06.12.2012 Views

View - The Applied Modelling and Computation Group (AMCG)

View - The Applied Modelling and Computation Group (AMCG)

View - The Applied Modelling and Computation Group (AMCG)

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Space-dependent kinetics simulation of a gas-cooled fluidized<br />

bed nuclear reactor<br />

C.C. Pain a , J.L.M.A. Gomes a , M.D. Eaton a , C.R.E. de Oliveira a, ,<br />

A.P. Umpleby a , A.J.H. Goddard a ,H. van Dam b , T.H.J.J. van der Hagen b ,<br />

D. Lathouwers b<br />

a <strong>Computation</strong>al Physics <strong>and</strong> Geophysics, T.H. Huxley School of the Environment, Earth Sciences <strong>and</strong> Engineering, Imperial College of<br />

Science, Technology <strong>and</strong> Medicine, Prince Consort Road, London SW7 2BP, UK<br />

b Interfaculty Reactor Institute (IRI), Delft University of Technology, Mekelweg 15, NL 2629 JB Delft, <strong>The</strong> Netherl<strong>and</strong>s<br />

Abstract<br />

Received 12 September 2001; received in revised form 16 April 2002; accepted 28 May 2002<br />

In this paper we present numerical simulations of a conceptual helium-cooled fluidized bed thermal nuclear reactor.<br />

<strong>The</strong> simulations are performed using the coupled neutronics/multi-phase computational fluid dynamics code finite<br />

element transient criticality which is capable of modelling all the relevant non-linear feedback mechanisms. <strong>The</strong><br />

conceptual reactor consists of an axi-symmetric bed surrounded by graphite moderator inside which 0.1 cm diameter<br />

TRISO-coated nuclear fuel particles are fluidized. Detailed spatial/temporal neutron flux <strong>and</strong> temperature profiles have<br />

been obtained providing valuable insight into the power distribution <strong>and</strong> fluid dynamics of this complex system. <strong>The</strong><br />

numerical simulations show that the unique mixing ability of the fluidized bed gives rise, as expected, to uniform<br />

temperature <strong>and</strong> particle distribution. This uniformity enhances the heat transfer <strong>and</strong> therefore the power produced by<br />

the reactor.<br />

# 2002 Elsevier Science B.V. All rights reserved.<br />

1. Introduction<br />

Nuclear reactor concepts based on gas fluidization<br />

of fine uranium fuel pellets have attracted<br />

considerable attention over the years. Reasons<br />

behind this interest lies in their excellent heat<br />

transfer capabilities (Molerus <strong>and</strong> Wirth, 1997)<br />

<strong>and</strong> the mixing ability of the fluidized bed. <strong>The</strong><br />

Corresponding author. Tel.: /44-20-7594-9319; fax: /44-<br />

20-7594-9341<br />

E-mail address: c.oliveira@ic.ac.uk (C.R.E. de Oliveira).<br />

Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245<br />

0029-5493/02/$ - see front matter # 2002 Elsevier Science B.V. All rights reserved.<br />

PII: S 0 0 2 9 - 5 4 9 3 ( 0 2 ) 0 0 2 1 5 - 7<br />

www.elsevier.com/locate/ned<br />

latter unifies the temperature of the bed, <strong>and</strong><br />

increases the active surface area from which heat<br />

transfer occurs. In addition, the constant mixing of<br />

the bed potentially leads to a uniform burnup of<br />

the uranium particles. A self-controlling feature is<br />

also present in that as the bed is fluidized <strong>and</strong> the<br />

gas flow increases the power achieves a maximum<br />

at a particular bed height. At this height, the<br />

power will be that at which heat production is<br />

balanced by heat losses.<br />

A possible disadvantage of such a reactor is the<br />

chaotic particle flow characteristics of the fluidized


226<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245<br />

bed in which large bubbles <strong>and</strong> slugs propagate<br />

through it (Davidson et al., 1985), changing the<br />

geometry <strong>and</strong> nuclear criticality. This will impact<br />

on the fission rate which will also be highly<br />

variable*/although it is possible that the power<br />

output obtained from the heated gases may not be<br />

as variable. This variability <strong>and</strong> chaotic unpredictability<br />

requires thorough investigation in order<br />

that the concept can be assessed.<br />

Power variability in such a system has been<br />

studied by van Dam et al., 1998 who investigated<br />

the sensitivity of the reactor to voidage fluctuations.<br />

<strong>The</strong>y concluded that, due to the slow<br />

neutron kinetics of the reactor (a consequence of<br />

the long neutron lifetime*/large scattering crosssections,<br />

Hetrick, 1993) the amplitude of the<br />

fission-power fluctuations would be small. <strong>The</strong><br />

present paper aims to check this conclusion with<br />

fully coupled transient fluidized bed simulations.<br />

<strong>The</strong> stability of the reactor is provided by the<br />

instantaneous negative reactivity temperature<br />

feedback in the coated fuel particles.<br />

Fluidized bed nuclear reactor concepts adopt<br />

some aspects of the pebble bed reactor Pebble bed<br />

reactor (Gerwin <strong>and</strong> Scherer, 1987) <strong>and</strong> the fuel<br />

particles are of a prefabricated design (Gulden <strong>and</strong><br />

Nickel, 1977). Optimization of this fuel particle is<br />

described in (Golovko et al., 1999). However, the<br />

concept investigated here is not the only fluidized<br />

concept; for example Sefidvash (1996) suggests<br />

fluidizing 0.2 cm diameter fuel particles with<br />

supercritical steam in a reactor designed to be<br />

non-fluctuating.<br />

<strong>The</strong> modelling approach we have developed<br />

applies detailed spatial/temporal modelling so<br />

that the reactor dynamics evolve naturally. This<br />

is in contrast to point kinetics models (Hetrick,<br />

1993) which, although often having adequate<br />

accuracy, require correlation with existing data<br />

when the material evolves within the transient,<br />

such as in fissile liquid transients, (Mather et al.,<br />

1994; Mather, 1991; Mather <strong>and</strong> Barbry, 1991)<br />

<strong>and</strong> nuclear fluidized beds.<br />

Others have used space-dependent kinetics models<br />

mostly to model transients in fissile liquids,<br />

such as the multi-region model of Kimpl<strong>and</strong> <strong>and</strong><br />

Korneich (1996), the finite difference model of<br />

Yamamoto (1995) <strong>and</strong> the nodal model of Rifat et<br />

al. (1993). However, there are a limited number<br />

point kinetics models available for powders, (see<br />

for example Rozain, 1991; Basoglu et al., 1994).<br />

Golovko et al. (2000a) investigated the nuclear<br />

fluidized bed (similar to the one studied here)<br />

using point kinetics models linked to expressions<br />

for heat loss <strong>and</strong> bed expansion, looking at start<br />

up transients of the reactor see <strong>and</strong> various<br />

accident scenarios such as loss of heat sink (coolant<br />

gas is not cooled adequately) <strong>and</strong> change of<br />

gas inlet temperature (Golovko et al., 2000c). <strong>The</strong><br />

model used in these studies is described in Golovko<br />

et al., (2000b).<br />

Without a doubt, the most satisfactory approach<br />

is an integrated neutrons/fluids/heat transfer<br />

method, such as that contained in the finite<br />

element transient criticality (FETCH) code (Pain<br />

et al., 2001b). <strong>The</strong> neutronics model solves the<br />

neutron Boltzmann transport equation in full<br />

phase-space using an second-order variational<br />

principle, (de Oliveira et al., 1998). <strong>The</strong> fluids<br />

algorithm is a multi-phase compressible/incompressible<br />

flow model which solves the conservation<br />

equations for both gas <strong>and</strong> solid particle phases.<br />

This unique fundamentally based combined methodology<br />

is potentially capable of modelling the<br />

complex non-linear reactivity feedback mechanisms<br />

which occur in nuclear reactor designs such as<br />

the one studied in this paper.<br />

Although no means of directly validating the<br />

overall FETCH model against experimental data<br />

is available, we have made every effort to validate<br />

the transient criticality (Pain et al., in press Pain et<br />

al., 2001b,d, 1998a) <strong>and</strong> the fluidized bed modelling<br />

(Pain et al., 2001a) individually, with careful<br />

comparison with experimental results for transient<br />

criticality in fissile solutions (Barbry, 1987; Ogawa<br />

et al., 1999). <strong>The</strong>se studies have provided a strong<br />

foundation from which to investigate a fluidized<br />

bed nuclear reactor.<br />

We have chosen to use the two-fluid granular<br />

temperature method of modelling which has a gas<br />

<strong>and</strong> a solid fluid phase. Within the solid phase,<br />

particle modelling is based on an analogy between<br />

the kinetic theory of gases <strong>and</strong> binary particle /<br />

particle collisions (Savage, 1983; Shahinpoor <strong>and</strong><br />

Ahmadi, 1983; Lun et al., 1984; Johnson <strong>and</strong><br />

Jackson, 1987; Jenkins <strong>and</strong> Savage, 1983; Chap-


man <strong>and</strong> Cowling, 1970). <strong>The</strong>se models are proving<br />

to be accurate for a wide range of gas /solid<br />

fluidization scenarios, (Cao <strong>and</strong> Ahmadi, 1995;<br />

Samuelsberg <strong>and</strong> Hjertager, 1996; Ding <strong>and</strong> Gidaspow,<br />

1990)<br />

<strong>The</strong> remainder of this paper is structured as<br />

follows: in the next section the FETCH coupled<br />

fluid dynamics/neutronics code is described. This<br />

is followed by a description of the reactor in<br />

Section 3 which also presents static modelling.<br />

Section 4 describes the transient modelling. Conclusions<br />

are drawn in the final section.<br />

2. <strong>The</strong> FETCH code<br />

<strong>The</strong> FETCH code is used here to simulate the<br />

dynamics of a nuclear fluidized bed. It is comprised<br />

of three modules: two transient 3D finite<br />

element modules*/the neutron transport code<br />

EVENT (de Oliveira, 1986) <strong>and</strong> the computational<br />

fluid dynamics (CFD)/multi-phase code FLUID-<br />

ITY (Mansoorzadeh et al., 1998), <strong>and</strong> an interface<br />

module which provides the coupling between<br />

neutronics <strong>and</strong> fluids.<br />

2.1. Neutronics<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245 227<br />

<strong>The</strong> Boltzmann neutron transport equation is<br />

solved using finite elements in space, spherical<br />

harmonics (PN) in angle, multi-group in energy<br />

<strong>and</strong> implicit two level time discretization methods.<br />

<strong>The</strong>se methods have been applied using the<br />

second-order even-parity variational principle in<br />

the EVENT computer code. Its lowest mode of<br />

angular resolution is equivalent to diffusion theory.<br />

Further details of the numerical formulation<br />

implemented in EVENT can be found in de<br />

Oliveira et al. (1998).<br />

At each time-step the interface module organizes<br />

the feedback from FLUIDITY of spatial temperature,<br />

density <strong>and</strong> delayed neutron precursor distributions<br />

into the EVENT neutronics module <strong>and</strong><br />

also, in the light of these fields, updates the spatial<br />

distribution of multi-group neutron cross-sections.<br />

For a given element of the finite element (FE)<br />

mesh, a cross-section set is obtained by interpolating<br />

in temperature <strong>and</strong> gas content a cross-section<br />

data-base. This database has been group-condensed<br />

taking into account resonance self shielding<br />

<strong>and</strong> thermal temperature effects, into six groups<br />

using the WIMS8A code (WIMS8A, 1999) <strong>and</strong> a<br />

representative geometry. <strong>The</strong> neutronics module<br />

generates for FLUIDITY spatial distributions of<br />

fission-power <strong>and</strong> delayed neutron generation<br />

rates.<br />

Material cross-sections are generated as follows<br />

using the lattice cell code WIMS8A. First the<br />

cross-sections were self-shielded using the equivalence<br />

theory method in WHEAD (part of WIMS)<br />

which relates the heterogeneous problem to an<br />

equivalent homogeneous model. A subgroup resonance<br />

calculation was then performed using the<br />

WPROC (part of WIMS) collision probability<br />

routine which calculates collision probabilities<br />

using a synthetic approximation for a system of<br />

spherical grains packed in annular geometry.<br />

<strong>Group</strong> cross-sections were then obtained for<br />

temperatures ranging from 550 to 2000 K by<br />

condensing to six groups the st<strong>and</strong>ard WIMS 69<br />

group library.<br />

2.2. Multi-phase fluids modelling<br />

Conservation equations for the particles <strong>and</strong> the<br />

helium gas are expressed in Eulerian form using a<br />

two phase continuum description. <strong>The</strong> momentum<br />

equations are discretized with an implicit nonlinear<br />

Petrov /Galerkin method, (Hughes <strong>and</strong><br />

Mallet, 1986), <strong>and</strong> the other conservation equations<br />

are solved using an implicit high resolution<br />

method which is globally second-order accurate in<br />

space <strong>and</strong> time, (Leonard, 1991). <strong>The</strong> second-order<br />

fluxes for the high resolution method are obtained<br />

from a finite element interpolation of the solution<br />

variables. <strong>The</strong>se methods are embodied in the<br />

CFD code FLUIDITY, (Pain et al., 2001c). <strong>The</strong><br />

delayed neutrons are solved for <strong>and</strong> transported in<br />

FLUIDITY <strong>and</strong> are passed to EVENT through<br />

the interface code.<br />

<strong>The</strong> governing equations which include delayed<br />

neutron precursor concentrations are listed in<br />

Table 1 <strong>and</strong> Table 2 <strong>and</strong> interfacial momentum<br />

<strong>and</strong> energy exchanges between phases are listed in<br />

Table 3. <strong>The</strong> convective <strong>and</strong> conductive heat<br />

transfer correlations used here are based on the


228<br />

Table 1<br />

Conservation equations used in the simulations<br />

Continuity equation @<br />

@t (okrk )<br />

@<br />

(okrk vki )<br />

@xi 0<br />

Momentum equation @<br />

@t (okrkn kt)<br />

@<br />

(okrkv kivkj) @xj @rg ok @xi okrkg i b(vki vki) @<br />

(tkij) @xi Gk <strong>The</strong>rmal energy<br />

equations<br />

DTg Cpg rgo g<br />

Dt<br />

pg @<br />

ogvgi @xi @<br />

osvsi @xi @<br />

@xi @Tg ogkg @xi a(Ts Tg) DTs Gwg; cps rsos Dt<br />

Granular energy<br />

equation<br />

3<br />

2<br />

@(oxr xU) @t<br />

@<br />

(osrs vsjU) @xj @vsi tsij @xj @qj @xj g 3bU<br />

Equation for d th<br />

delayed neutron<br />

group precursor<br />

oncentration<br />

@Cd (r; t)<br />

@t<br />

@vsj Cd (r; t)<br />

@xj ldCd (r; t) bdg0 v X<br />

f(r; E; t)dE<br />

f<br />

Neutron transport<br />

equation<br />

1 @c(r; V; E; t)<br />

v @t<br />

V×9c(r; V; E; t) Hc(r; V; E; t) S(r; V; E; t)<br />

work of Schmidt <strong>and</strong> Renz (1999), Molerus et al.<br />

(1995a,b), Natarajan <strong>and</strong> Hunt (1998), Hunt<br />

(1997), Hsiau (2000). Delayed neutron precursors<br />

are assumed to exist only in the solid phase <strong>and</strong> are<br />

in six delayed neutron precursor group form<br />

Duderstadt <strong>and</strong> Hamilton (1976). <strong>The</strong>rmal radiation<br />

heat transfer is neglected in the present study<br />

since its role would be to unify the temperature<br />

distribution in the reactor (Molerus et al., 1995a)<br />

<strong>and</strong> the calculated temperature distributions, see<br />

Section 4, havebeen found to be fairly homogeneous<br />

even without it. However, thermal radiative<br />

heat transfer may play a significant role in the<br />

course of the transient in accident scenarios in<br />

which large temperature differences could occur<br />

across the reactor.<br />

<strong>The</strong> fluids equations are solved only in the fluids<br />

occupied domain, shown in Fig. 1, which extends<br />

to a height of 500 cm. Because the particulate fuel<br />

does not exp<strong>and</strong> into the remainder of the 600 cm<br />

gas/fuel filled cavity, it is excluded from the fluids<br />

calculation domain. <strong>The</strong> boundary conditions at<br />

the inlet are, for the gas, a superficial velocity<br />

1<br />

normal to the inlet boundary of 120 cm s */<br />

Umf /25.0 cm s<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245<br />

1 at 6 MPa pressure <strong>and</strong><br />

230 8C according to the Ergun equation Table 2.<br />

Umf is the minimum fluidization gas velocity of the<br />

fuel particles. <strong>The</strong> gas was assumed to enter at<br />

230 8C <strong>and</strong> at a density dictated by a 6 MPa<br />

pressure. No heat loss conditions are applied at the<br />

vertical graphite walls of the reactor. Zero stress<br />

conditions are applied to the gas at the outlet<br />

boundary <strong>and</strong> on the walls slip <strong>and</strong> no normal<br />

flow conditions were applied. At the outlet (top<br />

plane of fluid domain) gas can enter depending on<br />

the evolving gas dynamics near the outlet, it is<br />

assumed that this gas is at 230 8C <strong>and</strong> 6 MPa<br />

pressure. To ensure the top boundary does not<br />

provide a external heat source the temperature <strong>and</strong><br />

pressure of any incomming gas at the top plane<br />

boundary must be set to equal to the inlet<br />

conditions. For the solid phase a specified shear<br />

stress condition was applied as described in Pain et<br />

al. (2001c) <strong>and</strong> no normal flow conditions were<br />

enforced. <strong>The</strong> granular temperature boundary<br />

conditions are described in Pain et al. (2001c)<br />

<strong>and</strong> we have assumed particle /particle, wall /<br />

particle <strong>and</strong> friction coefficients of 0.97, 0.9 <strong>and</strong><br />

0.1, respectively.<br />

All simulations are impulsively so that after the<br />

first time-step the gas inlet velocity is at 120.0<br />

cm s<br />

@<br />

@x i<br />

@Ts osks @xi a(T g T s) G wg S f<br />

1 . This is a stern test of the robustness of the<br />

nuclear fluidized bed because this initialization<br />

results in rapidly expansion of the bed <strong>and</strong> a<br />

corresponding rapid change in the nuclear criticality<br />

of the bed. That is the ramp reactivity


Table 2<br />

<strong>The</strong> two-fluid granular temperature constitutive equations used in the simulations<br />

Gas phase Newotnian viscous stress<br />

tgij 2ogmg 1<br />

2<br />

@vgi @xj @xi 1 @vgk 3 @xk Solid phase stress tensor<br />

tsij ps @vsk oszs @xk dij 2osms 1<br />

2<br />

Solids pressure ps osrs[1 2(1 e)osg0]U Solids shear viscosity<br />

zs 4<br />

3 osrsd sg0(1 e) U<br />

p<br />

0:5<br />

Radial distribution function<br />

1<br />

os 1<br />

3<br />

1<br />

Collisional energy dissipation<br />

Flux of fluctation energy<br />

insertion is potentially very large. In practice the<br />

reactor would be initiated with a gradual increase<br />

in fluidization velocity up to a maximum <strong>and</strong> one<br />

would therefore expect a smaller ramp reactivity<br />

insertion <strong>and</strong> so a smaller in magnitude initial<br />

fission response <strong>and</strong> therefore lower temperature.<br />

<strong>The</strong> finite element discretization <strong>and</strong> solution of<br />

the multi-phase flow equations are described in<br />

Pain et al. (2001c). In summary, this involves the<br />

use of a mixed finite element formulation with<br />

rectangular elements. Both velocity components<br />

are centered on the four nodes of the rectangle <strong>and</strong><br />

thus result in a bi-linear variation of velocity<br />

through out each element. Pressure, temperatures<br />

of both gas <strong>and</strong> solid phases, volume fractions,<br />

densities <strong>and</strong> delayed neutron precursor concentrations<br />

all have piece-wise constant variations,<br />

with a constant variation though out each element.<br />

An adaptive time-stepping method is used here<br />

<strong>and</strong> allows transient behavior of all fields to be<br />

resolved, (Pain et al., 2001d).<br />

3. <strong>The</strong> reactor<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245 229<br />

<strong>The</strong> reactor is drawn to scale in 3D in Fig. 1a<br />

with part of the reactor removed to reveal the<br />

g 0<br />

o i<br />

g 3(1 e 2 )o 2 r sg 0U 4<br />

q j<br />

2r so 2<br />

s g 0d s<br />

U<br />

p<br />

@v gj<br />

U<br />

p<br />

ds 0:5<br />

@U<br />

@xj 0:5<br />

@v s<br />

@x k<br />

@v si<br />

@x j<br />

@v sj<br />

@x i<br />

internal cavity. A schematic of the reactor is<br />

shown in Fig. 1b.<br />

<strong>The</strong> particles are formed in layers as detailed in<br />

Table 4. <strong>The</strong>y have a 300:1 moderator (in the form<br />

of carbon compounds) to uranium oxide fuel ratio.<br />

However, they are still under-moderated <strong>and</strong> thus<br />

the fuel responds with positive reactivity feedback<br />

to the additional moderation provided by the<br />

surrounding graphite walls. <strong>The</strong> reactors leakage<br />

<strong>and</strong> moderating properties must be such that as<br />

the bed height increases (due to fluidization) from<br />

maximum packing (under moderated) the criticality<br />

increases to a maximum <strong>and</strong> decreases on<br />

further expansion of the bed (over moderated).<br />

This provides a method of controlling the fission<br />

rate (power) with fluidizing flow rate <strong>and</strong> provides<br />

a safety mechanism for decreasing the criticality of<br />

the system in a rapid transient with rapid expansion<br />

of the gas along with bed height, (Golovko et<br />

al., 1999). In this demonstration we have chosen to<br />

impulsively start the gas flow to a fixed rate. This<br />

is likely to impulsively start the fission rate with a<br />

form that will depend in part on the level of the<br />

‘fixed’ neutron source. <strong>The</strong> power level might then<br />

expect to decrease as the bed temperature rises. No<br />

account is taken of fission product poisons influencing<br />

reactivity in this study.<br />

1<br />

3<br />

@v si<br />

@x k


230<br />

Table 3<br />

Correlations used in the simulations conducted<br />

Gas /solid friction coefficient<br />

Drag coefficient<br />

Convective heat transfer<br />

coefficient<br />

Conductive heat transfer<br />

coefficient<br />

<strong>The</strong> particles are chosen to be 0.1 cm in diameter<br />

which is small enough that the neutron flux<br />

distribution across each particle is fairly uniform<br />

resulting in uniform burning of the fuel, (Shmakov<br />

<strong>and</strong> Lyutov, 2000), <strong>and</strong> also allowing one to<br />

assume that the particles form a continuum for<br />

spatial homogenization <strong>and</strong> group collapsing purposes.<br />

This would be invalid for pebble bed<br />

reactors (Gerwin <strong>and</strong> Scherer, 1987) as the pebbles<br />

are typically of the order of 5.0 cm in diameter.<br />

<strong>The</strong> larger the particles in the bed, the larger the<br />

Table 4<br />

Material composition of TRISO fuel particle<br />

Material Density<br />

(g cm 3 )<br />

UO2 kernel 10.88 0.26<br />

Porus carbon buffer<br />

layer<br />

1.1 0.77<br />

PyC coating 1.9 0.85<br />

SiC coating 3.2 0.92<br />

PyC coating 1.9 1.00<br />

b<br />

C D<br />

o<br />

150<br />

2 s ms (1 os )d2 7 osrgjv g vsj s 4 ds 3<br />

4 C (1 os )osrg jvg vsj D<br />

(1 os )<br />

ds 2:65<br />

o<br />

20(0:225 os) 150<br />

2 s mg (1 os )d2 8<br />

><<br />

7 osrgjv g v<br />

>:<br />

sij<br />

15(os s 4 ds 24<br />

Rep (1 os ) f1 0:15[(1 os )Rep ]0:687 8<br />

><<br />

g if RepB1000 >: 0:44 if RepE1000 a 6o g<br />

Outer diameter<br />

(mm)<br />

<strong>The</strong> uranium is enriched to 16.76 wt.% with an overall<br />

particle density of 1.92 g cm 3 .<br />

o<br />

0:175)C<br />

srgjv g<br />

D<br />

hgs hgs dg kg [(7 10og 5o<br />

ds 2<br />

g )(I 0:7Re1=5 p Pr1=3 ) (1:33 2:4og 1:2o 2<br />

g )Re7=10 P Pr1=3 qffiffiffiffiffiffiffiffiffiffiffi<br />

]<br />

ogkg (1 1 og)kgas o s k s<br />

g? 0<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245<br />

ffiffiffiffiffiffiffiffiffi<br />

p<br />

osrs Cxds 3 p<br />

U<br />

16 7o s<br />

16(1 o s ) 2<br />

32g? 0<br />

flow rate required to fluidized them which can<br />

enhance heat transfer with the particles. However,<br />

larger particles have a smaller heat transfer rate<br />

because of the relatively small surface area per unit<br />

mass compared to small particles. Thus, some<br />

compromise is required. In addition, the particles<br />

must not be so large that the bed dynamics are<br />

those of very large slugs which will make the<br />

fission rate difficult to control. In fact the chosen<br />

particle are D particles in the Geldart classification<br />

<strong>and</strong> thus prone to producing large voids/slugs<br />

when fluidized with gas, (Geldart, 1986).<br />

3.1. Static modelling<br />

d s<br />

vsj (1 os) 1:65<br />

if o s<br />

0:225<br />

if o s 50:175<br />

if 0:175Bo s50:225<br />

<strong>The</strong> first step in obtaining an underst<strong>and</strong>ing of<br />

the reactor is to perform a series of Keff eigenvalue<br />

(criticality) calculations. <strong>The</strong> critical eigenvalue<br />

Keff provides an indication of the initial quantity<br />

of fuel required in the reactor <strong>and</strong> the feedback<br />

mechanism resulting from changes in temperature<br />

<strong>and</strong> fuel geometry. <strong>The</strong> mass of fuel particles used<br />

here is 1.619 /10 6 g which corresponds to a<br />

collapsed core height of 136.0 cm with a maximum<br />

packing factor of 0.62.


C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245 231<br />

Fig. 1. <strong>The</strong> fluidized bed nuclear reactor 3D domain <strong>and</strong> schematic: (a) 3D domain showing internal cavity; (b) 2D schematic of<br />

FLUBER reactor.<br />

Using the assumption that, as the fluidized bed<br />

exp<strong>and</strong>s, the fuel particles distribution remains<br />

uniform, Keff versus fluidized bed height is calculated<br />

<strong>and</strong> plotted in Fig. 2a. This graph confirms<br />

that changing flow rates, which change the bed<br />

expansion, may provide a means of controlling the<br />

power output of the reactor in addition to the<br />

inherent stabilization. <strong>The</strong> maximum temperature<br />

achievable would be that associated with the<br />

height at which Keff is at a maximum <strong>and</strong> the<br />

maximum power output would be at a height<br />

larger than this*/due to the power output being a<br />

function of the quantity of gas heated, that is the<br />

fluidization flow rate <strong>and</strong> therefore the height of<br />

the bed.<br />

<strong>The</strong> effect of changing temperature of the<br />

particles for a bed of height 172 cm is shown in<br />

Fig. 2b. <strong>The</strong> graph shows the strong negative<br />

reactivity feedback effect with increasing temperature<br />

which provides the main passive control of<br />

criticality. <strong>The</strong> temperature reactivity coefficient,<br />

which equals the gradient of the graph Fig. 2b at<br />

230 8C is /4.7 /10 5 K 1 . For neutronics<br />

purposes the temperature of the graphite moderator<br />

surrounding the inner core, is assumed to be<br />

230 8C in all static <strong>and</strong> transient calculations.


232<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245<br />

Fig. 2. Critical eigenvalue results with a fuel particle mass of 1.619 /10 6 g the height versus Keff (a) is for this constant mass (b) show<br />

the strong negative temperature coefficient: (a) K eff versus exp<strong>and</strong>ed core height; (b) K eff versus temperature.<br />

As well as providing information on reactivity<br />

feedback effects, the static calculations also provide<br />

an indication of the power distributions in the<br />

reactor. Since most of the fissions occur in the<br />

thermal groups, the scalar flux distribution of<br />

thermal group 6, for the eigenvalue calculation,<br />

shows that much of the fission energy is deposited<br />

next to the graphite walls from which thermalized<br />

Fig. 3. <strong>The</strong>rmal (a) <strong>and</strong> fast (b) neutron scalar flux contours for a fuel particle mass of 1.619 /10 6 g <strong>and</strong> a collapsed bed height of<br />

136.0 cm. (c) finite element mesh used in both transient <strong>and</strong> eigenvalue calculations. <strong>The</strong> whole computational domain consists of 2000<br />

quadrilateral elements <strong>and</strong> 2121 nodes, the fluids domain contains 750 quadrilateral elements <strong>and</strong> 656 nodes. <strong>The</strong> central axis of the<br />

axi-symmetric model is on the left h<strong>and</strong> side of the diagrams.


C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245 233<br />

Fig. 4. Fission rate <strong>and</strong> cumulative fissions for the three simulations with differing fissile mass content in the reactor: (a) low power; (b)<br />

intermediate power; (c) high power.<br />

neutrons emanate, see Fig. 3a. As one might<br />

expect this flux distribution is very similar to the<br />

particle importance map in the central cavity,<br />

shown in van der Hagen et al. (1997), which<br />

estimates the importance to criticality of a particle<br />

at a given position in the reactor. <strong>The</strong> fast group,<br />

group 1, flux distribution is shown in Fig. 3b.<br />

4. Transient modelling<br />

<strong>The</strong> aim of this section is to report the reactor<br />

dynamics when the power is allowed to evolve. To<br />

this end, we present three transient simulation<br />

results with differing particle mass (fuel mass)<br />

content in the reactor of 1.809 /10 6 , 1.735 /10 6<br />

Fig. 5. Maximum temperature <strong>and</strong> temperature at three sensors at the bottom of the reactor. High power simulation: (a) maximum<br />

<strong>and</strong> central temperature; (b) temperature at the two sensors.


234<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245<br />

Fig. 6. Selected fields at 6 s (just after fission spike) into the simulation with high power. <strong>The</strong> central axis of the axi-symmetric model is<br />

on the left h<strong>and</strong> side of the diagrams: (a) solids fraction; (b) gas temperature (8C); (c) third longest-lived delayed concentration (cm 3 );<br />

(d) shortest-lived delayed concentration (cm 3 ).<br />

<strong>and</strong> 1.661 /10 6 g which corresponds to a collapsed<br />

core height, assuming a maximum packing<br />

factor of 0.62, 152.0, 145.6 <strong>and</strong> 139.2 cm, respectively.<br />

<strong>The</strong>se three simulations will be referred to<br />

Fig. 7. Maximum pressure deviation from 6 MPa overpressure <strong>and</strong> velocity of both phases for the high power simulation: (a)<br />

maximum pressure deviation; (b) maximum velocity.


C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245 235<br />

Fig. 8. Maximum temperature <strong>and</strong> temperature at three sensors at the bottom of the reactor. Intermediate power simulation: (a)<br />

maximum <strong>and</strong> central temperature; (b) temperature at the two sensors.<br />

as high power, intermediate power <strong>and</strong> low power.<br />

All three transients are initiated with zero neutron<br />

fluxes <strong>and</strong> have a fixed source of 0.3 neutrons<br />

cm 3 1<br />

s in each of the six neutron energy groups<br />

<strong>and</strong> in the lower 172.0 cm of the inner cavity. Each<br />

simulation took approximately 2 weeks on a 500<br />

MHz Compaq AXP1000 workstation in single<br />

precision.<br />

Fig. 9. Maximum temperature <strong>and</strong> temperature at three sensors at the bottom of the reactor. Low power simulation: (a) maximum <strong>and</strong><br />

central temperature; (b) temperature at the two sensors.


236<br />

4.1. Fission-power <strong>and</strong> temperature<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245<br />

Fig. 10. Particle volume fraction at the three sensors versus time into the simulation for (a) low power; (b) intermediate power; <strong>and</strong> (c)<br />

high power reactor fuel loading.<br />

Fig. 4a /c show the fission-power in fissions per<br />

second (3.2 /10 11 J /1 fission) together with the<br />

cumulative fissions for the three cases. For the<br />

high power case, there is a large fission peak which<br />

occurs, 4 s after, initiation of gas flow. <strong>The</strong> large<br />

magnitude of the fission peak heats the fuel<br />

particles along with fluidizing gases to a maximum<br />

temperature of 1200 8C, see Fig. 5a. This results<br />

Fig. 11. A comparison of fission rate versus time for the high fuel loading case with a large <strong>and</strong> a small neutron source. <strong>The</strong><br />

corresponding maximum gas temperatures for the two simulations is also shown: (a) comparison of fission rates; (b) comparison of<br />

maximum temperatures.


C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245 237<br />

in strong negative temperature reactivity feedback<br />

(via spectral shift <strong>and</strong> Doppler broadening, (Duderstadt<br />

<strong>and</strong> Hamilton, 1976) which reduces the<br />

fission rate <strong>and</strong> thus the temperature gradually<br />

decreases, see Fig. 5a, as the cooling gas extracts<br />

heat from the particles. This rapid <strong>and</strong> large<br />

deposition of heat energy exp<strong>and</strong>s the cooling<br />

gas <strong>and</strong> results in a rapid expansion of the bed, see<br />

Fig. 6a, which also has a negative feedback effect,<br />

see Fig. 2a. Fig. 6 show the (a) solids fraction, (b)<br />

temperature, (c) third longest-lived delayed neutron<br />

concentration <strong>and</strong> (d) shortest-lived delayed<br />

neutron concentration at 6 s into the transient, at<br />

the bed’s most exp<strong>and</strong>ed state. <strong>The</strong> maximum<br />

velocity of the particles <strong>and</strong> gas appears to increase<br />

after the fission spike, see Fig. 7b, along with the<br />

maximum pressure deviation from the 6 MPa over<br />

pressure, Fig. 7a. <strong>The</strong> small pressure deviations<br />

suggest that gas density differences are mostly<br />

attributed to temperature changes.<br />

<strong>The</strong> intermediate power simulation shows a<br />

much smaller fission peak <strong>and</strong> maximum temperature,<br />

see Fig. 8a. This temperature quickly decreases<br />

as the reactor approaches a quasi steadystate.<br />

<strong>The</strong> frequency spectrum of the fission rate,<br />

for this intermediate power case, shows a dominant<br />

frequency of 1 Hz. Although, as with all the<br />

simulations, the fission rate (power) oscillates<br />

vigorously <strong>and</strong> by about an order of magnitude,<br />

the temperature of the bed varies smoothly, see<br />

maximum temperature versus time graphs Figs. 5<br />

<strong>and</strong> 8 <strong>and</strong> Fig. 9, which is a gauge of the steadiness<br />

of the energy output of the reactor. This suggests<br />

that the power extracted from the gas would be<br />

steady also. To generate the temperature versus<br />

time graphs, Figs. 5 <strong>and</strong> 8 <strong>and</strong> Fig. 9, <strong>and</strong> the<br />

particle volume fraction versus time graphs, Fig.<br />

10, three sensors were placed in the bottom of the<br />

reactor cavity. <strong>The</strong>se are labelled; bottom center<br />

which refers to the sensor situated along the<br />

Fig. 12. Selected fields at 40 s into the simulation with low power. <strong>The</strong> central axis of the axi-symmetric model is on the left h<strong>and</strong> side<br />

of the diagrams: (a) solids fraction; (b) gas temperature (8C); (c) third longest-lived delayed concentration (cm 3 ).


238<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245<br />

Fig. 13. Selected fields at 70 s into the simulation with intermediate power. <strong>The</strong> central axis of the axi-symmetric model is on the left<br />

h<strong>and</strong> side of the diagrams: (a) solids fraction; (b) Gas temperature (8C); (c) third longest-lived delayed concentration (cm 3 ).<br />

central axis, bottom corner which refers to the<br />

sensor in the corner of the domain where the<br />

vertical walls meet the floor of the cavity, <strong>and</strong><br />

bottom mid-center sensor which is positioned half<br />

way between the bottom center <strong>and</strong> bottom corner<br />

sensors.<br />

<strong>The</strong> lowest power simulation produces enough<br />

fission energy to heat up the reactor at about 20 s<br />

into the simulation. This is due to the smallness of<br />

criticality <strong>and</strong> suggests (as would usually be good<br />

practice) that a larger neutron source is required at<br />

reactor start up. <strong>The</strong> temperature has no large<br />

peak, see Fig. 4, <strong>and</strong> seems to quickly reach a quasi<br />

steady-state*/in a time averaged sense. It is<br />

recognized that, by analogy with a continuous<br />

filling fissile solution criticality the initial form of<br />

the power rise will depend strongly on the fixed<br />

source density, (Pain et al., 1998b). We investigate<br />

its effect here on the coarse of the high power<br />

transient by repeating the high powered case with<br />

the source strength increased to 3 /10 3 neutrons<br />

cm 3 1<br />

s . <strong>The</strong> resulting fission rate <strong>and</strong> maximum<br />

temperature versus time graphs are compared<br />

in Fig. 11a <strong>and</strong> b, respectively.<br />

Notice that the fission peak is much smaller for<br />

the case with the larger source. This is because the<br />

reactor undergoes a ramp reactivity insertion due<br />

to the movement of the particles in the bed. <strong>The</strong><br />

larger the source the quicker the neutron population<br />

builds up, during this ramp, to which<br />

increases the bed temperature. <strong>The</strong> negative temperature<br />

reactivity feedback effects then stabilize<br />

the temperature. Thus, the excess reactivity at the<br />

point at which the initial fission spike occurs<br />

governs the initial power of the system. As one<br />

would expect the temperature of the bed for the<br />

simulation with the source is much smaller, see<br />

Fig. 11b.<br />

<strong>The</strong> unsteadiness of the reactor is seen in the<br />

particle volume fractions at three sensors placed at


C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245 239<br />

Fig. 14. Selected fields at 40 s into the simulation with high power. <strong>The</strong> central axis of the axi-symmetric model is on the left h<strong>and</strong> side<br />

of the diagrams: (a) solids fraction; (b) gas temperature (8C); (c) third longest-lived delayed concentration (cm 3 ).<br />

Fig. 15. Relationship between outlet temperature <strong>and</strong> power output from the three simulations: (a) power versus outlet temperature;<br />

(b) reactor power output


240<br />

the bottom of the reactor, see Fig. 10. <strong>The</strong><br />

temperature for all three simulations is plotted at<br />

the sensors at the bottom of the bed, in Figs. 5 <strong>and</strong><br />

8 <strong>and</strong> Fig. 9. <strong>The</strong> quickest variability in temperature<br />

is observed at the bottom of the bed, due to<br />

the cooling influence of the incoming helium gas.<br />

<strong>The</strong> initial temperature rise is fairly rapid for all<br />

three cases <strong>and</strong> since much of the heat is deposited<br />

in the bottom outer edge of the reactor (as<br />

indicated by the thermal flux distribution, Fig.<br />

3a) this is the point at which the temperature is<br />

largest as the temperature initially rises, Fig. 5b,<br />

Fig. 8b <strong>and</strong> Fig. 9b. However, on larger time scales<br />

advection takes place <strong>and</strong> thus this is no longer the<br />

case*/as seen in these figures. <strong>The</strong> solid phase<br />

temperature (not shown here) is very similar to the<br />

gas phase temperature, indicating rapid gas /solid<br />

heat transfer rates.<br />

Increasing the power output of the reactor<br />

increases the height to which the particles fluidize,<br />

due to the exp<strong>and</strong>ing fluidizing gases with temperature,<br />

compare Fig. 12a, Fig. 13a <strong>and</strong> Fig. 14a.<br />

4.2. Gas power output<br />

<strong>The</strong> power (fission rate), although highly variable<br />

by an order of magnitude, deposits much of<br />

its energy into the particles as they move into the<br />

vicinity of the bottom outer edge of the reactor. In<br />

this way the maximum temperature for all cases is<br />

fairly steady <strong>and</strong> is another reason why the<br />

temperature distribution is fairly uniform. A<br />

consequence of this uniformity is that the heat<br />

transfer coefficient for the reactor as a whole,<br />

f Gout<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245<br />

r g C g DT g u z dG; which is a measure of the power<br />

output of the reactor, is fairly steady, see Fig. 15b,<br />

for all three reactors. In which Gout is the top<br />

outlet boundary of the fluids domain, DTg is the<br />

deviation of the gas outlet temperature from the<br />

inlet temperature <strong>and</strong> uz is the normal velocity<br />

component to the outlet boundary. Occasionally,<br />

the heat flux from the gas as shown in Fig. 15b<br />

oscillates because relatively cool gas (at 230 8C) is<br />

dragged into the domain along G out, increasing the<br />

hot gas flow rate out of the system <strong>and</strong> thus<br />

resulting in a peak in heat flux output. This peak is<br />

followed by a dip in the heat flux as these cool<br />

gases are expelled. Thus this oscillation is due to<br />

the restricted domain size <strong>and</strong> boundary conditions.<br />

<strong>The</strong> heat transfer rate out of a reactor system is<br />

perhaps more accurately estimated from the maximum<br />

temperature versus time graphs, shown in<br />

Fig. 5a, Fig. 8a <strong>and</strong> Fig. 9a. Using these temperatures<br />

combined with the graph of the power output<br />

of the reactor versus its temperature (Fig. 15),<br />

gives the steady heat flux out of the system. In a<br />

time averaged sense this will equal the heat flux<br />

out of the system given by Fig. 15b. At the end of<br />

the three simulations the power output is 23.0<br />

MWt (34.5 KW kgU 1 ), 11.0 MWt (17.3<br />

KW kgU 1 ) <strong>and</strong> 6.0 MWt (9.8 KW kgU 1 ) for<br />

the high power, intermediate power <strong>and</strong> low power<br />

simulations, respectively. Typical power outputs of<br />

commercial reactors are: 3600.0 MWt (37.9<br />

KW kgU 1 ) for pressurized water reactors;<br />

3579.0 MWt (25.9 KW kgU 1 ) for boiling water<br />

reactors <strong>and</strong> 3000.0 MWt (77.0 KW kgU 1 ) for<br />

high-temperature gas reactors (Duderstadt <strong>and</strong><br />

Hamilton, 1976). Due to the large variablility in<br />

the designs of these reactors these power outputs<br />

are meant only as a rough guide.<br />

4.3. Fission heat source<br />

<strong>The</strong> shortest-lived neutron precursor concentration<br />

distributions at a quasi steady-state, for all<br />

three simulations Fig. 12d, Fig. 13d <strong>and</strong> Fig. 14d<br />

provide an indication of the instantaneous power<br />

distribution which is at a maximum near the<br />

bottom outer edge <strong>and</strong> vertical wall of the reactor,<br />

again as indicated by the thermal flux distribution<br />

in static criticality, see Fig. 3a. <strong>The</strong> delayed<br />

neutron precursor concentrations, with half lives<br />

of 0.18, 0.50, 2.2, 6.0, 22 <strong>and</strong> 55 s (Duderstadt <strong>and</strong><br />

Hamilton, 1976), provide an indication of time<br />

averaged (averaged over time scale of half-life)<br />

heat source. Delayed neutron precursors are unstable<br />

fission products, which are advected with<br />

the particles <strong>and</strong> on decay result in a neutron<br />

emission. For modelling purposes it is convenient<br />

to lump the precursors into a small number of<br />

delayed groups each with a characteristic half-life.<br />

A small fraction, 0.7%, of fissions are delayed<br />

which provides a means of controlling the power<br />

variation of this <strong>and</strong> all other nuclear reactors.


C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245 241<br />

Fig. 16. Selected time-averaged fields-averaged over 50 /70 s of the intermediate power simulation. <strong>The</strong> central axis of the axisymmetric<br />

model is on the left h<strong>and</strong> side of the diagrams: (a) solids fraction; (b) vertical gas velocity (cm.s<br />

1<br />

); (c) vertical particle<br />

velocity (cm s<br />

1<br />

); (d) shortest-lived delayed concentration (cm<br />

3<br />

).<br />

Since the delayed precursor generation rate is<br />

approximately proportional to the fission rate<br />

(power), the delayed neutron concentration can<br />

be viewed as time averaged heat sources, over time<br />

scales associated with the decay rate.<br />

<strong>The</strong> third longest-lived delayed neutron concentration<br />

distributions (half-life of 6 s), Fig. 12c, Fig.<br />

13c <strong>and</strong> Fig. 14c provide an indication of the<br />

history of the particles <strong>and</strong> also evidence to suggest<br />

that in the three simulations all particles have been<br />

subject to approximately the same heat source<br />

from fissions over a time scale of 6 s. In the three<br />

simulations the second <strong>and</strong> longest-lived delayed<br />

precursor concentration distributions are also very<br />

similar to the particle concentrations. Thus the<br />

longest-lived delayed neutron concentrations will<br />

reflect the particle concentrations, as seen in Fig.<br />

12c, Fig. 13c <strong>and</strong> Fig. 14c when the particles are<br />

subject to the same heat source. This similarity<br />

between time averaged heat source <strong>and</strong> particle<br />

concentration can only come about from the<br />

movement of the particles around the bed <strong>and</strong><br />

through areas of large heat source. <strong>The</strong> uniformity<br />

of the gas phase temperature distribution throughout<br />

the bed, see Fig. 12b, Fig. 13b <strong>and</strong> Fig. 14b, is<br />

a result of the uniformity of this heating (over a 6 s<br />

time scale) <strong>and</strong> also the rapid gas movement<br />

through the bed.<br />

4.4. Time averaged results<br />

Particles, in a time averaged sense, move down<br />

the center of the reactor <strong>and</strong> up the sides, see Fig.<br />

16c. <strong>The</strong>se time averaged results were obtained<br />

from the intermediate power calculation <strong>and</strong><br />

averaged over the final 20 s of the simulation.<br />

This particle recirculation provides the global<br />

mixing mechanism. However, this flow is contrary<br />

to that typically observed in fluidized beds, <strong>and</strong> so<br />

is the large accumulation of particles near the


242<br />

center of the reactor, see Fig. 16a. <strong>The</strong>se are a<br />

result of imposing axi-symmetry on the flow. <strong>The</strong><br />

time averaged shortest-lived delayed neutron precursor,<br />

Fig. 16d, reflects the time averaged power<br />

distribution of the reactor <strong>and</strong> shows that as well<br />

as being peaked near the walls, the power is also<br />

peaked along the central axis, due to the large<br />

density of particles near the center. <strong>The</strong> time<br />

average vertical gas velocity distribution is shown<br />

in Fig. 16b, which shows that a gas circulation is<br />

set up in the bed which provides an additional<br />

method of transporting particle heat <strong>and</strong> unifying<br />

the temperature of the bed.<br />

5. Conclusions<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245<br />

In this work we have demonstrated how a<br />

nuclear fluidized bed reactor can be modelled<br />

using space-dependent kinetics. It was shown<br />

here how increasing the fuel content in the reactor<br />

increases the temperature <strong>and</strong> power output of the<br />

reactor <strong>and</strong> as a consequence exp<strong>and</strong>s the gas <strong>and</strong><br />

results in an increase in freeboard height. <strong>The</strong><br />

superb mixing abilities of the fluidized bed were<br />

demonstrated with the uniformity in the calculated<br />

temperature distributions. This uniformity enhances<br />

the heat transfer out of the bed <strong>and</strong><br />

therefore the power output of the reactor.<br />

<strong>The</strong> modelled fission-power varied by about an<br />

order of magnitude however, the gas temperature<br />

was fairly steady after the initial transient. Thus<br />

the power output extracted from the simulated<br />

beds would be relatively steady. <strong>The</strong> fission-power<br />

fluctuations are large <strong>and</strong> further work is required<br />

to eliminate the possibility that they might lead to<br />

an uncontrolled criticality excursion. <strong>The</strong> simulations<br />

were conducted over a relatively short period<br />

of time, namely tens of seconds, <strong>and</strong> to use more<br />

reasonable time variation of coolant gas inflow<br />

rate. <strong>The</strong>re is thus a need to conduct similar<br />

investigations over larger time scales*/minutes.<br />

In addition, it was observed that all particles were,<br />

remarkably, exposed to approximately the same<br />

heat source quantity, over a short time scale of 6 s.<br />

Although, the axis-symmetric model used in this<br />

investigation results in an unrealistic accumulation<br />

of particles along the central axis we believe the<br />

model does provide an insight into the complex<br />

dynamics of this reactor. Future work will involve<br />

using a 3D model as well as a range of other<br />

models, including this one, to further investigate<br />

the dynamics of the reactor. In the future these<br />

models should be useful in further optimizing the<br />

design of this <strong>and</strong> other reactors.<br />

Acknowledgements<br />

Mr Gomes is supported by CAPES/Brazil <strong>and</strong><br />

by UERJ(PROCASE)/Brazil.<br />

Appendix A: Nomenclature<br />

v velociy, m s<br />

t time, s<br />

r position vector<br />

x coordinate<br />

g gravitational constant, m s<br />

2<br />

p pressure, Pa (N m 2 )<br />

Cp specific heat capacity, J kg 1 K 1<br />

T temperature, K<br />

q flux of fluctuation energy,<br />

kg m 1 s<br />

3<br />

CD drag coefficient<br />

g0<br />

radial distribution function<br />

g 0?<br />

radial distribution function for effective<br />

conductivity<br />

e particle /particle restitution coefficient<br />

d diameter, m<br />

Re Reynolds number<br />

Pr Pr<strong>and</strong>tl number<br />

h fluid-particle heat transfer coefficient,<br />

W m 2 K 1<br />

Cd(r, t) dth delayed group precursor concentration<br />

ld<br />

decay constant (b decay) of dth<br />

precursor group, s<br />

1<br />

bd fraction of all fission neutrons (both<br />

prompt <strong>and</strong> delayed) emitted per<br />

fission that appear from the dth<br />

precursor group<br />

f(r, E, t) neutron scalar flux, cm 2 s<br />

1<br />

eV<br />

1<br />

1


f(r, V, E, t) neutron angular flux,<br />

cm 2 s<br />

1<br />

eV<br />

1<br />

sr<br />

1<br />

E neutron energy (eV)<br />

Sf(r, t) fission heat source, cm 3 s<br />

1<br />

S(r, V, E, t) neutron source,<br />

cm 3 s<br />

1<br />

eV<br />

1<br />

sr<br />

1<br />

Sf(r, t) macroscopic fission cross-section,<br />

cm 1<br />

/H/ scattering-removal operator<br />

MWt megawatt thermal<br />

KW kgU 1 kilowatt per kilogram of Uranium<br />

minimum fludization velocity<br />

Umf<br />

Greek symbols<br />

o volume fraction<br />

r density, kg m 3<br />

b interphase drag constant,<br />

kg m 3 s<br />

t viscous stress tensor, N m 2<br />

G frictional force exerted on the wall<br />

by the phase, N m 4 s<br />

U granular temperature, m 2 s<br />

2<br />

g collisional energy dissipation,<br />

kg m 1 s<br />

m viscosity, N s m 2<br />

z bulk viscosity, N s m 2<br />

k thermal conductivity, W m 1 K 1<br />

V direction of neutron travel<br />

Subscripts<br />

k phase (g, gas; s, solid)<br />

i, j x, y-directions<br />

p particle<br />

w wall<br />

gas pure gas<br />

References<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245 243<br />

3<br />

Barbry, F., 1987. Fissile solution criticality accidents-review of<br />

pressure wave measurements experiments in the SILENE<br />

reactor. Institut de Protection et de Surete Nucleaire,<br />

Technical note SRSC No. 87.96.<br />

Basoglu, B., Brewer, R.W., Haught, C.F., Hollenbach, Wilkenson,<br />

A.D., Dodds, H.L., Pasqua, P.F., 1994. Simulation<br />

of hypothetical criticality accidents involving homogeneous<br />

damped low-enriched UO2 powder systems. Nuclear Technology<br />

105, 14 /30.<br />

1<br />

Chapman, S., Cowling, T.G., 1970. <strong>The</strong> Mathematical <strong>The</strong>ory<br />

of Non-uniform Gases. Cambrige University Press, Cambridge,<br />

UK.<br />

Cao, J., Ahmadi, G., 1995. Gas-particle two-phase turbulent<br />

flow in a vertical duct. International Journal of Multiphase<br />

Flow 21, 1203 /1228.<br />

Davidson, J.F., Clift, R., Harrison, D., 1985. Fluidization.<br />

Academic Press, London.<br />

de Oliveira, C.R.E., 1986. An arbitrary geometry finite element<br />

method for multigroup neutron transport with anisotropic.<br />

Progress in Nuclear Energy 18, 227 /236.<br />

de Oliveira, C.R.E., Pain, C.C., Goddard, A.J.H., 1998. <strong>The</strong><br />

finite element method for time-dependent radiation transport<br />

applications. Proceedings of the 1998 Radiation<br />

Protection <strong>and</strong> Shielding Topical Conference, Nashville,<br />

USA, 343.<br />

Ding, J., Gidaspow, D., 1990. A bubbling fluidization model<br />

using kinetic theory of granular flow. A.I.Ch.E. 36, 523 /<br />

538.<br />

Duderstadt, J.J., Hamilton, L.J., 1976. Nuclear Reactor<br />

Analysis. Wiley, New York.<br />

Geldart, D., 1986. Gas Fluidization Technology. Wiley, Chichester,<br />

UK.<br />

Gerwin, H., Scherer, W., 1987. Treatment of the upper cavity in<br />

a pebble-bed high temperature gas-cooled reactor by diffusion<br />

theory. Nuclear Science <strong>and</strong> Engineering 97, 9 /19.<br />

Golovko, V.V., Kloosterman, J.L., van Dam, H., van der<br />

Hagen, T.H.J.J., 1999. Fuel particle design for a fluidized<br />

bed reactor. Proceedings of Jahrestagung Kerntechnik ’99,<br />

Annual Meeting on Nuclear Technology ’99, Karlsruhe,<br />

Germany, 625 /628.<br />

Golovko, V.V., Kloosterman, J.L., van Dam, H., van der<br />

Hagen, T.H.J.J., 2000a. Investigation of a hypothetical<br />

start-up transient of a fluidized bed nuclear reactor.<br />

Proceedings of Jahrestagung Kerntechnik 2000, Annual<br />

meeting on Nuclear Technology 2000, Bonn, Germany.<br />

Golovko, V.V., Kloosterman, J.L., van Dam, H., van der<br />

Hagen, T.H.J.J., 2000b. Dynamic core stability analysis of a<br />

fluidized bed nuclear reactor, PHYSOR 2000, Pittsburg,<br />

Pennsylvania, USA.<br />

Golovko, V.V., Kloosterman, J.L., van Dam, H., van der<br />

Hagen, T.H.J.J., 2000c. Analysis of transients in a fluidized<br />

bed nuclear reactor, PHYSOR 2000, Pittsburg, Pennsylvania,<br />

USA.<br />

Gulden, T.D., Nickel, H., 1977. Preface coated particle fuels.<br />

Nuclear Technology 35, 206 /213.<br />

Hetrick, D.L., 1993. Dynamics of nuclear reactors. 555 N,<br />

Kensington Avenue, La Grange Park, Illinois 60525 USA:<br />

American Nuclear Society.<br />

Hughes, T.J.R., Mallet, M., 1986. A new finite element<br />

formulation for computational fluid dynamics: IV. A<br />

discontinuity-capturing operator for multi-dimensional advection-diffusion<br />

systems. Computing Methods in <strong>Applied</strong><br />

Mechanics <strong>and</strong> Engineering 58, 329 /336.<br />

Hunt, M.L., 1997. Discrete element simulations for granular<br />

material flows: effective thermal conductivity <strong>and</strong> self-


244<br />

C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245<br />

diffusity. International Journal of Heat <strong>and</strong> Mass Transfer<br />

40, 3059 /3068.<br />

Hsiau, S.S., 2000. Effective thermal conductivities of a single<br />

species <strong>and</strong> a binary mixture of granular materials. International<br />

Journal of Multiphase Flow 26, 83 /97.<br />

Jenkins, J.T., Savage, S.B., 1983. A theory for the rapid flow of<br />

identical, smooth, nearly elastic spherical particles. Journal<br />

of Fluid Mechanics 130, 187 /202.<br />

Johnson, Jackson, P.C., 1987. Frictional-collisional constitutive<br />

relations for granular materials with application to plane<br />

shearing. Journal of Fluid Mechanics 176, 67 /93.<br />

Kimpl<strong>and</strong>, R.H., Korneich, D.E., 1996. A two-dimensional<br />

multi-region computer model for predicting nuclear excursions<br />

in aqueous homogeneous assemblies. Nuclear Science<br />

<strong>and</strong> Engineering 122, 204 /211.<br />

Leonard, B.P., 1991. <strong>The</strong> ULTIMATE conservative difference<br />

scheme applied to unsteady one-dimensional advection.<br />

Computing Methods in Apllied Mechanics <strong>and</strong> Engineering<br />

88, 17 /74.<br />

Lun, C.K.K., Savage, S.B., Jeffrey, D.J., Chepurniy, N., 1984.<br />

Kinetic theories for granular flow: inelastic particles in<br />

couette flow <strong>and</strong> slightly inelastic particles in a general flow<br />

field. International Journal of Multiphase Flow 140, 223 /<br />

256.<br />

Mansoorzadeh, S., Pain, C.C., de Oliveira, C.R.E., Goddard,<br />

A.J.H., 1998. Finite element simulations of incompressible<br />

flow past a heated/cooled sphere. International Journal for<br />

Numerical Methods in Fluids 28, 903.<br />

Mather, D., Barbry, F., 1991. Examination of some fissile<br />

solution scenarios using CRITEX, Proceedings of the<br />

Fourth International Conference on Nuclear Criticality<br />

Safety, Oxford, UK.<br />

Mather, D., 1991. Validation of the CRITEX code. Proceedings<br />

of the Fourth International Conference on Nuclear Criticality<br />

Safety, Oxford, UK.<br />

Mather, D., Buckley, A., Prescott, A., 1994. CRITEX-a code to<br />

calculate the fission release arising from transient criticality.<br />

AEA Report CS/R1007/R.<br />

Molerus, O., Burschka, A., Dietz, S., 1995aa. Particle migration<br />

at solid surfaces <strong>and</strong> heat transfer in bubbling fluidized<br />

beds-I: particle migration measurement systems. Chemical<br />

Engineering Science 50, 871 /877.<br />

Molerus, O., Burschka, A., Dietz, S., 1995bb. Particle migration<br />

at solid surfaces <strong>and</strong> heat transfer in bubbling fluidized<br />

beds-II: prediction of heat transfer in bubbling fluidized<br />

beds. Chemical Engineering Science 50, 879 /885.<br />

Molerus, O., Wirth, K.E., 1997. Heat Transfer in Fluidized<br />

Beds. Chapman & Hall, London.<br />

Natarajan, V.V.R., Hunt, M.L., 1998. Kinetic theory analysis<br />

of heat transfer in granular flows. International Journal of<br />

Heat <strong>and</strong> Mass Transfer 41, 1929 /1944.<br />

Ogawa, K., Nakajima, K., Yanagisawa, H., Sono, H., Aizawa,<br />

E., Morita, T., Sugawara, S., Sakuraba, K., Ohno, A., 1999.<br />

Measurment of power profile during nuclear excursions<br />

initiated by various reactivity additions using tracy. Proceedings<br />

of the Sixth International Conference on Nuclear<br />

Criticality Safety, Versailles, France.<br />

Pain, C.C, Mansoorzadeh, S., de Oliveira, C.R.E., 2001aa. A<br />

study of bubbling <strong>and</strong> slugging fluidized beds using the twofluid<br />

granular temperature model. International Journal of<br />

Multiphase Flow 27, 527 /551.<br />

Pain, C.C., de Oliveira, C.R.E., Goddard, A.J.H., Umpleby,<br />

A.P., Criticality behaviour of dilute plutonium solutions.<br />

Nuclear Science <strong>and</strong> Technology, in press.<br />

Pain, C.C., de Oliveira, C.R.E., Goddard, A.J.H., Umpleby,<br />

A.P., 2001bb. Transient criticality in fissile solutions*/<br />

compressibility effects. Nuclear Science <strong>and</strong> Engineering<br />

138, 78 /95.<br />

Pain, C.C., Mansoorzadeh, S., de Oliveira, C.R.E., Goddard,<br />

A.J.H., 2001cc. Numerical modelling of gas-solid fluidized<br />

beds using the two-fluid approach. International Journal of<br />

Numerical Methods in Fluids 36, 91 /124.<br />

Pain, C.C., de Oliveira, C.R.E., Goddard, A.J.H., 2001dd.<br />

Non-linear space-dependent kinetics for criticality assessment<br />

of fissile solutions. Progress in Nuclear Energy 39, 53 /<br />

114.<br />

Pain, C.C., Goddard, A.J.H., de Oliveira, C.R.E., 1998a. <strong>The</strong><br />

finite element transient criticality code FETCH-verification<br />

<strong>and</strong> validation. Proceedings of the Second NUCEF International<br />

Symposium on Nuclear Fuel Cycle, Hitachinaka,<br />

Ibaraki, Japan, 139.<br />

Pain, C.C., de Oliveira, C.R.E., Goddard, A.J.H., 1998b.<br />

<strong>Modelling</strong> the criticality consequences of free surface<br />

motion in fissile liquids. Proceedings of the Second NUCEF<br />

International Symposium on Nuclear Fuel Cycle, Hitachinaka,<br />

Ibaraki, Japan.<br />

Rifat, M., Al-Chalabi, R.M., Turinsky, P.J., Faure, F.X.,<br />

Sarsour, H.N., Engr<strong>and</strong>, P.R., 1993. NESTLE: a nodal<br />

kinetics code. Transactions of the American Nuclear Society<br />

68, 432 /433.<br />

Rozain, J., 1991. Criticality excursions in wetted powder.<br />

Proceedings of the Fourth International Conference on<br />

Nuclear Criticality Safety, Oxford, UK.<br />

Samuelsberg, A., Hjertager, B.H., 1996. An experimental <strong>and</strong><br />

numerical study of flow patterns in a circulating fluidized<br />

bed reactor. International Journal of Multiphase Flow 22,<br />

575 /591.<br />

Savage, S.B., 1983. Granular Flows at High Shear Rates.<br />

Academic Press, London.<br />

Schmidt, A., Renz, U., 1999. Eulerian computation of heat<br />

transfer in fluidized beds. Chemical Engineering Science 54,<br />

5515 /5522.<br />

Sefidvash, F., 1996. Status of the small modular fluidized bed<br />

light water nuclear reactor. Nuclear Engineering Design<br />

167, 203 /214.<br />

Shahinpoor, M., Ahmadi, G., 1983. A kinetic theory for the<br />

rapid flow of rough identical spherical particles <strong>and</strong> the<br />

evolution of fluctuation. In: Shahinpoor, M. (Ed.), Advances<br />

in Mechanics <strong>and</strong> the Flow of Granular Materials,<br />

II. Trans. Tech. Pub, Andermannsdorf, Switzerl<strong>and</strong>, pp.<br />

641 /667.<br />

Shmakov, V.M., Lyutov, V.D., 2000. Effective cross sections<br />

for calculations of criticality of dispersed media. PHYSOR<br />

2000, Pittsburg, Pennsylvania, USA.


C.C. Pain et al. / Nuclear Engineering <strong>and</strong> Design 219 (2002) 225 /245 245<br />

van der Hagen, T.H.J.J., van Dam, H., Harteveld, W.,<br />

Hoogenboom, J.E., Khotylev, V., Mudde, R.F., 1997.<br />

Studies on the inhomogeneous core density of a fluidized<br />

bed nuclear reactor. Proceedings of Global 97-International<br />

Conference on Future Nuclear Systems, Pacifico Yokohama,<br />

Yokohama, Japan, 1050 /1055.<br />

van Dam, H., van der Hagen, T.H.J.J., Hoogenboom, J.E.,<br />

Khotylev, V.A., Mudde, R.F., 1998. Statics <strong>and</strong> dynamics<br />

of a fluidized bed fission reactor. Proceedings of the<br />

International Conference of Emerging Nuclear Energy<br />

Systems, ICENES ’98 Tel-Aviv, Israel, 609 /616.<br />

WIMS8A: user guide for version 8, 1999. AEA Technology<br />

Report ANSWERS/WIMS (99)9.<br />

Yamamoto, Y., 1995. Space-dependent kinetics analysis of a<br />

hypothetical array criticality accident involving units of<br />

aqueous uranyl fluoride. Proceedings of the Fifth International<br />

Conference on Nuclear Criticality Safety, Albuquerque,<br />

New Mexico, 10 /19.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!