Skip to main content

ORIGINAL RESEARCH article

Front. Energy Res., 20 June 2022
Sec. Nuclear Energy
Volume 10 - 2022 | https://doi.org/10.3389/fenrg.2022.922169

Numerical Study of Low Pr Flow in a Bare 19-Rod Bundle Based on an Advanced Turbulent Heat Transfer Model

www.frontiersin.orgXianwen Li1,2 www.frontiersin.orgXingkang Su1,2* www.frontiersin.orgLong Gu1,2,3* www.frontiersin.orgLu Zhang1 www.frontiersin.orgXin Sheng1
  • 1Institute of Modern Physics, Chinese Academy of Sciences, Lanzhou, China
  • 2School of Nuclear Science and Technology, University of Chinese Academy of Sciences, Beijing, China
  • 3School of Nuclear Science and Technology, Lanzhou University, Lanzhou, China

Compared to assuming a constant turbulent Prandtl number model, an advanced four-equation model has the potential to improve the numerical heat transfer calculation accuracy of low–Prandtl number (Pr) fluids. Generally, a four-equation model consists of a two-equation kε turbulence model and a two-equation kθεθ heat transfer model. It is essential to analyze the influence of dissimilar turbulence models on the overall calculation accuracy of the four-equation model. The present study aims to study the effect of using different turbulence models on the same kθεθ heat transfer model. First, based on the open-source computational fluid dynamics software OpenFOAM, an advanced two-equation kθεθ heat transfer model was introduced into the solver buoyant2eqnFoam, which was developed based on the self-solver buoyantSimpleFoam of OpenFOAM. In the solver buoyant2eqnFoam, various turbulence models built into OpenFOAM can be conveniently called to close the Reynolds stress and an advanced two-equation heat transfer model can be utilized to calculate the Reynolds heat flux of low-Pr fluids. Subsequently, the solver buoyant2eqnFoam was employed to study the fully developed flow heat transfer of low-Pr fluids in a bare 19-rod bundle. The numerical results were compared and analyzed with the experimental correlations and other simulation results to validate the effectiveness and feasibility of the solver buoyant2eqnFoam. Furthermore, the influence of combining different turbulence models with the same two-equation kθεθ heat transfer model was also presented in this study. The results show that the turbulence model has a considerable influence on the prediction of turbulent heat transfer in the high Peclet number range, suggesting that it should be prudent when picking a turbulence model in the simulations of low-Pr fluids.

1 Introduction

Nuclear energy is playing an increasingly irreplaceable role in the future energy structure as the demand for energy increases rapidly (Gu and Su, 2021). Lead-cooled fast reactor (LFR) is one of the six types of innovative nuclear power systems proposed by the Generation IV International Forum (Abram and Ion, 2008; Pacio et al., 2015). Benefiting from the excellent performance in chemical inertness, neutron economy, and thermohydraulic properties, lead–bismuth eutectic (LBE) is considered as one of most promising coolants for LFR. It is indispensable to research the thermohydraulic behaviors of the LBE inner fuel assembly, which influences the security and economic performance of LFRs but is poorly understood (Martelli et al., 2017; Pacio et al., 2017).

Since it is expensive, parlous, and complicated to conduct an experiment with LBE under a high-temperature state, computational fluid dynamics (CFD) methods are widely employed to study the thermohydraulic characteristics of LBE. The CFD methods can be subdivided into three categories: direct numerical simulation (DNS), large eddy simulation (LES), and Reynolds-averaged Navier–Stokes simulation (RANS). Despite the high calculation accuracy of DNS and LES, they have a high demand for computational resources, and as a result, they are only suitable for some specific and straightforward geometric models (Kawamura et al., 1999). Since the computational cost of the RANS approach is much lower than that of the DNS and LES, the RANS approach is the most widely adopted CFD method in engineering calculation. In the RANS method, the linear eddy-viscosity kε or kω turbulence model is ordinarily sufficient to accurately predict the momentum transport of various fluids (Nagano, 2002). On the other hand, for reproducing the heat transfer, the Reynolds-analogy hypothesis assuming a constant turbulent Prandtl number Prt=0.850.9 is adopted in almost all commercial codes (Manservisi and Menghini, 2014a). For the simulation of ordinary fluids like water and air, having a relatively high Prandtl number, the rational results can be obtained with a constant Prt (He et al., 2021). However, LBE is characterized by high thermal diffusivity and low viscosity values, resulting in low Prandtl numbers (Pr0.010.03). Consequently, the Reynolds-analogy hypothesis is no longer appropriate to be employed to study the thermohydraulic characteristics of LBE by the CFD methods (Cheng and Tak, 2006). For this reason, some advanced turbulent heat transfer models which can reproduce the heat transfer behaviors of LBE with high precision are highly desirable.

In the past four decades, to improve the calculation accuracy of heat transfer for low-Pr fluids, various heat flux models to close the energy conservation equation in the framework of RANS have been developed.

1.1 Differential Heat Flux Model

DHFM is a full second-moment differential model for the transport of Reynolds heat fluxes. Compared with the constant Prt model, DHFM fully considers the convection, diffusion, generation, and dissipation terms of Reynolds heat flux in the differential equations. Carteciano (1995), Carteciano et al. (1997), Carteciano et al. (2001), and Carteciano and Grötzbach (2003) developed a kind of DHFM named turbulence model for buoyant flows (TMBF). The simulations of two-dimensional forced convection and mixed convection with different fluids were carried out to evaluate the accuracy of TMBF. The numerical results obtained by TMBF demonstrate that stratified flows and buoyant effects were well reproduced compared with the constant Prt model, especially in mixed convection conditions. Based on a summary of the various DHFM models developed in recent years, Shin et al. (2008) proposed a new set of DHFM models with an elliptic blending model. The new model was utilized to study the fully developed square duct flow, rotating, and nonrotating channel flow. The numerical results show good agreement with the LES and DNS results.

1.2 Algebraic Heat Flux Model

AHFM is a simplified second-moment form of DHFM, which transports Reynolds heat flux by establishing algebraic equations. Hanjalić et al. (1996), Kenjereš et al. (2005), Otić et al. (2005), and Otić and Grotzbach. (2007) developed and analyzed an implicit algebraic transport equation for the Reynolds heat flux term to close the energy equation. Evaluations and calibrations of AHFM for low-Pr fluids were implemented by Shams et al. (2014), Shams (2018), Shams et al. (2018), and De Santis et al. (2018) and De Santis and Shams (2018). AHFM has been validated in their works by comparison with the DNS data for turbulent flows in forced, mixed, and natural convection of different fluids. The numerical results obtained by AHFM illustrate that temperature, heat flux field, buoyant effects in-plane, backward-facing step, corium pool, rod bundle, etc. are well predicted.

1.3 A two-equation kθεθ model for Reynolds heat flux

The kθεθ model is a first-order 2-equation model for the calculation of Reynolds heat flux, which can be developed in a way similar to that of a first-order 2-equation kε model for the turbulent transport of momentum formulated. Compared with DHFM and AHFM, the two-equation model has been widely applied in recent years because of its lower calculation cost. To precisely reproduce the heat transfer of low-Pr fluids, extensive contributions of model coefficient and function, wall boundary conditions, and near-wall thermal turbulence effect of a two-equation kθεθ model had been made by Nagano and Kim (1988); Nagano and Shimada (1996); Nagano et al. (1997); Nagano (2002), Sommer et al. (1992), and Youssef et al. (1992); and Youssef (2006), Abe et al. (1994), Hattori et al. (1993), Hwang and Lin (1999), Deng et al. (2001), and Karcz and Badur (2005). In recent years, based on previous study, Manservisi and Menghini (2014a); Manservisi and Menghini (2014b); Manservisi and Menghini (2015), Cerroni et al. (2015), Cervone et al. (2020), Chierici et al. (2019), and Da Via et al. (2016); Da Vià and Manservisi (2019); and Da Vià et al. (2020) proposed a two-equation kθεθ model suitable for the LBE turbulent heat transfer simulation, and improved new numerical near-wall boundary conditions for turbulence variables. The numerical results obtained by the literature (Manservisi and Menghini, 2014a; Manservisi and Menghini, 2014b; Manservisi and Menghini, 2015; Su et al., 2022) indicate that turbulent heat transfer statistics such as in-plane, tube, backward-facing step, triangular rod bundle, square lattice bare rod bundle, and hexagonal rod bundle of forced convection of LBE are well reproduced based on their kθεθ turbulent heat transfer model.

In recent years, the interest in reliable CFD methods used to investigate the turbulent heat transfer of low-Pr fluids in complicated industrial configurations has increased dramatically. Nevertheless, commercial codes are still lacking, except for an AHFM model available on software STAR-CCM+ (Simcenter, 2016). A two-equation kε model utilized to calculate Reynolds stress with a two-equation kθεθ model used to calculate Reynolds heat flux is usually called a four-equation kεkθεθ model, which is expected to improve the numerical CFD accuracy of turbulent heat transfer for LBE. However, different turbulence models will have a certain impact on the time-scale transport of a two-equation kθεθ model. It is necessary to evaluate the sensitivity of various turbulence models to a two-equation kθεθ model.

Thus, in the present study, an improved CFD solver buoyant2eqnFoam, which introduces a two-equation kθεθ model to calculate the Reynolds heat flux and can directly call different turbulence models to calculate the Reynolds stress, was first developed based on the solver buoyantSimpleFoam of open-source CFD program OpenFOAM. The fully developed turbulent heat transfer results of LBE inner flow and a bare 19-rod bundle geometry with different Peclet numbers were investigated and compared with experimental relations to verify the effectiveness of the solver buoyant2eqnFoam and the numerical algorithm. Finally, the heat transfer sensitivity of different turbulence models to the two-equation kθεθ model was presented.

2 Mathematical Model

2.1 Physical Model

Thermal-hydraulic phenomena in a 19-rod bundle geometry are an essential research topic. In the past decades, numerous experimental and simulation researches have been conducted to precisely obtain flow characteristics and heat transfer correlations of coolant (Pacio et al., 2014; Martelli et al., 2017). In the present study, a bare 19-rod bundle with a fully developed turbulent LBE flow is considered. Figure 1A displays the cross section of the bare 19-rod bundle. Since the cross-flow in the bare 19-rod bundle is negligible and its construction is symmetrical, one-twelfth of the whole bundle is selected to carry out simulation for the sake of economizing computational cost. The computational domain is sketched in Figure 1B, together with the definitions of sub-channels and boundary regions. Sub1, Sub2, and Sub3 are defined as inner sub-channels, while Sub4 and Sub5 are the edge sub-channel and corner sub-channel, respectively. The more detailed geometric parameters are summarized in Table 1, which are consistent with Pacio’s experiment (Pacio et al., 2015), except that there are no grid spacers in the current study. Dh,bun and Dh,sub1 are the hydraulic diameter of the bundle and hydraulic diameter of Sub1, respectively. The length of the whole computational domain is set to 15 Dh,bun to eliminate the effect of employing periodic inlet boundary conditions. The flow parameters of LBE are reported in Table 2, where Pebun=PrUbunDh,bunρ/μ is the Peclet number of the bundle.

FIGURE 1
www.frontiersin.org

FIGURE 1. Diagrammatic sketch of the bare 19-rod bundle. (A) Cross section and (B) computational domain.

TABLE 1
www.frontiersin.org

TABLE 1. Geometric parameters of the bare 19-rod bundle.

TABLE 2
www.frontiersin.org

TABLE 2. Flow parameters of LBE.

2.2 Conservation Equations

For forced convection, the incompressible RANS equations with constant physical properties and no gravity are considered

uixi=0(1)
uit+ujuixj=1ρPxi+xj(vuixj)xjuiuj¯(2)

where ν, ui, and P are the molecular viscosity, Reynolds-averaged velocity, and the so-called average pressure, respectively. To obtain the unknown Reynolds stress uiuj¯, the linear eddy-viscosity model can be adopted as follows:

uiuj¯=vt(uixj+ujxi)+2k3δij(3)

where k and νt, both derived from the turbulence model, represent the turbulent kinetic energy and the turbulent viscosity, respectively.

It should be noted that, in OpenFOAM, the energy conservation equation can be expressed in terms of enthalpy (Darwish and Moukalled, 2021):

ht+ujhxj+Kt+ujKxj=1ρPt+xj((α+αt)hxj)(4)

where h, K=|U|2/2, α, and αt are the enthalpy per unit of mass, the kinetic energy per unit of mass, molecular thermal diffusivity, and the turbulent thermal diffusivity, respectively. The unknown αt needs to be derived from a two-equation kθεθ turbulent heat transfer model. After solving Eq. 4, the distribution of h can be obtained. Subsequently, the Reynolds-averaged temperature T can be calculated by using the function Thermo.T () coming with OpenFOAM. The derivation of periodic momentum and energy equations in OpenFOAM can be found in the reference (Ge et al., 2017).

2.3 Turbulence Model for Momentum Field

Benefiting from replacing the friction velocity uτ with Kolmogorov velocity uε, the turbulence model proposed by Abe et al. (1994), which can well reproduce the low Reynolds number and near-wall effects of both separated and attached flows, was widely adopted in the calculation of LBE (Manservisi and Menghini, 2014a; Manservisi and Menghini, 2014b; Cerroni et al., 2015; Manservisi and Menghini, 2015; Da Via et al., 2016; Chierici et al., 2019; Da Vià and Manservisi, 2019; Cervone et al., 2020; Da Vià et al., 2020). However, the Abe kε turbulence model does not exist in the current turbulence model library of OpenFOAM. Therefore, in the current study, the Abe kε turbulence model is compiled into the turbulence model library that comes with OpenFOAM so as to utilize the wall functions of OpenFOAM in this self-compiled turbulence model. In the Abe kε turbulence model, the turbulent viscosity νt is computed as follows:

vt=Cμfμk2ε(5)

where Cμ is a constant. fμ is the model function, defined as follows:

fμ=(1exp(y14))2(1+5Rt3/4exp((Rt200)2))(6)
y=uεδν(7)

where Rt=k2/νε and uε=(νε)1/4. Moreover, δ is the distance from the wall. The equations for k and its dissipation ε can be written as follows:

kt+ujkxj=xj((v+vtσk)kxj)+Pkε(8)
Pk=uiuj¯uixj=vt(uixj+ujxi)uixj2kδij3uixj(9)
εt+ujεxj=xj((v+vtσε)εxj)+C1εεkPkC2εfεε2k(10)
fε=(1exp(y3.1))2(10.3exp(Rt26.5))(11)

The model constants utilized in the Abe kε turbulence model are reported in Table 3.

TABLE 3
www.frontiersin.org

TABLE 3. The model constants of the Abe kε turbulence model.

2.4 Two-Equation Model for Thermal Field

In the current work, the kθεθ turbulent heat transfer model developed and improved by Manservisi and Menghini (2014a), Manservisi and Menghini (2014b), and Manservisi and Menghini (2015), which introduces the average square temperature fluctuation kθ and its dissipation εθ in order to well reproduce the near-wall turbulent heat transfer behaviors of LBE having Pr in the range of 0.01–0.03, is adopted to calculate the turbulent thermal diffusivity αt. In the Manservisi kθεθ model, αt is computed as follows:

αt=Cθkτlθ(12)

where Cθ is the constant empirical coefficient and τlθ is the local thermal characteristic time, modeled as follows:

τlθ=f1θB1θ+f2θB2θ(13)

with the appropriate functions set as follows:

f1θ=(1exp(Rδ19Pr))(1exp(Rδ14))(14)
B1θ=0.9τu=0.9kε(15)
f2θB2θ=τu(f2aθ2RR+Cγ+f2bθ2RPr1.3PrRt3/4)(16)
f2aθ=f1θexp((Rt500)2)(17)
f2bθ=f1θexp((Rδ200)2)(18)

where Rδ=δε1/4/ν3/4, τu=k/ε, and R=τθ/τu with the thermal turbulent characteristic time τθ=kθ/εθ. In addition, τu=k/ε represents the dynamical turbulent characteristic time. The equations for kθ and its dissipation εθ can be written as follows:

kθt+ujkθxj=xj((α+αtσkθ)kθxj)+Pθεθ(19)
Pθ=αt(Txj)(Txj)(20)
εθt+ujεθxj=xj((α+αtσεθ)εθxj)+εθkθ(CP1PθCd1εθ)+εθk(CP2PkCd2ε)(21)
Cd2=(1.9(10.3exp((Rt6.5)2))1)(1exp(Rδ5.7))2(22)

The constant empirical coefficients used in the Manservisi kθεθ turbulent heat transfer are reported in Table 4.

TABLE 4
www.frontiersin.org

TABLE 4. The constant empirical coefficients of the Manservisi kθεθ turbulent heat transfer model.

3 Solver and Boundary Conditions

To calculate the thermal-hydraulic characteristics of LBE, a CFD solver named buoyant2eqnFoam was developed on the OpenFOAM platform having user-friendly programming language features based on the turbulence model and the aforementioned turbulent heat transfer model. The SIMPLE algorithm is adopted to handle pressure–velocity coupling equations and the coupled multigrid iterations technique is utilized for matrix solutions. All calculations were performed using double precision on OpenFOAM and the convergence conditions of residual error are set as follows:

Max|Qn+1Qn1|<106(23)

where Q stands for ui,T,P,kθ,εθ,k, and ε. The index i represents the number of iterations. The framework of the buoyant2eqnFoam solver is presented in Figure 2. The buoyant2eqnFoam solver mainly includes main-program buoyant2eqnFoam.C, velocity equation UEqn.H, energy equation EEqn.H, pressure–Poisson equation pEqn.H, call function of turbulence model turbulence- > correct(), and a two-equation kθεθ heat transfer model HEqn.H. The boundary condition data, mesh data, physical property data, calculation control, discrete format of each differential operator, algebraic equation solver, and relaxation factor required by buoyant2eqnFoam to perform calculation are included in the 0 folder, constant/polyMesh, constant/thermophysicalProperties, system/controlDict, system/fvSchemes, and system/fvSolutions.

FIGURE 2
www.frontiersin.org

FIGURE 2. Framework of the buoyant2eqnFoam solver in OpenFOAM.

In the computational domain, periodic boundary conditions are set on the region of inlet and outlet, considering the fully developed turbulent inner flow in the bundle. It is worth noting that the energy source term needs to be added to the energy Eq. 4 in order to apply periodic boundary conditions to temperature variables. The calculation method of energy source term refers to this literature (Ge et al., 2017). For kθ and εθ, the boundary condition zeroGradient is employed on the wall under the uniform heat flux condition, according to the research of Deng et al. (2001). The boundary conditions imposed on each boundary are summarized in Table 5. Since the wall functions kLowReWallFunction for k and epsilonWallFunction for ε are both suitable for the low–Reynolds number turbulence model and can well reproduce the near-wall turbulence behaviors when y+ is very low (Darwish and Moukalled, 2021), they are employed in this study. Given that there are no wall functions accessible for kθ and εθ in OpenFOAM, y+ must be less than or equal to 1 in order to accurately reproduce the thermal turbulent behaviors near the wall (Manservisi and Menghini, 2014a).

TABLE 5
www.frontiersin.org

TABLE 5. Boundary conditions imposed on each boundary.

4 Results and Discussions

4.1 Mesh Independence Analysis

In this section, the buoyant2eqnFoam, which utilizes the Abe kε turbulence model for turbulence fields and uses the Manservisi kθεθ model for thermal fields, is employed to investigate the thermohydraulic characteristics of LBE inner flow in the bare 19-rod bundle in a wide range of Pebun. As shown in Figure 3, the computational domain was discretized by GAMBIT unstructured meshes (tetrahedral and hexahedral mesh blending). The first layer grid was set with a height of 0.001 mm in order to satisfy the criterion of the low–Reynolds number turbulence model for y+1. A total of 15 layers of boundary grids with a height ratio of 1.3 were designed.

FIGURE 3
www.frontiersin.org

FIGURE 3. The mesh of computational domain.

Three sets of mesh with different mesh numbers of 2.05 million, 2.64 million, and 3.11 million were adopted to analyze the mesh sensitivity. The dimensionless coolant temperature Θ is defined as follows:

Θ=(TTb,bun)λDh,bunqw(24)

where Tb,bun is the bulk temperature of the computational domain. The dimensionless temperature profiles along line ab (shown in Figure 1B) of three sets of mesh are displayed in Figure 4 under Pebun=1500. It is evident that the difference in the dimensionless temperature profile between three sets of mesh is negligible. Consequently, the mesh with a mesh number of 2.05 million is selected, taking the calculation cost into consideration.

FIGURE 4
www.frontiersin.org

FIGURE 4. The dimensionless temperature profiles along line ab of the three sets of mesh.

4.2 Solver Verification

The fully developed turbulent heat transfer characteristics of LBE inner flow in the bare 19-rod bundle were studied by Chierici et al. (2019), using a four-equation model in logarithmic specific dissipation form kΩkθΩθ, which was developed based on the Abe kε turbulence model and the Manservisi kθεθ model. The numerical results of Chierici et al. (2019) can provide some reference for developing a CFD solver of LBE turbulent heat transfer. Therefore, the simulation results of Chierici et al. (2019) and some experimental data are picked for comparison to verify the validity of the solver buoyant2eqnFoam. The Nusselt number is selected for comparison since it is a critical parameter in engineering. Table 6 presents some Nusselt number experimental correlations of the triangular rod bundle channel cooled by liquid metal, obtained by Subbotin et al. (1965), Mikityuk (2009), Gräber and Rieger (1972), and Mareska and Dwyer (1964), respectively.

TABLE 6
www.frontiersin.org

TABLE 6. Correlations of the Nusselt number for triangular lattices.

Because these reported correlations were developed for triangular lattices, the Nusselt number of inner sub-channel Sub1 is picked for comparison. The Nusub1 is calculated as follows:

Nusub1=qwDh,sub1λ(Tw,sub1Tb,sub1)(25)

where Tw,sub1 and Tb,sub1 are the mean wall temperature and mean coolant temperature of Sub1, respectively. Correspondingly, Pesub1=PrUsub1Dh,sub1ρ/μ is the Peclet number of the inner sub-channel Sub1. Figure 5 displays the comparison of the Nusselt number with Chierici simulation and experimental correlations. From this figure, it can be clearly observed that the tendency of Nusub1 is consistent with the simulation results of Chierici and shows good agreements with experimental data in a specific Peclet number range, illustrating that the rational prediction of LBE turbulent heat transfer can be obtained by the self-compiled solver buoyant2eqnFoam which can use the Abe kε turbulence model with wall functions for turbulence fields and the Manservisi kθεθ model with zero-gradient boundary for thermal fields.

FIGURE 5
www.frontiersin.org

FIGURE 5. Comparison of the Nusselt number with Chierici simulation and experimental correlations.

4.3 Flow and Heat Transfer Analysis

4.3.1 Velocity Field

The profiles of velocity magnitude on the computational domain of the bare 19-rod bundle are reported in Figure 6, with Pebun=500 and Pebun=1500, respectively. It is obvious that the mean velocity in the inner sub-channels is higher than that in the edge sub-channel Sub4 and corner sub-channel Sub5 because the hydraulic diameter of inner sub-channels is relatively higher, resulting in much lower flow resistance. For the same reason, the average velocity of Sub5 is much lower compared with that of other sub-channels, as summarized in Table 7.

FIGURE 6
www.frontiersin.org

FIGURE 6. Profiles of velocity magnitude on a computational domain. (A) Pebun=500 and (B) Pebun=1500.

TABLE 7
www.frontiersin.org

TABLE 7. Mean velocity of each sub-channel.

4.3.2 Dimensionless Temperature and Hot Spot Factor Distributions

Figure 7 shows the distribution of dimensionless temperature from where it can be seen that the maximum temperature is located in the corner sub-channel Sub5, which is mainly due to the lower mean coolant velocity of the corner sub-channel Sub5. Owing to the larger hydraulic diameter of the edge sub-channel Sub4 and the adiabatic boundary condition applied on the outer casing wall, the coolant with the lowest temperature can be found in the area of the edge sub-channel Sub4 near the outer casing wall. Comparing the four cases reported in Figure 7, it can be found that the convective heat transfer of the coolant in each sub-channel is enhanced as the Reynolds number increases, leading to the decrease of maximum temperature and the increase of bulk coolant temperature.

FIGURE 7
www.frontiersin.org

FIGURE 7. Profiles of dimensionless temperature on computational domain. (A) Pebun=250; (B) Pebun=500; (C) Pebun=1000; and (D) Pebun=3000.

The dimensionless hot spot factor characterizing the inhomogeneity of wall temperature is defined as follows:

ϕ=θwmax,subθb,subθwb,subθb,sub(26)

where θwmax,sub, θb,sub, and θwb,sub are the maximum wall temperature of the sub-channel, the bulk temperature of the sub-channel, and the mean wall temperature of the sub-channel, respectively. As ϕ=1, it means that the maximum wall temperature of the sub-channel is equal to the average wall temperature. The calculated results of the hot spot factor of each sub-channel are summarized in Table 8, from where it can be deduced that the wall temperature distribution of the inner sub-channel Sub1 is the most homogeneous. On the other hand, due to the coexistence of the heated rod wall and adiabatic wall in the edge sub-channel Sub4 and the corner sub-channel Sub5, the phenomenon of nonhomogeneous wall temperature distribution in these channels is more dramatic.

TABLE 8
www.frontiersin.org

TABLE 8. Results of hot spot factor for sub-channels.

4.3.3 Dimensionless Thermal Diffusivity Distribution

To analyze the dependence of heat transfer on Pebun, defining the dimensionless thermal diffusivity α+ as follows:

α+=αtα(27)

α+ is the ratio between turbulent thermal diffusivity and molecular thermal diffusivity. Figure 8 reports the calculated dimensionless thermal diffusivity distribution on the computational domain for Pebun=500 and Pebun=1500. From this figure, it can be clearly seen that the α+ in the center of each sub-channel is higher than that near the wall, where the heat is mainly derived by the molecular heat conduction. Moreover, in the center of the edge sub-channel Sub4, the maximum dimensionless thermal diffusivity can be found, indicating that the thermal diffusion caused by turbulent flow reaches its peak in this region. It should be mentioned that when Pebun=500, the α+ in the whole computational domain is less than 1, suggesting that the molecular heat conduction affects the entire computational domain dominantly. In addition, with the increase of Pebun, a region where turbulent thermal diffusion is stronger than molecular heat conduction begins to appear.

FIGURE 8
www.frontiersin.org

FIGURE 8. Dimensionless thermal diffusivity distribution on a computational domain. (A) Pebun=500 and (B) Pebun=1500.

4.3.4 Turbulent Prandtl Number Distribution

The distribution of Prt in the computational domain is plotted against different Pebun numbers and displayed in Figure 9. As revealed in this figure, the Prt is higher in the district close to the wall of the heated rod. Moreover, the overall turbulent Prandtl number in the computational domain decreases with the increase of Pebun. In particular, the turbulent Prandtl number in the turbulent core region decreases significantly with the increase of Pebun.

FIGURE 9
www.frontiersin.org

FIGURE 9. Turbulent Prandtl number distribution on a computational domain. (A) Pebun=1000 and (B) Pebun=3000.

The mean turbulent Prandtl number Prtm of the computational domain is defined as follows:

Prtm=APrtdAAdA(28)

In order to investigate the influence of Pebun on the Prtm, Figure 10 plots the Prtm against different Pebun. From this figure, it can be concluded that the Prtm tends to decrease as the Pebun increases. However, the rate of decline also decreases with the increase of Pebun. Furthermore, the Prtm of the computational domain is higher than 1, suggesting that the analogy about Prt=0.85 is not appropriate in such low-Pr fluids.

FIGURE 10
www.frontiersin.org

FIGURE 10. Mean turbulent Prandtl number of the bare 19-rod bundle.

4.4 Assessment of Different Turbulence Models and Turbulent Heat Transfer Models

To analyze the effect of the turbulence model on the simulation of heat transfer, in this sub-section, various turbulence models of OpenFOAM are also employed in buoyant2eqnFoam, including standard kε (Launder and Spalding, 1972; 1983), SST kω (Menter and Esch, 2001; Menter et al., 2003; Hellsten, 2012), and LaunderSharma kε (Launder and Sharma, 1974). The specific definitions of these turbulence models can be found in the literature (Launder and Spalding, 1972; Launder and Sharma, 1974; Launder and Spalding, 1983; Menter and Esch, 2001; Menter et al., 2003; Hellsten, 2012). A comparison of Nusub1 calculated by different turbulence models with heat transfer experimental correlations is displayed in Figure 11. As demonstrated in this figure, the Nusub1 calculated by four turbulence models is pretty close when Pesub1<1000, mainly because the molecular heat conduction is dominant in this Peclet number range. It should be noted, however, that the deviations of Nusub1 obtained by each turbulence model gradually increase as the Peclet number grows. As indicated in Figure 11, all Nusub1 predicted by the standard kε turbulence model is located between Mareska and Mikityuk correlations. However, for Abe kε, LaunderSharma kε, and SST kω, the calculated Nusub1 lies between the Graber and Subbotin correlations when Pesub1>1000. Although the Nusub1 results obtained by Abe kε, LaunderSharma kε, and SST kω are very similar, in general, Abe kε is the most conservative. In addition, the maximum deviation of the Nusselt number obtained by the Abe kε and the standard kε is close to 19%. It is worth mentioning that due to the significant deviation between the various Nusselt number experimental correlations, the quality of these turbulence models cannot be evaluated. Therefore, great care and caution should be exercised when selecting a turbulence model in simulation. More precise experimental and analytical studies are required in the future to identify the thermohydraulic characteristics of heavy liquid metals like LBE.

FIGURE 11
www.frontiersin.org

FIGURE 11. Comparison of Nusub1 calculated by different turbulence models.

Moreover, the Nusub1 calculated by the Manservisi kθεθ model and the Prt=0.85 model is reported in Figure 12. It is evident that compared with the experimental correlations plotted in Figure 12, the Nusub1 obtained by the Prt=0.85 heat transfer model is higher under almost all Peclet numbers. Oppositely, the Manservisi kθεθ heat transfer model provides the more conservative results of Nusub1.

FIGURE 12
www.frontiersin.org

FIGURE 12. Comparison of Nusub1 calculated by different turbulent heat transfer models.

5 Conclusion

In the current study, the Abe kε turbulence model was compiled into the turbulence model library coming with OpenFOAM. A CFD solver buoyant2eqnFoam, which introduces the Manservisi kθεθ turbulent heat transfer model, was developed. Subsequently, the Abe kε turbulence model with wall functions and Manservisi kθεθ turbulent heat transfer model with zero-gradient boundary were employed to analyze the thermohydraulic characteristics of LBE inner flow in the bare 19-rod bundle. In addition, the influence of the turbulence model on the prediction of turbulent heat transfer was investigated by employing various turbulence models in the self-compiled solver buoyant2eqnFoam, including Abe kε, standard kε, SST kω, and LaunderSharma kε. Based on the aforementioned discussions, conclusions obtained from the present work can be summarized as follows:

1) The Nusselt numbers obtained by the self-compiled solver buoyant2eqnFoam are in good agreement with experimental correlations and Chierici simulation research, indicating the validity and reliability of the self-compiled solver.

2) In the bare 19-rod bundle with P/D=1.4, the flow resistance of the corner sub-channel is higher than that of other sub-channels due to the smaller hydraulic diameter, leading to the appearance of higher temperature distribution and larger hot spot factor in this region.

3) Although the turbulent Prandtl number of LBE inner flow in the bare 19-rod bundle will decrease as the Peclet number increases, the overall turbulent Prandtl number is higher than 0.85, revealing that the Reynolds-analogy hypothesis about Prt=0.85 is not appropriate for low-Pr number fluids like LBE.

4) The turbulence model has a considerable influence on the calculation of turbulent heat transfer of low–Pr number fluids in the high Peclet number range, suggesting that it should be prudent and rigorous when picking a turbulence model in the simulations. Moreover, compared with the kθεθ turbulent heat transfer model, the Reynolds-analogy hypothesis about Prt=0.85 may give the much higher Nusselt numbers in the simulation of low–Pr number fluids.

The applicability of the solver developed in the present study for the more complicated geometry like fuel assembly with grid spacer or wire-wrapped configurations requires further verification.

Data Availability Statement

The original contributions presented in the study are included in the article/supplementary material, further inquiries can be directed to the corresponding authors.

Author Contributions

XL: concept, research, writing, editing, code, and data processing; XS: modification, concept, research, and code; LG: fund, project management, concept, and research; LZ: editing and research; and XS: editing and research.

Funding

This study was supported by the Research on key technology and safety verification of primary circuit, Grant No. 2020YFB1902104; the Experimental study on thermal hydraulics of fuel rod bundle, Grant No. Y828020XZ0; and the National Natural Science Foundation of China, Grant No.12122512.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Abe, K., Kondoh, T., and Nagano, Y. (1994). A New Turbulence Model for Predicting Fluid Flow and Heat Transfer in Separating and Reattaching Flows-I. Flow Field Calculations. Int. J. Heat Mass Transf. 37 (1), 139–151. doi:10.1016/0017-9310(94)90168-6

CrossRef Full Text | Google Scholar

Abram, T., and Ion, S. (2008). Generation-IV Nuclear Power: A Review of the State of the Science. Energy Policy 36 (12), 4323–4330. doi:10.1016/j.enpol.2008.09.059

CrossRef Full Text | Google Scholar

Carteciano, L., Dorr, B., Grötzbach, G., Olbrich, W., and Jin, X. J. F. (2001). Two-and Three-Dimensional Thermal and Fluid-Dynamical Analysis of the Complete MEGAPIE-Module with the Computer Code. FLUTAN 6653, 443–458.

Google Scholar

Carteciano, L. J. H. (1995). Analysis of a Turbulence Model for Buoyant Flows Implemented in the 3d Thermal-Hydraulic Computer Code FLUTAN and Comparison with the Standard Kε-Σt Model. Turbulence 1.

Google Scholar

Carteciano, L. N., and Grötzbach, G. (2003). Validation of Turbulence Models in the Computer Code FLUTAN for a Free Hot Sodium Jet in Different Buoyancy Flow Regimes. ETDEWEB.

Google Scholar

Carteciano, L., Weinberg, D., and Müller, U. (1997). “Development and Analysis of a Turbulence Model for Buoyant Flows,” in Proc. of the 4th World Conf. on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics, Bruxelles, Belgium, 2–6. June: Citeseer).

Google Scholar

Cerroni, D., Da Via, R., Manservisi, S., Menghini, F., Pozzetti, G., and Scardovelli, R. (2015). Numerical Validation of a κ-ω-κθ-ωθ Heat Transfer Turbulence Model for Heavy Liquid Metals. J. Phys. Conf. Ser. 655 (1), 012046. IOP Publishing). doi:10.1088/1742-6596/655/1/012046

CrossRef Full Text | Google Scholar

Cervone, A., Chierici, A., Chirco, L., Da Vià, R., Giovacchini, V., and Manservisi, S. (2020). CFD Simulation of Turbulent Flows over Wire-Wrapped Nuclear Reactor Bundles Using Immersed Boundary Method. J. Phys. Conf. Ser. 1599 (1), 012022. IOP Publishing). doi:10.1088/1742-6596/1599/1/012022

CrossRef Full Text | Google Scholar

Cheng, X., and Tak, N.-I. (2006). Investigation on Turbulent Heat Transfer to Lead-Bismuth Eutectic Flows in Circular Tubes for Nuclear Applications. Nucl. Eng. Des. 236 (4), 385–393. doi:10.1016/j.nucengdes.2005.09.006

CrossRef Full Text | Google Scholar

Chierici, A., Chirco, L., Da Vià, R., and Manservisi, S. (2019). Numerical Simulation of a Turbulent Lead Bismuth Eutectic Flow inside a 19 Pin Nuclear Reactor Bundle with a Four Logarithmic Parameter Turbulence Model. J. Phys. Conf. Ser. 1224 (1), 012030. IOP Publishing). doi:10.1088/1742-6596/1224/1/012030

CrossRef Full Text | Google Scholar

Da Vià, R., Giovacchini, V., and Manservisi, S. (2020). A Logarithmic Turbulent Heat Transfer Model in Applications with Liquid Metals for Pr= 0.01–0.025. Appl. Sci. 10 (12), 4337. doi:10.3390/app10124337

CrossRef Full Text | Google Scholar

Da Vià, R., Manservisi, S., and Menghini, F. (2016). A K-Ω-Kθ-Ωθ Four Parameter Logarithmic Turbulence Model for Liquid Metals. Int. J. Heat Mass Transf. 101, 1030–1041. doi:10.1016/j.ijheatmasstransfer.2016.05.084

CrossRef Full Text | Google Scholar

Da Vià, R., and Manservisi, S. (2019). Numerical Simulation of Forced and Mixed Convection Turbulent Liquid Sodium Flow over a Vertical Backward Facing Step with a Four Parameter Turbulence Model. Int. J. Heat Mass Transf. 135, 591–603. doi:10.1016/j.ijheatmasstransfer.2019.01.129

CrossRef Full Text | Google Scholar

Darwish, M., and Moukalled, F. (2021). The Finite Volume Method in Computational Fluid Dynamics: An Advanced Introduction with OpenFOAM® and Matlab®. Springer.

Google Scholar

De Santis, A., and Shams, A. (2018). Application of an Algebraic Turbulent Heat Flux Model to a Backward Facing Step Flow at Low Prandtl Number. Ann. Nucl. Energy 117, 32–44. doi:10.1016/j.anucene.2018.03.016

CrossRef Full Text | Google Scholar

De Santis, D., De Santis, A., Shams, A., and Kwiatkowski, T. (2018). The Influence of Low Prandtl Numbers on the Turbulent Mixed Convection in an Horizontal Channel Flow: DNS and Assessment of RANS Turbulence Models. Int. J. Heat Mass Transf. 127, 345–358. doi:10.1016/j.ijheatmasstransfer.2018.07.150

CrossRef Full Text | Google Scholar

Deng, B., Wu, W., and Xi, S. (2001). A Near-Wall Two-Equation Heat Transfer Model for Wall Turbulent Flows. Int. J. Heat Mass Transf. 44 (4), 691–698. doi:10.1016/s0017-9310(00)00131-9

CrossRef Full Text | Google Scholar

Ge, Z., Liu, J., Zhao, P., Nie, X., and Ye, M. (2017). Investigation on the Applicability of Turbulent-Prandtl-Number Models in Bare Rod Bundles for Heavy Liquid Metals. Nucl. Eng. Des. 314, 198–206. doi:10.1016/j.nucengdes.2017.01.032

CrossRef Full Text | Google Scholar

Gräber, H., and Rieger, M. J. A. (1972). Experimentelle Untersuchung des Wärmeübergangs an Flüssigmetalle (NaK) in parallel durchströmten Rohrbündeln bei konstanter und exponentieller Wärmeflussdichteverteilung. Atomkernenergie (ATKE) Bd. 19 23-40.

Google Scholar

Gu, L., and Su, X. (2021). Latest Research Progress for LBE Coolant Reactor of China Initiative Accelerator Driven System Project. Front. Energy 15 (4), 810–831. doi:10.1007/s11708-021-0760-1

CrossRef Full Text | Google Scholar

Hanjalić, K., Kenjereš, S., and Durst, F. (1996). Natural Convection in Partitioned Two-Dimensional Enclosures at Higher Rayleigh Numbers. Int. J. Heat Mass Transf. 39 (7), 1407–1427.

Google Scholar

Hattori, H., Nagano, Y., and Tagawa, M. (1993). Analysis of Turbulent Heat Transfer under Various Thermal Conditions with Two-Equation Models. In Engineering Turbulence Modelling and Experiments, 43–52. Elsevier

CrossRef Full Text | Google Scholar

He, S., Wang, M., Zhang, J., Tian, W., Qiu, S., Su, G. H., et al. (2021). Numerical Simulation of Three-Dimensional Flow and Heat Transfer Characteristics of Liquid Lead-Bismuth. Nucl. Eng. Technol. 53 (6), 1834–1845. doi:10.1016/j.net.2020.12.025

CrossRef Full Text | Google Scholar

Hellsten, A. (2012). “Some Improvements in Menter's K-Omega SST Turbulence Model,” in 29th AIAA, Fluid Dynamics Conference, 2554.

Google Scholar

Hwang, C. B., and Lin, C. A. (1999). A Low Reynolds Number Two-Equation Model to Predict Thermal Fields. Int. J. Heat Mass Transf. 42 (17), 3217–3230. doi:10.1016/s0017-9310(98)00382-2

CrossRef Full Text | Google Scholar

Karcz, M., and Badur, J. (2005). An Alternative Two-Equation Turbulent Heat Diffusivity Closure. Int. J. Heat Mass Transf. 48 (10), 2013–2022. doi:10.1016/j.ijheatmasstransfer.2004.12.006

CrossRef Full Text | Google Scholar

Kawamura, H., Abe, H., Matsuo, Y., and Flow, F. (1999). DNS of Turbulent Heat Transfer in Channel Flow with Respect to Reynolds and Prandtl Number Effects. Int. J. Heat Fluid Flow 20 (3), 196–207. doi:10.1016/s0142-727x(99)00014-4

CrossRef Full Text | Google Scholar

Kenjereš, S., Gunarjo, S., and Hanjalić, K. (2005). Contribution to Elliptic Relaxation Modelling of Turbulent Natural and Mixed Convection. Int. J. Heat Fluid Flow 26 (4), 569–586.

Google Scholar

Launder, B. E., and Sharma, B. I. (1974). Application of the Energy-Dissipation Model of Turbulence to the Calculation of Flow Near a Spinning Disc. Lett. Heat Mass Transf. 1 (2), 131–137. doi:10.1016/0094-4548(74)90150-7

CrossRef Full Text | Google Scholar

Launder, B. E., and Spalding, D. B. (1972). Lectures in Mathematical Models of Turbulence. Academic Press.

Google Scholar

Launder, B. E., and Spalding, D. B. (1983). “The Numerical Computation of Turbulent Flows,” in Numerical Prediction of Flow, Heat Transfer, Turbulence and Combustion (Elsevier), 96–116. doi:10.1016/b978-0-08-030937-8.50016-7

CrossRef Full Text | Google Scholar

Manservisi, S., and Menghini, F. (2014a). A CFD Four Parameter Heat Transfer Turbulence Model for Engineering Applications in Heavy Liquid Metals. Int. J. Heat Mass Transf. 69, 312–326. doi:10.1016/j.ijheatmasstransfer.2013.10.017

CrossRef Full Text | Google Scholar

Manservisi, S., and Menghini, F. (2015). CFD Simulations in Heavy Liquid Metal Flows for Square Lattice Bare Rod Bundle Geometries with a Four Parameter Heat Transfer Turbulence Model. Nucl. Eng. Des. 295, 251–260. doi:10.1016/j.nucengdes.2015.10.006

CrossRef Full Text | Google Scholar

Manservisi, S., and Menghini, F. (2014b). Triangular Rod Bundle Simulations of a CFD κ-ϵ-κ-ϵ Heat Transfer Turbulence Model for Heavy Liquid Metals. Nucl. Eng. Des. 273, 251–270. doi:10.1016/j.nucengdes.2014.03.022

CrossRef Full Text | Google Scholar

Mareska, M. W., and Dwyer, O. E. (1964). Heat Transfer in a Mercury Flow along Bundles of Cylindrical Rods. J. Heat. Transf. 86 (2), 180–186.)

Google Scholar

Martelli, D., Marinari, R., Barone, G., Di Piazza, I., and Tarantino, M. (2017). CFD Thermo-Hydraulic Analysis of the CIRCE Fuel Bundle. Ann. Nucl. Energy 103, 294–305. doi:10.1016/j.anucene.2017.01.031

CrossRef Full Text | Google Scholar

Menter, F., and Esch, T. (2001). “Elements of Industrial Heat Transfer Predictions,” in 16th Brazilian Congress of Mechanical Engineering (COBEM), 650.

Google Scholar

Menter, F. R., Kuntz, M., and Langtry, R. J. T. (2003). Ten Years of Industrial Experience with the SST Turbulence Model. Heat Mass Transf. 4 (1), 625–632.

Google Scholar

Mikityuk, K. (2009). Heat Transfer to Liquid Metal: Review of Data and Correlations for Tube Bundles. Nucl. Eng. Des. 239 (4), 680–687. doi:10.1016/j.nucengdes.2008.12.014

CrossRef Full Text | Google Scholar

Nagano, Y., Hattori, H., and Abe, K. (1997). Modeling the Turbulent Heat and Momentum Transfer in Flows under Different Thermal Conditions. Fluid Dyn. Res. 20 (1-6), 127–142. doi:10.1016/s0169-5983(96)00049-4

CrossRef Full Text | Google Scholar

Nagano, Y., and Kim, C. (1988). A Two-Equation Model for Heat Transport in Wall Turbulent Shear Flows. J. Heat. Transf. 110 (3), 583–589. doi:10.1115/1.3250532

CrossRef Full Text | Google Scholar

Nagano, Y. (2002). Modelling Heat Transfer in Near-Wall Flows, 188–247.

Google Scholar

Nagano, Y., and Shimada, M. (1996). Development of a Two‐equation Heat Transfer Model Based on Direct Simulations of Turbulent Flows with Different Prandtl Numbers. Phys. Fluids 8 (12), 3379–3402. doi:10.1063/1.869124

CrossRef Full Text | Google Scholar

Otić, I., and Grötzbach, G. (2007). Turbulent Heat Flux and Temperature Variance Dissipation Rate in Natural Convection in Lead-Bismuth. Nucl. Sci. Eng. 155 (3), 489–496.

Google Scholar

Otić, I., Grötzbach, G., and Wörner, M. (2005). Analysis and Modelling of the Temperature Variance Equation in Turbulent Natural Convection for Low-Prandtl-Number Fluids. J. Fluid Mech. 525, 237–261.

Google Scholar

Pacio, J., Daubner, M., Fellmoser, F., Litfin, K., Marocco, L., Stieglitz, R., et al. (2014). Heavy-liquid Metal Heat Transfer Experiment in a 19-rod Bundle with Grid Spacers. Nucl. Eng. Des. 273, 33–46. doi:10.1016/j.nucengdes.2014.02.020

CrossRef Full Text | Google Scholar

Pacio, J., Litfin, K., Batta, A., Viellieber, M., Class, A., Doolaard, H., et al. (2015). Heat Transfer to Liquid Metals in a Hexagonal Rod Bundle with Grid Spacers: Experimental and Simulation Results. Nucl. Eng. Des. 290, 27–39. doi:10.1016/j.nucengdes.2014.11.001

CrossRef Full Text | Google Scholar

Pacio, J., Wetzel, T., Doolaard, H., Roelofs, F., and Van Tichelen, K. (2017). Thermal-hydraulic Study of the LBE-Cooled Fuel Assembly in the MYRRHA Reactor: Experiments and Simulations. Nucl. Eng. Des. 312, 327–337. doi:10.1016/j.nucengdes.2016.08.023

CrossRef Full Text | Google Scholar

Shams, A., Mikuž, B., and Roelofs, F. (2018). Numerical Prediction of Flow and Heat Transfer in a Loosely Spaced Bare Rod Bundle. Int. J. Heat Fluid Flow 73, 42–62. doi:10.1016/j.ijheatfluidflow.2018.07.006

CrossRef Full Text | Google Scholar

Shams, A., Roelofs, F., Baglietto, E., Lardeau, S., and Kenjeres, S. (2014). Assessment and Calibration of an Algebraic Turbulent Heat Flux Model for Low-Prandtl Fluids. Int. J. Heat Mass Transf. 79, 589–601. doi:10.1016/j.ijheatmasstransfer.2014.08.018

CrossRef Full Text | Google Scholar

Shams, A. (2018). Towards the Accurate Numerical Prediction of Thermal Hydraulic Phenomena in Corium Pools. Ann. Nucl. Energy 117, 234–246. doi:10.1016/j.anucene.2018.03.031

CrossRef Full Text | Google Scholar

Shin, J. K., An, J. S., Choi, Y. D., Kim, Y. C., and Kim, M. S. (2008). Elliptic Relaxation Second Moment Closure for the Turbulent Heat Fluxes. J. Turbul. 9 (9), N3. doi:10.1080/14685240701823101

CrossRef Full Text | Google Scholar

Simcenter, (2016). STAR-CCM+ Documentation.

Google Scholar

Sommer, T. P., So, R. M. C., and Lai, Y. G. (1992). A Near-Wall Two-Equation Model for Turbulent Heat Fluxes. Int. J. Heat Mass Transf. 35 (12), 3375–3387. doi:10.1016/0017-9310(92)90224-g

CrossRef Full Text | Google Scholar

Su, X., Li, X., Wang, X., Liu, Y., Chen, Q., Shi, Q., et al. (2022). Development and Assessment of an Isotropic Four-Equation Model for Heat Transfer of Low Prandtl Number Fluids. Front. Energy Res. 57. doi:10.3389/fenrg.2022.816560

CrossRef Full Text | Google Scholar

Subbotin, V., Ushakov, P., Kirillov, P., Ibragimov, M., Ivanovski, M., Nomofilov, E., et al. (1965). “Heat Transfer in Elements of Reactors with a Liquid Metal Coolant,” in Proceedings of the 3rd International Conference on Peaceful Use of Nuclear Energy, 192–200.

Google Scholar

Youssef, M. S. (2006). A Two-Equation Heat Transfer Model for Wall Turbulent Shear Flows. JES. J. Eng. Sci. 34 (6), 1877–1903. doi:10.21608/jesaun.2006.111184

CrossRef Full Text | Google Scholar

Youssef, M. S., Nagano, Y., and Tagawa, M. (1992). A Two-Equation Heat Transfer Model for Predicting Turbulent Thermal Fields under Arbitrary Wall Thermal Conditions. Int. J. Heat Mass Transf. 35 (11), 3095–3104. doi:10.1016/0017-9310(92)90329-q

CrossRef Full Text | Google Scholar

Keywords: low Pr fluid, liquid metal, turbulence model, turbulent heat transfer model, OpenFOAM

Citation: Li X, Su X, Gu L, Zhang L and Sheng X (2022) Numerical Study of Low Pr Flow in a Bare 19-Rod Bundle Based on an Advanced Turbulent Heat Transfer Model. Front. Energy Res. 10:922169. doi: 10.3389/fenrg.2022.922169

Received: 17 April 2022; Accepted: 06 May 2022;
Published: 20 June 2022.

Edited by:

Yixiang Liao, Helmholtz Association of German Research Centres (HZ), Germany

Reviewed by:

Jinbiao Xiong, Shanghai Jiao Tong University, China
Hui Cheng, Sun Yat-sen University, China

Copyright © 2022 Li, Su, Gu, Zhang and Sheng. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Long Gu, gulong@impcas.ac.cn; Xingkang Su, suxingkang@impcas.ac.cn

Download