the total number of fast neutrons is now N.
 These neutrons start to slow down, but some are captured in non-productive reso-
nances. If we define p as the resonance-escape probability (p < 1), the number of
neutrons which survive to thermal energies is now Np.
 These thermal neutrons can be absorbed in fuel or in non-fuel components of the
medium. If we define f as the ratio of thermal neutrons absorbed in fuel nuclides to
the total number of thermal neutrons absorbed in this same generation, then the
number of neutrons absorbed in fuel in the current generation is Npf; f is called the
fuel utilization (f  1).
 Some of the neutrons absorbed in fuel may induce more fissions, and some may not.
Define  as the average number of fission neutrons released per thermal absorption
event in the fuel. Then the total number of fission neutrons born (in this new gen-
eration!) from these thermal fissions is now Npf.  is called the reproduction fac-
tor ( > 1).
 We have now gone around the cycle one full time, and we can compare the number
of neutrons in the same phase in successive generations: the ratio in successive gen-
erations is k∞, given by:
N  pf 
k    pf . (122)
N

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
34 The Essential CANDU

5.2 The Two-Energy-Group Model


The neutron energy range in the neutron cycle is very wide, eight or nine orders of magnitude
(from ~10 MeV to ~0.01 eV). From this it is obvious that the one-neutron-energy-group diffu-
sion treatment, while very instructive, cannot be very accurate. Because a very large majority
of fissions are induced by thermal neutrons and the fission rate is the most important rate to
calculate correctly (because it is essentially the heat-production rate), a two-energy-group
treatment is often sufficiently accurate in diffusion calculations. The nuclear properties in two
groups must of course be obtained by proper averaging of the detailed many-group properties
determined by a transport-theory calculation. Next, we will analyze the two-energy-group
model.

5.3 Neutron Diffusion Equation in Two Energy Groups


With two energy groups, we will need:
 
 Two fluxes 1  r  and 2  r  , with index 1 for the fast (sometimes called the slowing-
down) group and index 2 for the thermal group;
 
 Two absorption cross sections,  a1 r  and  a 2 r  ;
 
 Two neutron-yield cross sections,  f 1 r  and  f 2 r  , acting on the fast and thermal
fluxes respectively;
 
 Down-scattering and up-scattering cross sections, 12 r  and  21 r  respectively;
 
 Two diffusion coefficients, D1 r  and D2 r  .
The diffusion equation for two energy groups is actually two equations, one for each group.
There must be neutron balance in each group. As previously indicated, all neutrons are born in
the fast group:
     
 f 1  r  1  r    f 2  r  2  r   21  r  2  r 
      
 a1  r  1  r   1 2  r  1  r     D1  r   1  r   0,
(123)
        
12  r  1  r    a 2  r  2  r   21  r  2  r     D2  r  2  r   0. (124)

5.4 Uniform Infinite Medium in Two Energy Groups


As we did in one energy group, let us first assume a uniform reactor, infinite in size. Again, all
points in space are “equivalent”, which means that the neutron fluxes are constant through
space and there is no leakage from one point to another; therefore the leakage terms can be
dropped from the equations.
The diffusion equations (123) and (124) in the infinite uniform multiplying medium then be-
come:
 f 11   f 22   212  a11  121  0
121   a 22   212  0,
which can be written as a linear homogeneous system of two equations:

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 35

f1  a1  12  1   f 2  21  2  0 (125)

121   a 2  21  2  0. (126)

Now, this system can have a non-trivial flux solution if and only if the determinant of the flux
coefficients is zero, i.e.:
  f 1   a1  12    a 2   21    f 2  2 1  12  0. (127)

Satisfying the criticality condition, Eq. (127), requires a very fine balance between the nuclear
properties. What if that balance is not achieved and Eq. (127) is not satisfied? Then there is no
non-trivial solution, i.e., we do not have a real operating infinite-medium reactor. However, we
can ensure that there is always a non-trivial solution if we “tune” the neutron-yield cross
sections by dividing them by a parameter, k∞ (to be determined). The diffusion equations then
become:
  f 1    
  a1  12  1   f 2  21  2  0 (128)
 k   k 
121    a 2  21  2  0, (129)

and the criticality condition is:


  f 1   
  a1  12    a 2  21    f 2  21  12  0. (130)
 k   k 
From Eq. (130), we can determine the value that k∞ must have for criticality:
 f 1  a 2  21    f 212
k  . (131)
a1a 2  a121  a 212
As before, we can define the reactivity as
1
  1  .
k

The difference of k∞ from unity (or of ∞ from 0) tells us how far from critical the original
uniform medium is.
When the fast-fission and up-scattering cross sections are neglected, the form obtained for k∞ is
simpler and instructive:
 f 2 12
k, no up  scattering & no fast  . (132)
 a 2 a1  12
fission

The second factor in Eq. (132),


12
,
 a1  12
gives the probability that fast neutrons will down-scatter relative to their probability of being
absorbed or down-scattered; i.e., it is the probability of fast neutrons surviving to thermal

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
36 The Essential CANDU

energies, which is simply the resonance-escape probability p in Eq. (122). The first factor in Eq.
(132) is the number of fission neutrons produced per thermal absorption, i.e., it is the repro-
duction factor  .
Another important quantity to determine in the two-group model, which has no meaning in the
one-group treatment, is the ratio of group fluxes. In the uniform infinite medium, this can be
obtained most simply from Eq. (129):
2 1 2
 . (133)
1  a 2   21
Note that the criticality condition (130) ensures that the same value would be obtained from Eq.
(128).

5.5 Uniform Finite Reactors in Two Energy Groups


We now analyze uniform finite reactors. The diffusion equations are now:
     
 f 11  r    f 22  r   212  r   a11  r   121  r   D121  r   0 (134)
   
121  r    a 22  r   212  r   D2 22  r   0. (135)

We can try to find a solution for the flux which is separable in space and energy, i.e., where the
two-row flux vector

  1  r  
 r     
  r  
 2 
can be written as a group-dependent amplitude times a group-independent flux shape. If we
can solve the equation with such a solution, then it is a good solution:
 
 1  r    A1  r  
       . (136)

 2  r  A
  2   r  
Substituting Eq. (136) into Eq. (135), we get:
 
 12 A1  a 2 A2  21 A2   r   D2 A22  r   0, (137)

and if we divide by  r  :

 2  r  1 2 A1   a 2 A2   21 A2
  . (138)
 r  D2 A2

The right-hand side of Eq. (138) is a single number, independent of space, which we can write as
–B2. [Note: We would have reached a similar conclusion if we had substituted Eq. (136) into Eq.
(134).] Equation (138) is analogous to Eq. (61), i.e., B2 is a group-independent geometric
buckling that is applicable to both the fast and the thermal groups:
 
2g  r    B 2g  r  , g  1, 2. (139)

Because this is exactly the same equation as we found with one energy group, the flux distribu-
tion in two groups is exactly the same as in one energy group: i.e., a product of cosines in a

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 37

parallelepiped reactor, an axial cosine times a radial Bessel function in a cylindrical reactor, and
a sine-over-r function in a spherical reactor. The only new quantity is the ratio of the group
fluxes!
Let us now return to the criticality conditions for a finite reactor. Substituting Eq. (139) into Eqs.
(134) and (135) gives a linear homogeneous system, and therefore we can divide the yield cross
sections by keff as usual to ensure a non-trivial solution:
  f 1      
  a1  12  D1B 2  1  r    f 2  21  2  r  0 (140)
  k 
 keff   eff 
 
121  r   ( a 2  21  D2 B 2 )2  r   0. (141)

The criticality condition for finite reactors in two energy groups is found by equating the deter-
minant of this system to zero, which yields:
 f 1   a 2   21  D2 B 2    f 212
k eff  . (142)
 a1  D1 B 2   a 2   21  D2 B 2   12   a 2  D2 B 2 

If we consider the simpler case obtained by neglecting up-scattering and fast fission, as we did
following Eq. (131), we get:
 f 212
keff , no up  scattering & no fast fission 
 a1  D1B 2   a 2  D2 B 2   12  a 2  D2 B 2 
(143)
 f 212
 .
 a1  12  D1 B   a 2  D2 B 
2 2

The ratio of Eq. (143) to Eq. (132) gives the effect of leakage:
 f 212
keff   a1  12  D1B  a 2  D2 B 
2 2
a1  12 a 2
  . (144)
k   
f 2 12  a1  12  D1B a 2  D2 B 2
2

 a 2  a1  12 

The first factor is the ratio of (fast absorption + down-scattering) to (fast absorption + down-
scattering + fast leakage). Therefore, the first factor represents the fast non-leakage probability,
which we can call P1NL. In the same way, the second factor represents the thermal non-leakage
probability, P2NL. Therefore, Eq. (144) is equivalent to
keff  k P1NL P2 NL , (145)

and the total non-leakage probability is


PNL  P1NL P2 NL . (146)
If we use the four-factor formula for k∞, then Eq. (146) becomes a six-factor formula for keff:
keff   pf P1NL P2 NL . (147)

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
38 The Essential CANDU

6 Solution for Neutron Flux in a Non-Uniform Reactor


For uniform (homogeneous) reactors of various geometries, we were able to solve the diffusion
equation analytically and find closed forms for the neutron-flux distribution. However, real
reactors are not homogeneous. The steady-state neutron-diffusion equation must then be
solved numerically. This section shows one method for doing this in the two-neutron-energy
group case.
We start with the two-group neutron-diffusion-equation system, Eqs. (123) and (124), which we
rewrite with the required keff added as a divisor to the fission cross sections (6.1):
       
( f 1  r  1  r    f 2  r  2  r ) / keff  21  r  2  r   a1  r  1  r  
    
1 2  r  1  r     D1  r   1  r   0 (148)
   
12  r  1  r    a 2  r  2  r  
    
 21  r  2  r     D2  r   2  r   0. (149)

All cross sections in a lattice cell are homogenized (averaged spatially within the cell) and
condensed to two groups using a multi-group transport code for each lattice cell, but in a real
reactor, they must also incorporate the effects of reactivity devices superimposed upon the
basic lattice.

6.1 Finite-Difference Form of the Neutron-Diffusion Equation


The simplest method for solving the neutron-diffusion equation in a non-homogeneous reactor
starts by developing a finite-difference form of the equations.

Figure 11 Face view of very simple reactor model


Consider a two-dimensional view of a reactor model, where each spatial dimension has been
subdivided into intervals (see Figure 11). The model is then a collection of homogeneous cells
(which may generally all have different nuclear properties). These cells can be basic lattice cells
or subdivisions of these cells. The finite-difference method then consists of solving for the flux
distribution at a single point in each cell, the centre of the cell in the mesh-centred formulation,
which we will use here.

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 39

Let us look at one of these cells (which we label with a superscript C for Central) and its six
nearest neighbours in the three directions, labelled with a superscript n = 1 to 6 (1 to 4 in x-y
geometry). We get the finite-difference form of the diffusion equation if we integrate Eqs. (148)
and (149) over the volume of cell C:
 
Cf11  r   Cf 22  r     C   
C 
   C

21 2  r   C
a1  C

12 1 r     D1 
1  r  dr  0 (150)
 keff 
    
  1C 21  r     Ca 2   C2 1  2  r     D2C  2  r   dr  0. (151)
C  

where we have dropped the dependence on r for the properties because the cells are homo-
geneous.
There are two types of integrals in Eqs. (150) and (151). The first type does not involve the

divergence operator "". In these integrals, the (homogeneous) cross sections can be taken
out of the integral sign, for example:
 
  Ca 22  r    Ca 2  2  r  dr.
C C

For the integral on the right-hand side, the approximation is made that the integral is equal to
the value of the flux at the centre of the cell multiplied by the volume V C  x C y C z C of the
cell ( x C , y C , z C being the dimensions of the cell in x, y, and z). Then:

  Ca 22  r    Ca 22CV C . (152)
C

The second type of integral has the divergence operator in the integrand. This type of integral
simply represents the leakage out of cell C to its neighbours. By Gauss’s theorem, the volume
integral over the divergence is equal to a surface integral, for example,
    
   D1C  1  r  dr   D1C  1  r   nds
ˆ   J 1  nds
ˆ , (153)
C S S

where S is the surface of the cell, n̂ is the outer normal to the surface, and by Fick’s Law, J 1 is
the net outward current at a point on the surface of the cell.

Figure 12 Flux in a cell and in immediate neighbours


©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
40 The Essential CANDU

Figure 13 Linear treatment of flux in central cell and one neighbour


Because the cell has six faces, the surface integral is evaluated over the six interfaces between
cell C and its six neighbours, n = 1-6. If we take as an example the face towards the cell with the
lowest x-value (n = 1) for cell C, we can evaluate the surface integral of the current in the fast
group over that face (which we call face C1) as follows. First, denote the group-1 flux at the
centre of the face by 1C1 (see Figure 12). Then let us assume that the flux is linear between the
centre of the cell and the centre of the face (see Figure 13). Then the group-1 current from cell
C to its neighbouring cell 1, at the centre of the face, is:

d 1C  C  1C1 
C  1
J 1,C 1   D C
1   D1 . (154)
dx 1 C
x
2
Similarly, the group-1 current from cell 1 to cell C is:
1C1  11 
J 1,1C  D 1
1 . (155)
1 1
x
2
However, the current must be continuous at the face, i.e.,
J1,1C  J1,C1. (156)

Then by equating Eqs. (154) and (155) and solving for the interface flux, we find that:
D1C x11C  D11xC11
1C1  . (157)
D1C x1  D11x C
We can then substitute this into Eq. (155) and calculate the current at the centre of the face:
1C  1C1  2 D1C D11 1C  11 
J 1,C 1   D C
 C 1 . (158)
  
1
1 C D x D 1
x C
x 1 1
2
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 41

If we assume that the average current over the face is equal to the current at the centre of the
face, the total current over face C1 is:
 2 D1C D11y C z C C
C1 1
J  ˆ
nds  
D1C x1  D11xC
1  11   A1C1 1C  11  , (159)

where A1C1 is a coupling coefficient:

2 D1C D11y C z C
A  C 1
C1
. (160)
D1 x  D11x C
1

A very similar process can of course be performed for all six faces between cell C and its
neighbours. The total outward current in Eq. (153) is then:

ˆ   A1Cn 1C  1n  .
6

S
J 1  nds
n 1
(161)

Using all the above results, the finite-difference neutron-diffusion Eqs. (150) and (151) then
become:
Cf 11C   f 22C C
V  C212CV C   Ca1  1C2  1CV C   A1Cn 1C  1n   0
6
(162)
keff n 1

1C21CV C   Ca 2  C21  2CV C   A2Cn 2C  2n   0.


6
(163)
n 1

These equations couple cell C to its closest neighbours. These and similar equations for all the
other cells in the model make up a coupled system of linear homogeneous equations for the
fluxes at the centres of the cells. If there are N cells in the model, the system is composed of 2N
equations. To solve this coupled system, we need also the boundary conditions at the model
edges, which we look at now.
The edge cells have neighbour cells only towards the “interior” of the model. In directions
outward from the model, the diffusion boundary condition with vacuum is that the flux goes to
zero at the extrapolation distance beyond the boundary.
However, we can express the boundary condition in the same form as Eqs. (162) and (163) by
creating a “dummy” neighbour cell of width 2 dextr, where dextr is the extrapolation length =
2.1312 DC [see Eq. (33)] and forcing the flux to be zero at the extrapolation distance (see Figure
14).

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
42 The Essential CANDU

Figure 14 Flux in cell at reactor boundary

6.2 Iterative Solution of the Neutron-Diffusion Equation


The finite-difference neutron-diffusion-equation system described above is a coupled set of 2N
homogeneous equations, where N is the number of real cells in the model. However, actually
we seem to have (2N + 1) unknowns: two values of flux for each of the N cells, plus the value of
keff. In fact, however, there are only 2N unknowns because the overall flux normalization is
arbitrary in the homogeneous diffusion equation and only relative fluxes can be determined.
The absolute flux normalization factor must be determined using an extraneous condition such
as the total reactor fission power.
In real-life reactors, the number of cells N is very large (typically tens of thousands). Conse-
quently, the system of equations can be solved only by iterative techniques. The first step is
that a flux guess must be selected; this can be as simple as a “flat” flux distribution, i.e., the
same value in all cells. A first guess must also be made for keff; for instance, we can use the
guess keff = 1.
Once a flux guess has been established, the flux iterations can begin. At each iteration, the
computer program progresses from cell to cell in the x-direction, the y-direction, and finally the
z-direction, with each cell considered as the central cell C. The two equations (162) and (163)
are solved simultaneously to obtain the fluxes in cell C, 1C and 2C , using the latest values of the
fluxes in the neighbouring cells. As the sweep progresses through the cells, all the fluxes are
updated and used as the latest flux values when the cells are called upon as neighbours. During
this sweep, the value of keff is kept constant. However, when a full sweep over all reactor cells
(a “full iteration”) is completed, the value of keff is updated. It can be recalculated from the
definition of the reactor multiplication constant:

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 43

Neutron Production
keff 
Neutron Absorption  Neutron Leakage
    
r [ f 1  r  1  r    f 2  r  2  r ]dr
           ,

  a1 1
  r    r    a2  r   2  r     J 1  r     J 2  r  
 dr
r

where the latest fluxes from the latest iteration are used and the integrals are evaluated as
sums over the cells.
One iteration is not sufficient to obtain a self-consistent solution of the entire system of finite-
difference equations. We must repeat the iterations until the fluxes converge, i.e., until the
relative difference in flux in each cell from one iteration to the next is very small, smaller than
an accepted tolerance, typically ~10-5. A convergence criterion is also needed for keff; typically, a
difference of 0.001 or 0.01 mk from one iteration to the next is used. Once convergence has
been reached, we have the sought-after solution for the flux shape, i.e., we have the unnormal-
ized (relative) flux distribution. To find the absolute values for the flux, we can normalize the
flux distribution to the total reactor fission power, for example.

7 Energy Dependence of Neutron Flux


In Section 4, we studied the spatial dependence of the neutron flux in one energy group in
uniform reactors. In Sections 5 (uniform reactors) and 6 (non-uniform reactors), we studied
how to determine the neutron flux both in space and in two energy groups. But what can be
said about the variation of the neutron flux with energy when treated as a continuous variable
(not just in two groups)? We examine this in the present section.
When we consider the range of neutron energies from ~1 (or a few) MeV down to a fraction of
1 eV, we can subdivide the flux energy spectrum into three broad regions (note that the region
boundaries cannot be considered “sharp”):
 Fission-neutron energies (>~0.5 MeV)
 Slowing-down (or moderation) range (~0.5 MeV - ~0.625 eV)
 Thermal range (< ~0.625 eV)
We will look at the variation of neutron flux with energy in these three regions in turn.

7.1 The Fission-Neutron Energy Range


This range consists of energies above ~0.5 MeV, up to several MeV, with a maximum flux
magnitude around 0.7 MeV. The energy distribution (“spectrum”) of neutrons in this range is
determined by experimental measurement. It is found to be well approximated by:
  E   0.453e 1.036 E sinh 2.29 E , (164)

where E is in MeV (see Figure 15). [Note: In Chapter 3, the Watt spectrum is used.]
Note that this is a distribution of the number of neutrons, but the flux can be obtained simply
by multiplying by the neutron speed  (not to be confused with , the fission neutron yield).

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
44 The Essential CANDU

Figure 15 Energy distribution of fission neutrons

7.2 The Thermal Energy Range


Next, let us look at the thermal energy range, with neutron energies less than ~0.625 eV. Here,
the neutrons are in thermal-equilibrium balance with the ambient medium at temperature T.
Analogously to the atoms in an ideal gas at thermal equilibrium, the neutron population has a
Maxwellian (or approximately Maxwellian) distribution. In terms of flux:
E
E 
 E  e kT
. (165)
 kT 
2

Room temperature is by convention taken as T = 293.6 K = 20.4oC, which gives kT = 0.0253 eV.
The energy value E = kT is the most probable neutron energy in a Maxwellian flux distribution,
and the corresponding “thermal neutron” speed is:
v  kT  0.0253 eV   2200 m / s. (166)

7.3 The Slowing-Down Energy Range


The slowing-down energy range is intermediate between the fission-energy range and the
thermal-energy range. This is the most complex range because of its very broad size and the
highly complex resonance scheme presented to neutrons by heavy nuclides, e.g., 238U.
Therefore, especially when considering neutron absorption in the resonance range in fuel,
neutron slowing-down can be very difficult to calculate and is now mostly done by numerical
computation using complex lattice codes. However, under certain approximations, it is possible
to derive an analytic form for the energy distribution of neutrons as they slow down in the
moderator through collisions with light nuclei.

7.3.1 Slowing down in a hydrogen moderator


The simplest case is slowing down in hydrogen. Using analysis of the kinematics of neutron

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 45

collisions with hydrogen nuclei, the slowing-down flux (E) in hydrogen can be derived. It can
be shown that below the lower boundary Es of the fission-neutron energy range and neglecting
neutron absorption relative to scattering (a fairly good approximation), the slowing-down flux is
inversely proportional to energy E:
S
 E  , (167)
E s
where S is the fission source and s is the scattering cross section (assumed to be independent
of energy).
This provides an important, simple, basic formula for the slowing-down spectrum, even if it is
somewhat of an approximation. Another way of interpreting this relationship is that the prod-
uct E(E) is nearly constant with energy below Es.

7.3.2 Slowing down in a non-hydrogenous moderator


The analysis of the slowing-down spectrum in a non-hydrogenous moderator is more complex,
but under similar approximations, the same form of the slowing-down flux can be found:
S
 E  , (168)
E s

This has the same form as in hydrogen, with an additional factor of  in the denominator.  is
the average lethargy gain per collision, where the lethargy u is defined as:
E 
u  ln  s  . (169)
E
Refer to Chapter 3 for greater detail.

7.3.3 Relaxing the no-absorption approximation


If the absorption cross section is no longer neglected, but is assumed to vary smoothly with
energy (i.e., in the absence of resonances), the 1/E flux spectrum can be modified to:

S  Es   E '  dE ' 
 E  exp    a  . (170)
Et  E     E '  E '
 E t 
The exponential factor in Eq. (170) represents the probability that the neutron survives slowing
down to energy E; i.e., it is the resonance-escape probability to energy E, which we can denote
as p(E).
From the general form of the slowing-down flux, the following simplified statements about the
product E(E) can be deduced:
 If absorption is neglected, and under the assumption that the scattering cross sec-
tion does not depend on energy, E(E) is constant (flat) with E.
 If absorption is included, and assuming a smooth variation of the absorption cross
section with E, then E(E) will decrease smoothly for decreasing E.
These statements are shown in graphical form in Figure 16.

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
46 The Essential CANDU

Figure 16 Variation of flux with energy if absorption is smooth

7.3.4 Slowing down in the presence of resonances


Let us now consider absorption resonances. At a resonance energy, the neutron flux decreases
significantly because of the very high absorption cross section at that energy.
As a result of the dip in flux, the absorption rate in the resonance (which is proportional to the
flux) is reduced relative to the absorption rate with the “background” flux value at energies
above and below the resonance, i.e., the product a is smaller than if the flux were unaffected.
This is called “resonance self-shielding”.
On account of resonances, then, the slowing-down flux will be distorted relative to the smooth
curve in the previous subsection, as shown in Figure 17.

Figure 17 Variation of flux with energy in the presence of resonances

7.4 Neutron Flux over the Full Energy Range


With the results in the previous subsections, we are able to “piece together” the neutron flux
over the energy range from fission energies to the thermal range, using:
 the fission spectrum at energies above about 500 keV;
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 47

 the slowing-down spectrum to about 0.625 eV. [Note that in the thermal energy
range, neutrons can gain as well as lose energy in collisions. To be consistent with
the approximation of no up-scattering in the derivation of the slowing-down spec-
trum, the “boundary” between thermal and epithermal energies should be selected
sufficiently high to ensure negligible up-scattering from the thermal region to the
epithermal region. This is one reason for the typical choice of 0.625 eV, which is
about 25 times the most probable energy of 0.025 eV at room temperature, as the
lower energy boundary of the epithermal range.]
 the Maxwellian spectrum at thermal energies. [Note that it is not a perfect Maxwel-
lian, being distorted somewhat by neutron absorption].
The piecing together of the neutron-flux portions in the various energy regions is shown in the
sketch in Figure 18.

Figure 18 Sketch of flux variation over full energy range

8 Basic Design of Standard CANDU Reactors


In this section, we look at the basic features of the standard CANDU reactor design. The CANDU
design is based on the following design principles and the rationale for each:
 Use natural uranium (NU) as fuel. The reason for this choice is that Canada has substantial,
high-grade, domestic uranium mineral resources and, for a reactor cooled and moderated
by heavy water (D2O), it is not necessary to use enriched uranium fuel. Developing a ura-
nium-enrichment industry is extremely expensive, primarily because of the sophisticated
technology involved as well as the costs associated with the stringent materials safeguards,
physical security, and criticality safety requirements associated with enriched nuclear mate-
rials. Moreover, the production cost of uranium-isotope enrichment is also high using cur-
rent technology, in part due to the energy requirements. Consequently, natural-uranium
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
48 The Essential CANDU

fuel is much cheaper and simpler to fabricate than enriched-uranium fuel.


 Use heavy water (D2O) as neutron moderator and reactor coolant. The fission chain
reaction cannot be self-sustaining with light water (H2O) as moderator and NU as fuel. The
reason is that normal hydrogen (H) has a fairly large absorption cross section for thermal
neutrons. Hence, light water robs neutrons from the chain reaction, and the fuel must be
enriched to restore a self-sustaining balance between neutron production and capture. On
the other hand, deuterium (D), or heavy hydrogen, has a very small neutron-absorption
cross section and enables the use of NU fuel: heavy water provides high neutron economy.
Although D2O is very expensive to produce due to the low abundance of D relative to H in
nature, the large mass difference between D and H makes separation relatively easy to
achieve using chemical technology processes, and, unlike enriched fuel, the D2O retains its
favourable nuclear properties for the full life of the reactor.
 Use a pressure-tube rather than a pressure-vessel design. The coolant takes heat away
from the fuel (to create steam and produce electricity in a turbine generator) and becomes
very hot. If we want to keep the coolant liquid (or to limit boiling to a small amount), we
must pressurize it (to about 100 atmospheres, ~10 MPa). Therefore, a pressure boundary
must be provided. In light-water reactors (LWRs) a large, thick, steel pressure vessel is used
that surrounds the entire reactor core. The first prototype CANDU reactor [Nuclear Power
Demonstration (NPD), the first reactor to produce electricity in Canada (about 20 MWe)],
was initially conceived with a pressure-vessel design. However, using D2O as the moderator
results in a larger vessel than for an LWR, because the smaller scattering cross section of
D2O (compared to that of H2O) requires a larger moderator volume. Concerns were ex-
pressed about the huge size of the pressure vessels which would be required for larger
CANDU reactors, because such vessels would be difficult to fabricate and very expensive.
The design of NPD, and that of all later CANDU reactors, was therefore changed to a pres-
sure-tube design, in which the pressure boundary occurs in a small, relatively thin, fuel
channel that surrounds each fuel assembly module and its associated coolant. A pressure-
tube design was made possible only by emerging research on zirconium, which showed that
zirconium had a very small neutron-absorption cross section and would be a very good can-
didate metal for the pressure tubes. Using steel, with its much larger absorption cross sec-
tion, would not permit a working pressure-tube design for a reactor using natural-uranium
fuel. The pressure-tube design has the additional safety feature that the break of one pres-
sure tube is more manageable than the break of a large pressure vessel.
Some additional CANDU design features and some consequences of the design decisions are
explored here:
 The natural-uranium lattice has low excess reactivity because the fuel is not en-
riched. This is a safety advantage because it limits the amount of reactivity increase
available in an accident. It is also another rationale for adopting a pressure-tube de-
sign which allows on-power refuelling of the reactor to sustain the core reactivity
without shutting down the reactor. Otherwise, batch refuelling of a CANDU reactor
in a pressure vessel would be required every few months, which would lower its ca-
pacity factor on account of the lost operating time during refuelling and during the
associated, time-consuming, shutdown/cool-down and start-up/heat-up transients.
 CANDU NU fuel achieves a discharge burnup in the ~7,500–~9,500 MW.d/Mg(U)
range, depending on the specific reactor design. This is much lower than the dis-

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 49

charge burnup in LWRs (~30,000–50,000 MW.d/Mg(U)). However, considering fuel


enrichment for LWRs (about 3.5%–5% 235U) and the depleted-uranium tailings which
are discarded, CANDU reactors are more efficient than LWRs and extract ~25% more
energy per Mg of mined natural uranium.
 The pressure-tube design means that the main reactor vessel (the calandria) is not
pressurized. The moderator is insulated from the hot coolant by placing a calandria
tube concentric with each pressure tube, with an insulating gas in the intermediate
(annulus) space. The moderator can be maintained relatively cool at ~70°C by circu-
lating it in its own heat-exchanger circuit, which means that more of the neutrons
will have lower thermal energies than in an LWR and consequently will more readily
induce fission in fissile nuclides, e.g., 235U and 239Pu.
 CANDU reactivity devices are located interstitially between pressure tubes and are
therefore in a benign (low-temperature and unpressurized) environment, which is a
safety advantage. CANDU reactivity devices are not subject to being ejected from
the reactor by coolant pressure.
 Heavy water is used as coolant instead of light water in all operating CANDU reactors
for additional neutron economy. Light water was used as coolant in the Gentilly-1
prototype, and organic coolant was used in the Whiteshell Reactor (WR-1) engineer-
ing test unit, both of which are no longer in operation. Light water is also used as
coolant in the proposed Advanced CANDU Reactor (ACR) design, to reduce capital
costs associated with heavy water, and also in the Supercritical-Water-Cooled Reac-
tor (SCWR) design.
 All operating CANDU reactors have horizontal pressure tubes (the Gentilly-1 proto-
type and the WR-1 had vertical pressure tubes). This orientation promotes symme-
try because coolant can be circulated in opposite directions in alternate tubes (i.e.,
using bi-directional coolant flow), making average neutronics conditions essentially
identical at the two ends of the reactor (unlike the situation in LWRs). With vertical
pressure tubes, the gradient in coolant temperature and density as the coolant picks
up heat and moves upward makes the neutronic and thermal-hydraulic conditions
asymmetric.
 Horizontal pressure tubes also facilitate bi-directional refuelling (i.e., fuelling in op-
posite directions in adjacent channels), which further promotes symmetry. When
the fuel is manufactured as short (~50-cm-long) fuel bundles which can easily be
pushed through the pressure tube from either end, the associated fuelling machines
will also be short, less cumbersome, and require less radiation shielding. With verti-
cal tubes, the entire fuel string (consisting of a vertical stack of perhaps 12 or 13
bundles) would have to be removed from the top of the reactor, so that a very long
fuel assembly would have to be used (together with a long and heavy transfer flask),
or short fuel bundles would have to be connected to a common stringer.
 Whereas the loss of moderator shuts any (thermal) reactor down, the loss of coolant
in a CANDU reactor does not have the same effect because the coolant is separated
from the moderator. Instead, the loss of coolant in operating CANDU reactors gen-
erates a positive reactivity insertion (called the coolant void reactivity, or CVR); this is
due to a number of spectral effects which change mostly the fast-fission factor and
the resonance-escape probability. As a consequence, a loss-of-coolant accident in
CANDU results in a positive power pulse, which must be quickly turned around by

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
50 The Essential CANDU

shutdown-system action to avoid overheating the fuel. In fact, the postulated large-
loss-of-coolant accident (LLOCA) scenario is the reason for the adoption of redun-
dancy in CANDU shutdown capability design. CANDU reactors have two fast shut-
down systems, which are physically and logically independent of one another and
each fully capable of shutting the reactor down from any credible configuration.
Shutdown system 1 consists of cadmium shut-off rods which fall under gravity (ini-
tially spring-assisted) into the reactor from above. Shutdown system 2 consists of
the injection of a solution of neutron-absorbing gadolinium (a high neutron ab-
sorber) under high pressure through nozzles directly into the moderator.
 The lattice pitch (distance between the centres of neighbouring tubes) in all operat-
ing CANDU reactors is 28.575 cm. This is not the optimum value in the sense of
maximizing the lattice reactivity (therefore minimizing the refuelling rate and maxi-
mizing the average fuel discharge burnup), which is closer to 34 cm or so. However,
the larger volume of D2O moderator would result in a higher capital cost; the shorter
pitch was selected to minimize the levellized unit-energy cost. A shorter lattice pitch
of 24 cm was selected for the ACR (to reduce moderator cost, among other reasons),
but the reduced moderation would not allow the chain reaction to be self-sustaining
with natural-uranium fuel, and the ACR would need to use enriched fuel.
 The “workhorse” control devices in all operating CANDU reactors are “liquid” zone
controllers. These control 14 compartments in which amounts of light water (used
for its much higher neutron-absorption cross section than heavy water) can be var-
ied uniformly across all compartments to control reactivity or to shape the power
distribution differentially.

9 CANDU Reactor Physics Computational Scheme


Because of the strong heterogeneity of reactor lattices (see the example of the CANDU basic
lattice cell in Figure 19) and because nuclear fuel is a strong neutron absorber, the diffusion
equation cannot be used in reactors based on the detailed lattice geometry. For cases where it
is not desired to apply neutron transport in full-core reactor models, a two-stage or three-stage
process has been developed to enable calculation of the neutron flux and the power distribu-
tion using the diffusion equation. This process is illustrated here for CANDU reactors.

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 51

Figure 19 CANDU basic lattice cell


The computational scheme for CANDU neutronics consists of three stages. Separate computer
programs have been developed to perform the calculations corresponding to each stage. The
three stages are:
 lattice-cell calculation
 reactivity-device (supercell) calculation
 finite-core calculation.

9.1 Lattice Calculation


The first stage involves solving for the flux distribution in the basic lattice cell, which consists of
the fuel, coolant, pressure and calandria tubes, and moderator, but no reactivity devices (see
Figure 19). This calculation is performed using a deterministic multi-group neutron transport
code such as WIMS-AECL. Then the transport code averages the nuclear properties (for absorp-
tion, moderation, fission, etc.) over the cell according to the calculated reaction rates in each
sub-region of the lattice cells to determine the effective neutronic properties of the whole
lattice cell (i.e., few (typically 2)-group macroscopic cross sections and diffusion coefficients).
This averaging (homogenization) of the properties of the basic lattice cell dilutes the strong
absorption of the fuel with the much weaker absorption in the moderator. With these ho-
mogenized properties, Fick’s Law becomes a good approximation over most of the reactor
model, and the diffusion equation can be used to calculate the full-core neutron-flux and power
distributions. The lattice properties govern the neutron-multiplying behaviour of the reactor
lattice, and therefore they must be obtained for all ages of the reactor fuel, i.e., the transport
code must perform a “depletion” calculation to evolve the lattice-cell neutronic properties to
reflect the changes in the nuclide composition of the fuel with irradiation/burnup. The lattice
properties must also reflect any modified configurations of the lattice. For example, if the
coolant is assumed “voided” in a hypothetical loss-of-coolant accident, or the temperature of
the material in a sub-region is assumed to change, then the lattice properties must change
accordingly.

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
52 The Essential CANDU

9.2 Reactivity-Device Calculation


This type of calculation determines the effect of a reactivity device on the nuclear properties in
its vicinity. This effect is expressed in the form of “incremental” cross sections which are added
to the (unperturbed) homogenized properties of neighbouring lattice cells in a modelled vol-
ume around the device, called a “supercell”; see Figure 20.
For instance, shutoff-rod incremental cross sections dictate the local efficacy of the shutoff rods
in absorbing neutrons and shutting down the fission chain reaction. The incremental cross
sections of a device are obtained by calculating the differences in the supercell homogenized
properties between the case when the device is inserted into the supercell and the case when it
is withdrawn from it. The incremental cross sections are added to the lattice-cell cross sections
in the modelled volume around the device’s position in the reactor core, when that particular
device is inserted.

Figure 20 Supercell for calculating device incremental cross sections


Because the supercell also contains highly absorbing material, transport theory must again be
used. CANDU reactivity devices are perpendicular to the fuel channels. Therefore, realistic
supercell models are three-dimensional, and as a result supercell calculations are best done
using the DRAGON transport code, which can perform calculations in three dimensions.

9.3 Full-Core Calculation


In the final stage, a full-core reactor model is assembled. The full-core model must incorporate,
for each cell in the reactor, the homogenized unperturbed-lattice-cell cross sections which apply
there, together with the incremental cross sections of nearby reactivity devices in their appro-
priate locations. The model therefore consists of homogenized basic-lattice cells on which are
superimposed homogenized subcells representing reactivity devices. One of many such “device
model areas” is shown in Figure 21; in the white subcells, the nuclear cross sections are ob-
tained by summing the basic-cell properties and the appropriate device incremental properties.
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 53

A full, 3-D reactor model constructed in this way is used to calculate the three-dimensional flux
and power distributions in the core. Because homogenized properties are used, few-group
diffusion theory can be applied to the full-core reactor model.

Figure 21 Simple model with superimposed reactivity device

10 Evolution of Lattice Properties


When nuclear fuel is “burned” in a reactor, changes occur in the material composition of the
fuel. These changes are generally called “fuel (isotopic) depletion”. In this section, we look at
how the properties of the CANDU lattice evolve with fuel irradiation (or burnup).
The following changes occur cumulatively in CANDU nuclear fuel with time:
 The 235U depletes (i.e., its concentration, which starts at 0.72 atom% for fresh natural
uranium, decreases).
 Fission products accumulate; most of these are radioactive, and many have a signifi-
cant neutron-absorption cross section.
239
 Pu is produced by neutron absorption in 238U and two subsequent beta decays:
238
U  n  239 U*  239 Np  
239
Np 239 Pu  .
235
 Pu participates strongly in the fission chain reaction (because it is fissile, like 235U),
but while it continues to be created at about the same rate from 238U, its net rate of
increase slows.
 Further neutron absorptions lead from 239Pu to 240Pu (non-fissile), and then to 241Pu
(fissile).
 Other higher actinides are also formed (e.g., curium, americium).
 The total fissile fraction in the fuel (235U + 239Pu + 241Pu) decreases monotonically.
The evolution of the concentration of these three nuclides is shown in Figure 22.
A typical graph of reactivity ∞ of the infinite CANDU lattice versus fuel irradiation, obtained
from a cell calculation using a transport code, is shown in Figure 23. The following points are of
note:

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
54 The Essential CANDU

 The fresh-fuel infinite lattice (where the fuel has not yet received any irradiation) has
a high reactivity (~78 mk when the 135Xe and other saturating fission products have
built up). To achieve a steady state in the infinite lattice with fresh fuel, a corre-
sponding amount of negative reactivity must be added to the lattice [e.g., by dissolv-
ing a neutron poison (i.e., a material with a large neutron absorption cross section)
in the moderator] to suppress the initial supercriticality.
 The reactivity starts to decrease immediately on account of 235U depletion.
 It then starts to increase for a while, on account of production of 239Pu, which is
slightly more effective than 235U. Note the slight delay due to the 239Np ~2-day half-
life.
 However the rate of increase of reactivity slows (because the net rate of plutonium
production decreases), and the reactivity proceeds through a maximum, called the
plutonium peak, with increasing burnup (note that this is not a peak in 239Pu concen-
tration, but in lattice-cell reactivity!).
 Following the plutonium peak, the reactivity decreases monotonically on account of
the continuing depletion of 235U and the continuing accumulation of fission products.
 The infinite lattice reaches zero reactivity at an irradiation corresponding to a burnup
of ~6,700 MW.d/Mg(U).
 A homogeneous infinite lattice with fuel beyond that burnup would be subcritical.

Figure 22 Evolution of fuel isotopic densities

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 55

Figure 23 Infinite-lattice reactivity vs. irradiation


However, remember that the infinite lattice does not experience reactivity losses due to neu-
tron leakage; moreover, it does not account for neutron absorption in all the reactivity devices
within the core. Consequently, a homogeneous reactor with all fuel at the same irradiation
would reach zero reactivity at a much lower burnup than the infinite lattice.
In the CANDU-6 reactor, leakage is about -30 mk, and the average in-core reactivity-device load
is ~-18 mk. Therefore, an estimate of the reactivity eff of a finite homogeneous CANDU-6
reactor can be obtained by subtracting 48 mk from the reactivity of the infinite lattice; see
Figure 24.

Figure 24 Finite-reactor reactivity vs. irradiation


In a homogeneous reactor, a significant amount of additional negative reactivity (~30 mk) is
required to suppress the initial supercriticality. A homogeneous CANDU-6 would reach zero
reactivity at an irradiation corresponding to a burnup of ~4,000 MW.d/Mg(U). A homogeneous
CANDU-6 with fuel beyond that burnup would be subcritical [note that the argument is hypo-
thetical in any case because a homogeneous finite reactor would not remain homogeneous with
burnup]. Therefore, if the CANDU-6 were to be batch-refuelled, the discharge burnup would be
only ~4,000 MW.d/Mg(U).
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
56 The Essential CANDU

However, CANDU reactors are refuelled on-power, and therefore there is always (except near
start of life) a mixture of fresh fuel and fuel with high irradiation. The fuel with high irradiation
has negative “local reactivity”, but this is compensated for by the positive local reactivity of low-
irradiation fuel. The proper mixture of fuel in this inhomogeneous reactor (obtained by the
proper rate of daily refuelling) maintains the reactor critical day-to-day. Physically, the older
fuel does the job of reducing the high reactivity of the young fuel, a job which moderator poison
does in the batch-refuelled reactor. The difference is that whereas the poison is just a “para-
sitic” absorber, the older fuel does provide fissions and therefore additional energy.
The mixture of new and old fuel makes it possible to drive the discharge burnup to a much
higher value than that deduced from the “homogeneous” reactivity curve. We can guess (or
calculate) approximately how far we can drive the exit burnup by determining what value gives
equal “positive” and “negative” areas “under” the reactivity curve [this tells us where the
average eff would be 0]; see Figure 25.
From the figure, we can see that positive and negative areas are equal when the average exit (or
discharge) burnup to which we can take the fuel with daily refuelling is ~7,500 MW.d/Mg(U).
This is almost twice the discharge-to-burnup value attainable in the “batch-refuelled” reactor
and represents quite a benefit provided by on-power refuelling!

Figure 25 Averaging reactivity with daily refuelling

11 Summary of Relationship to Other Chapters


 The Neutron Physics chapter explains the interactions that neutrons have with matter and
defines all the basic quantities needed in the study of the fission chain reaction and the
analysis of reactors.
 The present chapter covers mostly the time-independent neutron-diffusion equation and
the analysis of steady-state (time-independent) neutron distributions, either in the presence
of external sources or in fission reactors. It has not covered time-dependent phenomena,
which are left for the following chapter.
 The Reactor Kinetics chapter covers the time-dependent neutron-diffusion equations and

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 57

studies the phenomena of fast-neutron kinetics, fission-product poisoning, reactivity coeffi-


cients, and others in reactors.

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
58 The Essential CANDU

12 Problems
Problem 1
There is an infinite homogeneous non-multiplying medium with diffusion length 8 cm and
diffusion coefficient 2 cm. There are four isotropic point sources of neutrons on the x-y plane:
 Source S1 = 108 s-1 at (x, y) = (10 cm, 0)
 Source S2 = 108 s-1 at (x, y) = (-10 cm, 0)
 Source S3 = 1010 s-1 at (x, y) = (0, 20 cm)
 Source S4 = 1010 s-1 at (x, y) = (0, -20 cm)

(a) Find the total flux at (x, y) = (0, 0)


(b) Find the magnitude and direction of the total current at (x, y) = (0 , 0)
(c) Find the value and the direction of the total current at (x, y) = (0, -10 cm).

Problem 2
An isotropic point source of strength S n.s-1 is located at the origin of axes in a homogeneous
non-multiplying material. The material is characterized by an absorption cross section a and a
diffusion constant D.
(a) Imagine a sphere of arbitrary radius R centred at the origin of axes. Calculate the integrated
absorption rate of neutrons (per s) within the sphere.
(b) What is happening to the remaining neutrons (the difference between the number emitted
and the number absorbed per s)? Prove this by calculation.

Problem 3
Suppose that you have an infinite plane source of neutrons in the y-z plane (i.e., at x = 0), which
emits a total of N neutrons per cm2 per s (half in the positive and half in the negative x-
direction).
Of course, we can think of the plane source as an infinite number of point sources, one at each
point of the plane. Let us take the emission from each point source as isotropic. If the point
source actually has differential area dA (which can be as small as we want), then the point-
source strength will be NdA neutrons per s.
Calculate the flux at any point x in space by integrating the flux from all point sources (by
symmetry, this will of course be the same for any values of y and z). Show that you get the
formula for the flux from the infinite plane source.
Problem 4
This is an exercise on the quantitative aspects of the neutron cycle.
Refer to the figure below which pertains to a critical reactor. Refer also to the notes in the
figure. You are asked to calculate how many thermal neutrons escape from the reactor per unit
time.
Remember that the two main things that can happen when a neutron is absorbed in the fuel

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 59

are capture (when the neutron is absorbed and a gamma ray is emitted) and fission (when the
neutron is absorbed and fission is induced). Therefore, the ratio of capture to fission, /f, is
an important parameter. To solve the problem, use the data given and find the right sequence
(up and/or down) for filling numbers into the boxes.
Note that numbers need not be exact integers. In each box, retain non-integer numbers to 3
decimal places.

1000 fast  1000 fast neutrons are born from ther-


neutrons mal fissions
 A thermal fission releases 2.38 fast neu-
Fast n from
trons
Fast Fissions
 For the fuel,    /f = 1.059
 30 fast neutrons are produced by fast
Fast Leakage fission
 32 thermal neutrons are absorbed else-
where than in the fuel
Epithermal and  9 fast neutrons escape from the reactor
Fast Absorption  95 non-thermal neutrons are absorbed

Thermal
Leakage

Non-Fuel Thermal
Absorption
Thermal
Absorptions
in Fuel
Thermal
Captures in
Fuel
Thermal
Fissions

Problem 5
A homogeneous, bare cylindrical reactor with extrapolated axial length 5.8 m is critical. It is
operated at a fission power of 900 MW. In one energy group, the reactor material is known to
have  = 2.38, f = 0.0042 cm-1, and D = 1.14 cm a = 0.0099 cm-1.
The leakage is 9.6 mk. [Note: neglect the extrapolation distance.]
(a) Calculate the reactor buckling and the material’s absorption cross section.
(b) Determine the reactor’s extrapolated diameter.
(c) What is the average flux in the reactor?
(d) What is the ratio of the flux on the cylindrical reactor axis at 50 cm from the reactor face to
the maximum flux in the reactor?

Problem 6
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
60 The Essential CANDU

Design the proportions of a cylindrical reactor which minimize leakage [neglect the extrapola-
tion distance].

Problem 7
A research reactor is in the shape of a parallelepiped with a square base of side 5.2 m and a
height of 6.8 m. The reactor is filled uniformly with a fuel of one-group properties f = 0.0072
cm-1 (and  = 2.45) and a = 0.0070 cm-1. The reactor operates steadily at a fission power of 15
MW. The average value of energy per fission Ef = 200 MeV, and 1 eV =1.6*10-19 J. [Neglect the
extrapolation distance.]
(a) What is the value of the diffusion coefficient?
(b) What is the average value of the neutron flux?
(c) What is the maximum value of the neutron flux?
(d) At what rate is the fuel consumed in the entire reactor (in nuclides.s-1) and at the centre of
the reactor (in nuclides.cm-3.s-1)?

Problem 8
A critical homogeneous reactor in the shape of a cube loses 4% of produced neutrons through
leakage.
a) Calculate k for an infinite reactor made of the same material.
b) The initial reactor is re-shaped into a sphere. Calculate the new keff .

NOTE: Use one-group diffusion theory and ignore the extrapolation length.

Problem 9
A reactor is made in the shape of a cone, with the base radius equal to the height and both
equal to 3 m. The reactor is homogeneous, with the following properties:
f = 0.002 cm-1
a = 0.0018 cm-1
Calculate the number of neutrons leaking out of the reactor per second knowing that the
reactor is critical and that the average flux in the reactor is 1013 n/cm2/s.

Problem 10
Imagine we have nuclear material with the following properties in two energy groups:
a1 = 0.0011 cm-1, 12 = 0.0068 cm-1, a2 = 0.0043 cm-1, f2 = 0.00528 cm-1
D1 = 1.07 cm, D2 = 0.92 cm.

(a) Calculate the reactivity of an infinite lattice made of this material.


©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 61

(b) In this infinite lattice, what fraction of neutrons is captured in resonances?


(c) We are asked to make a critical homogeneous reactor with this material, cylindrical in
shape, with a diameter 10% greater than the axial dimension [neglect the extrapolation dis-
tance]. What will be this reactor’s dimensions?
(d) Calculate the fast and thermal non-leakage probabilities for this reactor. Also, what is the
total leakage in mk?
(e) What is the ratio of the group-1 flux to the group-2 flux in the reactor?

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
62 The Essential CANDU

13 Appendix A: Reactor Statics


In this Appendix, we derive Fick’s Law.
We copy here the angle-integrated version of the time-independent neutron-transport equa-
tion:
    
S  r , E     E    f  r , E '    r , E '  dE '   s  r , E '  E    r , E '  dE ' 
E' E'

    
t  r , E   r , E     J  r , E   0

(presented earlier as Eq. (16)). Our overall objective is to rewrite the last term (the leakage

term) in terms of the total flux  r, E  only.
 
We first try to obtain an equation for the total current J r , E  by multiplying the transport
equation (7), which depends on angle, by ̂ and integrating over it, yielding Eq. (A.1):
1 
S r, E    ˆ   E
ˆ d   ˆ
 f  r , E '   r ,   
ˆ ' dE ' ˆ d
ˆ 
4 ˆ 4   
ˆ'
', E ' d  
ˆ
 E'  
 
   
    s r , E '  E , ˆ '  ˆ  r , ˆ ', E ' ˆ dE ' d ˆ ' d ˆ 
ˆ 
ˆ ' E'
(A.1)

  ˆ  

t  r , E    r ,   ˆ    J r, 
ˆ d ˆ ,E 
ˆ d 
ˆ 0 
ˆ

,E   ˆ

In the first two terms, which we can call T1 and T2, we used the assumption of isotropy in the
ˆ d
external and fission sources, so that these terms are proportional to the integral   ˆ . This is
ˆ

a vector quantity; for instance, the x component is  ˆ dˆ .


ˆ

x It is clear that this integral (and

similarly for the other components) must have a zero value because there are as many positive
ˆ ' s as there are negative. Therefore, the first two terms drop out of the equation:
x

T1  T2  0.
(A.2)
Let us jump to the fourth term, which by definition is:
 ˆ  
ˆ  J r, 

ˆ d  ˆ  J  r , E .
ˆ , E d  
  r , 
ˆ
, E   ˆ
 

Therefore, the fourth term, which we call T4, becomes:


  
T4   t  r , E  J  r , E 
(A.3)
Let us look at the third term, T3:
 ˆ  r , 
 ˆ '
T3      s r , E '  E , 
ˆ'
ˆ
ˆ 
ˆ d 
ˆ ', E ' d 
ˆ ' dE '  
E' 
 
      r , E '  E , ˆ '  ˆ 
ˆ'
E'  ˆ
s
ˆ ˆ '  ˆ  r , 
ˆ 'd ˆ ', E '  d 
ˆ ' dE '.

ˆ '
where we have introduced into the integral the unity factor  ˆ '  1 . Also, recall that s is not
©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
Reactor Statics 63

ˆ 
a function of absolute angles, but only of the cosine  ˆ '.

ˆ '   r, E '  E , 


 ˆ 
ˆ' 
ˆ  ˆ  r, 
ˆ ' d  
ˆ ', E ' d 
ˆ ' dE '. 
T3  E ' ˆ ˆ
 s
'  
The integral over ̂ can be written as
 
 sa  r , E '  E   2   s  r , E '  E ,    d  ,
1

1 (A.4)
and
  ˆ  
 T3    sa  r , E '  E    r ,   ˆ ' d  ' dE '    r , E '  E  J  r, E '  dE '.

', E '   sa
E' ˆ'
 E (A.5)
Note that if the scattering cross section were isotropic, s would not depend on , and there-
fore sa would be 0. Although this is not a good approximation, one that is more reasonable is

to neglect the energy change in anisotropic scattering, i.e., to assume that  sa  r , E '  E  is a

delta function    E' - E  sa  r , E  , which leads to
  
T3   sa  r , E  J  r , E  ,
(A.6)
where
  
1
 sa  r , E   2   s  r , E ,   d   s  r , E 
1 (A.7)
and  is the average cosine of the scattering angle.
Moving on to the fifth term, T5:
   

ˆ ,E 
T5      J r ,  ˆ      r, 
ˆ d  ˆ , E 
ˆ ˆ d
ˆ.  

ˆ
 ˆ
 (A.8)
To calculate the integral, we adopt again the assumption of weak angular dependence of the
angular flux, which leads to the following approximate expression [Eq. (27)] for the angular flux
in terms of the total flux and the total current expression; see derivation in Section 2.3:

 3J  ˆ
ˆ
  
4
 
4
.
(27)
Substituting this expression into Eq. (A.8), we get
    r, E  3  
T5        ˆ 
J  ˆ ˆ d
ˆ.
  
ˆ
  4 4  (A.9)
Now it can be shown (a reasonable result) that the integral of the product of two components
of ̂ , e.g., ij, is equal to 4/3 if i = j, but 0 if i  j. Similarly, the integral of the product of
three components is 0.
Using these two results, Eq. (A.9) becomes

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014
64 The Essential CANDU

1 
T5     r , E  .
3 (A.10)
Incorporating all these results for the various terms [Eqs. (A.2), (A.3), (A.6), (A.10)], Eq. (A.1)
becomes:
      1 
0  0   sa  r , E  J  r , E   t  r , E  J  r , E     r   0,
3 (A.11)
which yields the final approximation for the total current:
  1 
J r, E      r , E  ,
3  t  r , E    sa  r , E  
(A.12)
which is Fick’s Law, giving for the diffusion coefficient,
 1
D r, E    .
3  t  r , E    sa  r , E  
(A.13)

The quantity in parentheses in Eq. (A.13) is defined as the transport cross section, tr r  .

14 Acknowledgements
The following reviewers are gratefully acknowledged for their hard work and excellent com-
ments during the development of this Chapter. Their feedback has much improved it. Of course
the responsibility for any errors or omissions lies entirely with the authors.
Marv Gold
Ken Kozier
Guy Marleau
Bruce Wilkin
Thanks are also extended to Diana Bouchard for expertly editing and assembling the final copy.

©UNENE, all rights reserved. For educational use only, no assumed liability. Reactor Statics – September 2014

You might also like