You are on page 1of 27

Analysis of Natural Convection Heat Transfer and Solidification within a

Corium Simulant

ed
iew
ev
rr
ee
tp
t no
rin
ep
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
Analysis of Natural Convection Heat Transfer and Solidification within a Corium

ed
Simulant

Mahsa Rezaeea , Dijo Davida , Marilyn Lightstonea , Stephen Tullisa

iew
a McMaster University, 1280 Main St W, Hamilton, ON L8S 4L8

Abstract

ev
In the event of a severe nuclear accident such as that at the Fukushima Daiichi Nuclear Power Plant (NPP) in 2011, it is
essential to minimize the release of radioactive material into the environment. Studies on Canada Deuterium Uranium

r
(CANDU) reactors showed that thermal stresses are a potential threat to vessel wall integrity. There is thus a need
to analyze the heat transfer and fluid flow within the corium and determine the spatial distribution of heat flux at the

er
calandria vessel wall to ensure the ex-vessel cooling is sufficient. Canadian Nuclear Laboratories (CNL) designed and
conducted an experiment to analyze the heat transfer within the corium using a 1/5 scale CANDU calandria vessel
with molten salts as the corium simulant. The current study used computational fluid dynamics (CFD) to simulate
pe
the heat transfer and crust formation in the experiment to provide insight into the flow pattern and heat transfer and to
assess the adequacy of the CFD modelling. Unsteady Reynolds-Averaged Navier-Stokes (URANS) and enthalpy-based
solidification models were used to simulate the experiment.
Keywords: Severe accidents, CANDU corium, Unsteady Reynolds-Averaged Navier-Stokes (URANS) model,
ot

Solidification model, Canadian Nuclear Laboratories (CNL) experiment


tn

1. Introduction

In Canada’s province of Ontario, over 60% of the electricity generation is provided by CANDU nuclear power
rin

plants [1]. While the CANDU reactor has a long history of safe operation, there is a need to ensure that the reactor
vessel maintains its integrity in the event of a severe accident. Severe accidents with some radiological release have
occurred at Three Mile Island NPP, Chernobyl NPP, and most recently in the Fukushima Daiichi NPP. As a response,
regulatory bodies require utilities to ensure that severe accident management strategies such as external reactor vessel
ep

cooling are capable of preventing reactor vessel failure [2].

The CANDU reactor has built-in safety features that support in-vessel retention. Pressure tubes containing fuel
bundles are placed within calandria tubes in the reactor core. Calandria tubes are arranged horizontally in the CANDU
Pr

Email address: Rezaem3@McMaster.ca (Mahsa Rezaee)

Preprint submitted to Elsevier October 3, 2022

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
calandria vessel, as shown in figure 1. Heavy water surrounds the calandria tubes and acts as a passive heat sink which

ed
slows the progression of a severe accident. The calandria vessel is housed within a water-filled concrete vault which
provides additional heat sink capability [3].

iew
r ev
er
pe
Figure 1: CANDU reactors assembly [4]

A station blackout scenario can lead to a severe accident in a CANDU reactor if the operators do not take any
mitigating action within one hour [5]. In a severe accident, pumps are unavailable to remove heat from fuel bundles.
ot

Although the reactor shuts down automatically, decay heat generation continues. In the absence of adequate cooling
and the presence of the decay heat, the fuel bundles heat up, and the calandria tubes gradually fail. If mitigation
actions are not taken, the entire core eventually collapses to the bottom of the calandria vessel, forming a debris bed.
tn

The moderator liquid ultimately evaporates completely, and melting of the debris bed begins near the centre. After
approximately 11-26 hours, depending on the activation of the crash-cooling systems, the debris bed reaches a fully
molten state, forming the homogenous corium mixture. The primary components of the corium fluid are UO2 , Zr, and
rin

ZrO2 [6].

During a severe accident, the main challenge is to minimize radiological release to the environment. Retaining
the corium within the calandria vessel reduces the threat of a significant radiological release into the environment.
ep

Luxat and Luxat [7] showed that in the event that the calandria vessel wall fails, thermal stress concentrations at
the calandria vessel wall are expected to be the dominant failure mechanism. Thermal stresses depend on the wall
temperature distribution, which is controlled by the heat flux distribution at the calandria vessel wall. As a part of
Pr

in-vessel retention analyses, the study of heat transfer within the corium is important to predict the heat flux distribution
at the vessel wall for assurance of the vessel wall integrity.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
The water in the vault, which acts as a thermal reservoir, removes the generated heat from the reactor vessel. Energy

ed
is removed from the corium through radiation between the corium top surface and the calandria vessel interior surface.
Additionally, heat is conducted through the bottom vessel wall and is transferred to the water in the vault [8, 9]. As the
molten corium cools, it solidifies, and crusts form at the boundaries [8, 9].

iew
Heat transfers from the pool top surface and the vessel wall are defined as Qup and Q dn , respectively. The CANDU
corium geometry is cylindrical and shallow with a large top heat removal surface, unlike in pressurized water reactors
(PWR) where the corium geometry is hemispherical and deeper with a smaller top heat removal surface. Consequently,
in CANDU corium, most of the decay heat (approximately 80% of the total) is expected to be removed from the top
surface [8].

ev
Studies developed correlations to calculate the Nusselt numbers (Nu) in upward (Nuup ) and downward (Nudn )
directions as a function of the modified Rayleigh (equation 1) and Prandtl (Pr) numbers [10]. The Ra 0 is defined by

r
thermophysical properties and volumetric heat generation rate in W/m3 . The Pr and Ra 0 numbers for CANDU corium
range between 0.18-0.8 and 1013 -1014 , respectively. For PWR reactors, the Pr and Ra 0 numbers of the molten corium

er
are between 0.25-1 and 1014 -1016 , respectively [8]. The estimated Pr numbers are different since the percentage of Zr
oxidation is expected to be different for CANDU and PWR corium. The PWR corium’s volumetric heat generation
rate and molten pool height are greater, resulting in a higher Ra 0 number.
pe
gβq H 5
000

Ra =
0
(1)
ανk

CANDU corium is a homogenous molten pool; however, PWR corium may stratify into layers composed of a
ot

homogenous molten pool, a metallic layer on the top surface of the molten pool, and possibly a metallic layer at the
bottom of the calandria vessel [9]. In the molten PWR corium, heat transfer forms three main regions: an upper region
tn

with mostly convection heat transfer; a lower region with mainly conduction heat transfer; and boundary layers along
the vessel wall where heat transfer is by both convection and conduction [11]. Numerical analyses of heat transfer
showed a similar flow pattern forms in the molten CANDU corium [12, 13]. These studies, however, did not consider
the effects of both the 3D CANDU geometry and phase change at the boundaries of the domain. The step region in
rin

the calandria vessel (the connection of the main and sub shells) and an approximately 500 K temperature difference
between the corium solidus and liquidus temperatures add complexity to the prediction of the heat transfer and crust
formation within the corium.
ep

Nourgaliev et al. [14] showed that the Pr number influences the flow pattern and heat transfer within the PWR
corium. For lower Pr numbers, the boundary layers along the vessel wall are able to penetrate further toward the
bottom of the vessel due to the lower fluid viscosity. As a result, heat transfer in this region is influenced by convection.
Pr

At higher Pr numbers, the thermal boundary layers do not penetrate as far down resulting in mainly conduction heat
transfer in the lower region of the vessel.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
The majority of existing studies of heat transfer within the corium are for non-CANDU conditions. Experiments

ed
of the natural convection heat transfer in corium pools were summarized in a review work done by Zhang et al. [10].
Some of these experiments tested a 2D semicircular slice geometry. As this geometry approximates the 2D sliced
CANDU corium, the correlations developed for the Nu number can be used to estimate heat transfer within the 2D

iew
CANDU corium. Table 1 is a summary of these experiments.

Table 1: Summary of previous experimental studies that pertain to CANDU conditions [10]

Authors Geometry Ra 0 Pr Boundary conditions

1/4 circular slice

ev
Bonnet [11] Radius: 2 m 1013 − 1017 7 Isothermal top and bottom walls

Thickness: 0.15 m

r
Semicircular slice
Lee et al. [15] 107 − 1010 7 Isothermal top and bottom walls
Radius: 0.125 m

er
pe
1.1. CNL experiment details

Due to limited amount of experimental data for CANDU geometry, CNL designed an experiment to analyze heat
transfer and crust formation within a scaled CANDU corium geometry [16]. The experiment used a 1/5 scale of the
bottom section of the CANDU 6 calandria vessel, including the sub shells. Table 2 summarizes the scaled calandria
ot

vessel and molten pool dimensions.

Table 2: 1/5 scaled CANDU 6 calandria vessel and corium pool dimensions [16]
tn

Main shell radius 760 mm


Sub shell radius 676 mm
Main shell length 793 mm
rin

Sub shell length 197 mm


Molten pool depth 200 mm
ep

During the experiment, a grid of submerged electrical heating elements produced volumetric heating in the molten
pool. The heaters in the main shell were evenly spaced but placed away from the bottom of the vessel to not interfere
with the solidification at the boundaries and effectively create volumetric heat generation only in the liquid domain [16].
Pr

Several values for the heat generation rate were examined to explore the effects of heat generation rates on heat transfer.

In the experiment, the simulant was the molten equimolar solution of NaNO3 and KNO3 [16]. This salt solution

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
has a eutectic melting temperature of 495 K [17]. The salt’s thermophysical properties were estimated according to the

ed
correlations proposed by Serrano-Lopez et al. [18] and Iverson et al. [19], as listed in table 3. The Ra 0 number varies
between 1.2 × 1011 -5.4 × 1011 depending on the volumetric heat generation rates, and the Pr number is 17.4.

Table 3: Summary of the thermophysical properties of the salt solution at a temperature of 507.5 K (the average temperature of the

iew
molten salt pool)

Properties Value
Density (ρ) 1882 kg/m3
Specific heat capacity (Cp ) 1482.6 J/kg/K
Thermal conductivity (k) 0.46 W/(m.K)

ev
Viscosity (µ) 5 × 10−3 Pa.s
Thermal expansion (β) 10−4 1/K
Latent heat (h f ) 100.8 kJ/kg

r
Melting temperature (Tl ) 495 K

er
Cooling was applied to the boundaries of the molten salt pool. Argon gas above the molten salt pool removed heat
via natural convection. Additionally, radiation between the molten salt top surface and the top lid removed heat from
pe
the molten salt [16]. Water outside the vessel (figure 2) cooled the molten salt pool from the bottom wall. Water was
circulated through a chiller to keep it at an almost constant temperature. Cases with different water temperatures were
tested to determine their effects on heat transfer and crust formation within the molten salt pool. The experimental
results showed that variation in the water temperature does not have measurable impacts on the temperature distribution
ot

within the molten salt pool [16].


tn
rin
ep
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
ed
iew
r ev
Figure 2: Left: the experiment apparatus including the bottom of the vessel suspended in water in the tank, Right: cross-section of
the experiment apparatus [16]

er
In the experiment, the effects of variation in decay heat on heat transfer and crust formation were studied by
pe
testing different volumetric heat generation rates. For each volumetric heat generation rate, the system was brought
to a steady-state condition, which takes approximately 10000 seconds. The temperature distribution was measured
by thermocouples, and the crust thickness was measured by several mechanical probes. The current study simulates
two of the heat generation rates used in the CNL experiments. The total heat generation rates of 7.18 kW (average
53.3 kW/m3 ) and 31.64 kW (average 234.9 kW/m3 ) are modelled. These two cases have Ra 0 values of 1.2 × 1011 and
ot

5.4 × 1011 (based on the total depth of the salt pool), respectively.
tn

2. Methodology

The current study uses CFD to predict the heat transfer and fluid flow in the corium simulant. The goal is to
rin

assess the adequacy of the turbulence and the two-phase models by comparing model predictions to the experimental
data from the CNL experiment. The governing equations are mass conservation, momentum conservation with a
Boussinesq buoyancy model, and energy conservation as defined in equations 2, 3, and 4.
ep

∂Ui
ρ =0 (2)
∂ xi
Pr

∂Ui ∂Ui ∂P ∂ 2Ui ∂ui0u 0j


ρ + ρU j =− +µ 2 −ρ − ρβgi (T − Tl ) + AUi (3)
∂t ∂ xj ∂ xi ∂ xj ∂ xj

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
ed
∂T ∂T ∂ 2T 000
ρCp + ρCp U j = k 2 − ρCp u 0j T 0 + q + Sh (4)
∂t ∂ xj ∂ xj

Where U, P, and T are time-averaged velocity, pressure, and temperature. Velocity and temperature fluctuations are

iew
u 0 and T 0. Unsteady RANS turbulence modelling was applied. The k − ω model [20] was selected based on validation
studies described in this section.

∂(ρh f ) 
The energy conservation equation was modified with a source term Sh = − ∂t to include the phase change
effects. In the momentum equation, a damping source (AUi ) was introduced to model the liquid-solid interface. The

ev
coefficient (A) in the damping source is defined as equation 5 [21].

(1 − ωl )2
A = −C (5)
ωl3 − 

r
Where  is a very small number (0.001) to avoid division by zero. In equation 5, C is the mushy zone constant.

er
Kumar and Krishna [21] showed that the value of the mushy zone constant impacts the prediction of the liquid-solid
interface; so, different values of the mushy zone constant were tested during the validation to make sure the results are
pe
independent of this constant. At the liquid-solid interface, ωl is the liquid mass fraction and is calculated based on
the enthalpy difference between liquid and solid phases (equations 6). The equations are solved using the commercial
CFD code ANSYS-CFX 2019 R3.
ot

h − hs
ωl = (6)
hl − hs

A suite of preliminary validation studies was performed to assess the capability of the code. The validation test cases
tn

are summarized in table 4.


rin
ep
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
Table 4: Summary of validation steps

ed
Non dimensional
Validation cases Geometry Material Objective
numbers
Turbulent natural Assessment of the
Square cavity Air Ra=1.58 × 109

iew
convection experiment [22] turbulence model
0.75 m×0.75 m Pr=0.7
Simulant internal gravitated Assessment of
material apparatus Rectangular cavity modelling the
Air Ra 0 = 2.34 × 109
rectangular pool (SIGMA 0.5 m×0.5 m×0.16 m volumetric heat
Pr=0.7

ev
RP) experiment [23] generation
Laminar natural convection
Assessment of the
with solidification Rectangular cavity Molten tin Ra = 2.7 × 105
solidification model
experiment [24] 0.067 m×0.089 m Pr=0.03

r
er
The first validation case considered turbulent natural convection in a differentially heated cavity. The key parameters
and geometry were summarized in table 4. Results confirmed the previous studies’ results [25, 26] that the k − ω
turbulence model predicts the temperature distribution and velocity profile in the differentially heated cavity with the
pe
maximum error of 8.6%. Therefore, the transient k − ω turbulence model was used in the current study.

The second validation study explored turbulent natural convection with volumetric heat generation in a cavity. The
top surface was a copper plate with a thickness of 5 mm and was cooled by the flow of water on the outside. The other
walls were insulated [23]. Therefore, in the simulation, a conjugate heat transfer boundary condition was applied to
ot

the top surface, and the rest of the walls were set to adiabatic. Heat generation within the cavity was experimentally
simulated by forty small discrete heaters [23]. In the numerical simulation, two cases were modelled to ensure the
tn

heaters can be simulated as a uniform heat generation source within the entire fluid. In the first case, the discrete
heaters were modelled with the spacing shown in figure 3, and in the second case, a uniform volumetric heat generation
source was applied to the entire domain.
rin
ep
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
ed
iew
r ev
Figure 3: SIGMA RP geometry [23]

er
Figure 4 compares the experimental and numerical results. The simulation accurately predicted the experimental
results. A comparison between the results of the two simulation cases shows that the case with the discrete heaters
pe
predicted the experimental results more accurately. The simulation of the discrete heaters is computationally expensive.
The maximum difference between the predicted dimensionless temperatures from both simulations is approximately
0.004 (8% error). This error is less than the uncertainties of the experiment, which is about 12%; therefore, the discrete
heaters can be modelled as the uniform volumetric heat generation source to reduce the computational costs.
ot
tn
rin
ep
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
0.50

ed
0.40

iew
0.30
Height [m]

0.20

ev
0.10

r
Experiment [23]
Numerical- discrete heaters
Numerical- uniform volumetric heating
0.00
0.00 0.01 0.02
er 0.03 0.04
Dimensionless temperature
0.05 0.06
pe
Figure 4: A comparison between the SIGMA RP experimental and simulation results

The third validation test explored the ability of the CFD model to predict phase change. Wolf and Viskanta [24]
performed experiments on a differentially heated cavity with molten tin. Details of the experiment performed by Wolf
ot

and Viskanta [24] are listed in table 4. During the experiment, the side walls were kept at constant temperatures of
233°C and 229°C, and the top and bottom walls were insulated [24]. The molten tin started solidifying from the colder
boundary. As shown in figure 5, the crust formation was distorted by the convection circulation in the liquid domain.
tn

A comparison between the experimental and numerical results shows a good agreement in figure 5. A range of values
for the mushy zone constant was tested to check the independence of the solution from this constant. Numerical results
indicate that the predicted phase liquid-solid interface is independent of the mushy zone constant when the constant is
rin

107 kg/(m3 .s).


ep
Pr

10

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
0.07

ed
0.06

0.05

iew
0.04

Height [m]
0.03

0.02
Experiment [24]

ev
0.01 Numerical- C=107 kg/(m3 .s)
Numerical- C=108 kg/(m3 .s)
0.00
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
Width [m]

r
Figure 5: A comparison between the liquid-solid interface predicted by the simulation and experiment [24]

3. Simulation of the CNL experiment


er
pe
As discussed in section 1.1, the CNL experiment was performed on a 1/5 scale bottom section of the CANDU
6 calandria vessel. The experiment is able to capture the three-dimensional flow effects arising from the sub shell
regions of the vessel. Crust formation is emulated through the phase-change within the corium simulant. In the CNL
experiment, volumetric heating was applied to the central region, approximately 30 mm from the vessel wall, to avoid
ot

interfering with the boundary layer and crust formation at the vessel wall. Similarly, in the simulation, the volumetric
heat generation source was applied to the centre region of the main shell (30 mm away from the main shell vessel
tn

wall) and the entire region of the sub shells. Volumetric heating might have influenced the near vessel wall region. A
second case was simulated with a uniform volumetric heat generation source applied to the entire domain to address
this uncertainty. Table 5 presents a summary of the simulation cases.
rin
ep
Pr

11

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
Table 5: Summary of simulation cases

ed
Total heat generation Non dimensional
Case number Volumetric heating model
rate numbers
Uniform volumetric heat

iew
7.18 kW Ra 0 = 1.2 × 1011 Case 1 generation source in the central
Pr = 17.4 region (q 000=64.6 kW/m3 )
Uniform volumetric heat
Case 2 generation source in the entire
domain (q 000=53.3 kW/m3 )

ev
Uniform volumetric heat
31.64 kW Ra 0 = 5.4 × 1011 Case 3 generation source in the central
Pr = 17.4 region (q 000=285 kW/m3 )

r
Uniform volumetric heat
Case 4 generation source in the entire

er domain (q 000=234.9 kW/m3 )


pe
At the top surface, a combination of radiation and free convection removed heat from the molten salt pool. The
molten salt is well mixed at the top surface with almost uniform measured temperatures at several points when steady
state was reached. The experimentally measured temperatures are used as the boundary conditions at the top surface of
the molten salt pool and are 520 K (for the volumetric heat generation rate of 7.18 kW) and 597 K (for the volumetric
ot

heat generation rate of 31.64 kW).

Heat removal from the top surface was a combination of radiation and natural convection heat transfer. The radiation
tn

between the molten salt top surface, top lid, and side walls was estimated by assuming three enclosed surfaces [17].
The radiation network is shown in figure 6. The net radiation from the molten salt pool top surface was calculated
by equations 7-9. The measured temperatures of the top lid and side walls (T2 and T3 ) were 403 K throughout the
experiments. Table 6 summarizes the properties of the three surfaces.
rin
ep
Pr

12

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
ed
iew
ev
Figure 6: Radiation network on the top surface of the molten salt pool

r
σT14 − J1 J1 − J2 J1 − J3
= + (7)
(1 − 1 )/1 A1 1/A1 F12 1/A1 F13

σT24 − J2
=
er
J2 − J1
+
J2 − J3
(1 − 2 )/2 A2 1/A2 F21 1/A2 F23

σT34 − J3
=
J3 − J1
+
J3 − J2
(8)

(9)
pe
(1 − 3 )/3 A3 1/A3 F31 1/A3 F32

Table 6: Properties of the surfaces above the molten salt pool

Wall Emissivity () Area


1.11 m2
ot

Molten salt top surface 0.96


Side walls 0.22 1.24 m2
Top lid 0.22 1.11 m2
tn

Argon gas was injected above the top surface at room temperature and removed heat from the molten salt pool.
Although the argon gas flowed over the top surface, its velocity was very low; hence the heat loss due to convection
rin

can be estimated by natural convection heat transfer correlations above the heated surface (equation 10) [27].

Nu = 0.15Ra1/3 (10)
ep

The bottom wall thermal boundary condition is an effective heat transfer coefficient based on the bottom wall heat
flux. This bottom wall heat flux was estimated by calculating the top surface heat flux and subtracting that from the
total heat input. Details on the calculation of the effective heat transfer coefficient are given below.
Pr

The heat transfer coefficient for the bottom boundary (Ub ) was estimated by equation 11, where q and qtop are the

13

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
volumetric heat generation rate and heat removal from the top surface, and Avw is the area of the vessel wall (1.49 m2 ).

ed
The temperature at the vessel wall was not uniform; hence, the average of measured temperatures at the vessel wall
(T vw ) was used. The water temperatures (Tw ) were 332 K and 328 K for the total volumetric heat generation rates of
7.18 kW and 31.64 kW. A heat flux boundary condition is applied to the entire bottom vessel walls. The heat flux is

iew
calculated using a loss coefficient (equation 11 and table 7) and the difference between the measured water temperature
and the locally calculated domain boundary temperature. As such, the wall heat flux varies both circumferentially and
axially on the bottom surface of the domain.

q − qtop
Ub = (11)

ev
Avw (T vw − Tw )

Table 7 summarizes the boundary conditions used in the simulation.

Table 7: Summary of simulation boundary condition

r
Momentum
q Boundary Heat transfer

7.18 kW Top surface


Bottom walls Ub =15.2
er
Constant temperature=520 K
W/(m2 .K) Tw =332 K
condition
Free slip wall
No slip wall
pe
31.64 kW Top surface Constant temperature=597 K Free slip wall
Bottom walls Ub =105.6 W/(m2 .K) Tw =328 K No slip wall

In the simulation, the unsteady k − ω turbulence model and the solidification model with the momentum source
ot

constant of 107 kg/(m3 .s) based on the validation cases presented in section 2. Time step and grid independence were
tested. Results were independent with a time step of 0.1 seconds and a grid with 1.7 million nodes, as shown in
tn

figure 7. The simulations were run for 10000 seconds to ensure that a quasi-steady-state condition was reached. This
time is also consistent with the experimental procedure. Monitored temperature points in the simulation reached steady
temperatures with a maximum standard deviation of 1 K.
rin
ep
Pr

14

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
ed
iew
r ev
er
Figure 7: Computational domain and the grid
pe
4. Results

Figures 8 and 9 show the fluid velocity vectors and temperature at the mid-vessel cross-section. As can be seen in the
figures, there are three main zones in the liquid domain. Natural convection circulation cells form in the upper portion
of the vessel and act to mix the fluid yielding a nearly uniform temperature, as shown in figure 9. The temperature in
ot

this zone is mainly influenced by cooling from the top surface and the volumetric heat generation rate. Boundary layers
form along the vessel walls. The descending boundary layer grows to roughly a thickness of 7 mm with a maximum
tn

velocity of 0.008 m/s. As the boundary layer fluid penetrates lower in the vessel, it locally melts some of the crust; as
a result, there is a jump in the crust thickness at the lower portion of the boundary layer region, as seen in figure 8.
This jump in the crust thickness was independent of the grid resolution and was also consistent with the non-uniform
crust thickness seen experimentally. The sharpness is most likely, however, a result of the shape of the grid (rectangle
rin

grids) imposed in that region. The crust formation along the vessel wall in the upper region depends on the temperature
distribution in the top zone, the formation of the boundary layer, and the cooling from the vessel wall. In the bottom
region, the fluid is relatively stationary and is thermally stratified, as shown in figures 8 and 9. The temperature in the
ep

lower region is affected by cooling from the vessel wall.


Pr

15

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
ed
iew
ev
Figure 8: Instantaneous velocity vectors at the mid-vessel cross-section for case 1, the solid blue area is the crust

r
er
pe
ot

Figure 9: Instantaneous temperature isotherms at the mid-vessel cross-section for case 1


tn

Figures 10 and 11 show the velocity vectors and temperature distribution at the cross-section midway between the
step and end shield. The molten pool is shallower in the sub shell region; hence, the two boundaries along the vessel
rin

wall penetrate to the bottom of the vessel and mix on top of the crust, as seen in figure 10. Since the boundary layers
extend to the bottom of the vessel, there is no jump in the crust thickness. The mixing of the boundary layers and
convection circulations in the top region form a uniform high-temperature region on the top of the crust, as shown in
ep

figure 11; as a result, the crust thickness is lower in the sub shell compared to the main shell.
Pr

16

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
ed
iew
ev
Figure 10: Instantaneous velocity vectors at the sub shell cross-section for case 1, the solid blue area is the crust

r
er
pe
ot
tn

Figure 11: Instantaneous temperature isotherms at the sub shell cross-section for case 1

The instantaneous velocity and temperature distributions at the mid-vessel axial section of the molten salt are
rin

shown in figures 12 and 13, respectively. There are circulations in the upper half of the main shell, as seen in figure 12.
The overall axial circulation is towards the end shields and returns at roughly mid-depth of the pool. This circulation
results in a relatively constant higher velocity at the edge of the step region, as seen in figure 12. In the main shell, the
temperature in the upper region is relatively uniform due to the convection mixing of the fluid, and in the lower region,
ep

temperature stratifies, as seen in figure 13. In the sub shell, the molten salt pool is shallower; hence, there is only the
convection zone with a uniform temperature on the top of the crust. The lowest temperatures are seen at the bottom
corner of the main shell, where it connects to the sub shell. This is due to the higher heat transfer surface area due to
Pr

the vertical step. In the step region, cooling occurs both from the bottom and vertical walls.

17

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
ed
iew
ev
Figure 12: Instantaneous velocity vectors at the mid-vessel axial section for case 1, the solid blue area is the crust

r
er
pe
ot

Figure 13: Instantaneous temperature isotherms at the mid-vessel axial section for case 1
tn

The impact of an increased volumetric heat generation rate is explored in cases 3 and 4, which have a total heat
generation rate of 31.64 kW. In case 3, similar to case 1, the volumetric heat generation source was applied to the
rin

central region of the molten salt pool; while, in case 4 (similar to case 2), the heating source was applied uniformly
to the entire domain. The increase in the volumetric heat generation rate decreases the crust thickness. Although the
temperatures are higher in cases 3 and 4 in comparison to the lower heat generation rate studies of cases 1 and 2, the
ep

overall profile shape is similar. This is due to the similar flow patterns with natural convection mixing dominating the
top section (figure 14) and yielding a fairly flat temperature profile (figure 15). Conduction heat transfer dominates in
the lower region of the molten pool. Similar to case 1, a boundary layer forms along the vessel wall with an increased
crust thickness at the bottom portion of the boundary layer region. Overall, the crust is approximately 25% thinner
Pr

compared with the lower heat generation rate cases.

18

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
ed
iew
ev
Figure 14: Instantaneous velocity vectors at the mid-vessel cross-section for case 3, the solid blue area is the crust

r
er
pe
ot
tn

Figure 15: Instantaneous temperature isotherms at the mid-vessel cross-section for case 3

Figure 16 shows the predicted and measured location of the crust boundary on the two symmetry plane profiles
rin

through the vessel. As expected, the crust thickness for case 1 (with applying the heating in the centeral region) is
slightly thicker than for case 2 (with uniform heating throughout the domain). The predicted crust thicknesses for both
cases are close to the experimental results. Both cases show the crust is thinner along the vessel wall in the upper
ep

zone since the temperature is high in the top region and the formation of the boundary layer along the vessel wall, as
discussed previously. The predicted crust thickness remains almost constant along the main shell in the axial direction;
however, it increases near the step region, similar to the measured crust thickness in the experiment. At the base of
the step region, the crust thickness is maximum and reaches a minimum of 3 mm at the lip of the step region due to
Pr

the protrusion of the step geometry into the domain. In the sub shell, the crust thickness is lower, as discussed earlier
(figures 10 and 11), and uniform in the axial direction. Crust does not form along the end shield since the heat removal

19

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
from the molten salt pool is inadequate.

ed
0.20

iew
Numerical result, case 1
Numerical result, case 2
Molten pool height (m)

0.15 Experimental result (q = 7.18 kW) [16]


Vessel wall

0.10

ev
0.05

0.00
−0.55 −0.45 −0.35 −0.25 −0.15 −0.05 0.05 0.15 0.25 0.35 0.45 0.55

r
Distance from the axial centreline (m)
0.20

er
Molten pool height (m)

0.15

0.10
pe
0.05

0.00
−0.6 −0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4 0.5 0.6
ot

Axial position (m)

Figure 16: A comparison between liquid-solid interface predicted by simulation and experiment for the total heat generation rate
of 7.18 kW, top: axial-section, bottom: cross-section; case 1: volumetric heat generation source applied to the central region, case
tn

2: volumetric heat generation source applied to the entire domain

Measured and predicted temperature profiles along a vertical line in the centre of the domain are shown in figure 17.
rin

The predicted temperature is roughly uniform in the upper half of the molten salt pool, which agrees well with the
experimental results. Below this well-mixed region, the temperature stratifies in the bottom of the molten salt pool
above the crust. As expected, there is a large temperature gradient within the solid crust. Case 1, with its volumetric heat
generation approximating the CNL heater locations, has lower temperatures near the bottom wall (and thicker crust)
ep

than the uniformly heated case 2; hence, case 1 shows a better agreement with the lowest experimental temperature
measurement. The location of this temperature measurement point is very close to the measured crust depth, although it
is substantially below the melting temperature (495 K), which shows uncertainties in the crust thickness measurements.
Pr

20

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
530

ed
520
510
500
490
Temperature (K)

480

iew
470
460
450
440
crust (case1)
430
crust (case2) Numerical result, case 1
420

ev
Numerical result, case 2
410
Experimental result (q = 7.18 kW) [16]
400
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20
Molten pool height (m)

r
Figure 17: A comparison between temperature predicted by the simulation and measured in the experiment for the total heat
generation rate of 7.18 kW; case 1: volumetric heat generation source applied to the central region, case 2: volumetric heat
generation source applied to the entire domain

er
Figures 18 and 19 show the comparison between the experimental and numerical results for the highest volumetric
pe
heat generation rate (31.64 kW). The predicted crust thickness in the middle of the main shell (approximately 35
mm) is substantially lower than the crust thickness (40 mm) when the heat generation rate is 7.18kW. Throughout
the vessel, the crust profiles are similar to the lower heat generation rate cases. Notably, there is no crust formation
at the upper edge of the step region. The discrepancy between predicted crust thickness in cases 3 and 4 is more
ot

prominent than in cases with the lower heat generation rate. Case 4 (uniform heat generation source) underpredicts
the crust thickness, particularly along the axial centreline. Similarly, the predicted temperatures by case 3 are in better
agreement with the near-wall temperature measurement, as seen in figure 19. Both cases show similar temperature
tn

stratification immediately above the crust and a well-mixed uniform temperature in the top half of the pool.
rin
ep
Pr

21

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
0.20
Numerical result, case 3

ed
Numerical result, case 4
Molten pool height (m)

0.15 Experimental result (q =31.64 kW) [16]


Vessel wall

0.10

iew
0.05

0.00
−0.55 −0.45 −0.35 −0.25 −0.15 −0.05 0.05 0.15 0.25 0.35 0.45 0.55
Distance from the axial centreline (m)

ev
0.20
Molten pool height (m)

0.15

r
0.10

0.05

0.00
er 0 0.1 0.2 0.3 0.4 0.5 0.6
pe
−0.6 −0.5 −0.4 −0.3 −0.2 −0.1
Axial position (m)

Figure 18: A comparison between liquid-solid interface predicted by simulation and experiment for the total heat generation rate
of 31.64 kW, top: axial direction, bottom: cross section; case 3: volumetric heat generation source applied to the central region,
case 4: volumetric heat generation source applied to the entire domain
ot
tn
rin
ep
Pr

22

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
590

ed
570
550
530
510
Temperature (K)

iew
490
470
450
430
410 crust (case3)
390 Numerical result, case 3

ev
Numerical result, case 4
370 crust (case4)
Experimental result (q = 31.64 kW) [16]
350
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20
Molten pool height (m)

r
Figure 19: A comparison between temperature predicted by the simulation and measured in the experiment for the total heat
generation rate of 31.64 kW; case 3: volumetric heat generation source applied to the central region, case 4: volumetric heat
generation source applied to the entire domain

er
pe
5. Conclusion

Numerical results indicate that three main regions form in the molten salt pool; the upper region with uniform
temperature, the lower region with stratified temperature, and the boundary layers along the vessel wall. The crust
thickness is almost uniform along the axial direction in the main shell and is higher at the step region since the cooling
ot

area increases; however, it decreases in the sub shell since the temperature is higher in this region. In the cross-section,
the crust thickness decreases closer to the top surface. This is mainly because of the high temperature and the formation
tn

of the boundary layer along the vessel wall, which melts the formed crust in the upper region. The increase in the
volumetric heat generation rate decreases the crust thickness and increases the temperature in the molten pool; however,
the flow patterns are similar in the two cases. The comparison between the numerical and experimental results indicates
that the CFD model accurately predicted the temperature distribution within the molten pool and the crust formation
rin

at the boundaries.

The Canadian Nuclear Laboratories experiment explored natural convection heat transfer and crust formation in
ep

a 1/5 scale CANDU-6 reactor vessel using a corium simulant. There are some significant differences between the
Canadian Nuclear Laboratories experiment and CANDU corium natural convection. These include the Pr number,
non-eutectic solidification, and the uniform volumetric heat generation in the corium. Therefore, further simulations
and analyses are needed to assess the heat removal from the corium under a severe accident.
Pr

23

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
6. Acknowledgements

ed
The authors would like to thank the University Network of Excellence in Nuclear Engineering (UNENE) and
National Science and Engineering Research Council (NSERC) for funding this research. The authors also thank Justin

iew
Spencer and Canadian Nuclear Laboratories for sharing the details of the experiment and results.

References

ev
[1] Independent Electricity System Operator, Year-End Data Report, http://www.ieso.ca, 2018.

[2] R. McLean, In-Vessel Retention (IVR) Design Features Phenomenology and Technical Basis in the CANDU
Reactor, Proceeding of the Technical Meeting on Phenomenology and Technologies Relevant to In-Vessel Melt

r
Retention and Ex-Vessel Corium Cooling, Shanghai, China, 2016.

[3] S. K. Gupta, R. Jaitly, Y. H. Jin, D. H. Kim, S. Lee, S. Y. Park, A. Viktorov, A. White, Analysis of Severe

er
Accidents in Pressurized Heavy Water Reactors, International Atomic Energy Agency (IAEA), 2008.

[4] P. M. Mathew, T. Nitheanandan, S. Bushby, Severe Core Damage Accident Progression within a CANDU 6
pe
Calandria Vessel, Proceeding of the 3rd European Review Meeting on Severe Accident Research, Nessebar,
Bulgaria, 2008.

[5] Canadian Nuclear Safety Commission Staff, Severe Accident Progression without Operator Action,
http://nuclearsafety.gc.ca/, 2015.
ot

[6] F. Zhou, D. Novog, Mechanistic Modelling of Station Blackout Accidents for a Generic 900 MW CANDU Plant
Using the Modified RELAP/SCDAPSIM/MOD3.6 code, Nuclear Engineering and Design, Vol. 335, pp. 71-93,
tn

2018.

[7] D. L. Luxat, J. C. Luxat, Calandria Vessel Integrity under Severe Accident Loads, Proceeding of the 19th
International Conference on Structural Mechanics in Reactor Technology, Toronto, Canada, 2007.
rin

[8] A. Muzumdar, P. Mathew, J. Rogers, M. Lamari, Core Melt Retention Capability of CANDU Reactors, Proceed-
ings of the 11th Pacific Basin Nuclear Conference, Banff, Canada, 1998.
ep

[9] D. Home, M. Chai, Determination of In-Vessel Retention under Molten Corium Pool Attack, Proceeding of
the 16th International Topical Meeting on Nuclear Reactor Thermal Hydraulics, Chicago, USA, pp. 1196-1209,
2015.
Pr

[10] L. Zhang, Y. Zhou, Y. Zhang, W. Tian, S. Qiu, G. Su, Natural Convection Heat Transfer in Corium Pools: A
Review Work of Experimental Studies, Progress in Nuclear Energy, Vol. 79, pp. 167–181, 2015.

24

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
[11] J. Bonnet, Thermal Hydraulic Phenomena in Corium Pools: the BALI Experiment, Proceedings of the 7th

ed
International Conference on Nuclear Engineering, Ankara, Turkey, 1999.

[12] S. Nicolici, D. Dupleac, I. Prisecaru, Numerical Analysis of Debris Melting Phenomena During Late Phase
CANDU 6 Severe Accident, Nuclear Engineering and Design, Vol. 254, pp. 272–279, 2013.

iew
[13] R. David, Computational Fluid Dynamics Study of a Corium Pool in a CANDU Calandria, CNL Nuclear Review,
Vol. 6, No. 2, pp. 175-182, 2017.

[14] R. Nourgaliev, T. Dinh, B. Sehgal, Effect of Fluid Prandtl Number on Heat Transfer Characteristics in Internally
Heated Liquid Pools with Rayleigh Numbers up to 1012 , Nuclear Engineering and Design, Vol 169, No. 1-3, pp.

ev
165-184, 1997.

[15] J. K. Lee, S. D. Lee, K. Y. Suh, Boundary Dependent Natural Convection Heat Transfer in a Circular Slice Water
Pool, Proceedings of the 5th Symposium on Nuclear Thermal Hydraulics and Safety, Jeju Island, South Korea,

r
2006.

er
[16] J. Spencer, Canadian Nuclear Laboratories Corium Convection Experiment Report, 2021.

[17] A. Palagin, A. Miassoedov, X. Gaus-Liu, M. Buck, C. Tran, P. Kudinov, L. Carenini, C. Koellein, W. Luther,
V. Chudanov, Analysis and Interpretation of the LIVE-L6 Experiment, Proceeding of the 5th European Review
pe
Meeting on Severe Accident Research, pp. 21-23, 2012.

[18] R. Serrano-López, J. Fradera, S. Cuesta-López, Molten Salts Database for Energy Applications, Chemical
Engineering and Processing: Process Intensification, Vol. 73, pp. 87-102, 2013.
ot

[19] B. D. Iverson, S. T. Broome, A. M. Kruizenga, J. G. Cordaro, Thermal and Mechanical Properties of Nitrate
Thermal Storage Salts in the Solid-Phase, Solar Energy, Vol. 86, No. 10, pp. 2897-2911, 2012.
tn

[20] P. A. Davidson, Turbulence: an Introduction for Scientists and Engineers, Oxford University Press, 2015.

[21] M. Kumar, D. J. Krishna, Influence of Mushy Zone Constant on Thermohydraulics of a PCM, Energy Procedia,
Vol. 109, pp. 314-321, 2017.
rin

[22] F. Ampofo, T. Karayiannis, Experimental Benchmark Data for Turbulent Natural Convection in an Air Filled
Square Cavity, International Journal of Heat and Mass Transfer, Vol. 46, No. 19, pp. 3551–3572, 2003.

[23] S. D. Lee, J. K. Lee, K. Y. Suh, Natural Convection Thermo Fluid Dynamics in a Volumetrically Heated
ep

Rectangular Pool, Nuclear Engineering and Design, Vol. 237, No. 5, pp. 473–483, 2007.

[24] F. Wolff, R. Viskanta, Solidification of a Pure Metal at a Vertical Wall in the Presence of Liquid Superheat,
International Journal of Heat and Mass Transfer, Vol. 31, No. 8, pp. 1735–1744, 1988.
Pr

[25] A. T. Kitagawa, Numerical Analysis of Fluid Flow and Heat Transfer in Atria Geometries, Master’s Thesis,
McMaster University, 2012.

25

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803
[26] J. Yingchun, CFD Modelling of Natural Convection in Air Cavities, CFD letters, Vol. 6, No. 1, pp. 15–31, 2014.

ed
[27] F. P. Incropera, D. P. DeWitt, T. L. Bergman, A. S. Lavine, et al., Fundamentals of Heat and Mass Transfer, Wiley
New York, 1996.

iew
r ev
er
pe
ot
tn
rin
ep
Pr

26

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4268803

You might also like