This section explores the mathematical model of turbulent heat transfer and the numerical implementation of the simulated data related to the heat source generated by the proton beam.
3.1. Turbulent Heat Transfer in Liquid Metals
Liquid metals are ideal fluids for new energy applications due to their thermal properties that greatly enhance heat transfer. The thermophysical properties of liquid lead as a function of temperature are characterized by high thermal conductivity
k and low viscosity
. In this paper, we assume [
15]
for the density
, molecular viscosity
, heat conductivity
k and specific heat
, respectively. These physical properties have been obtained by fitting experimental data. Their trends, as a function of temperature
T, are shown in
Figure 3. The interested reader can consult the Handbook on liquid metals in [
15], where experiments and measurement uncertainties are discussed in detail. Other relevant data about lead are the boiling point, at 2021 K, and the melting point, at 600 K [
15].
The accuracy required for the thermal management of a circulating HLM target raises the problem of CFD models for low Prandtl numbers. In fact, when dealing with liquid metals, the thermal diffusivity dominates the momentum diffusivity, leading to a Prandtl number in the range
–
. This problem is common in many industrial and research applications, such as nuclear reactors with fast spectrum and MHD (Magneto-Hydro-Dynamics) energy conversion facilities, where the high thermal conductance and heat capacity are desirable to remove high quantities of heat. When trying to model turbulent heat flows in liquid metals, standard models implemented in commercial codes such as Fluent and CFX are generally not accurate enough to reproduce the heat transfer in such regimes. The turbulent Prandtl number
is defined as the ratio between the diffusivity
and the turbulent viscosity
, and it is commonly assumed to be constant: this is not the case for these fluids [
16]. Commercial codes are, in fact, calibrated on materials such as water and air where the hypothesis of similarity between the dynamics of thermal and viscous turbulence holds and the calculation of turbulent diffusivity is set by taking
and therefore
. A more sophisticated model is required to reproduce the heat exchange when dealing with liquid metals with low Prandtl numbers. The model development requires extensive experimental data for all the regimes of interest. It must be taken into account that these experiments are very expensive and require ad-hoc measuring instruments developed to manage these types of fluids. When it is possible to perform CFD studies utilizing DNS (Direct Numerical Simulations), the correct heat transfer coefficients can be calculated directly with the simulation, avoiding the costs of the experimental facility. However, even these are limited to small spatial domains and simple geometries since the computational cost of DNS is very high, even on modern parallel architectures. Only a handful of geometries have been successfully explored with DNS using low Prandtl numbers. Various models have been developed in recent years based on RANS (Reynolds Averaged Navier–Stokes) models for different fluids. In particular, recent interest in turbulence models for heat transfer has increased because some designs for fourth-generation nuclear reactors utilize liquid metals (sodium and lead, in particular) as the refrigerant. This has spurred the development of effective and tailored computational tools for these new reactors [
5,
17].
Many strategies have been explored to tackle the turbulent heat exchange in liquid metals. A simple model can be obtained by specifying the turbulent Prandtl number
as a function of the geometry and other dimensionless quantities, such as the distance to the wall or the viscosity [
18]. The problem with this approach is that this function depends on the case under consideration and the specific geometry. The advantage of this approach is the simple implementation, and it does not significantly affect the system of partial differential equations to be solved. This is crucial when using commercial codes that leave a reduced set of options for the user to manipulate the mathematical model. Experimental evidence measuring integral heat transfer can be used to determine such functions that must be explored for a comprehensive set of Reynolds and Péclet numbers. For this purpose, experiments, not yet fully available, are necessary with advanced measurement techniques to estimate the mechanism of turbulent heat exchange that provides the heat and velocity flows near the walls, together with the fluctuations of the thermal and velocity fields. Furthermore, additional experiments are needed to define the thermal wall functions for liquid metals valid in the nonlinear region above
.
Heat transfer in liquid metals with low Prandtl numbers is characterized by a high contribution of molecular conduction compared to conventional heat transfer with high Prandtl numbers for fluids such as water, where the inertial contribution and convective exchange have a dominant role. The temperature drop in liquid metals is limited within the laminar substrates. The models, based on global heat exchange and tailored to simple experimental geometry, show great accuracy when applied within the limits where they were obtained. On the other hand, they perform poorly when adapted to different regimes or geometries. Theoretical models are more complex to develop and often do not perform to the same level of accuracy as empirical models. However, they lead to a significant and deeper understanding of the heat transfer mechanism. Among the empirical models, we can mention the Aoki model, the Dwyer model, and the Reynolds model, which we report below [
5].
All these models give comparable results in cylindrical geometries with diameters of 10 mm at 300 °C with velocities in the range from to m/s, which gives values for the Pèclet number ranging from 140 to 2800. In these models, the value of is assumed constant.
Numerous DNS simulations in flat channel geometry have been performed to study the spatial behavior of the value of the turbulent Prandtl number. These calculations show that the near-wall value is close to unity regardless of the Prandtl and Reynolds number if the Prandtl number is larger than
[
18]. For small values of Prandtl, such as the range
–
characteristic of liquid metals, the value of
tends to 2 at the wall. The value increases moving away from the wall and reaches its maximum value at a distance of
. The turbulent Prandtl number ranges from 1 to 3, which also agrees with the above correlations.
When using the Ansys CFX code, starting from the results of these correlations in [
5], the following expression for the value of the turbulent Prandtl number is recommended
where
b is calculated by the following formula
With the values defined above, the Kirillov correlation [
19]
is reproduced quite well with the liquid metal LBE for low values, while for high values, it comes close to that of Stromquist [
20]
The correlation proposed in [
5] is therefore
where
b is defined by (
6). The (
5) can be used for Péclet values greater than 1000 and a constant value of the turbulent Prandtl number equal to 4 for Péclet values less than 1000.
3.2. CFD Codes for Thermo-Hydraulic Simulation
The dynamic and thermal analysis of the target device is carried out with two Ansys commercial codes: CFX and Fluent. The two codes are used in parallel to confirm the feasibility of the project and to compare the results obtained with different implementations of the turbulent Prandtl model.
The CFX and Fluent solvers were used and compared to check their suitability for performing the simulation of the BDF facility. The two codes adopt both the finite volume method and evaluate the quantities of interest (velocity, density, pressure, and temperature) in control volumes by using flux interpolations. The main difference between the two solvers lies in the definition of the single control volume: CFX implements the vertex-centered method (VCM), while Fluent uses the cell-centered method (CCM), as shown in
Figure 4 [
13,
14].
In the CFX code, the variables associated with the control volume are defined by the middle lines passing through each face of the grid cells. Conversely, in the cell-centered method, the variables are stored in the centroids of the computational cells, and the control volume is defined by the mesh triangulation. Both schemes have advantages and disadvantages in terms of accuracy and computational cost. The CCM has fewer flow unknowns but more degrees of freedom in total. This makes it more computationally expensive, using approximately double the memory and resources compared to VCM. On the other hand, it is less sensitive to mesh quality, in particular to the non-orthogonality and skewness of the discretization. Both solvers have different choices with different spatial discretization schemes, i.e., first-order upwind difference, second-order central difference, a high-resolution scheme in CFX, and first-order upwind, second-order upwind, and second-order central differencing in Fluent.
The CFD simulations performed with the CFX software v2021-R1 use the
SST turbulence model. The
and the
SST turbulence model are considered the best approximations among RANS models for the simulation of heavy liquid metals, see references [
5,
16]. The boundary condition at the inlet is an imposed mass flow rate for the velocity and fixed temperature. The pressure is fixed, and the temperature gradient is set to zero at the outlet. All other boundaries are modeled as adiabatic walls, so we impose a no-slip condition for the velocity and no heat flux for the temperature. The computational domain is an unstructured mesh of
million nodes with a
on all the walls. The time step for the transient is set to
s. The lead inlet temperature is 400 °C. The thermal transient analysis solves only the energy equation with imposed motion obtained from a previous isothermal simulation. In CFX, three classes of simulations are performed with different mass flow rates: 185 kg/s (A),
kg/s (B), and
kg/s (C), respectively. All cases have the same mesh with the same boundary conditions; the only variation is the mass flow rate at the inlet. Thermal power is implemented by an interpolation function using standard tools in CFX from an appropriate window selector. The heat source implementation in CFX using this approach has an error of
if compared to the average steady-state power integrated directly from the FLUKA results. The maximum temperature, reached at the end of the thermal deposition, is 660 °C (case A) and 1163 °C (case C), while the maximum wall temperatures are 460 °C (case A) and 520 °C (case C). The target flow rate value is
kg/s. With this flow rate, a CFD analysis is performed spanning two subsequent beams to analyze the temperature profile dependence on the frequency of the beams, i.e., to understand what influences the residual heat can impose on the next heat cycle. This analysis shows an increase in peak temperature between two successive beams of only 2 °C, so the thermal capacity of the lead can almost entirely remove the heat of a single beam before a new one delivers its heat payload.
From the point of view of the implemented models, Fluent presents more complex physical and numerical models. It can simulate almost all fields of physics, such as magnetohydrodynamics, acoustics, radiative heat exchange, combustion, and motion in porous materials. The implementation and solution of additional physical models in Fluent is also possible thanks to the implementation of UDFs.
The setup of the simulations in Fluent follows the CFX setup as close as possible to minimize numerical differences. The same turbulence SST model is selected. The boundary conditions have a fixed mass flow rate at the inlet, pressure at the outlet and zero velocity on walls. The temperature is fixed at the inlet. Adiabatic conditions are used on all other boundaries. The mesh is a polyhedral unstructured mesh of million nodes with a . The time step for the transient is s. The mesh size is selected based on the intensity of the time-dependent heat source. The mesh has an inflation layer (a layer of cells very close to the wall, whose dimensions grow further away from it) of mm.
The study of the thermal transient is carried out with different subsequent simulations aimed at analyzing the problem at increasing levels of complexity and computational cost. The classes of simulations carried out can be grouped into three cases labeled by (1) En_T, where only the energy equation with imposed motion is solved; (2) En_Tm, where only the energy equation with variable physical properties is solved; (3) En_full, where the complete simulation with variable properties, equation of motion, and energy are solved.
3.3. Thermal Source Modeling
From a thermo-hydraulic point of view, we study the turbulent heat transport generated by a volumetric thermal source inside an axial symmetric computational domain. The thermal source is generated by the interaction of the proton beam with the target, which is liquid lead entering at 400 °C. The chosen inlet temperature is typical for circulating liquid lead facilities [
3,
4,
17,
19], as it provides a robust margin against accidental localized solidification (occurring at 327 °C), and prevents significant corrosion on commercial-grade stainless-steel pipe walls [
3]. The computational domain consists of three ducts, a middle one that is 2 m long and
m in diameter, and two smaller 76 mm ducts for the inlet and the outlet, transverse to the central duct. The dimensions of the main duct reflect the interaction volume calculated with Monte Carlo simulations, and the transverse ducts reproduce the internal diameter of commercial 3 in pipes. The proton beam distribution provided on the coordinates
is axisymmetric and peaks along the
z-axis. The size of the thermal source is determined by the beam source using the Monte Carlo code Fluka [
11,
12]. Fluka is a general-purpose tool for particle transport and interactions with matter, covering an extended range of applications from proton and electron accelerator shielding to target design in different fields, such as calorimetry, dosimetry, and radiotherapy. The Fluka code, based on microscopic models, is consistent with the reaction and conservation laws. As a result, final predictions are obtained with a minimal set of free parameters fixed for all energy and target combinations. Fluka code can determine the interaction and propagation of many particles, such as photons, electrons, and neutrons.
The thermal source, which fills the entire volume of the middle conduit, is represented in
Figure 5 as a function of the three spatial coordinates. The distribution is Gaussian on the
x and
y axis and exponential on the
z direction.
Table 1 summarizes the variation in the source data within the domain. The values reported in
Table 1 indicate the minimum (Min), maximum (Max), and average (Ave) power density of the source; the quantities
,
indicate the variation in spatial coordinates in cylindrical coordinates, where
is the increment over the
z coordinate.
The implementation of the thermal source is performed differently for each software. CFX allows the implementation of the source via a more user-friendly graphical interface. The implementation of the thermal source in CFX is, in fact, achieved by defining a volume within the domain in which the thermal source will be interpolated, using a function containing the heat map data.
The heat source for the
k-th centroid is given by
where
is the distance between the
k-th centroid and the
i-th point of the heat source,
the value of the heat source at
k-th centroid,
n the number of heat source points and
N is the number of mesh centroids.
and
are the
j-th components of the
k-th centroid and of the
i-th point of the thermal source, respectively. The value of the heat source
, to insert into the energy balance equation, is expressed in W/m
3. In our case, given the viscosity of lead, the heat-generating contribution due to viscous effects can be neglected. In both solvers, the source is interpolated for each cell centroid in the computational domain via (
10). This value is then averaged on the inverse distance between the three thermal sources closest to the centroid in the exam.
The thermal source generated by the proton beam is time-dependent and has characteristics summarized in
Table 2. To verify the effects of the interpolation of the heat map on the mesh, a stationary simulation of the En_T type was conducted. In this case, only the energy equation was solved with constant physical properties for the lead and with an imposed velocity field. This simulation results in an average deposited power of
W, against the
W provided by CERN, with an error of
, and a temperature at the outlet of 711 K. The heat map implementation is calibrated with this error of
, the same one present in CFX simulations.
The thermal source implementation in Fluent requires the use of User-Defined Functions (UDFs), written in C or C++. These allow the manipulation and implementation of equations and solution models directly in the Fluent solver loop. The source intensity, shown in
Figure 5, significantly increases in a narrow cylinder close to the axis in the region near the beam entrance (
m). This small volume can be reasonably assumed to be a coaxial cylinder inside the central duct with a one-meter length and two-centimeter diameter. Inside this volume, the thermal source has values exceeding
W/m
3. This narrow distribution can affect the source interpolation based on (
10).
To see this, we consider a C++ code, which creates a 2D uniform square mesh in the
x,
y axes.
Figure 6 shows the centroid points (in red), which are carried out from the C++ code, and the source points (in blue) on a
z constant plane. Two cases with different mesh sizes are considered to understand the importance of the discretization of the source term: a mesh with a pitch of 2 mm (on the left) and a finer one with a pitch of 0.5 mm (on the right). We observe that in
Figure 6 on the left, there are no mesh points inside the inner circle where the interpolation may smooth the energy intensity contained in the central points.
In the C++ code, an algorithm has been implemented to improve the source discretization, obtaining a redistribution of the source near the center of the cylindrical duct. The algorithm redefines the coordinates of the source points () with a distance from the duct axis under 4 mm.
The coordinates
of these new points were obtained through a transformation into polar coordinates. In doing so, three new distinct circles were defined, starting from the point of origin, to spatially redistribute the points uniformly.
Figure 7 shows the source distribution before and after applying the algorithm just discussed. We decided to keep the peak of the heat source within the first 2.5 mm (see
Figure 5) and interpolate the thermal source on the centroid based on the coordinates of the closest point of the original source instead of averaging it with the inverse of the distance according to the appropriate function.
The thermal source obtained with our interpolation strategy is displayed in
Figure 8, where we can see that the thermal source is well interpolated in both meshes. With the
mm mesh, the heat source points are spaced along the entire
x interval, while with the 2 mm mesh (on the left), these points may lead to a poor approximation near the axis and affect the overall thermal source input in the CFD code.
A polyhedral mesh was used for the CFD simulations in Fluent, with a cell size of 8 mm. In particular, after trying different configurations, it was decided to implement the heat map on the mesh through the two body of influence (BOI) mesh technique. These are regions within the volume of interest that can be defined by the user to modify the dimensions of the mesh in the area on which they are defined. They are, therefore, used for modeling effects such as obstacles, for example, or heat sources in our specific case. We chose to size them as two cylinders coaxial to the central duct of one meter long: the first with a 20 mm diameter and a 2 mm refinement, and the other with a 10 mm diameter and a 1 mm refinement.
UDFs are scripts in C++ or C that, taking advantage of Fluent’s native MACROS, allow you to solve scalar equations (User Define Scalar, UDS), define properties of materials and sources, execute commands during the solver wheels or implement new boundary conditions. For the simulations shown in this paper, two C scripts were developed. The first, property.c, is used to add the physical properties of lead as a function of the temperature, while the second, source.c, implements the heat source.
The
source.c script approximates the heatmap provided by CERN in the fluid domain. The script contains two macros:
DEFINE ON_DEMAND and
DEFINE SOURCE. The
DEFINE SOURCE is shown in
Figure 9. The macros can handle various types of data and the native C data structures, which are useful for mesh manipulation. The first calculates the value of the volumetric heat source for each centroid of the mesh and stores this value in a User Defined Memory (UDM) structure. UDMs associate variables of interest to a group of mesh centroids, where the thermal source interpolation is based. The second macro adds the source term to the equation at runtime. The
source.c code is parallelized using native Fluent structures and commands to allow code execution in parallel.
Figure 10 shows the resulting thermal source implemented in Fluent using the UDF and the original Fluka heat source over the corresponding mesh points.
Figure 10 shows that the interpolation can reproduce the central value. This is important to obtain a solution that does not change with internal refinement.