You are on page 1of 47

Introductory materials for reactor laboratory exercises at ISIS

reactor at CEA Saclay, France.

Reactor Physics Course at Royal Institute of Technology - KTH

Prof. Waclaw Gudowski

Material has been prepared based on:

“Reactor Physics Course at VR-1 Reactor”

Prepared by: Kolros A., Kropík M., Rataj J.6NOHQNDďŠkoda R. and 0LODQ7ČãtQVNê

Reactor Exercises at VTT

Prepared by: Jarmo Ala-Heikkilä, Tom Serén and Calle Berglöf

“Reactor Exercises at ISIS reactor at Saclay”

Prepared by CEA-team
1. Introduction

THE SACLAY CENTRE is one of the 9 research sites of the French Atomic Energy
Commission (CEA). It is a top-ranking innovation and research centre at the European level.
More than 5000 people work in the centre. The centre is multidisciplinary, with activities in
fields such as nuclear energy, life sciences, material sciences, climatology and the
environment, technological research and teaching.

The ISIS Reactor is called a neutron model of the bigger reactor Osiris. The criticality of
ISIS was reached on April 28, 1966, after four years of construction work and OSIRIS
reached criticality on September 8 of the same year. After two years of OSIRIS operation at
50 MW, the nominal thermal
power of 70 MW was reached in 1968. From 1966 until the start of 1980, the reactor operated
with an U-Al fuel enriched to 93% and from 1980 to 1994 with a UO2 fuel enriched to 7%.
The gradual conversion of the reactor to using U3Si2 Al (called silicide) fuel enriched to
19.75% began in January 1995 and was completed in April 1997.

The core of the ISIS reactor.


ISIS is a reactor with the same geometry as Osiris reactor but a thermal power limited to 700
kW, compared to 70 MW of Osiris. The Isis reactor is particularly dedicated to neutronic
measurements of the reactor and the irradiations. ISIS is used to carry out tests of new core
configurations, new fuel elements or irradiation experiments, by measuring the effects of
reactivity, the levels of neutron flux or gamma heating. The reactor is also used to teaching
and train a large number of trainees.
The reactor is characterized by the possibility of loading the core with irradiated elements
from OSIRIS. It is equipped with an installation enabling it to operate by adjusting a weak
boron concentration in the water of the primary circuit. The control-command of the reactor is
designed to meet the requirements specific to this reactor, such as:
x various approaches under criticalities and divergences,
x fine measurements over a wide range of power,
x the use of detectors in various nuclear environments

Exercises comprise the following activities:

Lab I:
1. Reactor presentation
2. Introduction of the last 4 fuel assemblies in the core

Lab II:
1. Approach to criticality
2. Reactor start up
3. Reactivity effects around criticality
4. Reactivity effect of devices placed in the core

Lab III:
1. Control rod calibration
2. Control rod global worth by the control rod drop method
3. Delayed neutrons precursor effect
4. Temperature effect

The ISIS reactor core loading for the exercises is presented at Figure 1

Figure 1 The ISIS reactor core layout.


B1-B6 denote 6 control (scram) rods
with hafnium as neutron absorber.
Be-denotes the Beryllium reflector at
West-side of the core. Water
surrounding and filling the reactor
core plays a neutron moderator role.

Before reading the specific descriptions of the ISIS exercises it is recommended to study the
following chapters on neutron detection, physics of criticality, delayed neutron fractions and
temperature effects.

2. Neutron Detection

Regarding the operation and use of the research nuclear reactor, the problems of the neutron
detection and detectors are related to the problems of reactor control (measuring the power)
and to the experimental specification of the neutron field parameters (flux density, fluency,
and spectrum). The operation of the experimental and training research reactors is
characterized with frequent and quick changes of the power and with a relatively wide range
of power. Therefore, to control the reactor, the neutron gas detectors are exclusively used.
They enable obtaining immediate information about the status of the reactor. For the
experimental measurement of the neutron field, the gas detectors or the activation detectors
are used, depending on the problems.

2.1. Gas Filled Neutron Detectors


Generally, neutron detectors can be divided into two groups: passive neutron detectors and
active neutron detectors.
The passive neutron detectors - the response of the detector is proportional to neutron flux
LQWHJUDOWKHHYDOXDWLRQLVGHOD\HG§PLQXWHV«KRXUV«PRQWKV
x activation detectors – thin foils or wire,
x thermoluminescent detectors.
The active neutron detectors - a change of neutron flux produce prompt response of the output
signal:
x scintillation detectors,
x VHOISRZHUHGGHWHFWRUV GHOD\§-30s),
x gas filled detectors.

Generally, the gas detectors use the fact that, when the ionizing particle passes through the
detector content, the gas filling is ionized. Thus, the detector acts as a condenser and its
electrodes are under electric voltage. Positively charged ions and the electrons formed are
accelerated with the electric field and they both impinge on the cathode or the anode of the
detector. An electric pulse is created at the detector output and then it is processed.
Depending on the structure and the geometry of the detector, supply voltage and kind and
pressure of the gas filling, three types of detectors are distinguished: ionization chamber,
proportional detector, Geiger-Müller detector and corona detector. If the response of the
detector is proportional to the energy of the impinged particles, it is said that the detector has
spectrometric properties.
The detector can be connected in a pulse or in a current mode. In the pulse mode, the number
of pulses is proportional to the number of impinged particles. In the current mode, the current
response is measured. The size of it is proportional to the charge that is formed with the
ionization during the time unit. The current mode (Figure 2) is advantageous at high rates of
impinged particles when the individual pulses are overlapped. A general diagram of the gas
detector connection in the pulse mode is shown in Figure 3Error! Reference source not
found..

+ HV + HV

Rp
pA Amplifier

CV

Detector Detector

Figure 2. Schematic diagram of the Figure 3. Schematic diagram of the


gas filled detector (current mode). gas filled detector (pulse mode).

Note: sometimes the third possible mode – Mean Square Voltage Mode (Fluctuation,
Campbell) - is used to analyze the noise fluctuation of the direct current from the detector. It
is suitable only for fission chambers which are employed in reactor power instrumentation.
Neutrons belong among indirectly ionizing particles. The interaction of neutrons with
common materials is poor; the cross section is very low. The neutron detection by gas
detectors comes from the fact that when neutrons pass through a suitable material, secondary
charged particles are formed and cause ionization. Therefore, the walls of the detectors are
covered by the 10B isotope or the 235U isotope. The detector can be filled with a special gas
too – for example BF3, 3He, or 1H.
To detect thermal neutrons, the reactions below are frequently used:
B  01n o 7 m3 Li  24D
10
5 Q 2,31MeV ( 94%) , (1)
7m
3Li o 37Li  J EJ 0,48 MeV , (2)
10
5 B  01n o 37Li  24D Q 2,79MeV (6%) . (3)
Two ionizing particles are emitted in opposite directions with kinetic energy of
ELi ~ 0.84 MeV and EĮ ~ 1.47 MeV respectively. The non-uniqueness of the reaction
10
% QĮ 7 Li (94% to the excited state of the 7mLi nucleus and 6% to the ground state) and the
rather high energy of the reaction prevent using the detection for common spectrometric
purposes. The 10B isotope has a large cross section, 3840 .10-28 m2, in the thermal neutron area
which decreases with increasing energy of the neutrons according to 1/v. In nature, boron is
represented with the 10B (19%) and 11B (81%) isotopes.
1,0E+06

1,0E+05
5315,7 U235
3839,4 B10
m]

1,0E+04
2

He3
-28

thermal cross section


Neutron Cross Section [10

1,0E+03
585,3
1,0E+02
thermal neutron energy 0,025 eV

1,0E+01

1,0E+00

1,0E-01

1,0E-02
1,0E-05 1,0E-03 1,0E-01 1,0E+01 1,0E+03 1,0E+05 1,0E+07 1,0E+09
En [eV]

Figure 4. Cross-section neutron energy relationship of selected interactions.

For detecting thermal neutrons fission chambers with the isotope 235U are also used. In a
fission chamber, a thin layer ~ 1 mg/cm2 of 235U (enriched to > 90%) is deposited onto the
inner surface of the cathode cylinder. The formed fission fragments, FP1 and FP2 have a large
ionization effect, as heavy ions with average kinetic energies of ~95 MeV and ~65 MeV are
produced.
U  01n o
235
92
236
92 U o A1
FP1  A2 FP2  Q . n ... Q 195 MeV . (4)
For the neutron detection and for the fast neutron spectrometry (En > 1 MeV) with
proportional detectors, the following nuclear reaction is used:
3
2 He  01n o 13 H  10p  0.765 MeV . (5)
The course of this reaction is unique and this reaction has a favourable reaction energy. The
cross section for the thermal neutrons is 5330.10-28 m2 and decreases proportionally to 1/v.
Although sensitivity of the neutron detectors to the gamma radiation is small, the influence on
the output response is not insignificant. However, using the amplitude discrimination, the
neutron response can be distinguished from the detector noise and from the accompanying
radiation that has small ionization losses (e.g. gamma and electrons) compared to ĮSDUWLFOHV
or fission fragments. The differential pulse height characteristic of a corona detector of the
SNM type with optimal setting of the pulse height discrimination for elimination of gamma
and noise influence is shown in Figure 5. Another important basic characteristic of the
detector is depicted in Figure 6. A region of minimum slope of integral distribution is defined
as a counting plateau and it represents the area of operation in which minimum sensitivity to
HV supply drift is achieved.
40000
10000

8000 30000
counting plateaus

Number of Counts
6000
Number of pulse

20000

4000
discrimination level

noise + gamma neutrons 10000

2000
optimal operating point

0
0 0 250 500 750 1000 1250 1500 1750 2000 2250
0 50 100 150 200 250
HV [V]
Pulse height [channel]

Figure 5. Differential pulse height Figure 6. Counting characteristic of the


characteristic of the SNM 10 neutron SNM 10 neutron corona detector.
corona detector.

3. Delayed Neutrons Detection

3.1. Introduction
The great majority (over 99%) of the neutrons produced in fission are released within very
short time after the actual fission event. These are called prompt neutrons. A small portion of
neutrons are produced with some delay after the fission process. These are called delayed
neutrons.

Although the delayed neutrons comprise only a very small part of the total number of
neutrons generated from fission, they play an important role during the nuclear reactor control
and operation. The reactor operation would become impossible without their existence. In
addition to the influence on a reactor control and operation the delayed neutrons are important
for probe and analyses of fissionable materials tool.
The exercises described in this chapter are focused on the delayed neutrons detection. Some
basic parameters of the delayed neutrons will be determined in the first part of the
measurement. The second part demonstrates the delayed neutrons utilisation for determining
the uranium content in an unknown sample.

3.2. Prompt and Delayed Neutrons Generation


In the fission reaction the incident neutron enters the heavy target nucleus, forming a
compound nucleus that is excited to such a high energy level (Eexc > Ecrit 1) that the nucleus
"splits" (fissions) into, usually, two large fragments plus some neutrons.
The primary fission fragments 2 directly after fission are unstable, because of neutrons excess.
Their transition into the more stable state (before eventually undergoing a subsequent beta-
decay) is accompanied by the release of one, two or more neutrons (depending on the mother
nucleus, e.g. for the 235U the average number of emitted neutrons is Ȟp3 = 2.43 per fission
event). Generation time of these neutrons – being identical to the generation of the fragments
- is 10-14 y 10-12 s. In relation to this generation time, these neutrons are called prompt
neutrons. The energy of the prompt neutrons ranges approximately from hundreds eV to ten
MeV. The average prompt neutrons energy is E 2 MeV .
It is necessary to note that there are many ways for the compound nucleus to fission. An
example of a typical fission reaction is shown below

1
0 n  235
92 U o 236
92
*
U o 140
55 Cs  37 Rb  3 0 n .
93 1

The number of possible fission products is very large. Figure 7 shows the typical mass
distribution of the fission products in the case of 235U thermal neutron fission.
After the emission of the prompt neutrons usually there is no direct neutron emission
anymore, the fission products undergo several successive E  decays to reduce their neutron-
excess. However, in some cases a daughter nucleus is formed after a E  decay, where the
excitation energy is higher than the neutron-separation energy. This nucleus will emit again a
neutron, nearly promptly after its formation. As these neutrons are generated from several
milliseconds to several minutes after the fission event, they are called delayed neutrons. The
delayed neutron generation can be described as follows:
E A 1
A
Z X N o Z 1YN 1 o Z 1TN  2  n ,
A
(1)

where
A – is mass number,
Z – is number of protons,
N – is number of neutrons.
The X nucleus is called delayed-neutron precursor, the Y nucleus is called delayed neutron
emitter. The time of delayed neutrons generation depends on the half-life of the precursor

                                                                                                                         
1
The measure of how far the energy level of a nucleus is above its ground state is called the excitation energy -
Eexc. For fission to occur, the excitation energy must be above a particular value for that nuclide. The critical
energy Ecrit is the minimum excitation energy required for fission to occur.
2
  The  nuclei  that  are  formed  just  after  the  fission  process  (within  10-­‐14s)  are  called  fission  fragments.  These  fast  
moving  nuclei  slow  down  by  colliding  with  the  atoms  of  the  fuel  material,  then  pick-­‐up  electrons,  and  finally  
become   neutral   atoms.   Since   they   are   radioactive,   they   undergo   several   decay   processes,   and   form   other  
atoms.  These  latter  are  called  fission  products.  
3
The value Ȟp is called prompt neutrons yield and represents average number of prompt neutrons released in a
fission event.
nucleus X. An example of delayed neutron precursor and delayed neutron generation is shown
in Figure 8.

Figure 7. Cumulative fission yield versus mass number for thermal fission of 235U
(data from JEFF 3.1. – Joint Evaluated Fission and Fusion File).

Figure 8. Origin of delayed neutrons from 87Br.


As the excitation energy of the delayed-neutron emitter nucleus is usually much lower than
the excitation energies of the direct fission fragments, the average energy of the delayed
neutrons is also smaller (200÷600 keV) than that of the prompt neutrons (2 MeV). Therefore,
in a thermal system, the delayed neutrons are more efficient in causing further fission than the
prompt neutrons since the delayed neutrons have smaller probability for absorption in 238U
during slowing-down process.

3.3. Delayed Neutron Groups and Their Properties


A large number of delayed neutron precursors with various half-life were identified so far. A
correct treatment of the delayed neutrons in reactor dynamics calculations would consider
each precursor with its own half-life and yield. However, there are two problems for
proceeding this way:
x The calculation scheme becomes complicated because of the large number of the
precursors,
x The basic properties, i.e. the decay scheme, half-life and yield are not well known for
every precursor.
For that reason, it is convenient to combine the known precursors into groups with
appropriately averaged properties. For most applications, the delayed neutron precursors and
delayed neutron properties are divided into six groups. The delayed neutrons are characterised
by the half-life (T1/2) and/or decay constant (Ȝ), average energy ( E ), yield (Ȟd) and by the
fraction (ȕ). Delayed neutron fraction is related to the total yield of neutrons from fission

Qd
E , (2)
Q tot
where

Ȟd – is delayed neutrons yield,

Ȟtot – is total neutrons yield from fission (i.e. Ȟtot = Ȟp + Ȟd).


The same formula as (2) can be written for each delayed neutrons group:

Q di
Ei , (3)
Q tot
where Ȟdi is delayed neutrons yield of ith group.
Table 1 lists the characteristics for six precursor groups (delayed neutrons groups) resulting
from thermal fission of 235U.
Table 1. Delayed neutron characteristics for thermal fission of 235U [8], [10].
Group Half-life Decay constant 4 Average energy Yield Ȟi Fraction

T1/2 [s] Ȝi [s-1] E [keV] [n/fissions] ȕi [%]

1 55.72 0.0124 250 0.00052 0.021

2 22.72 0.0305 560 0.00346 0.140

3 6.22 0.111 430 0.00310 0.126

4 2.30 0.301 620 0.00624 0.252

5 0.610 1.14 420 0.00182 0.074

6 0.230 3.01 - 0.00066 0.027

total 0.01580 0.64

The delayed neutrons parameters vary considerably depending on the fissile material in use.
For example, the delayed neutron fraction is 0.0064 in the case of 235U fission and 0.0021 in
the case of 239Pu fission.
As mentioned in the previous chapter, the delayed neutrons are more efficient in causing
further fission than the prompt neutrons considering theirs smaller energy. This effect is
described in the reactor theory by the parameter which is called the effective delayed neutron
fraction (ȕeff). The parameter ȕeff is defined as the fraction of neutrons at thermal energies
which were born delayed. The effective delayed neutron fraction is the product of the average
delayed neutron fraction and the importance factor

E eff E ˜I , (4)

where

E – is average delayed neutron fraction,

I – is importance factor.
In a small reactor with highly enriched fuel, the increase in fast non-leakage probability will
dominate the decrease in fast fission factor, and the importance factor I will be greater than
one. In a large reactor with low enriched fuel, the decrease in the fast fission factor will

                                                                                                                         
ln( 2 )
4
The values are calculated from the half-lives through the relation: Oi .
T1 / 2
dominate the increase in the fast non-leakage probability and the importance factor I will be
less than one (e.g. about 0.97 for a commercial PWR).

3.4. Influence of Delayed Neutrons on the Reactor Dynamics


The influence of the delayed neutrons on the reactor dynamics can be shown by following
analyses of reactor period and neutron lifetime.
Let’s assume that it is possible to describe the reactor time behaviour by a kinetic equation
with no influence of delayed neutrons:

dn keff ( t )  1
n( t ) , (5)
dt A
where

n – is neutron density,

keff – is multiplication factor,

Ɛ – is prompt neutron lifetime 5.


Supposing that keff is constant in time, than the solution of equation (5) is:
'k eff
˜t
n( t ) n0 ˜ e A
, (6)

where n0 is the initial neutron density and ǻkeff = keff – 1.


Under some simplifying assumptions, the neutron density in equation (6) can be substituted
for the reactor power P
'k eff
˜t
P( t ) P0 ˜ e A
. (7)

The reactor period T is defined as the time required for reactor power to change by a factor of
"e," where "e" is the base of the natural logarithm and is equal to about 2.718. The reactor
period is usually expressed in units of seconds. From previous definition and equation (7) it is
possible to determine an expression for the reactor period
'k eff
˜T
P( t ) P( t ) A
if e then e e A
and T , (8)
P0 P0 'keff

The relationship between reactor power P and reactor period T is expressed by equation
t

P( t ) P0 ˜ e . T
(9)

                                                                                                                         
5
Depending on the reactor type, the value of prompt neutron lifetime usually ranges between 10-5 and 10-3 s.
The importance of delayed neutrons can be shown by the following simple example focused
on determination of reactor period while only prompt neutrons are considered
Let’s assume the initial reactor state is the critical state (i.e. keff = 1) and reactor operated at
the stationary power (P0 = const), after that the multiplication factor is changed to keff = 1.001
(i.e. ǻkeff = 0.001) by insertion of positive reactivity ȡ §. If only prompt neutrons are
considered it is possible to calculate the reactor period from equation (8)

A 9.97 ˜ 10 5
T | 0.1 s . (10)
'k eff 1 ˜ 10 3

This means that the power increases during one second from initial value P0 to the value
2.104.P0:
1s

P( 1s ) P0 ˜ e 0.1 s
P0 ˜ e 10 # 2 ˜ 10 4 ˜ P0 . (11)

Obviously, there is no control mechanism (containing mechanical parts) that could follow this
very rapid increase. However, the delayed neutrons could (under special circumstances) “slow
down” this pace, and so they provide the control of the reactor.
The average lifetime of delayed neutrons (total delay time) is equal to
6

¦E
i 1
i ˜W i . (12)

where

ȕi – is delayed neutron fraction of the ith group,

IJi – is lifetime of the ith delayed neutrons group. IJi = Ȝi-1


And consequently, the total average lifetime of neutrons (prompt and delayed) is equal to a
sum
6
A A  ¦ E i ˜W i . (13)
i 1

The total delay time for 235U equals 0.0942 s. Prompt neutrons average lifetime is

approximately A # 10 3 y 10 5 s in general and in comparison with the total delay time it is
negligible. Therefore, as a total average lifetime of neutrons, the total delay time can be
considered.
The one-point kinetics equations are used to explain the effect of delayed neutrons

dn t k eff ˜ 1  E  1 6
˜ n( t )  ¦ Oi ˜ ci t , (14)
dt A i 1

dci t k eff
Ei ˜ ˜ n( t )  Oi ˜ ci ( t ) i = 1, 2, ..., 6 (15)
dt A
where

n(t) – is neutron density,


ci(t) – is precursor concentration of the ith delayed neutrons group
kef – is effective multiplication factor,
Ɛ – is prompt neutron lifetime,
ȕ – is delayed neutrons fraction,
ȕi – is delayed neutrons fraction of the ith delayed neutron group.
For simplification, only one (averaged) delayed neutron group is taken into account. The
average half-live of delayed neutrons (in the case of 235U) from Table 1 is approximately 9 s.
The six decay constants can be substituted by one averaged decay constant

log 2
O 0.077 s 1 . (16)
9s
Equations (14) can be modified

dn t k eff ˜ 1  E  1
˜ n t  O ˜c t . (17)
dt A
The first term on the right side expresses the neutron-multiplication due to the prompt
neutrons, whereas the second term shows the effect of the delayed neutrons.
The equation describing the variation of precursor concentration (15) can be rewritten

dc t k eff
E˜ ˜ n t  O ˜ c t . (18)
dt A
The first term on the right side describes the fact that the precursor nuclei are created by the
fission events; the second term describes the decrease due to the decay. The solution of the
equations system (17) and (18) is searched in the form
t t
n t n0 ˜ e T
and c t c0 ˜ e T . (19)

Substituting these solutions into (17) and (18), the reactor period T would be

U A 1 1
˜  , (20)
E Ek eff T 1  OT

k eff  1
where the reactivity is defined as U .
k eff

Equation (20) is a quadratic equation for the period T (supposing constant reactivity):

U 2 ¨§ U AO · A
OT  ¨   1¸T  0. (21)
E ¸
© E k eff E ¹ k eff E
Equation (21) has two roots T1 and T2 and the solution – the time-dependence of the neutron
density - can be found in the following form
t t

n t n1 e T1
 n2 e ,
T2
(22)

where the constants n1 and n2 depend on the initial conditions.


If the reactivity U is negative (i.e. keff < 1, the reactor is subcritical), all three coefficients in
Equation (21) are negative. This means that both roots are negative which implies that the
neutron density – irrespective of the (22) initial conditions – decreases in time. If the
reactivity ȡ is positive (i.e. keff > 1, the reactor is supercritical), the first coefficient in
Equation (21) is positive, the third one is still negative. Depending on the actual value of the
reactivity the second coefficients can be negative or positive. However, in both cases one of
the roots is negative and the other is positive. Denoting the positive root by T1, after
sufficiently long time (t) Equation (22) can be written in following form

n( t ) | n1 ˜ e t / T1 (t >> T2), (23)

since the term with the negative root vanishes. The number of neutrons increases
exponentially in a supercritical reactor even in the presence of the delayed neutrons. However,
the time-constant of this exponential increase is crucial. Assuming the following numerical
values

E = 0.0064; A 9.97 ˜ 10 5 s ; O 0.077 s 1 ;


keff = 1.001, U § 0.001;
and substituting these values into Equation (21), one gets T1 = 70.2 s, which is substantially
larger than the result that we calculated without the presence of the delayed neutrons (0.1 s).
This numerical example shows clearly the essential feature: the delayed neutrons cause an
increase in the reactor period so the safe control of the reactor with external devices is
possible.
However, this statement holds only for small reactivities, more precisely while U/ȕ < 1. If this
holds, the terms containing Ɛ in Equation (21) can be neglected. The positive period can be
written approximately as

E U
T1 | . (U/ȕ < 1). (24)
UO
This approximation gives T1 = 70.12 s in the above numerical example, i.e. the approximation
is sufficiently correct. The situation changes dramatically when U/ȕ > 1. In this case the
coefficient of the second term in Equation (21) becomes positive, and the positive root of the
equation should be approximated by a formula

A
T1 | . (U/ȕ > 1). (25)
k eff U  E
For example, if U/ȕ = 1.1 (keff = 1.0071) according to the formula (25), we obtain
T1 = 0.152 s, which means that the reactor becomes uncontrollable again. These results show
that the reactor is controllable while U/ȕ < 1, i.e. keff < 1+E (approximately). In other words,
the necessary condition of the controllability is that the reactor must be subcritical without the
delayed neutrons. When the reactor becomes supercritical without the delayed neutrons (i.e. if
U/ȕ > 1), the reactor becomes uncontrollable and a reactor excursion occurs. Such a reactor
state is called prompt supercritical state, and it is considered as a severe reactor accident.

4. Reactivity Measurement

4.1. Introduction
Reactivity is one of the most important parameters related to the nuclear reactor dynamics.
The reactivity variations with time and its absolute value have a direct influence on the reactor
operation and safety and therefore they are strictly limited. Following parameters related to
the reactivity are usually determined at the reactor operation:
x Reactor subcriticality (quantity of negative reactivity before the reactor startup),
x Maximal reactivity excess of the reactor core,
x Control rod worth,
x Reactivity changes caused by insertion/removal of a fuel element or some experimental
device to/from the reactor core.

4.2. Reactivity Definition


If there are N0 neutrons in the preceding generation, then there are N 0 ˜ k eff neutrons in the
present generation. The numerical change in neutron population is N 0 ˜ k eff  N 0 . The gain or
loss in neutron population N 0 ˜ ( k eff  1) , expressed as a fraction of the present generation
N 0 ˜ k eff , is the following

N 0 ˜ ( k eff  1)
, (1)
N 0 ˜ k eff

where keff is the effective multiplication factor.


This relationship represents the fractional change in neutron population per generation and is
referred to as reactivity, ȡ. Cancelling out the term N0 from the numerator and denominator in
formula (1), the reactivity is determined as shown in the equation below

k eff  1
U . (2)
k eff

It is obvious from Equation (2) that ȡ may be positive, zero, or negative, depending upon the
value of keff. The larger the absolute value of reactivity in the reactor core, the further is the
reactor from criticality. It may be convenient to think of reactivity as a measure of reactor
deviation from criticality.
Equation (2) clearly implies that the reactivity is a dimensionless number. It does not have
dimensions of time, length, mass, or any combination of these dimensions. It is simply a ratio
of two quantities that are dimensionless.
Example:

Calculate the reactivity in the reactor core when keff is equal to 1.003 and 0.997.
Solution:

The reactivity for each case is determined by substituting the value of keff into Equation (2):

1.003  1 0.997  1
U 0.00299 U 0.00301 .
1.003 0.0097
As shown in the example, the value of reactivity is often a small decimal value. In order to
make this value easier to express, artificial units are defined. By definition, the value for
reactivity that results directly from the calculation of Equation (2) is in units of k/k.
Alternative units for reactivity are % k/k and pcm (percent mȡ). Another unit of reactivity
that is used at some reactors is equivalent to 10-4 k/k. This unit of reactivity does not have a
unique name. In nuclear engineering a unit commonly used to measure reactivity, is the dollar
($) and even its hundredth part - cent. The dollar, is measured in units of the effective delayed
neutron fraction ȕeff, and is defined as

U
reactivity in dollars { .
E eff

The reason for this unit of measure is due to the prompt critical condition that occurs within
the reactor when it is critical on prompt neutrons alone. A disadvantage is the fact that for the
change back to the absolute scale the numeric value of ȕeff is needed. To determine ȕeff is
difficult.

4.3. Point Kinetics Theory


If the thermal power of the reactor is insignificantly small, the influence of thermal changes
on physical properties of the core is not present and such a reactor type is designated as the
zero power reactor. Time behaviour of the system is then described by a solution of the
reactor kinetics equations.

Kinetics equations describe behaviour of a large number of neutrons in the matter where the
neutron scattering, absorption and multiplication occur. Since the cross sections of nuclear
reactions depend on the neutron energy in a complicated way, some simplifications for
description of neutron transport in the matter are required. The most general approach to the
neutron transport is based on the transport theory that respects anisotropy of neutron
scattering. If the angular distribution of the neutron velocity vector is isotropic or nearly
isotropic or it depends neither on the energy nor the location, then the neutron behaviour can
be described by diffusion theory.
Next simplification is related to the spatial dependency of the observed quantity. In the
majority of the cases point-kinetic model is used that enables to describe the time behaviour of
a reactor by a system of ordinary differential equations of the first order. Besides the spatial
effects also the whole process of neutron slowing down is neglected.

Using the diffusion equation, it is possible to derive the kinetic equation with NO influence of
delayed neutrons

dn k ef ( t )  1
n( t ) . (3)
dt A
The quantity A is the neutron lifetime (prompt neutron lifetime in this case).
The influence of the delayed neutrons on the reactor behaviour cannot be neglected except for
large reactivity changes (keff – •ȕ). Delayed neutrons are produced by radioactive decay of
certain fission products (precursors). For that reason, it is necessary to include the delayed
neutron production into the kinetic equation (3) and describe the time variation of precursor
concentration. If six groups of delayed neutrons are assumed, reactor kinetics can be
described by a system of seven differential equations

dn k eff ˜ 1  E  1 6
˜ n( t )  ¦ Oi ˜ ci t , (4)
dt A i 1

dci k eff
Ei ˜ ˜ n( t )  Oi ˜ ci ( t ) i = 1, 2, ..., 6
dt A
or rewritten using the reactivity

dn UE 6
˜ n( t )  ¦ Oi ˜ ci t , (5)
dt / i 1

dci Ei
˜ n( t )  Oi ˜ ci ( t ) i = 1, 2, ..., 6 (6)
dt /
where

n(t) – is neutron density,


ci(t) – is precursor concentration for the ith delayed neutrons group
keff – is effective multiplication factor,
Ɛ – is neutron lifetime,
ȕ – is delayed neutrons fraction,
ȕi – is delayed neutrons fraction for the ith delayed neutron group,
ȡ – is reactivity,
A
ȁ – is mean neutron generation lifetime ( / ).
k eff
The above mentioned equations are frequently called one-point kinetics equations and fully
describe the time behaviour of a zero power reactor.
4.4.   The  Rod  Drop  and  the  Source  Jerk  Methods  

The Rod Drop method is based on a dynamics study of a critical system response to the
negative reactivity insertion (e.g. by a rod drop to the core). The Source Jerk method is based
on a dynamics study of a subcritical system with an external neutron source and its response
to the source jerk. The Rod Drop method is usually used for determination of control rods
worth, while the Source Jerk method is used for determination of the absolute value of the
subcritical reactor reactivity and for determination of control rods worth as well.
Although both are based on a different initial reactor state (critical or subcritical), they are
related to each other and both are derived from the one-point equations (5) and (6). The
quantity describing the external neutron source (S(t)) is added to the equation (5) in the case
of the Source Jerk method. The analyses come out from study of reactor behaviour after a
dynamic change (rod drop or source jerk). Equation (5) is initially integrated from 0 (time of
the dynamics change by rod drop or source jerk) to f

k eff E
f f f f f
k eff - 1 6

³ dn(t) =
0 A
³ n(t)dt -
0 A
³ n(t)dt + ¦ O i ³ ci ( t )dt + ³ S ( t )d ,
0 i=1 0 0
(7)

f
k eff E i f f

³ dci ( t )
0 A
³ n( t )dt  Oi ³ ci ( t )dt ,
0 0
(8)

and let’s assume that


n(t) o 0 and ci(t) o 0 for t o f,
S=0 after a source jerk or in the case of a rod drop,
f f

³ dn(t) = n( f ) - n(0) and ³ dc (t) = c ( f ) - C (0) .


0 0
i i i

Integrating one-point kinetics equation, using the previous assumption, one obtains
k eff E
f f f
k eff - 1 6
 n( 0 ) =
A
³ n(t)dt -
0 A
³ n(t)dt + ¦ O ³ c ( t )dt ,
0 i=1
i
0
i (9)

k eff E i f f
 ci ( 0 )
A
³ n( t )dt  O ³ c ( t )dt .
0
i
0
i (10)

Substitution of the last two expressions on the left side of equation (9) for equation (10) leads
to
f
k eff - 1 6
 n( 0 ) =
A
³ n(t)dt + ¦ ci ( 0 ) .
0 i=1
(11)

The parameters Ci(0) are determined from equation (6) for t=0
d ci (0) k eff E i k eff E i
=0= n( 0 ) - O i ci ( 0 ) Ÿ ci ( 0 ) = n( 0 ) , (12)
dt A Oi A
Replacing the parameters Ci(0) in equation (11) by formula (12) re-arranges (11) to the
following form
§ A 6
Ei ·
¨ ¸n( 0 )
¨¨  ¦ ¸¸
k eff  1 k eff i 1 Oi ¹
© f
, (13)
k eff
³ n( t )dt
0

and one obtains the formula for reactivity in the units of ȕ (or ȕeff)
§ A 6
E ·
n( 0 ) ˜ ¨ +¦ i ¸
¨ ¸
U © k ef E i=1 E O i ¹
= f
. (14)
E
³ n(t)dt
0

Equation (14) is the basic formula for determination of the absolute value of reactivity by the
Rod Drop and the Source Jerk methods. The expression in parenthesis can be simplified and
replaced by a constant, because it contains only physical parameters invariable during the zero
power rector operation.
A 6 E effi
The value of is approximately equal to 10-2 s, while the value of ¦ is about
k effo E ef i=1 E eff O i
A
10 s thereby it is possible to neglect .
k effo E ef
Finally, the formula for measurement of reactivity is
U 13.01 ˜ n( 0 )
= f . (15)
E
³ n(t)dt
0

Typical time evolution of the neutron rate is stated in Figure 9 for the case of the Rod Drop
(Source Jerk) method.

Figure 9. Course of the neutron rate during the reactivity measurement


by the Rod Drop/Source Jerk method.
4.5. The Positive Period Method
The method utilises period measurements in a slightly super-critical region. By measuring the
reactor period following a reactivity change, the reactivity change is determined.
Equations (6) and (7) may be solved for the case of an initially critical reactor in which the
properties are changed at t=0 in such a way that a reactivity ȡ0 is introduced and stays
constant over time. Using the Laplace transform, or equivalently assuming an exponential
time dependence e-ȦW, the equations for the time-dependent parts of n and ci are
UE 6
  Zn( Z ) ˜ n( w )  ¦ Oi ˜ ci Z  n0 ,   (16)  
/ i 1

Ei
  Zc i ( Z ) ˜ n( Z )  Oi ˜ ci ( Z )  ci 0      i  =  1,  2,  ...,  6   (17)  
/
which can be reduced to
f ( Z , n0 , ci 0 )
n( Z ) , (18)
g( Z )
where
§ 6
Ei ·
g( Z ) U0  Z¨ /  ¦ ¸. (19)
¨ 1 Z O
¸
© i
i ¹

The roots of J Ȧ = 0 determine the time dependence of the neutron and precursor
populations. J Ȧ   is seventh-order equation and is better known as the inhour equation
§ 6
Ei ·
U0 Z¨ /  ¦ ¸. (20)
¨ 1 Z O
¸
© i
i ¹

The solution of the equation is graphically indicated in Figure 10.

ª 6
º
Figure  10.  Plot  of  the  function   U ( Z ) Z «/  ¦ E i / Z  Oi » .  
¬ i 1 ¼
The left-hand side of the equation (20), ȡ0, would plot as a straight horizontal line and the points
at which it crosses the right-hand side are the roots of the equation (20). For ȡ0 > 0 there are one
positive and six negative roots (for ȡ0 < 0 all roots are negative).

The solution for the time dependent neutron flux is of the form
6

¦A e
Z jt
n( t ) j , (21)
j 0

where Ȧj are the roots of g(Ȧ)=0.


After a sufficient time, the solution will be determined by the largest root Ȧ0

n( t ) # A0 e Z0t A0 e t / T , (22)

where T Z01 is referred to as the asymptotic period (Tas). Therefore, measurement of the
asymptotic period allows determining the reactivity

§ ·
¨ ¸
1¨ 6
Ei ¸.
U0 / ¦ (23)
¨ i 1 1
¸
T¨  Oi ¸
© T ¹

5. Control Rod Calibration


5.1. Introduction
Most reactors contain control rods that are used to provide precise, adjustable control of
reactivity by neutron absorption. Control rods can be designed and used for coarse control,
fine control and fast shutdown. These rods can be moved into or out of the reactor core and
typically contain elements such as cadmium, boron or hafnium. The material used for the
control rods varies depending on the reactor design. Generally, the selected material should
have a good absorption cross section for neutrons and have a long lifetime as an absorber (i.e.
not burn up rapidly). Hafniumis used as an absorbing material in the case of the ISIS reactor.
The effectiveness of a control rod depends largely upon the value of the ratio of the neutron
flux at the location of the rod to the average neutron flux in the reactor. Control rods have
maximum effect (i.e. insert largest negative reactivity) if they are placed in the reactor where
the flux is maximal. The exact effect of control rods on reactivity determines safety reactor
operation.

5.2. Types of Control Rods


There are several ways to classify the types of control rods. Typical classification is based on
the purposes of a control rod. A reactor is usually equipped with three types of control rods:
x The safety rods (scram rods) are used for a fast shutdown of the reactor. A large
amount of negative reactivity is rapidly inserted into the reactor core by this control rod
types. Usually it is possible to operate the reactor only when the safety rods are
completely withdrawn from the reactor core.
x The shim rods are used for coarse control and/or to remove reactivity in relatively large
amounts.
x The regulating rods are used for fine reactor power level adjustments and for
compensating occasional reactivity changes.

5.3. Integral and Differential Control Rod Worth


The exact effect of control rods on reactivity drives safety reactor operation. This influence
can be determined experimentally; e.g. a control rod can be withdrawn in small increments
and the change of reactivity can be determined following each increment of withdrawal.
When the full control rod has been calibrated, the graph presenting the control rod calibration
curve can be depicted. Two types of graphs can be created based on the control rod
calibration.
The graphs can present:
x integral rod worth,
x differential rod worth.
The integral control rod worth (see Figure 11) is the total reactivity worth of the rod at that
particular degree of withdrawal and is usually defined to be the greatest when the rod is fully
withdrawn. The slope of the curve (ǻȡ/ǻx), and therefore the amount of reactivity inserted per
unit of withdrawal, is usually greatest when the control rod is about midway out of the core.
This occurs because the area of the greatest neutron flux tends to be near the centre of the
core; therefore, the amount of change in neutron absorption is the greatest in this area.
A plot of the slope of the integral rod worth curve, also called the differential control rod
worth, is shown in Figure 12. At the bottom of the core, where there are not so many
neutrons, rod movement has little effect so the change in the rod worth per mm varies little.
As the rod approaches the core centre, its effect becomes greater, and the change in rod worth
per mm is greater. At the centre of the core the differential rod worth is the greatest and varies
little with rod motion. From the centre of the core to the top, the rod worth per mm is
basically the inverse of the rod worth per mm from the centre to the bottom.
In summary, the differential control rod worth is the reactivity change per unit movement of a
rod and it is normally expressed as ȡ/mm, ǻk/k per mm, ȕeff($)/mm or pcm/mm. The integral
rod worth at a given withdrawal is merely the summation of all the differential rod worths up
to that point of withdrawal. It is also the area under the differential rod worth curve at any
given withdrawal position.

 
 

Figure 11. Integral control rod worth.

Figure 12. Differential control rod worth.

5.4. Control Rod Calibration by the Inverse Rate Method


The control rod calibration can be made by various methods, e.g. the Rod Drop method, the
Positive Period method or the Inverse Rate method. The control rod calibration by using the
inverse rate method is explained in this chapter. The method follows from the theory of
neutron multiplication in sub critical reactors (i.e. keff < 1) with external neutron source. The
reactor is in a sub critical state during the whole process of control rod calibration. The
multiplication of neutrons emitted from the external neutron source in sub critical reactors can
be described by the following equation
1  k effm
N H ˜S ˜ ,   (1)  
1  k eff

where

N – is measured neutron rate,

İ – is detection efficiency,

S – is emission rate of the external neutron source,

m – is number of neutron generations,

  keff – is multiplication factor.


for keff < 1 and m o f the equation (1) can be adapted to the formula

1
  N H ˜S ˜ ,   (2)  
1  k eff

and from that it is possible to express keff

H .S
k eff 1 . (3)
N
The reactivity change depending on the control rod position (ǻȡ(x)) can be derived from the
Figure 13

U( x )  U p
'U ( x ) U0 ˜ ,   (4)  
Un  Up
where

ȡ0 – control rod worth

ȡ(x) – reactivity for the actual position x of the calibrated control rod,

Up – reactivity when the calibrated control rod is in the bottom position,

Un – reactivity when the calibrated control rod is in the top position.


 

Figure 13. Control rod calibration curve at the typical reactor and the basic
parameters for the curve determination using the inverse rate
method

Using the definition of the reactivity

k eff  1
U , (5)
k eff

one can rewrite the equation (4) to the following form

k eff ( x )  1 k eff p  1

k eff ( x ) k eff p
'U ( x ) U0 ˜ . (6)
k eff n  1 k eff p  1

k eff n k eff p

Substituting expression (3) for keff in the previous equation, one obtains

1 1

Np N ( x ) k eff n
'U ( x ) U0 ˜ ˜ . (7)
1 1 k ( x )
 eff
Np Nn
Because the fraction

k eff n
| 1, (8)
k eff ( x )

it can be neglected and the final expression for ǻȡ(x) determined from the inverse neutron rate has the
following form

1 1

N( x ) Np
'U ( x ) U0 ˜ , (9)
1 1

Nn Np

where

N(x) – is detected neutron rate at the actual position x of the calibrated control rod,

Np – is detected neutron rate when the calibrated control rod is in the bottom
position,

Nn – is detected neutron rate when the calibrated control rod is in the top position.

6. Study of the Reactor Dynamics

6.1. Introduction
For the safety operation of the reactor, very deep knowledge about its kinetics and dynamics
in various operating modes is essential. The exercises are used to demonstrate the basic
phenomena, which you can meet in the reactor.

6.2. Zero Power Reactor Without Delayed Neutrons


Behaviour of the reactor as a finite multiplying system without external neutron sources and
without delayed neutrons could be described by the following equation
dn( t ) k eff  1
˜ n( t ) , (1)
dt A
where
n(t) – is total number of neutrons at the time t,
Ɛ – is prompt neutron lifetime,
keff – is effective multiplication factor.
The reactor is in a critical state if there is a time independent chain reaction in absence of
external neutron sources. Strictly speaking, the reactor is in critical state, when keff is equal to
one. When effective multiplication factor keff is larger than one, the reactor is in supercritical
state and when keff is lower than one, the reactor is in sub critical state. Total number of
neutrons in the reactor (which corresponds to the number of fissions and thus to the power of
the reactor) depends on effective multiplication factor keff and could be described by following
equation
§ k eff  1 ·
n( t ) n0 ˜ exp¨¨ ˜ t ¸¸ , (2)
© A ¹
which is a solution of equation (1) for the case when at time t = 0 is n(0) > 0. From
equation (2) it is evident that when keff is equal to one, n(t) = n0. Similarly when keff < 1, the
numerator in the fraction is negative and n(t) is decreasing exponentially. Finally, when
keff > 1, the numerator in the fraction is positive and n(t) is increasing exponentially (refer to
the Figure 14 where n0 = 1). Considering the case where a time-dependent source is in the
reactor, note that the definition of criticality depends only on the effective multiplication
factor keff, and not on the external neutron source. In that case equation (1) should include the
source S(t) in the following form
dn( t ) k eff  1
˜ n( t )  S ( t ) . (3)
dt A
The total number of neutrons in the reactor could be described by the following equation

A ˜ S0 § § k eff  1 · ·
n( t ) ˜ ¨ exp¨¨ ˜ t ¸¸  1 ¸ , (4)
¨ ¸
k eff 1 © © A ¹ ¹
which solves equation (3) for the case when S(t) ĺ S0 (external neutron source is constant)
and n0 = 0. Behaviour of the reactor is not very evident from equation (4), let us analyses it
closer. In the case of keff > 1, the numerator in the fraction in the exponent is positive and n(t)
is increasing exponentially. The behaviour of a subcritical reactor becomes more transparent
if equation (4) is rewritten in the form

A ˜ S0 § §  1  k eff · ·
n( t ) ˜ ¨ 1  exp¨¨ ˜ t ¸¸ ¸ . (5)
¨ ¸
1  k eff © © A ¹¹
Let’s assume keff < 1, the fraction is positive and n(t) is increasing at first, but then as the
exponential term decays away, n(t) stabilises at long time to a time-independent value Q ’
A ˜ S0
n( f ) . (6)
1  k eff
In the critical reactor the denominator in the first fraction of equation (5) is equal to zero. In
this case when keff ĺ 1, the exponential part of equation (5) could be expanded as a power
series and one can write
§ 1 k eff  1 2 ·
2
A ˜ S0 k eff  1
n( t ) ¨
˜ 1 ˜t  ˜ ˜ t  ...  1 ¸ . (7)
1  k eff ¨ A 2 A2 ¸
© ¹
Cancelling terms and taking the limit keff ĺ 1 results in n(t) = S0.t which means that in a
critical reactor with an external neutron source n(t) increases linearly. Finally, Figure 15
shows in a simple graphic form how the total number of neutrons in sub critical, critical and
supercritical reactors with an external neutron source is changing.
Figure 14. Power response in subcritical, critical and supercritical
reactors without an external neutron source.

Figure 15. Power response in sub critical, critical and supercritical


reactors with external neutron source.

Even a very small deviation of the effective multiplication factor keff from one causes very
large and fast changes of number of neutrons in the reactor. It is so because the prompt
neutron lifetime is rather short (typical values for LWRs range in Ɛ -5 - 10-4 s). That is
why delayed neutrons are necessary for the safe reactor operation. Delayed neutrons,
however, make reactor kinetics and dynamics more complicated. Influence of the delay
neutrons on the reactor operation is described in the experiment “Delay neutrons detection”.

6.3. Reactivity Feedback and Reactor Dynamics


In the previous chapter the reactor has been studied as a system without feedback. Reactor
parameters (temperature, pressure, reactor power...) have been constant or they have only
slightly changed with no effect on reactivity. In real reactors during the standard operation the
parameters of the core are changing and cause the changes of the core properties and thus
affect the reactivity of the reactor. These changes are very important in an emergency
operation during the power excursion. To make understanding of these changes easier and to
model the dynamic and transient processes, reactivity coefficients are defined as
wU
x
x
a y , (8)
wy
where x is a certain part of the core (fuel, moderator...) and y is a reactor parameter
(temperature, power...). The most important reactivity coefficients are:
x fuel temperature coefficient,
x moderator temperature coefficient,
x power coefficient,
x void coefficient.
As changes of the core parameters cause directly the reactivity change, reactivity coefficients
work as a feedback. The basic requirement for the safety operation of the reactor is its
dynamic stability and one of the basic requirements to form a stable system is the negative
feedback of the system. Therefore, it is required for the reactivity coefficients to be negative,
as a total, for the reactor. In research and training reactors the power coefficient is negligible;
only the temperature and void coefficients are important.
Fuel temperature coefficient is a typical negative coefficient caused mainly by the Doppler
broadening of the resonance capture cross section in thermal power reactors. In the reactor the
effect appears as a decrease in the resonance escape probability with increased fuel
temperature.
Moderator temperature coefficient and void coefficient could be positive or negative and they
depend on moderator - fuel ratio (also called water - uranium ratio). If the water - uranium
ratio is lower than the “optimal value” the core is so-called “sub moderated” and both
coefficients are negative. In this case, the density of the moderator is decreasing and the
moderation in the core is worse which cause the negative reactivity change. If the water -
uranium ratio is higher than the “optimal value” the core is so-called “over moderated” or
“over absorbed” and both coefficients are positive. In this case, the density of the moderator is
decreasing, but the moderation in the core is better because less absorption in moderator
causes the positive reactivity change. “Optimal value” of the water - uranium ratio and sub
moderated and over moderated states of the core are shown in Figure 16.

Figure 16. Water - uranium ratio and temperature/void reactivity


changes.
7. The Effect of Temperature on Reactivity
7.1 General Considerations
The power of a nuclear reactor stays constant when the multiplication factor keff = 1. At
constant power the cooling system of the reactor removes all the heat that is generated. If the
multiplLFDWLRQIDFWRULVLQFUHDVHGE\ǻkeff = keffí!WKHSRZHURIWKHUHDFWRUVWDUWVWRJURZ
at a rate GHWHUPLQHG E\ WKH PDJQLWXGH RI ǻkeff. The cooling system cannot, at least
immediately, absorb all the thermal energy released in fission, but part of it will raise the
temperature of the reactor core. In the absence of any limiting factors the temperature would
rise indefinitely, and finally the core would melt. In cases like this the temperature
dependence of the reactivity ȡ ǻkeff /keff and the feedback caused by it usually constitute a
limiting factor.
The dependence of the reactivity on temperature is described by the temperature
coefficient of the reactivity. For this we have to assume that the whole reactor core can be
characterized E\ RQH VLQJOH WHPSHUDWXUH 7 7KH WHPSHUDWXUH FRHIILFLHQW Į LV GHILQHG DV D
derivative:
dU
D . (1)
dT

Usually it is assumed that when ȡ is expanded into a series of (TíT0), the series can be
truncated after the first-degree term, i.e., we have a linear relationship. Here T0 is some
suitable reference temperature. Į is usually expressed in units °C-1, less frequently $/°C.
If the temperature coefficient Į is negative, i.e., the reactivity decreases as the temperature
increases, the reactor will behave stably. When the multiplication factor for some reason or
other undergoes a positive change the neutron flux and thus also the power starts to rise. Since
all the heat generated is not removed from the reactor the core temperature will rise. This
reduces the reactivity and slows down the increase in power. Finally the temperature reaches a
level where it is sufficient to compensate the whole original reactivity insertion. The power
will stop increasing and with still rising temperature will start decreasing. Finally, the system
will stabilize at a new power level between the original level and the power peak. Such a
behavior is, of course, highly desirable for safety reasons. It is easy to accomplish reactor
designs which lead to increased reactivity with rising temperature and, as a consequence, are
inherently unstable. Chernobyl provides a horrendous example.
However, merely a negative temperature coefficient is not a sufficient guarantee for safe
operation. Often there is a delay in the feedback between power increase and reduced
reactivity. This delay is usually caused by the time needed for heat transfer or, e.g., for
expelling the excess water volume created by thermal expansion from the core. If a reactor
achieves strong prompt criticality the power increase will be extremely fast. The heat transfer
from the fuel elements to the coolant and moderator is then far too slow to have an influence
on phenomena whose time behavior is determined by the neutron life-time in the core.
The phenomena associated with the expulsion of matter from the core also require a time
which is of the same order as the reactor dimensions divided by the speed of sound. For a
reactor to be safe even in a prompt critical state one must, therefore, require that the
temperature coefficient of reactivity is not only negative but also fast. By fast we here mean a
time which is short compared to the neutron life-time.

7.2 The Fast Temperature Coefficient


The fast or prompt temperature coefficient stems from phenomena associated with the
changes in the reactor physical properties caused by heating of the fuel. Nearly all hitherto
constructed reactors have a fast negative temperature coefficient of reactivity, and licensing
authorities do not generally allow the construction of reactors without this property.
In the some reactor (e.g. TRIGA) only the fast temperature coefficient is negative. Thus, if
the core and cooling water are heated homogeneously and simultaneously, one can observe an
increase rather than a decrease in reactivity. The fact that during normal operating conditions
the reactivity decreases with increasing power is thus exclusively a consequence of the rise in
temperature difference between the fuel and coolant caused by the increased heat flow. This
type of behavior is very desirable from the reactor technical point of view, since no excess
reactivity is needed to compensate reactivity loss when going from a cold reactor core to the
operational temperature.
Two effects contribute mainly to the fast negative temperature coefficient:
1. Doppler broadening
2. Increased neutron leakage

Doppler broadening
With rising fuel temperature the thermal motion of the nuclei increases. As a consequence a
neutron moving with constant speed in the fuel will encounter fuel nuclei whose movement in
the direction of the neutrons movement will follow a Maxwellian distribution. This leads to a
broadening of the resonance peaks in the absorption cross sections, so-called Doppler
broadening. The probability of parasitic absorption of the neutron during slowing-down
increases. This in turn decreases the resonance escape probability, the multiplication factor
and the reactivity. This phenomenon is prompt.

Increased neutron leakage


The increase in neutron leakage is caused by the dependence of the thermal cross sections on
the fuel temperature. With rising temperature the diffusion length will increase
The temperature dependence of the cross sections can be treated quantitatively if the
neutron spectrum in the reactor can be determined. The shape of the spectrum is determined
by the slowing-down properties of the water. Slowing down is well-known, although
complicated to handle theoretically.

The basic equations


In the following we shall try to analyze in more depth the behavior of a reactor that has been
running at constant power for a long time after which its multiplication factor is quickly
increased, e.g., by raising a control rod. The fast temperature coefficient of reactivity is
assumed to be negative.
For this purpose the kinetic behavior of the reactor must be described with some model. In
practice this can usually be accomplished with the aid of the so-called point kinetic equations,
originally derived for an infinite reactor. Moving from neutron density to power (P) the
kinetic equations can be cast into the form (with the usual six groups of delayed neutrons):

dP keff 1  E eff  1 6
P  ¦ Oi Ci (t )
dt l i 1 . (2)
dCi Ei keff
Oi Ci  P, (i 1, 2...6)
dt l
keff  1
U (3)
keff

and

l
keff (4)
/

have been utilized.


The multiplication factor is written in the form:

keff 1  'keff  D T  T0 # 1  U  D'T (5)

ZKHUH ǻ7 7í70. T0 is the value of the temperature T with the reactor running at constant
power P0. The temperature coefficient Į has been assumed negative in Eq. (5).
Also the cooling system of the reactor must be modeled in some way. In the following
treatment the reactor core is assumed to consist only of fuel. It is assumed that during the time
interval under consideration the cooling system removes heat from the fuel elements with a
power Ȗ 7í7e) proportional to the difference between the core temperature and an external
temperature Te which is assumed to remain constant. Te mainly describes the temperature of
the cooling system, and the proportionality factor Ȗ, which is assumed constant, is then the
heat transfer coefficient between the core and the cooling system. Since at constant power the
cooling system removes all the heat generated in the core we have:
P0
J (6)
T0  Te

The heat capacity of the core, C, is also assumed constant. Under these assumptions we obtain
the following equation for the temperature change:

d 'T P  J T  Te P  P0  J'T
(7)
dt C C

When the reactivity injection is large enough to make the reactor prompt critical, the so-called
Fuchs model can be applied. However, since no real pulses will be fired in this exercise, the
discussion is omitted here.

7.3 The Influence of the Temperature Coefficient on the Prompt Jump


When the multiplication factor of the reactor is increased moderately, so that it does not
become prompt critical, the power increase remains manageable even without the aid of the
temperature coefficient. The power initially grows quickly (the so-called prompt jump), but
very soon it starts to exponentially approach its asymptotic value:
E
Pas P0 . (8)
E U

Because of the delayed neutrons the power will, however, not settle at this value, but
continues to grow slowly. The influence of the temperature coefficient becomes visible if P0
is large or ȕíȡ is very small. The power then after a certain time reaches a maximum, after
which it will fall somewhat and finally stabilize at a level determined by
J
Pf P0  P. (9)
D

7.4 Determining the Fast Temperature Coefficient of Reactivity

7.4.1 Determining the temperature coefficient


We assume that the reactivity ȡ is a function of the time t and the core temperature T:
U U (t , T ) . (10)

By differentiating we obtain:

§ wU · § wU · § wU ·
dU ¨ ¸ dt  ¨ ¸ dT ¨ ¸ dt  D'T . (11)
© wt ¹T © wT ¹t © wt ¹T

Integrating Eq. (11) from 0 to time t and assuming that Į remains constant during this time
interval, and denoting T(t íT   ǻT(t), we obtain:

§ wU ·
t
U (t )  U (0) ³ ¨© wt ¸¹
0 T
dt  D'T (t ) . (12)

Let us now further assume that ȡ(t, T(t)) can be written as the sum of a time-dependent part
and a temperature-dependent part:
U (t , T ) Ut  UT . (13)

From Eqs. (12) and (13) we then obtain:


U (t )  U (0) GUt  D'T (t ) . (14)

where įȡt is the increase in the time-dependent reactivity (injected with the control rod)
during time 0 ĺ t. How this change is brought about is not important. When measuring the
temperature coefficient we have ȡ(0) = ȡt(0) = 0. The time-dependent reactivity ȡt receives its
whole constant value įȡt DOPRVWLQVWDQWDQHRXVO\DWW :KHQZHPHDVXUHǻT(t0), the change
in T going from t = 0 to t0, we obtain Į from the equation:
U (t0 )  GUt
D . (15)
'T (t0 )

It is advantageous to choose t0 as the maximum point of the power curve.

7.4.2 Determining ȡ(t0)


a) A coarse estimate. A very coarse estimate for ȡ(t0) is obtained by assuming that all
neutrons emitted in fission are prompt, i.e., ȕ = 0. We thus assume that the power increases so
slowly that the neutron mean lifetime can be considered a very small quantity. According to
Eq. (1) we then obtain:
dP U
P, (16)
dt l

and at the peak of the pulse:


U (t0 ) 0 . (17)
Naturally, the slower the pulse, the better this approximation will be, but on the other hand the
heat transfer models become less and less useful as the duration of the phenomenon increases.

b) A more accurate estimate. The delayed neutrons can also be accounted for exactly. From
Eq. (2) we can solve Ci(t):

ª E t º
Ci (t ) e  Oi t «Ci (0)  ³ eOi t keff P (t )dt » . (18)
¬ l 0 ¼

When this is inserted into Eq. 2, and taking into account that
6
E
¦ O C (0)
i 1
i i
l
P0 , (19)

we have:

dP 1 E E 6
E t
U P (t )  P (t )  P0  ¦ i e  Oi t ³ eOi t keff P (t )  P0 Oi dt . (20)
dt l l i 1 l 0

At the maximum, this expression equals zero. Inside the integral we can set keff = 1 to obtain:

1 § 6 t0
·
¨¨ E eff Pmax  P0  ¦ E i e ³ e i P (t )  P0 Oi dt ¸¸ .
 Oi t0 Ot
U (t0 ) (21)
1  E Peff max © i 1 0 ¹

In evaluating the integral in the above formula one must resort to approximate methods. It is
then worth noting that the major part of the integrals value comes from the region near the
peak of the pulse.

7.4.3 Determining ǻT(t0)


2QHFRXOGWU\WRPHDVXUHWKHULVHLQWHPSHUDWXUHǻT(t0) directly if one could unambiguously
establish what temperature the quantity T PRVWFORVHO\GHVFULEHV+RZHYHUǻT(t0) can also be
calculated from the heat transfer model by integrating Eq. (7):
t0
1
³ e P(t )  P P dt ,
 P t0 P
'T (t0 ) e t0
(22)
J
0
0

where we have denoted ȝ = Ȗ/C. However, the value of the coefficient Ȗ is usually not known.
In principle it could be obtained as the ratio between P0 and (T0íTe), but because of the
hypothetical nature of these temperatures their magnitude is difficult to determine. Usually
one must assume the cooling system to be slow (as is done in the Fuchs model), in which case
we can set Ȗ 7KHQǻT(t0) can be solved from the equation:
t
1 0
P(t )  P0 dt ,
C ³0
'T (t0 ) (23)

The significance of leaving out Ȗ can be evaluated by assuming that Te represents the average
temperature of the cooling water in the tank and T represents the spatial average of the fuel
temperature.
8. Approaching Critical State
8.1. Introduction
The critical state of the reactor is normally reached when the reactor is put successfully into
operation. In fact, every change of the power is the short-time deviation from the critical state
and returning to a new critical state. However, this process is realized at the known
arrangement of the core. In the case of a new reactor and the reactor start-up after changing
the core configuration approaching to the critical state is always connected with an
uncertainty factor. Neither the operators and the reactor physicists experience nor the most
perfect physical calculations can guarantee the exact determination of the core critical size,
the control rod positions in the critical state or the exact concentration of the absorber.
Therefore, the critical experiment must be conducted at all reactors with almost identical
methods.

8.2. Theory
The critical experiment is an experimental check of the calculated geometry and composition
of the core. From the comparison between the experimental results and the calculation, the
necessary corrections of the calculation methods, the constants used, etc. can be derived. The
knowledge of the real critical core is important not only during starting the reactor into
operation at the first time but it also determines the safety of the core, the quantity of loaded
fuel to reach the operational reactivity, and a number of other parameters.
Let us assume that we have in our disposal a core with the effective multiplication factor
keff < 1. There are several options how to approach the critical state:
x change of the fuel quantity,
x change of the moderator water-level,
x change of the neutron absorption (movement of the control absorption rods),
x change of the absorber concentration in the coolant/moderator.
In practice, keff depends on one or more of the above mentioned parameters. In addition, let us
assume that the parameter change is discontinuous in steps that are numbered and marked
with the index i (or during the continuous parameter change, we will realize the measurement
step-by-step at certain parameter values, i.e. pulling out the rods to a certain position or
reaching a certain moderator concentration). The reactor is in a sub-critical state and the
external neutron source is inserted in the core. Further, we assume that the reactor can be
described by a single point approximation. It means that the thermal neutron flux densities in
the core as well as in the reflector are mutually proportional at every moment. Then, in any
position of the reactor, a detector measures a value that is directly proportional to the reactor
power.
Let us analyze the case of approaching the critical state with pulling out the control rod step-
by-step. In the step zero, the rod is in the lowest position. The value that is proportional to the
total number of neutrons in the core (ni) is measured with the detector. After the first
movement of the rod, keff is changed and the number of neutrons is
2 3 m
n1 = no + no ˜ k ef1 + no ˜ ( k ef1 ) + no ˜ ( k ef1 ) + ....  no ˜ ( k ef1 ) , (1)
where m is the number of neutron generations. As the reactor is sub-critical, keff < 1, and the
final number of neutrons is given with the sum of the geometric series with the quotient of keff.
Thus, it is
m
1 - ( k ef1 )
n1 = no ˜ . (2)
1 - k ef1
Regarding the reactors with light water, the lifetime of one neutron generation is of the order
of 10-4 to 10-5 s. Therefore, it is possible to neglect keff in the nominator. Then the final
equation has the form of
1
n1 = no ˜ . (3)
1 - k ef1
In the next steps, we can apply a similar assumption and generally, the following can be
written
1
ni = n o ˜ . (4)
1 - k efi
If we are approaching the critical state, the value of keff is approaching one and the value of
the fraction in (4) increases to infinity; its inverse value approaching null. When the inverse
value of (4) is plotted to the graph together with keff (or generally, with the variable parameter
of the core), the curve intersects the x-axis (i.e. 1/ni =0) when the critical state is reached (see
Figure 17). Thus, with the extrapolation of this curve, we can foresee the size of the variable
parameter in the moment of reaching the criticality.

Figure 17. Procedure at the critical experiment.

As a rule, the value of 1/ni is entered at the x-axis. When reaching the criticality, 1/ni comes
near null. Thus, any constant multiple of it comes near null, as well. Therefore, it is suitable to
enter the values of n0/ni into the graph; the initial value of the scale is 1 and the problems to
adjust the suitable scale are omitted.
When the measurement in the (i-1)th step is finished, the positive reactivity change is realized
(keff is approaching the value 1). The measurement of the ith step is done after stabilizing the
reactor power. The value of n0/ni is entered and the extrapolation is plotted. The extrapolated
value, where the criticality can be expected with the same curve slope, is determined. This
value is compared with the value that was determined by the calculation. For safety reasons, it
is required that the modified parameter can be increased by no more than 1/2 of the difference
between the present state and the smaller value of the position that were determined from the
previous extrapolation and from the calculation (refer to Figure 18). After applying this
check, the requirement on the next step is made more accurate, the value is entered in the
graph and the whole process is repeated to reach the value of n0/ni = 0.1 - 0.15. Then, the last
extrapolation is realized and the operator can put the reactor to the critical state.

Figure 18. Approaching to the critical state

During the approach to the critical state three different cases - curve shapes - can occur. In
Figure 18, the ideal course is shown by the curve 3. From the point of view of the nuclear
safety, the curve 1 is disadvantageous because the extrapolated value is higher than the
subsequent real value. Similarly, the curve 2 is more advantageous, but the angle under which
it intersects the x-axis leads to a non-accurate point of intersection and thus the les accurate
forecast of the critical state.
The form of the curves depends on a number of factors - the most important factor is the
mutual position of the detector, the neutron source, and the fuel and its change during the core
assembling, the distance of the source from the detector, and finally, the way in which the
change of the multiplication factor depends on the change of the variable parameter.
Basic
Data Handling
in
Nuclear and Reactor
Applications

1.8

1.6

1.4

1.2

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

&DOOH%HUJO|I
'HSDUWPHQWRI5HDFWRU3K\VLFV
5R\DO,QVWLWXWHRI7HFKQRORJ\

FDOOH#QHXWURQNWKVH
ZZZQHXWURQNWKVH


Table of Contents

1 The Poisson Distribution...................................................................................................................3


2 The Propagation of Error Formula..................................................................................................3
3 Linear Fitting........................................................................................................................................4
4 The Ʒ2-test ............................................................................................................................................5
5 Mean Values.........................................................................................................................................8
6 Number of Accurate Figures.............................................................................................................8
7 Preparation Tasks................................................................................................................................9
8 References ............................................................................................................................................9
Appendix 1: Solutions to the Preparation Tasks ...................................................................................10

Basic Data Handling in Nuclear and Reactor Applications


Calle Berglöf, Department of Reactor Physics, KTH
2(10)
Basic Data Handling in Nuclear and Reactor Applications

When analyzing experimental data, it is of essential weight to treat the collected data correctly. It may
happen that different results are obtained from the same raw data depending on how the data is treated
and which methods that have been applied. Such situations can be avoided by following some general
statistical methods for data handling. It is recommended to carefully read this document before starting
any data analysis. It is not the aim of this document to give a complete derivation of all statistical relations
and distributions; the document will rather serve as a short handbook for an engineer or experimentalist.

1 The Poisson Distribution


An event that occurs with a constant probability in time is said to be Poisson distributed. Radioactive
decay is a typical example of a process obeying Poisson distribution. Each decay is completely random in
time and, therefore, also the subsequent detection of the related particle or gamma. Consequently, the
collected data from a measurement performed with radioactive samples should be treated as Poisson
distributed.
The standard deviation, ıi, of a measurement performed on a process obeying Poisson distribution is
given by

Vi yi , (1)

where yi is the number of events detected during the measurement i. The relative error, erel, will
consequently decrease with the square-root of the number of counts:

Vi yi 1
erel . (2)
yi yi yi

2 The Propagation of Error Formula


Assume that a measurement has been performed with the results yi and ıi. Now the standard deviation
ı(f(yi)) is to be calculated, where f(yi) is a quantity that is a function of the measured values. This can be
done by using the Propagation of Error Formula:

2
§ wf ·
V ( f ( yi )) ¦i ¨ wy V ( yi ) ¸ . (3)
© i ¹
Example 1: The prompt neutron decay constant, Į, has been measured in a subcritical reactor. In
subcritical systems, this is an important quantity describing the subcriticality of the core through the
relation

U D/  E eff , (4)

where ȡ is the reactivity, ȁ is the neutron mean generation time and ȕeff is the effective delayed neutron
fraction. Experimentally, Į was determined to -420±15 s-1 in an arbitrary subcritical core, and by
simulations, the neutron mean generation time and the effective delayed neutron fraction was estimated to
100±10 µs and 750±30 pcm respectively. By employing the Propagation of Error Formula, Eq. (3), the
standard deviation of the reactivity can be calculated as

2
· § wU ·
2 2
§ wU · § wU
V (U ) ¨ V (D ) ¸  ¨ V (/) ¸  ¨ V ( E eff ) ¸
© wD ¹ © w/ ¹ ¨© wE eff ¸ (5)
¹
/V (D )  DV (/)  V 2 ( E eff )
2 2
447 pcm.

Basic Data Handling in Nuclear and Reactor Applications


Calle Berglöf, Department of Reactor Physics, KTH
3(10)
The final result is therefore ȡ=-3450±450 pcm. This corresponds to an effective multiplication factor, keff,
of

1
keff 0.9667 (6)
1 U
with the standard deviation

2
§ wkeff · wkeff V (U )
V (keff ) ¨ V (U ) ¸ V (U ) 0.0042 . (7)
© wU wU 1  U
2
¹

The final result should then be given as keff=0.967±0.004, taking into account the number of relevant
digits with respect to the error.

3 Linear Fitting
Let us assume that a set of data points yi has been measured for different conditions xi (for instance space
or time) and that there exists a linear dependence between these data according to

y ax  b . (8)

The task is to find a and b and their standard deviations ı(a) and ı(b) in this linear relation. By assuming
all measurements having the same error this can be solved by minimizing the function

¦ y  ax  b
2
S LS i i (9)
i

with respect to a and b. This is generally known as the Least Square Method. However, since we are dealing
with Poisson distributed data, we know that the standard deviation of each measurement is

V i ( yi ) yi (10)

and that yi and consequently ıi(yi) will be different for each measurement. Therefore, measured data with
higher accuracy should be given higher weight in the linear fitting procedure, according to the following
function:

yi  axi  b
2

S ¦i V i2
. (11)

Minimizing S with respect to a and b gives a system of two equations that can easily be solved:

wS
2¦
yi  axi  b xi 0
wa i V i2
(12)
wS
2¦
yi  axi  b 0.
wb i V i2
By introducing the notation

Basic Data Handling in Nuclear and Reactor Applications


Calle Berglöf, Department of Reactor Physics, KTH
4(10)
xi 1
A ¦V 2
B ¦V
i
2
i i i

yi xi2
C ¦V
i
2
D ¦V
i
2
(13)
i i

xi yi
E ¦V
i
2
i

the result can be written as

EB  CA
a (14)
BD  A2
DC  EA
b . (15)
BD  A2
The standard deviations of these parameters are here given without further explanation as

B
V (a) (16)
BD  A2
and

D
V (b) . (17)
BD  A2

4 The ǒ2-test
When performing fitting of functions to experimental data, it is important to know how good the assumed
model describes the outcome of the experiment. It might happen that the proposed function do not
describe the reality well and another model must be used. If the previously defined quantity S is divided by
the degree of freedom, Ȟ, of the problem, the reduced Ȥ2 (““chi-squared”” or ““chi-two””) is obtained:

F2 S
. (18)
Q Q
The degree of freedom is given by

Q N m (19)

where N is the number of measurements and m is the number of parameters of the model. In the case of a
linear fit, m=2. If the reduced Ȥ2 is close to one, the quality of the fit is good. A more rigorous way to
evaluate whether the quality of the fit is acceptable or not is to calculate the probability

P( F 2 t S ) .

If this probability is larger than 5% the fit is considered to be acceptable. This is done by integrating the
Ȥ2-distribution

(u / 2)Q / 2 1 e  u / 2
P (u )du du (20)
2*(Q / 2)

according to

Basic Data Handling in Nuclear and Reactor Applications


Calle Berglöf, Department of Reactor Physics, KTH
5(10)
(u / 2)Q / 2 1 e  u / 2
S
P( F 2 t S ) 1  ³ du . (21)
0
2*(Q / 2)

In Eq. (20) and Eq.(21) u=Ȥ2 (just to avoid confusion with the exponent) and
f

³t
z 1  t
*( z ) e dt (22)
0

is the Gamma function 1 . The Ȥ2-test and the P(Ȥ2•S)-test serve as excellent tools when judging the quality
of a function fit. Most often, to check whether the reduced Ȥ2-value is close to 1 is a sufficient verification.
However, a visual control by plotting the data together with the fitted function is essential for the final
judgement and to make sure that the calculations are correct.

Example 2: The data in Table 1 have been obtained experimentally and a linear dependence between x
and y is assumed. Eq. (14) to Eq. (18) give the values a=7.0±0.3, b=10.7±1.2 with Ȥ2/Ȟ§0.6 and
P(Ȥ2•S)§68%. The reduced Ȥ2 is somewhat low, but P(Ȥ2•S) is well above 5% which tells us that the fit is
good. Since the reduced Ȥ2-value is less than unity it is expected that the uncertainties of the
measurements are lower than expected. Therefore, one might suspect that the data set is somehow biased.

Table 1. Example of experimental data.


x       
y       
ı(y)       

y=ax+b
65

60

55

50

45

40
y

35

30

25

20

15
0 1 2 3 4 5 6 7 8
x
Figure 1. Linear fit to experimental data.

Often, the linear fit cannot be performed directly on the measured values. For instance, if the
measured quantity is of exponential nature, the problem must be transformed to a linear problem.
Consider the decay of a radioactive source:

N H N 0 e  Ot . (23)

where N is the number of detected decays, İ is the efficiency of the detector, N0 is the decay rate at time
zero and Ȝ is the decay constant. This relation can be transformed to a linear problem by evaluating the
logarithm on both sides:

ln N ln H N 0  Ot . (24)

1 Numerical problems may appear when calculating this integral for large values of z.

Basic Data Handling in Nuclear and Reactor Applications


Calle Berglöf, Department of Reactor Physics, KTH
6(10)
This is equivalent with the previously used linear model with the transformations

y ln( N )
x t
. (25)
a O
b ln H N 0

It is important to remember that the standard deviation, ıi(Ni), must be transformed as well. Using the
Propagation of Error Formula, Eq. (3), the transformation of the standard deviation is given by

2
§ w ln( N ) · 1 1 1
V ln( N ) ¨ V (N ) ¸ V (N ) N . (26)
© wN ¹ N N N

Example 3: Consider a measurement of a radioactive decay. The data in Table 2 have been collected.
What is the half-life of the sample?
Table 2. Example of radioactive decay data.
t [s]       
N [cps] 2       

After performing the linearization (Eq. (24)) and transformation (Eq. (25)), the parameters a and b can
be calculated using Eq. (14) to Eq. (17). Note that Eq. (26) must be used when calculating the
components of Eqs. (13). The results are a=-0.029±0.001 and b=7.15±0.03 and hence the decay rate is
given by Ȝ=-a=0.029±0.001 s-1. The reduced Ȥ2 is found to be 1.7 which is somewhat high. On the other
hand, P(Ȥ2•S)§13% is above the limit of 5% and the fit is acceptable. The half-life is found from the
relation

ln 2
t1/ 2 23.7 s (27)
O
and by using the Propagation of Error Formula, Eq. (3), the standard deviation is
2
§ wt1/ 2 · ln 2
V (t1/ 2 ) ¨ V (O ) ¸ V (O ) 0.8 s . (28)
© wO ¹ O2
y=ax+b
Exponential 7
1100
6.8
1000
6.6
900
6.4
800
6.2
700
N [cps]

ln(y)

6
600
5.8
500

400 5.6

300 5.4

200 5.2

100 5
0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
t [s] x

Figure 2. Experimental data and function fitting in linear and logarithmic scale.

2 Counts per second (s-1).

Basic Data Handling in Nuclear and Reactor Applications


Calle Berglöf, Department of Reactor Physics, KTH
7(10)
5 Mean Values
Let yi, i=1,2,3……N, be the results of N measurements, all of them performed under the same conditions
and with the same error. The arithmetic mean value, y , is then given by
N
1
y
N
¦y
i 1
i . (29)

If the values yi are Gaussian distributed, for instance when performing successive measurements on
distances etc, the best estimate of the error of the arithmetic mean value is the standard error

V
V ( y) , (30)
N
where

¦ y  y
2
i
V i 1
(31)
N 1
is the ordinary definition of the standard deviation. On the other hand, if the data is Poisson distributed,
the best estimate of the error of the arithmetic mean value is

y
V ( y) . (32)
N

If the measurements were performed under different conditions, with different standard deviation, ıi,
for each measurement, the weighted mean value, Pˆˆ , should be used.
N

¦ y /V i i
2

Pˆˆ i 1
N
(33)
¦1/ V
i 1
i
2

Such situation might occur for instance if the same observable is measured using different techniques with
different accuracy. The standard deviation of the weighted mean value is given by

1
V ( Pˆˆ ) N
. (34)
¦1/ V
i 1
i
2

Note that if ıi is the same for all measurements, Eq. (33) transforms into Eq. (29).
Eqs. (33) and (34) are valid only for Gaussian distributed data. However, the mean value of Poisson
distributed data is Gaussian distributed.

6 Number of Accurate Figures


A common mistake is to include all figures from the last calculation step in the final presentation of the
result and, thereby, give false information concerning the accuracy of the measurement and its result. If
the calculation has been performed without error estimation of the included parameters, the number of
figures in the final result may not be more than given in the input data. For instance:
175/16.2 = 10.8024…… § 10.8

Basic Data Handling in Nuclear and Reactor Applications


Calle Berglöf, Department of Reactor Physics, KTH
8(10)
If the calculation has been performed on quantities with known errors, the result may not have more
accurate figures than described by the error. Moreover, the error itself should be left with one accurate
figure, maximum two if further analysis of the result is expected. Examples:
(175±1)/(16.2±0.1) = 10.8024……±0.0908…… § 10.80±0.09
(175±30)/(16.2±0.1) = 10.8024……±1.8530…… § 11±2

7 Preparation Tasks
1. Create a function, in for instance Matlab, that takes x, y and ı as input parameters and delivers a,
ı(a), b, ı(b), Ȥ2/Ȟ and P(Ȥ2•S). Try to reproduce the calculations in the examples 2 and 3 above.
Hint: In Matlab, a function is defined by writing
function [a, a_std, b, b_std, red_chisq, P] = lsq(x,y,std)
The function is then recalled by writing
[a, a_std, b, b_std, red_chisq, P] = lsq(x,y,std)
from another m-file or from the command line. Note: the filename of the function file must be
the same as the function name (in this case lsq.m).
2. In reactor applications, cosine functions are frequently used to describe flux distributions etc.
How do the standard deviation, ıi(y), transform when transforming the cosine function

y Amax cos B ( x  x0 ) (35)

to a linear relation? Assume that y is Poisson distributed and that the maximum amplitude, Amax,
is known. Hint:

w 1
arccos( x)  . (36)
wx 1  x2
3. Consider a fission chamber that is measuring the neutron flux in a reactor. Signals are collected
during 10 s and the total number of events during this time is displayed. In order to achieve
sufficient low statistical error, three identical measurements are performed according to Table 3.
What is the number of counts per second and the corresponding statistical error?
Table 3. Experimental data.
Measurement number   
y [counts per 10 s]   

8 References
The information in this document has been collected mainly from:
1. W.R. Leo, Techniques for Nuclear and Particle Physics Experiments, Springer-Verlag, 1994.
2. G. Blom, Sannolikhetsteori och statistikteori med tillämpningar, Studentlitteratur, 1989.

Further reading:
3. W.T Eadie, Statistical methods in experimental physics, Amsterdam, 1971.
4. D.L Smith et al., Probability, Statistics and Data Uncertainties in Nuclear Science and Technology,
OCDE/OECD, American Nuclear Society, LaGrange Park, Illinois, USA, 1991.

Basic Data Handling in Nuclear and Reactor Applications


Calle Berglöf, Department of Reactor Physics, KTH
9(10)
Appendix 1: Solutions to the Preparation Tasks

1. function [a, a_std, b, b_std, red_chisq, P] = lsq(x,y,std)

A=sum(x./std.^2);
B=sum(1./std.^2);
C=sum(y./std.^2);
D=sum(x.^2./std.^2);
E=sum(x.*y./std.^2);

a=(E*B-C*A)/(B*D-A^2);
b=(D*C-E*A)/(B*D-A^2);

a_std=sqrt(B/(B*D-A^2));
b_std=sqrt(D/(B*D-A^2));

S=sum((y-a*x-b).^2./std.^2);
N=length(x);
nu=N-2;
red_chisq=S/(nu);

F=@(u)((u./2).^((nu./2)-1).*exp(-u./2))./(2.*gamma(nu/2));
P=1-quad(F,0,S);

2. Transformation:

§ y ·
arccos ¨ ¸ Bx  Bx0 (37)
© Amax ¹
The error is transformed acoording to
2
§ § y ·· § w § § y ·· ·
V ¨¨ arccos ¨ ¸ ¸¸ ¨ ¨¨ arccos ¨ ¸ ¸¸ V ( y ) ¸¸
© Amax ¹¹ ¨ wy © Amax ¹¹
© © © ¹
w § § y ·· 1 1 . (38)
¨¨ arccos ¨ ¸ ¸¸ V ( y ) y
wy © © Amax ¹¹ § y · max
2 A

1 ¨ ¸
© Amax ¹

3. Eq.(29) gives the mean value of the three measurements, y =270. Eq.(32) then gives the
associated error, V y §9. The number of counts per second is consequently 27±0.9. Note that
the error must be calculated before the values are transformed to counts per second.
Alternatively, the total number of counts during 30 can be calculated and the error of this value is
simply its square root. Both values are then divided by 30 to get the correct unit.

Basic Data Handling in Nuclear and Reactor Applications


Calle Berglöf, Department of Reactor Physics, KTH
10(10)

You might also like