You are on page 1of 60

MEng Mechanical Engineering with Nuclear Technology

Investigating the impact of Above Core


Structures towards the flow behaviour in
the upper plenum of E-SCAPE facility

Indraneel Singh

Supervisor: Prof. Shuisheng He

Co-Supervisor: Dr. Ashish Saxena

A report submitted in fulfilment of the requirements


for the degree of
MENG
in the
Department of Mechanical Engineering

May 3rd, 2023


Declaration

All sentences or passages quoted in this document from other people’s work have been specif-
ically acknowledged by clear cross-referencing to author, work and page(s). I understand that
failure to do this amounts to plagiarism and will be considered grounds for failure.

Name: Indraneel Singh

Date: 03/04/2023

i
Abstract

ESCAPE is a 1/6th scaled model of the MYRRHA reactor in Belgium. It was built to better
understand the thermal-hydraulic phenomena in the reactor. Experiments in ESCAPE give
insight into the flow behaviour and heat transfer in the various regions of the reactor. The
experimental results also serve as a benchmark to validate various system codes and CFD
simulations. This helps in the development of simulation models for low-Prandtl fluid-like
liquid metals. Various CFD simulations have been performed on ESCAPE to gain further
understanding and validate the methodologies. A high-fidelity simulation performed on ES-
CAPE used the porous zone approach to represent the ACS in the barrels.

In this study, the flow behaviour in the upper plenum of ESCAPE was modelled along with
the Above Core Structures to understand their impact on overall flow in the upper plenum.
Commercial CFD code ANSYS Fluent was used for RANS modelling. The results for forced
and mixed convection cases were presented and compared with LES results. The temperature
in the upper plenum from simulations showed good agreement with the experimental data.
The porous zone approach is good at modelling the effects of ACS for forced convection
cases. Noticeable differences were observed during mixed convection cases. The porous
zone approach does not capture the impact of ACS on the hot fluid from the inlet, which
leads to differences in barrel jet behaviour.

ii
Acknowledgement

This work has been made possible due to the help and guidance provided to me by various
people.

Firstly I would like to thank my supervisor Professor Shuisheng He and my co-supervisor


Dr Ashish Saxena for their continued guidance and support throughout the duration of the
project. Without their continued feedback, I would not have been able to produce this work.
The weekly meetings were vital and kept me on track with my project. The discussions on
the results were vital in helping me develop a better understanding of this topic and helped
me substantially in writing this dissertation.

I would also like to thank my colleagues, from my time working as an engineer. Special
thanks to Dr Stephen Silvester for imparting to me some of the critical skills required in
CFD. These helped me substantially in running the simulations for this study.

I would also like to thank the University of Sheffield’s High-Performance Computing team.
Getting pilot access to the new Stanage facility helped me accelerate my work. Their support
in fixing IT issues saved me great time for my simulation.

Also, thanks to my friends whose continuous feedback helped me improve my writing for
this dissertation.

iii
Contents

1 INTRODUCTION 1

1.1 Liquid Metal as Coolant . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

2 TECHNICAL BACKGROUND 3

2.1 MYRRHA and Scaling Approach . . . . . . . . . . . . . . . . . . . . . . 3

2.2 Thermal Hydraulic Challenges . . . . . . . . . . . . . . . . . . . . . . . . 3

2.3 ESCAPE Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.3.1 ESCAPE Operation . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.4 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.4.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3 CFD THEORY 12

3.1 The Navier-Stokes Equations and RANS . . . . . . . . . . . . . . . . . . . 12

3.2 Closure Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

3.3 Low Prandtl number flows and the limitation of Reynolds Analogy: . . . . 14

3.4 Two-Equation Turbulence Model: . . . . . . . . . . . . . . . . . . . . . . 14

3.5 Solution Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

4 CFD METHODOLOGY DEVELOPMENT 17

4.1 CFD Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

4.2 Meshing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

4.3 Simulation Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

4.4 Computational Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

5 RESULTS and DISCUSSION 25

5.1 Simulation Results Validity . . . . . . . . . . . . . . . . . . . . . . . . . . 25

5.2 Comparison with Porous Zone Approach . . . . . . . . . . . . . . . . . . . 27

iv
5.2.1 Forced convection case F80-P80-BP0 . . . . . . . . . . . . . . . . 27

5.2.2 Mixed convection case F20-P80-BP0 . . . . . . . . . . . . . . . . 31

5.3 Flow behaviour inside the barrel . . . . . . . . . . . . . . . . . . . . . . . 33

6 CONCLUSION and FUTURE WORK 39

6.1 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

6.1.1 Further geometric detail . . . . . . . . . . . . . . . . . . . . . . . 40

6.1.2 Improving the accuracy of the RANS model . . . . . . . . . . . . . 41

6.1.3 Further thermal hydraulic investigation . . . . . . . . . . . . . . . 42

7 Bibliography 43

Appendices 48

A Appendix-A 48

B Appendix-B 50

List of Figures

1.1 Temperature and velocity profile for fluids at different Prandtl numbers. [1] 2

2.1 The E-SCAPE facility(left and middle) and the core (right) [2] . . . . . . . 6

2.2 Coupled CFD and system-code schematic [3] . . . . . . . . . . . . . . . . 9

3.1 Pressure-based solver [4] . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

4.1 ESCAPE CFD Model. 1. IVHMs 2. Barrel 3. Above Core Structure 4. SDs
5. Pumps 6. HX outlet 7. Bypass-inlet 8. Active-inlet 9. HX-slots . . . . . 17

4.2 Above core structure inside the barrel . . . . . . . . . . . . . . . . . . . . 18

4.3 Outlet types tested. Meshed outlet pipe(left), surface outlet (right) . . . . . 19

4.4 Features of Poly-hexcore mesh . . . . . . . . . . . . . . . . . . . . . . . . 21

4.5 Hydraulic diameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

v
4.6 Poly-hexcore mesh for ESCAPE model, showing mesh size inside and out-
side the barrel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

4.7 Location of monitoring points in the CFD model . . . . . . . . . . . . . . 24

5.1 Residuals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

5.2 Temperature and Velocity on monitoring points . . . . . . . . . . . . . . . 25

5.3 Recirculation in F80-P80-BP80 vs Stratification in F20-P80-BP0 . . . . . . 26

5.4 The temperature in the upper plenum for forced and mixed convection. . . . 26

5.5 Temperature along IVHM(top) SDD(bottom), LES(left) RANS(right) . . . 27

5.6 Velocity along IVHM(top) SDD(bottom) plane, LES(left) RANS(right) . . 28

5.7 Temperature and TKE in fifth barrel jet, LES(left), RANS(middle), RANS(right) 29

5.8 TKE along IVHM(top) SDD(bottom), LES(left) RANS(right) . . . . . . . 29

5.9 Fourth barrel jet, LES(left) RANS(right) . . . . . . . . . . . . . . . . . . . 30

5.10 Third jet along plane-1 (left) and plane-2 (right) . . . . . . . . . . . . . . . 30

5.11 Temperature along IVHM(top) SDD(bottom), LES(left) RANS(right) . . . 31

5.12 Velocity along IVHM(top) SDD(bottom) plane, LES(left) RANS(right) . . 32

5.13 Free-jet along two planes. Plane-1(left), Plane-2(right) . . . . . . . . . . . 32

5.14 Upper plenum temperature along free jet, LES(left) RANS(right) . . . . . . 33

5.15 TKE along IVHM and SDDs planes . . . . . . . . . . . . . . . . . . . . . 34

5.16 Velocity inside barrel, LES(top) RANS(bottom) . . . . . . . . . . . . . . . 34

5.17 Temperature inside barrel, LES(top) RANS(bottom) . . . . . . . . . . . . 35

5.18 Velocity along SDD plane, LES(top) RANS(bottom) . . . . . . . . . . . . 36

5.19 TKE inside the barrel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

5.20 Velocities before, at, and after the grips on the Above Core Structures . . . 37

5.21 Top barrel jet along different free jet directions. . . . . . . . . . . . . . . . 38

6.1 Core outlet jets interacting with ACS [3] . . . . . . . . . . . . . . . . . . . 41

vi
List of Tables

2.1 ESCAPE core heat transfer at 80% power [2] . . . . . . . . . . . . . . . . 6

2.2 ESCAPE flow operation conditions . . . . . . . . . . . . . . . . . . . . . 7

4.1 Mesh Types and Cell counts . . . . . . . . . . . . . . . . . . . . . . . . . 21

4.2 Mass flow rate calculations for CFD model . . . . . . . . . . . . . . . . . 22

4.3 Simulation run times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

vii
List of abbreviations and symbols

ACS - Above Core Structures

IVHM - In Vessel Fuel Handling Machines

LBE - Lead Bismuth Eutectic Alloy

LES - Large Eddy Simulation

RANS - Reynolds Averaged Navier Stokes Equations

SDD - Silicon Doping Device

TKE - turbulent kinetic energy

k - kinetic energy term

- dissipation energy

ω - specific dissipation rate

viii
1 INTRODUCTION

Nuclear Power has been known to our civilization for a few decades now. In December 1951,
a reactor at Argonne National Laboratory produced the first usable electric power[5]. Today,
around 442 reactors are operating in 33 countries, producing about 10% of the world’s total
power[6]. With the ever-increasing energy consumption, the current focus on low-emission
energy sources, and the recent Geo-political events, nuclear power stands as one of the leading
contenders for power supply.

Much progress has been made in reactor technologies. The latest concepts of reactor tech-
nologies are classified as Gen-IV reactors. These reactors promise increased efficiency, gen-
eration, and heat capture [7]. Some of the proposed reactor technologies include Supercritical
water-cooled reactors, liquid metal-cooled fast reactors(including lead and sodium), Molten
Salt reactors, Gas fast reactors, and the very-high-temperature reactors [8]. These reactors
make more efficient use of Uranium and minimize waste production, factors that are key
concerns in the nuclear industry.

Liquid metal-cooled fast reactors use metals like Lead(Pb), Sodium(Na), Lithium(Li), etc as
a coolant instead of water, as is used in traditional reactors. It is however not a new concept to
use liquid metal as a coolant in reactors, the experience of operating sodium-cooled reactors
is about 200 reactor years, and for lead-cooled reactors, it is about 80 reactor years[9]. These
reactors provide significant improvements over their predecessors. These reactors are termed
fast reactors as they make use of fast neutrons for reactions, hence the lack of moderators in
these reactors. They have the ability to breed their own fuel from fertile isotopes of Uranium
U238 and Thorium T h238 [10]. This is very advantageous. Uranium has two isotopes, U235 ,
mainly used as fuel in most reactors, and is less common naturally, while U238 makes up for
99.3% of natural uranium deposits. Since fast reactors can make use of U238 , they promise
to be economically feasible.

1.1 Liquid Metal as Coolant

There are many liquid metal candidates to be used in reactors, but the most common ones
are Sodium, Lead, and eutectic alloys like Na-K and Pb-Bi. Sodium and Lead cooled fast
reactors are part of the proposed GEN-IV reactors[11]. Liquid metals provide excellent heat
transfer at the core and have a large margin for boiling. For sodium, the melting and boiling
point is 97.72°C and 883°C[12]. For Lead bismuth eutectic(LBE) alloy, it is 123.5°C and
1670°C[12]. The reactors during operation are not expected to go close to these boiling tem-
peratures and so there is no risk of coolant boiling, thus removing any need for pressurization
of the reactor. These reactors, therefore, operate at atmospheric pressure, which makes Loss

1
of coolant accidents (LOCA) virtually impossible [11]. These features make them safer than
the older designs of reactors. Liquid metals are weak absorbers of neutrons which makes
them ideal for use in fast reactors. Fast reactors do not need moderators as they utilize the
fast neutrons spectrum.

Liquid metals like Sodium and LBE are classified as low-Prandtl number fluids. A Prandtl
number is a dimensionless number that compares momentum transport to thermal transport
in a fluid. Liquid metals have Prandtl number in the range of 10−2 − 10−3 . Conventional
fluids like air and water have Prandtl numbers in the order 1. Comparatively, oils have higher
Prandtl numbers, in the ranges of 102 - 106 . The properties of fluids with different Prandtl
number is shown in Figure-1.1.

Looking at conventional fluids like air and water, middle of Figure-1.1, the momentum
boundary layer is of the same height as the thermal boundary layer. The difference in bound-
ary layers can be seen when the Prandtl number is changed.

Figure 1.1: Temperature and velocity profile for fluids at different Prandtl num-
bers. [1]

Fluids like air water and oils have lower thermal conductivity relative to viscosity. In turbu-
lent flows, the heat transfer process includes eddy and momentum conduction. For normal
fluids like air, molecular conduction is important only in the viscous regions of the boundary
layer, which constitutes a very small region close to the wall. This can hence be ignored. For
liquid metals, molecular conduction is much more dominant. The conduction process not
only affects the boundary layer but the effects are seen throughout the flow field. This differ-
ence in heat transfer mechanism introduces some challenges in the simulation of low-Prandtl
fluid in computational fluid dynamics. This is discussed in Section-3.

In this paper, the ESCAPE facility is investigated, which uses Lead Bismuth eutectic(LBE)
alloy as a coolant. Compared to sodium, LBE offers higher boiling point temperatures, which
are well above the cladding failure temperature. They also have weak oxidation reactions
to the surrounding environment hence no risk of violent reactions during coolant exposure.
Another advantage seen is that lead coolant has a better capacity to remove heat from reactors
during Unprotected loss-of-flow cases[13].

2
2 TECHNICAL BACKGROUND

2.1 MYRRHA and Scaling Approach

The ESCAPE facility was constructed at the SCK.CEN site in Mol, Belgium, to better study
the thermal hydraulics phenomena inside a liquid metal-cooled fast reactor. The design is
based on MYRRHA(Multi-Purpose Hybrid Research Reactor for high-tech applications), a
fast-spectrum research reactor constructed by SCK.CEN. It is a pool-type reactor that uses
Lead-Bismuth eutectic as a coolant. It also makes use of passive cooling phenomena like
natural circulation to remove decay heat during shutdown/accident conditions. A good un-
derstanding of thermal hydraulics characteristics is thus key to achieving a safe and reliable
design.

ESCAPE was created to study these phenomena in detail experimentally. A non-dimensional


scaling analysis was done to preserve thermal-hydraulic behaviour and design E-SCAPE.
There are many challenges in such scenarios of scaling as various flow features are present.

Reynolds number scaling is not used in this case. The resultant flow rates would be very high
to be practical. As long as flow remains in the same regime, i.e. fully turbulent, Reynolds
number preservation can be ignored.

ESCAPE is a 1/6.3 isotopically geometric scaled model of MYRRHA [14]. For the forced
flow conditions, the conservation of the Froude number was chosen. The froude number
Inertia
represents the ratio between inertial and gravitational forces, F r = Gravity
. Its conservation is
important in cases where free-surface and jet behaviour characteristics need to be maintained.
To investigate decay heat removal in natural convection case, it is important to maintain the
value of ∆T . Richardson number, which represents the ratio between buoyancy effects vs
turbulence generation, is preserved in this case since buoyancy effects are dominant [15].
It is shown by Marino et al [14] that preserving Froude or Richardson number provides the
same results and so Froude scaling can be applied to this case as well.

It is to be noted that it is not practical to produce a scaled model of components like the
core, and heat exchangers. As such they were assumed to be black boxes, in which the
flow behaviour inside them is not considered and only their overall effects are taken into
consideration.

2.2 Thermal Hydraulic Challenges

Thermal Hydraulics is a field of study that looks at flow and heat transfer in various fluids.
Liquid metals are low-Prandtl number fluids, which makes their heat transfer properties very

3
different from conventional fluids like water and air. A better understanding of thermal hy-
draulics is essential for better operation of reactors as well as for designing safety systems
for accidental scenarios. The thermal hydraulics challenges inside a reactor can be broadly
divided into three regions, the core, the upper plenum, and the lower plenum. All these re-
gions need to be better understood for the safe and reliable performance of MYRRHA and
other liquid-metal-cooled reactors in general.

The core region is the most crucial part of the reactor. Heat transfer is looked at in great detail
here in order to obtain uniform core outlet temperature at nominal power[16]. Interwrapper
flow, i.e. the flow between spacer grids is also an important phenomenon as it can impact
the global core thermal hydraulics [17]. Core subassembly in fast reactors consists of a fuel
bundle with helical wire spacers. Total pressure drop across the fuel subassembly as such
needs to be considered along with peak temperature distributions to ensure the maximum
cladding temperature limit is not reached. Mixing induced by wire-wrapped spacers is also
important to understand the overall temperature in these regions.

In this study, the upper plenum region has been focused on. The upper plenum is considered
the noblest part of the reactor since many challenges are concentrated in that region. These
challenges are common across a range of reactor designs, for both Sodium and Lead-based
reactors. A better understanding of these phenomena is crucial to predict and prepare for
accident scenarios and for better performance of reactors in general. It is important to under-
stand the degree of mixing in the upper plenum as well as the possibilities of stratification
in this region during normal and decay-heat removal scenarios. Some of the areas of con-
cern are jet behaviour in the core outlet, flow-induced vibrations in the above core structure,
gas entrainment, and thermal stratification. Thermal stratification is an important problem
that reactors have to account for. During low flow rates, the liquid metal inside the reactor
can form stratified layers of temperature, which leads to a large temperature gradient. This
reduces the heat transfer inside the reactor and leads to thermal stress on the surrounding
structures[18]. These stratified layers are unstable which might lead to further neutronics
and thermal-hydraulic instabilities in the reactor[18]. Thermal stripping is another major
problem in the upper plena. When jets of liquid metal mix at two different temperatures,
high-frequency temperature fluctuation is seen, which leads to thermal fatigue in the sur-
rounding materials, thus lowering the life expectancy of reactor components[19]. The upper
plenum forms a free surface during operation, which brings about its own challenges. Free
surface oscillation, which occurs due to the flow behaviour in the upper plenum, can also
cause thermal fatigue and hence needs to be reduced[16]. Gas entrainment is another issue in
these types of reactors. The free surface has an argon gas layer above it maintained at a pres-
sure of 4 bar[20]. Looking at the interaction between the free surface level and with argon
gas layer is key to investigating if gas entertainment takes place during operating conditions,
and what possible effects this could have on the flow behaviour.

4
The lower plenum is another area of interest, although it has fewer thermal hydraulic chal-
lenges. The pumps in the reactor, like MYRRHA, introduce low-pressure and high-pressure
regions in the lower plenum, which may put high stresses on the surrounding structures. The
temperature difference between the upper and lower plenum can put thermal stresses on the
structures separating them.

2.3 ESCAPE Design

ESCAPE is a 1/6 scaled model of MYRRHA. The design of E-SCAPE has important fea-
tures from MYRRHA. All the key geometrical features are replicated to maintain geometrical
similarity. Although, certain components like heat-exchanger, pumps, etc are not included
inside ESCAPE due to size constraints. Instead, ESCAPE has an external cooling circuit that
contains pumps and heat exchangers.

The vessel contains Lead Bismuth Eutectic, the same working fluid as MYRRHA. The facil-
ity has a melting and storage tank for LBE. LBE is pumped by the external circuit, which also
contains external oil circuits and air coolers to remove the heat from the core. From the exter-
nal circuit, LBE enters the lower plenum and goes into the core where it is heated. As the flow
exits the core, it goes through the pipes attached to core outlet plate. These pipes push jets
of fluid to the core barrel(see Figure-2) which contains the above core structures(ACS)(see
Figure-3). The flow exiting the core interacts with ACS and exits radially through the barrel.
The core barrel is a cylinder with an external diameter of 0.228m and contains around 120
holes of 0.03184m in diameter.

The fluid leaves the barrel through these holes and goes into the upper plenum. The fluid
of lower temperature exits through the bottom holes, while higher temperature flow exits
through the top holes in the barrel[21]. In the upper plenum, re-circulation is seen dur-
ing high-flowrate cases and stratification can be seen during low-flowrate operating condi-
tions[21]. In the upper plenum, there are other components, the Silicon Doping device(SDDs)
are cylindrical volumes filled with LBE. There are also the In-vessel fuel handling ma-
chines(IVHMs) which are large cylinders in thermal contact with the upper and intermediate
plena and in open contact with the lower plenum[20]. Unlike the MYRRHA, there are no heat
exchangers in the upper plenum, instead, there are four pipes that take the fluid to the exter-
nal circuit which has the heat exchanger. The external circuit is connected to pumps(PP) that
push fluid back into the lower plenum. Figure-1 shows the details of the ESCAPE facility.

To mimic the core of a reactor, the core region in ESCAPE contains electric heaters which
provide the required pressure drop and heat to the coolant as calculated by scaling analysis.
The core has the capacity to provide 100kW of heat. The core itself is a complicated struc-
ture(see Figure-2.1). The inlet grid of the core is designed to provide the required pressure

5
Figure 2.1: The E-SCAPE facility(left and middle) and the core (right) [2]

drop as calculated by scaling analysis. The core consists of seven annular layers, each sep-
arated from the other. The outermost layer is the bypass layer. It does not offer any heat
energy to the coolant. The inner six layers are the active region, where the heat transfer oc-
curs. Details on heat transfer taking place within each layer have been mentioned in the table
below.

Power layer 1 2 3 4 5 6 7
Heat transfer 9.9 kW 9.6 kW 13.1kW 16.2kW 17.6kW 0.0kW 0.0kW
Table 2.1: ESCAPE core heat transfer at 80% power [2]

ESCAPE is heavily instrumented to extract data from tests. It includes 300 thermocouples
to provide a detailed temperature distribution and possibly look into thermal stratification.
There are two radar level sensors that determine the level of lower and upper plena. It is also
equipped with three pressure sensors and 45 Ultrasonic Dopper sensors in the lower plena
for velocity.

2.3.1 ESCAPE Operation

ESCAPE can be operated at forced and natural convection cases, depending on whether the
pump is operational. Various test conditions are simulated in ESCAPE that corresponds to
scenarios expected in MYRRHA. The various tests done in ESCAPE correspond to shutdown
conditions, reduced power, pump and heat exchanger failures, and other steady-state and
transient scenarios[20].

6
In this study, two forced convection conditions and one mixed convection condition will be
investigated, all three working at 80% power and no heating from the by-pass layer. The
Forced condition at 80% power corresponds to normal operating conditions and is important
to investigate thermal-hydraulic phenomena for this case. The case with 40% power repre-
sents a case with reduced power from the pump but forced convection still dominates. At
20% power, buoyancy effects become a significant factor. This case is important to simulate
in order to investigate the heat transfer phenomena that can happen during the loss of coolant
situations and understand the capacity of LBE to remove heat through natural convection
during such scenarios. The configurations simulated are shown below with respect to Flow
rate,(F) Power(P), and By-pass power(BP).

Operation condition F-80 P-80 BP-0 F-40 P-80 BP-0 F-20 P-80 BP-0
Mass flow rate (kg/s) 93.2 46.6 23.4
Core inlet temperature 180 180 180
Total Core Power 66.4 66.4 66.4
Case Forced convection Forced Convection Mixed convection
Table 2.2: ESCAPE flow operation conditions

2.4 Literature Review

Much work has been done in developing reactor concepts and designs for Lead cooled reac-
tors. The designs can be broadly classified into forced circulation type and natural circulation
type, depending on whether a pump is employed in the primary system. BREST-300 and
SVBR-100 in Russia are some examples of forced circulation reactors[22]. SSTAR and SU-
PERSTAR designed by Argonne national laboratory[23], and ELECTRA designed by the
Royal Institute of Technology and the Chalmers University of Technology [24] are some
examples of natural circulation reactors.

For the simulation of liquid metal, an appropriate model for turbulent Prandtl number is
required (see section 3.3). Investigation of turbulent heat transfer in channel flow was done
for P r = 0.005 using DNS by Chai et al[25]. Based on reference results, an algebraic
turbulent heat transfer model was employed which showed potential for accurate prediction
of temperature in a low-Prandtl fluid using RANS. A key observation noted was that for
channel flows, peaks of turbulent Prandtl number are located in the near-wall region and
they slightly decrease while approaching the channel centre. Chen et al looked at different
turbulent-Prandtl number models for accuracy[26]. They found that Cheng’s P rt model[27]
is suitable for constant heat flux boundary conditions while Jishca’s model[28] is suitable for
constant wall temperature boundary conditions. Assessment of RANS modelling was done
for forced convection flows in low-Prandtl number fluids using LES simulation by Dupncheel

7
et al [29]. It was noted that neither linear law nor log law is accurate for resolving the near-
wall temperature profile for low Prandtl fluids. It was found that even at high Reynolds
numbers, the heat transfer by molecular conduction is of the same order as turbulent heat
transfer. A mixed law wall function was developed that provides a good temperature profile
in y + = 70 − 300 It is also recommended that Kays correlation for turbulent Prandtl number
is good for RANS modelling for y + > 100. Using the mixed law proposed in the paper
along with the Kays correlation for Turbulent Prandtl number provides results with accuracy
comparable to that of LES. Although, it is recommended to use these models cautiously
for complex flow phenomena like separation. It is also recommended that near the wall, a
fixed turbulent Prandtl number value of P rt = 2.0 for fluids with P r = 0.01.....025 can be
assumed.

To understand the thermal hydraulics of liquid metal fast reactors, efforts have been made
to develop numerical tools. Italian National Agency for New Technologies, Energy and
Sustainable Economic Development (ENEA) has employed many research efforts to de-
velop numerical tools, focusing on system thermal hydraulics(RELAP5), fuel pin perfor-
mance(TRANSURANUS), fuel element thermal hydraulics(ANTEO+) and CFD codes [30].
RELAP5 is a popular system code developed by Idaho National Laboratory which can per-
form simulations with liquid metals like LBE, Pb, and PbLi as working fluids [31]. System
codes employ a 1D-level system method which comes with its limitations, although using
a full-scale CFD model is not always feasible. There have been approaches to do coupled
approach between CFD and system codes [3] [32].

As mentioned previously, the upper plenum has many thermal hydraulics challenges. Much
work has been done to gain insight into these challenges. Analysis of thermal stratification
has been done on MONJU, a prototype fast breeder reactor in Japan[33] and on a Thermal
stratification testing facility(TSTF) at the University of Wisconsin[34], both of which use
sodium as a coolant. Thermal stripping has also been looked at, a challenge even for CFD
codes, as RANS equations are not able to accurately capture the temperature fluctuation seen.
LES and DNS are needed to resolve the temperature fluctuation seen [19] [35], but these usu-
ally come with a very high computational cost, making them unfeasible for most practical
purposes. Work was done by Kolozara et al [36]to develop a CFD platform MYRRHAFoam
based on OpenFOAM, an open-source CFD solver to simulate the flow inside MYRRHA.
MyrrhaFOAM can simulate buoyancy-corrected steady and unsteady flow using the temper-
ature dependant properties of LBE fluid. It uses a porous zone approach for the core, ACS,
and Heat exchanger.

Work has also been done on experimentation and simulation of the ESCAPE facility. Work
done by Alessandro Marino provided support to the design of the E-SCAPE facility sup-
ported by a simulation using RELAP5 system code [14]. The scaling analysis done in the

8
paper recommends Froude number criteria to mimic real plant behaviour in E-SCAPE. Using
system code RELAP 5, a few scenarios of simulations were run to better understand thermal
hydraulics and validate the scaling analysis. The scenarios included were as follows:

• Start-up transient: To analyze the system during the start-up scenario until operational
conditions are achieved.

• Symmetric and Asymmetric loss of heat sink: Looking at free surface height due to
loss of cooling with heat power shutdown.

• Symmetric main pump failure: Effect on free surface height due to failure of main
pumps.

Limitations were noted for system codes, for example, the mixing phenomena in the lower
plenum were not simulated in a reliable way in Marino’s work, and using CFD was recom-
mended.

Toti et al presented a coupling between system code and CFD [3]. RELAP 5 was paired with
ANSYS Fluent, a commercial CFD code, to investigate transient flow. The upper plenum
model was created and meshed for CFD while the lower plenum and external cooling circuit
were simulated using the RELAP 5 system code(see Figure-2.2. For the CFD domain, the
SST k-ω turbulence model was used along with the VOF model to simulate the free surface
of the upper plenum. Through their work, the benefits of using a coupled system code - CFD
approach over simple 1D code was noted.

Figure 2.2: Coupled CFD and system-code schematic [3]

Work has been done in validating CFD simulation against experimental results to develop
CFD methodology for liquid-metal simulations rigorously. It is to be noted that all the work
mentioned below has been done within the framework of the MYRTE Horizon 2020 project.
The accuracy of the CFD methodology was henceforth inferred from these papers and helped

9
in the methodology developed for this study. The experiment results used as a benchmark
were provided by results gathered from Van Tichelen and Mirelli [37]. The thermocouples
fixed at various locations of the upper plenum in the experiments provided temperature data
during various operating conditions. The data from these experiments were used in the veri-
fication of results in this study.

Three independent simulations were performed and validated against the above experiment.
These were done by NRG[2], SCK.CEN[38] and VKI[39], using STAR CCM+, ANSYS
CFX, and OpenFOAM respectively. In the three different CFD models mentioned above,
there were differences in the mesh size and type used. All three of them used the k- model
with wall functions with a y + > 30. There were differences in the extent of the ESCAPE
simulated in the three papers. CFD of the model of NRG did not include the gas layer above
the upper plenum but instead accounted for the effects as a slip boundary condition. The
external cooling circuit was also not modelled. The VKI CFD model did not include the
external circuit but did model the gas region by using the Volume of Fluid approach. The
free surface was not fixed and hence was formed depending on the flow conditions. CFD
model generated by SCK.CEN included both the external cooling circuit as well as the gas
phase.

Comparing the three simulations, the pressure drop from NRG was lowest while it was high-
est in the VKI model. Similar temperature results were obtained from all three simulations.
It is to be noted these CFD models overpredicted the pressure drop. This difference was at-
tributed to the difference in the model of the core and the true geometry of the core. A small
deviation in geometry seems to have a large impact on the pressure drop.

2.4.1 Motivation

Detailed simulation of flow in the upper plenum of ESCAPE has been done by Ashish et al
at the University of Sheffield [21]. For this study, Large Eddy Simulation was used. Unlike
most commercial applications of CFD that use RANS equations, LES directly solves Navier
Stokes equations to resolve large-scale eddies. As such it provides a greater degree of ac-
curacy. The downside however is that it requires high computational time, several orders
of magnitude higher than RANS. A high-quality mesh preferably a structured grid is also
recommended to get an accurate solution, which further adds to the difficulty of using LES.
To make it more feasible, some simplifications were made by Ashish et al, mainly using a
porous zone approach for the flow inside the barrel.

As mentioned before, the barrel contains the above core structures which makes meshing
inside the barrel challenging by highly increasing the number of mesh elements. The porous
zone approach is an efficient way to simulate flow resistance in a domain of a CFD model.

10
The porous media is a simplified model used in CFD applications. Using this model, a zone
is defined where the effects of a porous zone can be simulated without having to model
the details of the geometry. The pressure drop and heat transfer across the defined zone are
determined by the inputs provided. Given that it is a simplified model, it assumes the porosity
to be homogeneous. This approach was also used in the work of Koloszar et al where they
developed a CFD model MYRRHAFoam for simulating the MYRRHA reactor facility [36].
In the paper, the derivation of porous media properties of ACS was taken from the book:
Nuclear Systems[7]. Porosity values and radial and axial resistances were calculated for
ACS and applied to the region inside the barrel

In this study, using RANS modelling, the effects of ACS on the overall flow in the upper
plenum are investigated. This is done to verify the accuracy of the porous zone assumption
used for LES simulation and analyze flow features inside the barrel with the ACS present.
This study attempts to determine if the porous zone approach can accurately represent the
porosity of ACS in such reactor designs. This will help in the further development of CFD
models of such reactors. The aims and objectives for this study are thus as follows:

Aim:

To investigate the effects of the above core structures(ACS) on the flow behaviour in the
upper plenum of the E-SCAPE facility.

Objectives:

1. Investigate liquid metal thermal hydraulics phenomena in GEN IV reactors to develop an


understanding of the benefits and challenges of liquid metal fast reactors.

2. Understand the limitation of RANS modelling in CFD in simulating low-Prandtl number


liquids to understand the theoretical problems associated with modelling low-Prandtl flows.

3. Perform a literature review to understand the work done in simulating the upper plenum
of the E-SCAPE facility and build upon them.

4. Develop a CFD model and methodology including finding a suitable meshing strategy, tur-
bulence model, boundary treatment, and turbulent Prandtl number value for the flow problem.

5. Carry out steady-state simulations for various flow rates that represent both forced and
mixed convection cases.

6. Analyse the results to understand the flow behaviour around the ACS and its impact on
the overall flow in the upper plenum

7. Compare the results with the LES simulations to compare the accuracy and limitations of
the porous zone approach taken for the above core structure in Ashish et al[21].

11
3 CFD THEORY

ANSYS Fluent, a commercial CFD package was used for simulating E-SCAPE. In this chap-
ter, the underlying theory of computational fluid dynamics is discussed.

3.1 The Navier-Stokes Equations and RANS

Fluid Flows are represented using the Navier-Stokes equation and continuity equation.

∂ρ ∂(ρui )
+ =0 (3.1)
∂t ∂xi

∂(ρui ) ∂[ρui uj ] ∂p ∂τij


+ =− + + ρfi (3.2)
∂t ∂xj ∂xi ∂xj

where τij represents the viscous stress which can be represented as:

∂ui
τij = ρν (3.3)
∂xj

These equations, although, don’t have a general analytical solution, except for limited flow
scenarios. Commercial CFD codes use the RANS equation to solve fluid flow problems.
RANS equations are formed by taking the time average of the Navier-Stokes equation. Every
flow quantity can be decomposed into time-averaged quantity and the fluctuation about the
time-averaged quantity:
Φ = Φ + Φ0 (3.4)

where

Z
1
Φ≡ Φ(t)dt (3.5)
T T

In a scenario with the in-compressible flow and ignoring the body forces, the continuity,
and the Navier Stokes equation can be written in its time-averaged form, making the RANS
equations.
∂(ρui )
=0 (3.6)
∂xi

∂(ρui ) ∂ ∂p ∂τij
+ [(ρui uj ) + (ρu0i u0j ) = − + (3.7)
∂t ∂xj ∂xi ∂xj

12
The momentum equations total 3 equations along with one continuity. It can be seen that
the number of unknown quantities is more than the number of equations. The ρu0i u0j term is
called Reynolds Stresses.

The Reynolds stress terms cannot be decomposed into time-averaged quantities and hence
models are needed to describe these terms. This is the closure problem of RANS equations.

3.2 Closure Models

The closure models were introduced to relate the Reynolds stresses to the velocity field.
Boussinesq, in 1877, introduced such a relation through the introduction of eddy viscosity.
The relation was as follows:

∂U
ρu0i u0j = µt (3.8)
∂y

where µt is the eddy viscosity, hence, this is also referred to as the eddy viscosity model.

In commercial CFD codes, the eddy viscosity is most commonly modelled using the two-
equation model which will be discussed in the later section.

Similar to Reynolds stress, turbulent scalar flux −ρu0i φ0j represents the transport of a scalar
quantity φj .

The gradient diffusion theorem is used to model the turbulent scalar flux, which represents
the turbulent scalar flux as a function of the gradient of the mean quantity:

∂φ
ρu0i φ0j = Γt (3.9)
∂y

where Γt is the turbulent diffusivity.

Using the above equations to look at temperature as the turbulent scalar flux, Γt becomes the
turbulent heat diffusivity. Reynolds Analogy is used to define the turbulent heat diffusivity
which introduces a relation between turbulent heat diffusivity Γt and eddy viscosity µt as
follows:

µt
Γt = (3.10)
P rt

wherein the P rt is the turbulent Prandtl number. Reynolds Analogy assumes that µt = Γt

13
which gives P rt = 1. This value seems to work well for most liquids with the Prandtl
numbers 0.1 and above. Most commercial codes use a value of P rt = 0.85.

3.3 Low Prandtl number flows and the limitation of Reynolds Analogy:

For liquid metals used in Nuclear reactors, like sodium and lead-bismuth eutectic, the Prandtl
number is much smaller, often less than 0.01. In such cases, the turbulent Prandtl number
value of 0.85, based on the Reynolds analogy is no longer valid. Reynolds’s Analogy shows
that the momentum boundary layer and thermal boundary layer are of the same height. In
low-Prandtl flows, the thermal boundary layer is much larger as the heat transfer is mostly
dominated by thermal conductivity.

As per the recommendation provided in the work of Duponcheel et al [29], the turbulent
Prandtl number was changed to P rt = 2. While this assumption provides a better represen-
tation of low Prandtl fluid than the default value, it still is not completely accurate.

3.4 Two-Equation Turbulence Model:

As previously discussed, eddy viscosity is modelled using turbulence models, termed eddy
viscosity models. They include zero, one, two, and more equation models. The most com-
monly used turbulence models in commercial CFD codes are the family of k −  and k − ω
models. These models solve transport equations for the terms k, , and ω.

For both these models, the first term k represents the turbulent kinetic energy and has its
own transport equation. Turbulent kinetic energy can be simply stated as the mean of kinetic
energy in a turbulent flow. The transport equation of k represents the rate of generation and
dissipation of turbulence.

 represents the dissipation rate of turbulence. The k −  model works on the following
relation:
µt k2
= Cµ (3.11)
ρ 

where k is the turbulent kinetic energy and  is the rate of dissipation. the Cµ is an empir-
ical constant derived from experiments. k and  are solved using their respective transport
equation.

For the standard k −  model, this constant is 0.09. The k −  model is a family of models,
each having different values for constants. In this study, the Realisable k −  model is used.
First proposed by Shih et al it is a recent development in turbulence models. This model

14
provides significant improvements over the standard k −  model. In the Realisable model,
the constant Cµ in Eq-3.11 is no longer a constant but is computed. The Realisable model
performs better in rotational flow, separation, recirculation scenarios, which typically are
considered to be the weaknesses of the standard k- model [40] .

The ω represents a specific dissipation rate, which is dissipation per unit of turbulent kinetic
energy. The k-ω models are considered low-Re models and are good for resolving near-
wall flows. They perform well for boundary layer flows under adverse pressure gradients
and separation. The k-ω SST model proposed by Menter in 1993 is a very popular model
because of its accuracy in resolving near-wall flows as well as its flexibility with the near-
wall resolution value of mesh.

3.5 Solution Procedure

The CFD solver solves the RANS and other equations in a sequential manner depending on
the type of solver used. For incompressible flow pressure based solver is used, while density
based solver is used for compressible flows. Figure-3.1 shows the process Fluent solver
follows depending on the algorithm used.

Fluent has two classes of algorithms for pressure-based solvers. Segregated and Coupled
Solver are two options and they differ in how they solve the pressure and momentum equa-
tions. The segregated class of solvers which include SIMPLE, SIMPLEC, and PISO solves
the momentum and pressure equation sequentially. The COUPLED solver, which comes
under the class of Coupled Solver, solves the pressure and momentum equations simulta-
neously. They then solve the energy equation followed by turbulence and other equations
given the flow physics. The solver keeps performing iterations until the convergence criteria
is reached.

The COUPLED solver offers superior performance but since the equations are solved simul-
taneously, it has a higher memory requirement. These equations are solved and stored in the
centre of the mesh grids. These are then interpolated to the faces of the mesh through var-
ious interpolation schemes for different scalars like pressure, momentum, turbulent kinetic
energy, etc. By default, using the second-order discretization scheme provides higher accu-
racy. In some cases, the pressure discretization scheme needs to be changed, especially when
dealing with flow scenarios involving large body forces, like natural convection.

15
Figure 3.1: Pressure-based solver [4]

16
4 CFD METHODOLOGY DEVELOPMENT

4.1 CFD Model

A 3D CAD model was first prepared in ANSYS SpaceClaim for the simulation of the ES-
CAPE facility. Information on the dimensions of various components and their location was
gathered from a variety of sources. The geometry is mainly based on the model prepared by
Visser et al [2], although information from Toti et al [3]and Marino et al [14] was also taken.
The exact dimensions of the Above Core Structure could not be found in the literature. The
ACS model was thus prepared from scaling images from the paper of Toti et al [3]. Because
of the limitation of image resolution and human error, some discrepancies are expected in the
dimensions of the ACS model to the actual geometry.

It is to be noted that the aim of this project focuses on the upper plenum and hence only the
upper plenum was modeled. The lower plenum is thermally separated from the upper plenum
in the reactor and hence does not affect the flow in the upper plenum. The heat transfer and
flow rate through the core are known and hence can be directly applied to the upper plenum
through the inlet without needing to model the complex geometry of the core. The model is
shown in the figure below along with the above core structures in Figure-4.1 and Figure-4.2.

Figure 4.1: ESCAPE CFD Model. 1. IVHMs 2. Barrel 3. Above Core Structure
4. SDs 5. Pumps 6. HX outlet 7. Bypass-inlet 8. Active-inlet 9. HX-slots

Please note that the height of the upper plenum was calculated from Visser’s model [2] by
subtracting the height of the core and outlet pipes from an image in the paper. As such, slight
differences in height may appear.

The extent of the model is from the outlet of the core to the free surface level of the upper
plenum, thus encompassing the whole upper plenum region. The core itself is not modelled.
The outlet of the core is considered an inlet for the CFD model. The height of the established

17
Figure 4.2: Above core structure inside the barrel

free surface depends on the flow conditions. For this model, the height was chosen similar
to Visser et al [2], 745mm from the core, which corresponds to 370mm from the core outlet.

The core outlet of the ESCAPE facility serves as the inlet for the prepared CFD model. The
core outlet is only modelled as an inlet face for simplicity, but in reality, it consists of pipes
that insert jets of fluid into the barrel. The inlet has been divided into two sections to represent
the bypass and active regions of the core(see Figure-4.1). The inlet flow rate and temperature
will depend on the flow scenario being simulated(refer to section 2.3.1). While the by-pass
inlet only refers to the seventh layer in the literature, for the CFD model, the sixth layer is also
included in the by-pass layer since it does not add any heat to the fluid(refer Table-1). The
active inlet for this model includes the first five annular layers. This has been done because
the mass flow rate is calculated based on the temperature increase across the core(see section
4.3).

The heat exchanger outlet pipes of the ESCAPE facility serve as the outlet for our CFD model.
These pipes are placed at an elevation of 488mm from the outlet of the core. The faces at the
end of these pipes were assigned an outflow boundary condition. The outflow condition is a
type of boundary condition in ANSYS Fluent which is used when flow conditions at the exit
are not known. The conditions at the outflow are extrapolated by the solver from within the
fluid domain and do not impact the flow within the model.

A simplified version of the outlet was also tested to try and reduce the mesh count. As shown
in Figure-4.3, a simplified model of the outlet, where the whole cylindrical surface was used
as an outlet. This meant, that the heat exchanger pipes did not need to be meshed, which
led to a reduction in overall mesh size. Incorporating such a design proved to be inaccurate.
Treating the whole cylindrical face as an outlet imparted backflow from the outlet into the
upper plenum, which is not expected in the ESCAPE during operation. This also led to

18
instabilities in residuals leading to inaccurate simulation. As such, the heat exchanger pipes
were incorporated into the design.

Figure 4.3: Outlet types tested. Meshed outlet pipe(left), surface outlet (right)

Simplifications were made to the free surface region of the upper plenum since the exact
shape and perturbations of the free surface are not of interest. Since above LBE lies argon
gas, the interaction at the interface of two fluids can be assumed as zero shear. The free
surface of the upper plenum is simulated as a zero shear slip wall and the gas region above
is not modeled. Symmetry boundary conditions were used to model this zero slip wall. The
symmetry boundary condition assumes a zero normal velocity at the assigned face and zero
normal gradients of all other variables.

The components and the reactor body are modelled as walls. They were assigned as stationary
walls with no slip and zero roughness for simplicity. The heat exchange taking place at the
walls of these components was not included in the simulations. The heat transfer taking place
between the upper plenum and the components is low as is noticed by Visser et al [2] and can
thus be ignored.

In Figure-4.2, the Above core structures(ACS) inside the barrel are shown. These consist of
rods running along the length of the barrel. There are 36 rods in total. They are held together
by grips that run perpendicular to the flow direction. There are three sets of grips in total for
each rod, distributed along the length of the rod. The central rod is not connected to other
rods and runs the entire length of the barrel, connected to the inlet face of the CFD model.
It is to be noted that no information could be found on the exact dimensions of the Above
Core Structure. Since the CAD was constructed using image interpolation, some level of
uncertainty is expected.

19
4.2 Meshing

Meshing is one of the most important aspects of CFD simulations. It dictates the accuracy
of the results as well as the overall time and resources required to run the simulation. The
length scale in the gaps between ACS rods inside the barrel is much smaller than the length
scales outside the barrel. As such, the mesh was sufficiently refined inside the barrel while
less refinement was allowed in the outer region. The mesh element type was selected based
on a balance between achieving good accuracy and maintaining a lower cell count.

Traditionally, tetrahedral and hexahedral meshes have been used in CFD. Hex cells offer
more accuracy as the mesh has low numerical diffusion, it is however difficult to generate for
complex geometries. Tetrahedral cells offer flexibility in meshing around complex objects
but the numerical diffusion is significantly higher.

Polyhedral cells offer the same ease of mesh generation as tetrahedral cells but lower nu-
merical diffusion. This is because they have more neighbour cells and the gradient can be
well approximated. One of the biggest advantages is that they significantly reduce the num-
ber of cells required in the CFD model[41]. This results in lower run times. It is, however,
to be noted that while poly cells result in a lower number of cells, it increases the number
of interpolations carried out per cell since they have a higher number of faces, this also in-
creases the memory requirement. However, the increase in the number of interpolations is
still insignificant over the high reduction in cell count offered.

Poly-hexcore is a mesh style unique to ANSYS Fluent, patent-pending, sometimes referred


to as Mosaic Meshing[42]. They offer hexahedral dominant mesh in the bulk region along
with an inflation layer at the wall. The bulk region and the inflation layer are stitched together
with polyhedra cells. The combination of polyhedra cells with hexahedral cells in the bulk
region offers many advantages. The poly cells offer flexibility at lower cell counts and better
numerical diffusion. The hex cells offer accuracy, as they are the most preferred type of
mesh. Also, not having poly cells as the dominant cell significantly reduces the face count.
Figure-4.4 shows poly-hexcore mesh created for our geometry. As seen in Figure-4.4, the
walls are captured using poly cells, while the core is mostly hexahedral.

Tetrahedral, Polyhedral, as well as poly-hexcore mesh types, were tested for this simulation.
Tetrahedral cells resulted in an overall increased number of cells and hence they were not
selected. While Polyhedral cells had the lowest meshing time and the smallest meshing sizes,
they produced highly skewed cells in small gaps of ACS rods. This lowered the overall
quality of the mesh. Finally poly-hexcore mesh was chosen because of its balance between
superior performance and low cell count. The cell counts for each mesh are noted in Table-4.1
below:

20
Figure 4.4: Fea-
tures of Poly- Figure 4.5: Hy-
hexcore mesh draulic diameter

Mesh type Approx. Cell Count


Tetrahedral 24 million
Polyhedra 15 million
Poly-hexcore 17million

Table 4.1: Mesh Types and Cell counts

The mesh generated has the following features. In the area of interest, which is the ACS and
the barrel, a target of y+ >30 was set. To achieve this target, the first cell height was calcu-
lated on the basis of hydraulic diameter for a triangular sub-channel. The ACS resembles
a triangular fuel assembly design and hence the formula shown in Figure-4.5 was used to
determine the equivalent diameter, which was about 78mm.

The first cell height was set to 0.01m with a total of 2 layers and a growth rate of 1.5. In the
components outside the barrel, automatically generated boundary layers were used which
produced 3 layers providing a smooth transition to the core mesh. This approach of two dif-
ferent types of y+ values could only be used because of the features of enhanced wall treat-
ment as discussed in section 4.3. Further parameters specific to poly-hexcore meshing were
also changed. Two parameters, buffer layer, and peel layers are specific to poly-hexcore.
Buffer layers represent the growth of hex cells in the mesh, whereas peel layers represent the
size of polycells joining the boundary layer with core hex cells. A buffer layer value of 4 was
chosen so that the mesh can slowly transition from within the barrel to a bigger size away
from the barrel. The Peel layer was kept at 0 to reduce the size of poly cells. The total cell
count was around 17 million cells.

Figure-4.6 shows the mesh size difference inside and outside of the barrel. As can be seen,
inside the barrel, the mesh sizes are small enough to fit a few cells between the ACS gaps.
Outside the barrel, the mesh starts to grow. On the barrel walls and the ACS rods, boundary
layer prism cells can be seen.

21
Figure 4.6: Poly-hexcore mesh for ESCAPE model, showing mesh size inside and
outside the barrel

4.3 Simulation Parameters

The mass flow inlet applied to the model was calculated by the temperature difference across
the active zone of the core (Q̇ = ṁcp ∆T ), shown in the table below.

Mass flow rate (kg/s) ∆T ṁ(active)(kg/s) ṁ(bypass)(kg/s)


93.2 10(190 − 180◦ C)

45 48.2
46.6 18(205◦ − 180◦ C) 25 21.6
23.4 37(217◦ − 180◦ C) 12.2 11.2
Table 4.2: Mass flow rate calculations for CFD model

The values above were then applied to the mass flow inlet in the CFD model. For mass inlet
specification values of turbulence parameters were also specified. Turbulence intensity is de-
fined as the ratio of deviation of velocity with mean velocity I = u0 /U . The value is specified
to define the level of turbulence entering the CFD domain. The turbulence intensity was cal-
culated based on the formula for intensity in a fully developed pipe flow: I = 0.16ReDh −1/8 .
This gives us a value of around 4% turbulence intensity. Although it is to be noted that this
formula assumes fully developed pipe flow, in reality, the jets exiting the barrel may have
higher turbulence values due to the sudden change in cross-section area.

The data on various fluid material properties were taken from the LBE handbook [1]. For
density and viscosity, the relations are mentioned below:

ρLBE = 11096 − 1.3236 ∗ T (4.1)

µLBE = 4.94 ∗ 10−4 e754.1/T (4.2)

This relation was implemented into the solver. A gravity acceleration of 9.81m/s2 was also
applied to look at buoyancy-induced effects in the flow. As mentioned before, the default

22
turbulent Prandtl number is not applicable for LBE, as such the turbulent Prandtl number
value was changed. There are a few turbulent Prandtl numbers in Fluent for k, , energy and
wall. Energy and wall turbulent Prandtl numbers were changed since these values affect the
energy equation in the solver. A value of 2 was applied as recommended by Duponcheel et
al [29].

Pressure-based solver is used with the Realisable k− turbulence model along with Enhanced
wall treatment. The enhanced wall treatment option was chosen because of the flexibility it
offers with near-wall mesh. As mentioned before, it is difficult to get a uniform y + value
all throughout the domain. The enhanced wall treatment is near wall modelling method that
uses a two-layer approach. Depending on the near-wall mesh, it can resolve the near-wall
phenomena, i.e. viscous sublayer if a fine mesh is present. If not, it uses a wall function for
coarse mesh. In the turbulent region, it uses the k −  model, whereas, in the near wall region,
it switches to one equation model. This method can handle a range of y + meshes, making it
suitable for a broad range of applications.

Within the solver, there are a few options available to couple the pressure and velocity fields.
For pressure-velocity coupling, the pressure-based coupled solver ’COUPLED’ was used
with pseudo transient time stepping applied. COUPLED solver offers better performance
over segregated solver schemes. Here the momentum equations and pressure correction
equations are solved together unlike the segregated solver where they are solved separately.
Psuedo-time stepping introduces implicit relaxation to the equations being solved. This re-
sults in a controlled change of variables which helps make the simulation stable and achieve
a converged solution faster.

The interpolation schemes for pressure were changed for the forced and mixed convection
cases. For the forced convection cases, a second-order scheme was used for its accurate in-
terpolation. For mixed convection where buoyancy effects are significant, PRESTO! scheme
was used for pressure interpolation. PRESTO! is known to perform better when body forces
are dominant in the flow, which is the case for buoyancy-driven flow.

To ensure convergence of simulations, residual values for various variables were set to order
10−4 while for energy it was set to 10−6 . Monitoring points were also created across various
areas in the model to check the convergence of variables like velocity and temperature. These
points were placed immediately outside barrel holes and three points in the far field region
in the upper plenum. The location of monitoring points is shown in Figure-4.7.

23
Figure 4.7: Location of monitoring points in the CFD model

4.4 Computational Setup

Given the high mesh size, the University of Sheffield’s High-performance computer(HPC)
”STANAGE” was used to run the simulations. Jobs on HPC are run in batch mode, which
requires a batch script, which requests resources and allocates time for each job. A total of 60
cores were requested with 120k Mb of memory. Run times of around 25 hrs were requested.
Although the cases were expected to finish much faster, extra hours were assigned so that the
HPC system would not close the job prematurely. Journal files were also written to set up
Fluent case in batch mode. Journal files are native to Fluent and are a text-based interface to
interact with Fluent. Through the journal file, Fluent was instructed to do the following:

1. Read the particular case file.


2. Set the pressure velocity coupling method to COUPLED.
3. Initialise the case using values from the inlet.
4. Run 3000 iterations.
5. Save the data file with the provided name.

The run times for each simulation case along with total iterations until convergence was
reached are listed in Table-4.3.

Simulation Case Total Time Iterations


F80 12 hrs 34 min 3000
F40 5 hrs 5 min 1003
F20 14 hrs 26 min 2797

Table 4.3: Simulation run times

The batch script and the journal file have been attached in Appendix A.

24
5 RESULTS and DISCUSSION

5.1 Simulation Results Validity

The residuals and monitoring points were checked to make sure that the results for all the
cases have converged. Figure-5.1 shows the residuals for various variables for the F20 case.
The residuals have dipped sufficiently low so that the differences in values with further iter-
ation are negligible.

Figure 5.1: Residuals

The values of monitoring points were also checked. Since this is a steady-state solution,
the variables at each of these monitoring points should not change significantly. Figure-5.2
shows the velocity and temperature values at the monitoring points for the F20 case.

Figure 5.2: Temperature and Velocity on monitoring points

Once the steady state solution was achieved, the flow behaviour in the upper plenum was
inspected to see if the behaviour of the flow was as expected in forced convection and mixed
convection case. For the forced convection case, re-circulation is expected which would lead

25
to a homogeneous distribution of temperature. For the mixed convection case, thermal strati-
fication is expected which would produce stratified layers of LBE with distinct temperatures.
The results were as expected, as can be seen in Figure-5.3.

Figure 5.3: Recirculation in F80-P80-BP80 vs Stratification in F20-P80-BP0

The temperature in the upper plenum was compared to the experimental data obtained from
Visser et al [37]. The ESCAPE facility is equipped with sensors at various heights in the
upper plenum. The temperature reading from these sensors was used to compare results for
the RANS simulation. Corresponding to the location of the sensor, a line was drawn and
temperature data on that line was extracted and plotted, as shown in Figure-5.4.

Figure 5.4: The temperature in the upper plenum for forced and mixed convection.

26
5.2 Comparison with Porous Zone Approach

In this section, the flow behaviour in the upper plenum is compared to the results obtained
from the LES simulation by Ashish et al [21]. The forced convection case F80-P80-BP0 and
mixed convection case F20-P80-BP0 are compared. For the analysis, various planes were
inspected along with the jets exiting the barrel. For the purpose of comparison, the results for
this simulation will be termed as RANS simulation, while the results from Ashish et al will be
referred to as LES. The fluid is referred to as LBE. The IVHM plane corresponds to the cross-
section along the In vessel fuel handling machine. The SDD plane is the cross section along
the Silicon Doping device and the pumps(see Appendix B). The freejet/freestream direction
represents a plane where barrel jets don’t interact with any components.

5.2.1 Forced convection case F80-P80-BP0

The temperature and flow velocities along the planes of IVHMs and SDDs are compared.
The first, third, and fifth jets exiting the barrel are along the IVHM plane, while the second
and fourth jets is along the SDDs and pump plane.

Figure 5.5: Temperature along IVHM(top) SDD(bottom), LES(left) RANS(right)

In Figure-5.5, the temperatures of LBE jets coming out of the barrel and the temperature
profile inside the barrel are the same for RANS and LES. For both simulations, hot LBE
of 190◦ C from the active inlet flows to the top inside the barrel and then exits through the
upper holes of the barrel, The cold LBE of 180◦ C from the by-pass inlet flows out of the
barrel through the lower holes. The middle jet along the IVHM plane shows a mix of hot
and cold LBE fluid. Away from the barrel, the mixing of the jets results in a mostly uniform
temperature in the upper plenum. The angles of the jets, mainly in the lower exits of the

27
barrel are higher for RANS when compared to LES.

Figure 5.6: Velocity along IVHM(top) SDD(bottom) plane, LES(left) RANS(right)

In Figure-5.6, the peak velocities noticed in the jets are slightly different, mainly for the top
jets exiting the barrel. For LES, the peak velocity leaving the top jet is 0.49m/s while for
RANS it is 0.41m/s. The peak velocities noticed in lower jets, mainly the first and second
jets are similar, although, in RANS, the jets diffuse faster. Along the SDD plane, there is a
difference between the two models. The space between the barrel and SDD is lower for LES.
The jet in LES thus impinges directly on the SDD while for RANS, the jet has time to diffuse
before it reaches the SDD.

The differences in velocities can be attributed to the different sizes of the inlets for both
CFD models. The LES simulation has a higher area for active inlet since it considers the six
annular layers as the active inlet whereas, in RANS, the sixth annular layer is grouped in the
by-pass layer (refer to Section 4.1). Furthermore, the heterogeneity of the porosity offered
by the ACS also plays a factor in the different velocities of the jets. In LES, the porous zone
approach assumes the porosity to be homogeneous.

Looking at the different heights of the upper plenum on a horizontal plane, differences in the
upper plenum away from the barrel are noticed.

In Figure-5.7, the temperature distribution in the upper plenum along the topmost barrel jet
is shown, along with the Turbulent kinetic energy on the same level as seen in RANS. Ex-
amining the top-most jet, i.e. the fifth jet, the porous zone approach which uses LES shows
less mixing when compared to RANS modelling. The jets in RANS lose much momentum
early on due to diffusion and as a result, there is less mixing of fluid further away from the
barrel. This results in cold zones behind the components in the upper plenum. As seen in
Figure-5.7, colder fluid regions, less than 183◦ C exists behind and around the SDDs, pumps,

28
and IVHMS. In LES, there is a much more uniform temperature distribution, mostly around
185◦ − 186◦ C. The high turbulent kinetic energy around the barrel explains the enhanced
diffusion of barrel jets in RANS. The temperature along other planes is also compared in
Appendix B.

Figure 5.7: Temperature and TKE in fifth barrel jet, LES(left), RANS(middle),
RANS(right)

The turbulent kinetic energy distribution for both simulations is compared in Figure-5.8 to
understand the enhanced mixing noticed. In the RANS simulation, higher kinetic energy
is noticed in the jets coming out of the barrel. It is to be noted that the colour band in the
images has been made consistent. The RANS simulation notices a higher peak value of
turbulent kinetic energy value of 0.0174 m2 /s2 . Looking at the jets exiting the barrel, there is
higher turbulence when compared to the LES. The turbulent kinetic energy of the jets exiting
the barrel can be attributed to the flow interaction with ACS. Turbulence can be seen being
generated inside the barrel. This generation is not included in the porous zone approach. This
is further discussed in section-5.3.

Figure 5.8: TKE along IVHM(top) SDD(bottom), LES(left) RANS(right)

Along the various planes of the upper plenum, the jets leaving the barrel were investigated.

29
The LES simulation with a porous zone approach produces a uniform jet behaviour at a par-
ticular height of the barrel. The presence of ACS in the RANS simulation changes this, as can
be seen in Figure-5.9. The figure shows the 4th jet leaving the barrel. In RANS, different jet
behaviours are noticed along different directions. In Figure-5.9 The jets along the direction
of the arrows have a higher radial component compared to other jets.

Figure 5.9: Fourth barrel jet, LES(left) RANS(right)

This kind of non-symmetric nature is not captured by the porous zone approach which applies
an homogenous porosity inside the barrel. The presence of ACS provides a varying cross-
section for jets leaving along various directions through the barrel. In Figure-5.10, examining
the third jet along different free-stream directions, this non-uniform behaviour is seen again.
In Figure-5.10, it is seen that along the different free-stream planes, the temperature of the
jet leaving the barrel is different. A difference in the peak velocity of jets along different
directions is also seen. This non-uniform nature of jets can be attributed to the presence of

Figure 5.10: Third jet along plane-1 (left) and plane-2 (right)

30
the Above Core Structure. The configuration on the above core structure is such that the
fluid from the inlet will face different porosity along the different planes. Along the different
planes, before exiting the barrel holes, the jet faces different cross-section areas due to the
ACS placement. This impacts the velocity at which the jet exits the barrel.

5.2.2 Mixed convection case F20-P80-BP0

Plotting temperature and velocities along IVHM and SDD planes, differences in flow be-
haviour between the two simulations are seen. Figure-5.11 shows the temperature in the
upper plenum. In both cases, hot LBE, 217◦ C, from the active inlet rises up due to buoyancy

Figure 5.11: Temperature along IVHM(top) SDD(bottom), LES(left) RANS(right)

and collects at the top. The colder fluid from the by-pass inlet, 180◦ C, occupies the lower
region of the barrel and exits through the four lower barrel holes. The hotter LBE leaves
through the topmost hole in the barrel. Along the SDD plane, it can is clearly seen that the
hot buoyant plume(hot LBE from the active inlet) rises up and collects at the top.

The difference noticed here is in the angle of the top jet. It is angled downwards in the LES
simulation whereas, in RANS, the jets exit straight. This shows that in LES, the hot plume
accumulates at the top and then rushes down and out from the 5th hole in the barrel, termed
”buoyant jet” in Ashish et al [21]. In RANS, the top jet behaves similarly to that in the forced
convection case along these planes.

Looking at the velocities in Figure-5.12, there are some differences. The topmost jet exiting
the barrel has the same peak velocities, although similar to the forced convection case, the
jet diffuses faster in RANS. Although along the IVHM plane, the first jet is weaker in RANS
when compared to LES. The second barrel jet, as seen in the SDD plane, has higher momen-
tum compared to LES. The cold fluid from the bypass inlet is accelerated due to ACS when

31
Figure 5.12: Velocity along IVHM(top) SDD(bottom) plane, LES(left)
RANS(right)

seen along the SDD plans reaching velocities > 0.1m/s. Along the IVHM plane, this accel-
eration is not seen as the rods and grips provide higher resistance to flow along this plane. In
LES simulation, there is no such increase in velocity for bypass fluid throughout the length
of the barrel.

The overall differences in the results for mixed convection cases are attributed to two factors.
Firstly, the overall height of the model is greater for the LES case. In RANS, the height is
lower which results in more accumulation of hot LBE at the top. There is the accumulation
of hot LBE behind the topmost barrel hole and as such the jet is straight. In LES, the accu-
mulation is further up which forces the hot LBE downwards and out from the topmost jet.
Secondly, the presence of ACS impacts the behaviour of hot buoyant plumes. This will be
discussed in section-5.3.

Figure 5.13: Free-jet along two planes. Plane-1(left), Plane-2(right)

Similar to the forced convection case, a non-symmetric nature is captured from the jets leav-

32
ing the barrel in different directions as seen in Figure-5.13. As seen in the figure, two differ-
ent free jet planes were plotted. All three jets behave differently. The topmost jet is angled
slightly downwards in plane 2 compared to plane 1. The third jet is weaker in plane 2 when
compared to plane 1.

Figure 5.14: Upper plenum temperature along free jet, LES(left) RANS(right)

Investigating the temperature along the free stream direction, the difference in temperature
and hence the stratified layer is seen in the upper plenum, as plotted in Figure-5.14. In RANS,
a higher temperature of around 217◦ C is observed on the top region of the upper plenum
compared to LES which sees around 210◦ C. As can be seen from the graph in Figure-13, the
RANS simulation produces a temperature profile similar to that of the experimental results.
The LES result does not produce the peak temperatures as observed at the top of the upper
plenum in experiments. The temperature along other planes is also compared in Appendix
B. The higher temperature in the upper plenum can be attributed to the topmost jet that exits
straight into the upper plenum. In the LES, this topmost jet is angled downwards as can be
seen in Figure-5.13. In both cases, there is some hot LBE that seeps out of the 4th barrel hole.

Similar to the forced convection case, turbulent kinetic energy was also plotted along the
IVHM and SDD planes in Figure-5.15. The effects of ACS in generating turbulence within
the barrel as is seen in the forced convection case.

5.3 Flow behaviour inside the barrel

The effects of ACS on the flow coming from the core are further investigated in this sec-
tion. The flow behaviour noticed inside the barrel as the inlet flow interacts with the ACS is
discussed and compared with the porous zone approach assumptions.

Velocity and temperature profiles are plotted along various heights inside the barrel in Figure-
5.16 and Figure-5.17. Please note that because of the difference in the heights of the upper
plenum in the two simulations, the various data points may not correspond to the same level
for LES and RANS.

33
Figure 5.15: TKE along IVHM and SDDs planes

Figure 5.16: Velocity inside barrel, LES(top) RANS(bottom)

Looking at velocity at various heights inside the barrel in Figure-26, stark differences between
the porous zone approach and the modelling of ACS are seen. There are high velocities
> 0.4m/s inside the barrel in RANS, whereas, in LES, they peak around 0.2m/s. The peaks

34
Figure 5.17: Temperature inside barrel, LES(top) RANS(bottom)

in velocities occur in the channels between the ACS rods which accelerate the flow. The
porous zone approach produces a constant velocity profile at a particular height inside the
barrel. When modelling the ACS, fluctuations in velocity are seen in the spaces between the
rods. For the area around the central rod, particularly high velocities are seen. This is because
the central rod is not connected to the rest of the ACS through grips. As such the grips don’t
provide any flow blockage around the central rod. The temperature profile inside the barrel
however is similar for both simulations, as seen in Figure-27.

The velocity profile along the SDD plane was also compared. An interesting profile shape
was observed as is seen in Figure-5.18. The difference in profile when compared to the
IVHM plane is due to the fact that the ACS along this plane is arranged differently. We
see fewer rods along this plane and the grips provide axial resistance to the flow. So, the
spaces between the grips see a higher velocity, whereas immediately above the grips, the
fluid is slowed down because of the axial resistance. As mentioned before, this type of three-
dimensionality of flow is not captured by the porous zone approach because of the isotropic
porosity assumption.

Turbulent kinetic energy is of interest, especially since the porous zone approach shows lam-

35
Figure 5.18: Velocity along SDD plane, LES(top) RANS(bottom)

inar behaviour inside the barrel. The ACS inside the barrel induces turbulence which can be
seen when comparing the profiles along various heights inside the barrel. Figure-5.19 shows
TKE along SDD and IVHM planes, all scaled to a maximum value of 0.01. While the ab-
solute value of turbulent kinetic energy depends on the assumptions for the inlet of the CFD
model, the generation can still be looked at to understand the effects of ACS. Higher values
of turbulent kinetic energy are observed above the grips of ACS. As the flow from the inlet
interacts and is slowed down by the grips, turbulence is generated. This generation becomes
less prominent along the height of the barrel as mixing becomes more prominent, as can be
seen along the SDD plane in Figure-5.19.

To understand the impact of the grips on the flow, three planes were examined in Figure-5.20.
These planes were placed immediately before and after the first grip and one across the grip
itself. In the plane that cuts across the grips, we see an acceleration of velocity due to reduced
cross-section area. The flow results are shown below for the three plans. Before the grip,
the velocity is mostly uniform. Immediately after the grips, the flow velocity is significantly

36
Figure 5.19: TKE inside the barrel

reduced. As mentioned before, the flow around the central rod is not reduced because it does
not contain any grips, thus no resistance to flow. This difference between the maximum and
minimum velocity, as seen below, slowly diminishes the further up we go along the barrel.

Figure 5.20: Velocities before, at, and after the grips on the Above Core Struc-
tures

While the above-mentioned impacts of ACS are valid for both forced and mixed convection
cases, there are other effects as well, as seen in mixed convection cases. The ACS impacts
the behaviour of the hot LBE(buoyant plume) inside the barrel.

As can be seen in Figure-5.21, the buoyant plume has a different profile depending on how
much of the ACS it interacts with. The right plane behaves similarly to the forced convection
case, as the hot plume rises and goes straight out through the topmost barrel hole. In the left
plane, we see a more buoyant behaviour. The plume rises to the top, accumulates, and then
comes down and exits through the topmost barrel hole. This explains why the angle of the
topmost barrel jet in the left plane of Figure-5.21 is tilted downwards.

37
Figure 5.21: Top barrel jet along different free jet directions.

38
6 CONCLUSION and FUTURE WORK

From the experiments performed on the ESCAPE facility, the temperature in the upper plenum
was compared with RANS results. A good agreement was noted between the two. For the
forced convection case, a uniform temperature distribution is seen in the upper plenum with a
temperature difference of less than 1◦ C. This is in agreement with the experimental data. For
the mixed convection case, the minimum and maximum temperature values are observed in
RANS similar to the experiment. In the LES results, overall a slightly higher temperature was
observed in the forced convection case. In the mixed convection case, LES does not capture
the peak temperature values on the top of the upper plenum as is noticed in experiments.

Comparing results with LES for forced convection case, a similar temperature profile is no-
ticed both inside the barrel as well as in the jets exiting the barrel. The peak velocities of jets
exiting the barrel are similar, although for RANS these jets diffuse faster. The diffusion is
attributed to higher turbulent kinetic energy observed in RANS. High TKE was noted inside
the barrel due to the flow interaction with ACS. High TKE is also seen in jets coming out of
the barrel. High momentum jets in LES induced greater mixing in the upper plenum, whereas
in RANS, cold spots are observed behind the components as the jets in RANS fail to induce
much mixing far away.

Significant differences were observed between RANS and LES for mixed convection cases.
The topmost barrel jet in LES is angeled downwards. In RANS, the same jet exits straight into
the upper plenum, similar to the forced convection case. Inside the barrel, the hot LBE rises
to the top for both cases. Although for RANS, it occupies a larger area at the top whereas,
in LES, it accumulates higher forming the free surface and turns down and out through the
topmost barrel holes. The velocities of jets are similar for both, although similar to forced
convection, the jets diffuse faster in RANS. The difference between LES and RANS arises
due to the difference in the height of the upper plenum between the two, as well as the effects
of ACS on the buoyant LBE.

Effects of ACS on barrel jets are seen in both cases. Interaction with ACS leads to a non-
symmetric jet behaviour exiting the barrel. The velocity profile inside the barrel is non-
uniform for RANS, unlike LES. High velocities are observed in regions around the core rod
of ACS in RANS compared to LES. The ACS contributes to generating turbulence, which
is not included in the porous zone modelling. For mixed convection cases, ACS impacts
the behaviour of a hot plume rising inside the barrel. On one plane, the hot plume rises and
exits straight out of the topmost barrel hole, similar to the forced convection case. Along
another plane, the hot plume rises up and then rushes downwards and out at a downwards
angle, similar to the LES case. This sort of varied behaviour is not captured by the porous
zone approach.

39
While porous zone approach simulation captures the flow resistance of ACS to a certain
extent, it does not capture the flow behaviour inside the barrel. Homogeneous assumptions
produce symmetric barrel jet behaviour, which is not true in real scenarios. Although, for
capturing the behaviour in the upper plenum for the forced convection case, the porous zone
approach produces similar results, and the homogeneous assumptions seem to have a low
impact on the overall flow in the upper plenum. Although in cases where detailed flow inside
the barrel is important, for eg: flow-induced vibrations on ACS, a porous zone approach is
unsuitable. For the mixed convection case, it is found that the ACS has a substantial impact
on the hot fluid coming from the inlet. The interaction produces a non-symmetric nature that
impacts the overall temperature in the upper plenum. Since the porous zone approach does
not capture the non-symmetric flow behaviour, it is not suitable for use in mixed convection
cases. This is seen by the lower temperatures noticed in the upper plenum for the LES case,
showing that the porous zone approach affects the temperature profile of the upper plenum,
and fails to capture the peak temperature as noticed in experiments.

It is however to be noted that a conclusive statement cannot be made at this stage since the
height of the upper plenum is different in the two models and thus may impact the flow
behaviour.

6.1 Future work

Time constraints limited the depth of investigation that could be done through the means
of simulation. The incorporation of further details in the construction of the model may
provide further insight into the upper plenum flow behaviour. Further work can be done
on developing more accurate models and investigating other thermal-hydraulic phenomena
occurring in ESCAPE.

6.1.1 Further geometric detail

In building the CFD model for this work, certain simplifications were made due to the lim-
itation of time and resources. Including further details in the geometry will provide a more
accurate insight into the flow behaviour in the upper plenum. There exist various areas that
need further investigation to better grasp the flow in ESCAPE. Firstly, the core outlet in this
work was treated as a surface with uniform flow through it. In reality, there are jets that
release fluid into the barrel. As such the velocity of LBE into the barrel is much higher than
what is used, >5m/s. The interaction between these jets from the core and ACS is of interest
and needs further investigation. It is expected that such an interaction will impart unsteady
behaviour to the flow. A transient simulation in such a case will be important. This will also

40
require better mesh refinement in the region along with proper resolving of the boundary
layer. Some work on this has been done by Toti et al, although since a coupling approach
between RANS and system code was used, there were certain limitations to the accuracy of
results noted in the paper[3]. In the paper, cold LBE re-entering the by-pass outlet of the
core is noted. Since they do not model the core, an accurate representation of this re-entering
of LBE in the core is thus not obtained. Full-scale RANS modelling of the upper plenum
including the core and core outlet jets is required to look at the recirculation of flow.

Figure 6.1: Core outlet jets interacting with ACS [3]

The height of the upper plenum was fixed in this model to save time on meshing. It is however
to be noted that at lower pump power, the height of the upper plenum is higher. So the mixed
convection case should have a higher upper plenum height. The Volume of Fluid approach
can be used in conjunction with transient simulation to let the upper plenum establish its own
free surface. In such a case, we can also examine if the free surface level sees any fluctuation.

6.1.2 Improving the accuracy of the RANS model


As discussed earlier, there are limitations to modelling low-Prandtl fluid with the RANS
approach. Building upon the literature review, there are further implementations that can
be made to the RANS model to improve the accuracy of heat transfer modelling for lead-
bismuth eutectic fluid and provide a similar representation of heat transfer phenomena as
LES simulations.

Currently, a log law profile is being used for temperature, similar to the velocity profile. This
does not provide an accurate representation of the temperature profile. Implementing mixed
law developed by Dupencheel et al will provide a better representation of the temperature
profile in the region of y + = 70 − 300 [29]. Furthermore, the Prandtl number has been fixed
to P rt = 2.0. This value is accurate in the near-wall region as recommended by Dupencheel
et al. Although, Kay’s model should be implemented in the bulk of the flow to provide a

41
better representation of turbulent Prandtl number [29].

Inside the barrel. flow recirculation is observed as the flow interacts with ACS. While the
Realisable k −  model is an improvement over k −  model for such flow scenarios, it is
known that k − ω based models perform better for such flow scenarios. These models require
a highly refined mesh near the wall, with a y + < 1. A consistent y + is difficult to achieve,
especially when modelling ACS. In such a case, the k − ω SST model is recommended. It
can work with a range of y + values.

It is to be noted that the above recommendations for heat transfer models and turbulence
models can be used together to provide a better representation of both the flow behaviour
as well as heat transfer. Although, as mentioned by Duponcheel et al, the mixed law model
needs to be used cautiously for complex flow phenomena like separation.

6.1.3 Further thermal hydraulic investigation


There are many other phenomena of interest in the upper plenum that can be investigated to
get a more holistic understanding of the conditions in the upper plena.

As mentioned before, the interaction of jets with ACS produces some interesting flow be-
haviour. Transient simulation can help find flow-induced stresses faced by ACS and the
barrel. This analysis will be useful to assess the long-term structural integrity of the compo-
nents.

As ACS comes in contact with the mixing of cold and hot LBE from by-pass and active inlet
respectively, it becomes important to investigate thermal stripping phenomena. This phe-
nomenon is of interest for all the structural components in the upper plenum since it results
in overall thermal fatigue and reduces the life of the material. Thermal stripping phenomena
cannot be captured by RANS simulation. LES simulation has the capability to capture these
high-frequency fluctuations of temperature and hence should be used to investigate this phe-
nomenon. It will however increase overall resource requirements, so the time required to run
these simulations should be taken into consideration.

Another phenomenon important to investigate is free surface entrainment. The free surface
has an argon gas layer above it maintained at a pressure of 4 bar. Looking at the interaction
between the free surface level and with argon gas layer is key to investigating if gas enter-
tainment takes place during operating conditions, and what possible effects this could have
on the flow behaviour.

Overall there are various avenues that need further investigation for proper design and un-
derstanding of liquid metal fast reactors. Computational Fluid Dynamics can be a key tool
for these investigations. Further development of CFD models can help increase the accuracy
and feasibility of these simulations for wider use in the nuclear industry.

42
7 Bibliography
[1] OECD and Nuclear Energy Agency, “Handbook on Lead-bismuth Eutectic Alloy
and Lead Properties, Materials Compatibility, Thermalhydraulics and Technolo-
gies,” Nov. 2015. [Online]. Available: https://www.oecd-ilibrary.org/nuclear-
energy/handbook-on-lead-bismuth-eutectic-alloy-and-lead-properties-materials-
compatibility-thermalhydraulics-and-technologies_42dcd531-en

[2] D. C. Visser, F. Roelofs, F. Mirelli, and K. Van Tichelen, “Validation of CFD analyses
against pool experiments ESCAPE,” Nuclear Engineering and Design, vol. 369,
p. 110864, Dec. 2020. [Online]. Available: https://www.sciencedirect.com/science/
article/pii/S0029549320303587

[3] A. Toti, J. Vierendeels, and F. Belloni, “Extension and application on a pool-type


test facility of a system thermal-hydraulic/CFD coupling method for transient flow
analyses,” Nuclear Engineering and Design, vol. 331, May 2018. [Online]. Available:
https://www.sciencedirect.com/science/article/pii/S0029549318300815

[4] “ANSYS FLUENT 12.0 Theory Guide - 18.1.1 Pressure-Based Solver.” [Online].
Available: https://www.afs.enea.it/project/neptunius/docs/fluent/html/th/node361.htm

[5] “The first nuclear reactor, explained | University of Chicago News.” [Online].
Available: https://news.uchicago.edu/explainer/first-nuclear-reactor-explained

[6] “Nuclear Power Today | Nuclear Energy - World Nuclear Association.” [On-
line]. Available: https://world-nuclear.org/information-library/current-and-future-
generation/nuclear-power-in-the-world-today.aspx

[7] I. Dincer, Ed., Comprehensive Energy Systems, iSBN: 9780128149256. [On-


line]. Available: http://www.sciencedirect.com:5070/referencework/9780128149256/
comprehensive-energy-systems

[8] A. S. Arnold, J. S. Wilson, and M. G. Boshier, “A simple extended-cavity


diode laser,” Review of Scientific Instruments, vol. 69, no. 3, pp. 1236–1239,
Mar. 1998, publisher: American Institute of Physics. [Online]. Available: https:
//aip.scitation.org/doi/10.1063/1.1148756

[9] Liquid Metal Coolants for Fast Reactors Cooled by Sodium, Lead and
Lead-Bismuth Eutectic, ser. Nuclear Energy Series. Vienna: INTERNA-
TIONAL ATOMIC ENERGY AGENCY, 2012, no. NP-T-1.6. [Online]. Avail-
able: https://www.iaea.org/publications/8589/liquid-metal-coolants-for-fast-reactors-
cooled-by-sodium-lead-and-lead-bismuth-eutectic

43
[10] K. D. Huff, “Chapter One - Economics of Advanced Reactors and Fuel Cycles,”
in Storage and Hybridization of Nuclear Energy, H. Bindra and S. Revankar, Eds.
Academic Press, Jan. 2019, pp. 1–20. [Online]. Available: https://www.sciencedirect.
com/science/article/pii/B9780128139752000016

[11] “Generation IV Nuclear Reactors: WNA - World Nuclear Association.” [Online].


Available: https://world-nuclear.org/information-library/nuclear-fuel-cycle/nuclear-
power-reactors/generation-iv-nuclear-reactors.aspx

[12] V. Sobolev, J. V. d. Bosch, and P. Schuurmans, “Database of thermophysical


properties of liquid metal coolants for GEN-IV,” Dec. 2011, publisher: SCK CEN.
[Online]. Available: https://inis.iaea.org/collection/NCLCollectionStore/_Public/43/
095/43095088.pdf

[13] A.-D. Pérez-Valseca, S. Quezada-García, A.-M. Gómez-Torres, A. Vázquez-


Rodríguez, and G. Espinosa-Paredes, “Reactor behavior comparisons for two
liquid metal-cooled fast reactors during an event of loss of coolant,” Case
Studies in Thermal Engineering, vol. 16, Dec. 2019. [Online]. Available: https:
//www.sciencedirect.com/science/article/pii/S2214157X19303934

[14] A. MARINO, “Thermal-Hydraulic analysis of liquid metal pool experiment E-SCAPE,”


Jul. 2011. [Online]. Available: https://etd.adm.unipi.it/t/etd-06172011-131323/

[15] “The three-level scaling approach with application to the Purdue University Multi-
Dimensional Integral Test Assembly (PUMA) - ScienceDirect.” [Online]. Available:
https://www.sciencedirect.com/science/article/pii/S0029549398002222?via\%3Dihub

[16] D. Tenchine, “Some thermal hydraulic challenges in sodium cooled fast reactors,”
Nuclear Engineering and Design, vol. 240, no. 5, pp. 1195–1217, May 2010. [Online].
Available: https://www.sciencedirect.com/science/article/pii/S0029549310000300

[17] X. Luo, W. Duan, R. Pan, K. Zhang, T. Ding, and H. Chen, “Whole core
thermal-hydraulic analysis considering inter-wrapper flow phenomena in the liquid
metal cooled fast reactor,” Progress in Nuclear Energy, vol. 154, Dec. 2022. [Online].
Available: https://www.sciencedirect.com/science/article/pii/S014919702200350X

[18] M. Azarian, M. Astegiano, M. Tenchine, M. Lacroix, and M. Vidard, “Sodium


thermal-hydraulics in the pool LMFBR primary vessel,” Nuclear Engineering and
Design, vol. 124, no. 3, Dec. 1990. [Online]. Available: https://www.sciencedirect.
com/science/article/pii/002954939090305H

[19] S. Suyambazhahan, T. Sundararajan, and S. K. Das, “Computational Analysis of


Thermal Striping in Primary Sodium System of Liquid Metal Fast Breeder Reactor

44
Using Finite Volume Method,” Nuclear Science and Engineering, vol. 197, no. 3,
Mar. 2023. [Online]. Available: https://doi.org/10.1080/00295639.2022.2116380

[20] K. Van Tichelen, F. Mirelli, M. Greco, and G. Viviani, “E-SCAPE: A scale


facility for liquid-metal, pool-type reactor thermal hydraulic investigations,” Nuclear
Engineering and Design, vol. 290, pp. 65–77, Aug. 2015. [Online]. Available:
https://www.sciencedirect.com/science/article/pii/S0029549314006104

[21] A. Saxena, M. Falcone, X. Huang, and S. He, “INVESTIGATION OF FORCED AND


NATURAL CONVECTION IN THE UPPER PLENUM OF A POOL-TYPE NU-
CLEAR FACILITY,” 17th UK Heat Transfer Conference (UKHTC2021), Apr. 2022.

[22] A. Filin, V. Orlov, V. Leonov, A. Sila-Novitskij, V. Smirnov, and V. Tsikunov, “De-


sign features of BREST reactors Experimental work to advance the concept of BREST
reactors Results and plans,” pp. 75–90, 2001, iWGFR–104 INIS Reference Number:
32021974.

[23] J. J. Sienicki, A. V. Moisseytsev, P. A. Pfeiffer, W. S. Yang, M. A. Smith, S. J. Kim,


Y. D. Bodnar, D. C. Wade, and L. L. Leibowitz, “SSTAR LEAD-COOLED, SMALL
MODULAR FAST REACTOR WITH NITRIDE FUEL.”

[24] J. Wallenius, E. Suvdantsetseg, and A. Fokau, “ELECTRA: European Lead-Cooled


Training Reactor,” Nuclear Technology, vol. 177, no. 3, pp. 303–313, Mar. 2012,
publisher: Taylor & Francis _eprint: https://doi.org/10.13182/NT12-A13477. [Online].
Available: https://doi.org/10.13182/NT12-A13477

[25] X. Chai, X. Liu, J. Xiong, and X. Cheng, “Numerical Investigation of Turbulent


Heat Transfer Properties at Low Prandtl Number,” Frontiers in Energy Research,
vol. 8, 2020. [Online]. Available: https://www.frontiersin.org/articles/10.3389/fenrg.
2020.00112

[26] F. Chen, X. Huai, J. Cai, X. Li, and R. Meng, “Investigation on the applicability
of turbulent-Prandtl-number models for liquid lead-bismuth eutectic,” Nuclear
Engineering and Design, vol. 257, pp. 128–133, Apr. 2013. [Online]. Available:
https://www.sciencedirect.com/science/article/pii/S0029549313000204

[27] X. Cheng and N.-i. Tak, “Investigation on turbulent heat transfer to lead–bismuth
eutectic flows in circular tubes for nuclear applications,” Nuclear Engineering and
Design, vol. 236, no. 4, Feb. 2006. [Online]. Available: https://www.sciencedirect.
com/science/article/pii/S0029549305003584

[28] M. Jischa and H. B. Rieke, “About the prediction of turbulent prandtl and
schmidt numbers from modeled transport equations,” International Journal of

45
Heat and Mass Transfer, vol. 22, no. 11, Nov. 1979. [Online]. Available:
https://www.sciencedirect.com/science/article/pii/0017931079901340

[29] M. Duponcheel, L. Bricteux, M. Manconi, G. Winckelmans, and Y. Bartosiewicz, “As-


sessment of RANS and improved near-wall modeling for forced convection at low
Prandtl numbers based on LES up to Reτ=2000Reτ=2000,” International Journal of
Heat and Mass Transfer, vol. 75, Aug. 2014.

[30] M. Tarantino, M. Angiolini, S. Bassini, S. Cataldo, C. Ciantelli, C. Cristalli,


A. Del Nevo, I. Di Piazza, D. Diamanti, M. Eboli, A. Fiore, G. Grasso,
F. Lodi, P. Lorusso, R. Marinari, D. Martelli, F. Papa, C. Sartorio, M. Utili,
and A. Venturini, “Overview on Lead-Cooled Fast Reactor Design and Related
Technologies Development in ENEA,” Energies, vol. 14, no. 16, Jan. 2021, number:
16 Publisher: Multidisciplinary Digital Publishing Institute. [Online]. Available:
https://www.mdpi.com/1996-1073/14/16/5157

[31] P. Balestra, F. Giannetti, G. Caruso, and A. Alfonsi, “New RELAP5-3D lead and LBE
thermophysical properties implementation for safety analysis of Gen IV reactors,” Sci-
ence and Technology of Nuclear Installations, vol. 2016, pp. 1–15, Mar. 2016.

[32] T. H. Fanning and J. W. Thomas, “Integration of CFD into systems analysis codes for
modeling thermal stratification during SFR transients,” Jul. 2011. [Online]. Available:
https://www.osti.gov/etdeweb/biblio/22501106

[33] H. Mochizuki and H. Yao, “Analysis of thermal stratification in the upper plenum of
the “Monju” reactor,” Nuclear Engineering and Design, vol. 270, Apr. 2014. [Online].
Available: https://www.sciencedirect.com/science/article/pii/S0029549314000211

[34] J. Schneider, M. Anderson, E. Baglietto, S. B. y Leon, M. Bucknor, S. Mor-


gan, M. Weathered, Z. Wu, and L. Xu, “Thermal Stratification Modeling
and Analysis for Sodium Fast Reactor Technology,” january 2018. [Online].
Available: https://www.academia.edu/75580767/Thermal_Stratification_Modeling_
and_Analysis_for_Sodium_Fast_Reactor_Technology

[35] K. Velusamy, K. Natesan, P. Selvaraj, P. Chellapandi, S. C. Chetal, T. Sundarara-


jan, and S. Suyambazhahan, “CFD STUDIES IN THE PREDICTION OF THERMAL
STRIPING IN AN LMFBR,” 2006.

[36] L. Koloszar, S. Buckingham, P. Planquart, and S. Keijers, “MyrrhaFoam: A CFD


model for the study of the thermal hydraulic behavior of MYRRHA,” Nuclear
Engineering and Design, vol. 312, pp. 256–265, Feb. 2017. [Online]. Available:
https://www.sciencedirect.com/science/article/pii/S0029549316301042

46
[37] K. V. Tichelen and F. Mirelli, “SESAME INTERNATIONAL WORKSHOP THER-
MAL HYDRAULIC EXPERIMENTS IN THE LBE-COOLED SCALED POOL FA-
CILITY E-SCAPE,” SESAME International Workshop Petten, The Netherlands, 2019.

[38] D. C. Visser, S. Keijers, S. Lopes, F. Roelofs, K. Van Tichelen, and L. Koloszar,


“CFD analyses of the European scaled pool experiment E-SCAPE,” Nuclear
Engineering and Design, vol. 358, p. 110436, Mar. 2020. [Online]. Available:
https://www.sciencedirect.com/science/article/pii/S0029549319304674

[39] S. Lopes, L. Koloszar, P. Planquart, K. V. Tichelen, and S. Keijers, “CFD simulation of


the heavy liquid European scaled pool experiment (E-SCAPE),” The 19th International
Topical Meeting on Nuclear Reactor Thermal Hydraulics (NURETH-19), Mar. 2022,
publisher: SCK CEN. [Online]. Available: https://researchportal.sckcen.be/en/
publications/cfd-simulation-of-the-heavy-liquid-european-scaled-pool-experimen

[40] R. Shaheed, A. Mohammadian, and H. Kheirkhah Gildeh, “A comparison of


standard k–ε and realizable k–ε turbulence models in curved and confluent channels,”
Environmental Fluid Mechanics, vol. 19, no. 2, pp. 543–568, Apr. 2019. [Online].
Available: https://doi.org/10.1007/s10652-018-9637-1

[41] M. Sosnowski, J. Krzywanski, K. Grabowska, and R. Gnatowska, “Polyhe-


dral meshing in numerical analysis of conjugate heat transfer,” EPJ Web
of Conferences, vol. 180, p. 02096, 2018, publisher: EDP Sciences.
[Online]. Available: https://www.epj-conferences.org/articles/epjconf/abs/2018/15/
epjconf_efm2018_02096/epjconf_efm2018_02096.html

[42] “Ansys Fluent Mosaic Meshing | CFD Mesh.” [Online]. Available: https://www.ansys.
com/en-gb/products/fluids/ansys-fluent/mosaic-meshing

47
Appendices

A Appendix-A

The bash file written is shown in the figure below:

Bash file for HPC job submission.

The Fluent journal file is in the figure below:

48
Fluent journal file

49
B Appendix-B

The figure below shows the direction of IVHM and SDD planes.

IVHM and SDD planes.

The figure below shows the temperature distribution in the upper plenum for forced convec-
tion condition (F80-P80-BP0) and for mixed convection condition (F20-P80-BP0).

50
Temperature in upper plenum along 5th, 4th and 3rd, RANS(top) LES(bottom) for
F80-P80-BP0

Temperature in upper plenum along 5th, 4th and 3rd, RANS(top) LES(bottom) for
F20-P80-BP0

51

You might also like