heat transfer in nuclear reactor

download heat transfer in nuclear reactor

of 18

Transcript of heat transfer in nuclear reactor

  • 8/10/2019 heat transfer in nuclear reactor

    1/18

  • 8/10/2019 heat transfer in nuclear reactor

    2/18

    2

    hand, significant efforts have been directed by investigators (Elligot et al., 2006; Jackson et al.,

    2012) in developing a criterion for the onset of HTD, which has been helpful in identifying the

    conditions favorable for HTD to occur. Although an overwhelming number of articles areavailable in this area, currently there still exists a lack of clear understanding on:

    The pressure dependency of HTE and HTD, and

    An HTD onset criterion that can be successfully applied to rod bundles.

    Each of these phenomena (HTD, HTE) has a distinct character and exhibit variability between

    the cases. This, in turn, presents a challenge for existing correlations (that were generally

    developed using a certain set of experiments) to successfully predict the flow and heat transfercharacteristics for SC flows. In recent years, CFD has been increasingly used for simulation of

    supercritical flows and has been assessed against data for smooth pipes (Cheng and Schulenberg,

    2001; Saha et al., 2013). Similar to subcritical flows, the case-by-case suitability of the

    turbulence model to be used for SC flows still remains a topic of active research (Cheng andSchulenberg, 2001). Due to the availability of the experimental data for smooth pipe flows,

    several investigators (He et al., 2008; Zhu, 2010; Gang et al., 2011; Zhang et al., 2012; Liu et al.,

    2013; Angelucci et al., 2013) have tested the suitability of the existing turbulence models to

    predict the experimentally reported HTD phenomenon. It was found that the SST k-turbulencemodel (Menter, 1994), along with wally

    +set to lower than one, was able to capture the reported

    HTD trends to a certain extent. Here it is worthwhile to point out that, although the SST k-turbulence model could capture the qualitative trends of the HTD reasonably well, the ability topredict the exact quantitative degree of temperature rise varied amongst investigations. Apart

    from using the original formulation of SST k-, attempts have been made at University of

    Stuttgart (Zhu, 2010) to modify this model to be better suited for modeling SC flows. Recently,

    with an aim to find suitable turbulence models and develop guidelines for using CFD forsimulation of SC flows, international consortiums such as THINS project in EU (Schulenberg

    and Visser, 2013) have tested the suitability of the STAR-CCM+CFD code. The outcome of

    their attempts was documented by various investigators (Ambrosini, 2009; De Rosa et al., 2011;Angelucci et al., 2013) which provided some guidance on the applicability of the existingturbulence models and potential areas of development in order to improve the CFD capability for

    SC flows. In spite of these efforts, a fit-to-purpose turbulence model for SC flows could not

    be formulated and remains an area for improvement.

    Understanding the possibility of occurrence of HTD in the proposed SCWR fuel bundles is ofkey interest to the conceptual fuel design. At AECL, our primary focus is to understand the fluid

    flow and heat transfer characteristics for the proposedbarefuel bundle design. The overall work

    is divided into phases. Phase 1 of the investigation focuses on the development of a bundle heattransfer correlation suitable for the proposed bundle design. This is an ongoing task which is

    performed by mainly using the subchannel code ASSERT-PV (Rao et al, 2013). Since HTD is

    a boundary layer phenomenon, which cannot be dealt with by the current ASSERT-PV, theSTAR-CCM+CFD code is used for simulating the fuel bundle to determine the conditions

    favorable for occurrence of HTD. This forms the subject of the current investigation.

    In order to accomplish the objective, a broad set of flow conditions over three operational

    pressures (23.5, 25 and 28 MPa) are simulated using a fine computational mesh with wally+< 1.

    Two mass flow rates, five inlet fluid temperatures, along with bundle powers proportional to the

    inlet mass flow rate based on the nominal operating power/flow conditions, are considered in the

  • 8/10/2019 heat transfer in nuclear reactor

    3/18

    3

    current simulations. Taking advantage of the un-staggered bundle design, only 1/32 section and

    0.4m long bundle is simulated in the current work to limit computational time and obtain

    relatively quick solutions. A step-wise methodology along with ASME CFD best practiceguidelines was adopted to execute this investigation. The work presented in this manuscript

    presents the first step in the application of CFD approach to the newly developed fuel bundle

    design by AECL. In this investigation, sensitivity analysis on the effect of the existingturbulence models was not performed. Instead, the aim was to use the recommendations madeby the previous investigators as a starting point to judge the capability of the suggested model to

    predict HTD for rod bundles. Unlike majority of the previous investigations that simulated

    smooth pipe flows, rod bundle geometry was considered in the present study, a subject that is ofimportance to the nuclear industry for which publications are scarce. As discussed earlier, the

    GEN-IV project at AECL is currently in its conceptual phase (Phase 1) for which the

    experimental data does not exists. Hence, the assessment of CFD predictions against

    measurements is currently not possible. In this paper, the mesh generation and solutionmethodology adopted for the current simulation is discussed in section 2, followed by sensitivity

    analysis and CFD predictions in sections 3 and 4 respectively.

    2.

    Methodology

    STAR-CCM+ v 7.04(referred to as STAR hereinafter) was used in this study instead of the

    ANSYS CFDcode due to its relative ease in meshing complex geometries, and the flexibility of

    power session license to execute on multiple nodes. In addition, participants involved ininternational consortiums such as CASL (CASL, 2013) and THINS (Angelucci et al., 2013;

    THINS, 2013) have tested the suitability of this code for modeling rod bundle geometries with

    complex spacers in both subcritical and SC flows within their respective project frame work.

    2.1 Development of computational domain

    The fuel bundle geometry comprised of 64 elements in two un-staggered rings and one large

    centre flow tube. The surface for the sub-assembly geometries was created using ANSYSDesign Modeler, with the model dimensions listed inTable 1. As discussed earlier in

    section 1, to reduce the computational effort, the geometry used in the current investigation wasa 1/32 section with an overall length of 0.4 m (Figure 1).

    Table 1 Geometric model dimensions used for developing CAD

    Pressure tube inner diameter 0.144 m

    Centre flow tube diameter 0.094 m

    # of elements in inner and outer ring 32 each

    Inner ring element diameter, Pitch circle diameter 0.0095 m, 0.108 mOuter ring element diameter, Pitch circle diameter 0.001m, 0.1315 m

    Length of the simulation domain 0.40 m

    The bare rod bundle geometry was created by extruding the two dimensional circular cross

    section in the flow direction. This solid geometry of fuel elements was exported into a parasolid

    format. The solid fuel elements model developed in CAD was then imported into STAR to

  • 8/10/2019 heat transfer in nuclear reactor

    4/18

    4

    extract internal fluid volume and mesh the fluid domain. The fluid model of the bare rod

    geometry was subsequently used for meshing as shown inFigure 2.

    The meshing domains (fluid part) for the geometry were developed by extracting the internal

    fluid volume from the CAD representation of the fuel elements. The resulting surfaces from thisoperation were automatically triangulated and used as the base for discretizing the computational

    domain. The initial surface mesh was improved using a surface remesher during volumemeshing.

    Considerable computational efforts were required to resolve the near wall region and reduce thecomputational uncertainty in capturing HTD. The wall thickness option, which enables a STAR

    user to specify the first-node-point distance away from the wall, was selected as the stretching

    mode. Boundary layers were used on the walls (fuel elements, pressure tube and centre tube),with the first node point set to respect the constraint of wally

    +< 1 (typically in the range of 0.6

    to 0.9). This resulted in the first node point to be typically set at ~2.7 to 3.2 m away from the

    wall. The mesh growth factor was maintained at 1.3 as recommended by STAR (CD-Adapco,

    2013).

    A minimum of ten boundary layers were used to capture the large variation of fluid properties in

    the near-wall region that may contribute to the occurrence of heat transfer deterioration

    phenomenon. Three volume meshing models, prism layer mesher, surface remesher and trimmer

    option, were used to generate hexahedral cells with boundary layers on the walls of the fuelelements (seeFigure 2). Using this meshing strategy, three computational meshes were

    developed for the current investigation, which typically comprised of a total cell count of 3.5 to

    11.3 million cells. For the current investigation, the computational mesh to be used was basedon a detailed mesh independence study (described in section 2.4) based on ASME CFD best

    practise guidelines (ASME, 2009).

    Figure 1 Computational domain for bare bundle(colors are used only to distinguish model

    elements)

    Figure 2 Mesh on a section for bare bundleshowing boundary layers on the

    elements

    2.2 Implementation of supercritical fluid properties

    Unlike subcritical flows, the thermophysical properties of supercritical water may vary drastically with

    temperature at a given pressure (Figure 3). For all properties, specific heat in particular, it can be seeninFigure 3 that an increase in pressure results in reduction in magnitude of property variation.

  • 8/10/2019 heat transfer in nuclear reactor

    5/18

    5

    STAR has built-in steam tables (IAPWS-IF97) to define the thermophysical properties of water which

    are only valid for sub-critical conditions, i.e., conditions below the critical point (22.1 MPa and 374 C).Consequently, STAR cannot be directly used to simulate flows subjected to supercritical conditions.

    Hencefor the three system pressures (23.5, 25 and 28 MPa) tested, the SCW properties were obtained fromthe NIST online database (Lemmon et al., 2013) and implemented in STAR. Similar to the previousinvestigations (Palko and Anglart, 2008; Ambrosini, 2009; Angelucci et al., 2013), properties of SCW

    were assumed to be only temperature dependent (under a specific system pressure). The dependencyon the pressure (under a specific system pressure) was usually small and hence neglected. Currently,

    most CFD solvers do not include the option to consider thermophysical properties as a function of bothpressure and temperature. In order for these properties to vary simultaneously with both pressure and

    temperature, a different approach such as a look-up table may be considered in future analyses.

    The SCW properties obtained from the online NIST database were fitted using higher orderpolynomials (cubic splines) as a function of temperature. In order to ensure accuracy, the polynomials

    were specified in a piece-wise fashion. They were divided into many temperature interval ranges,

    including small ranges for the pseudo-critical region which required significant effort. For density and

    specific heat, the resulting polynomial equations were introduced in STAR through a GUI hook-up.The variations of the thermal conductivity and viscosity with the temperature were introduced through

    tables as STAR does not currently provide a GUI hook-up capability for incorporating them. Theproperties implemented in the current work are valid only for the three system pressures tested. The

    procedure stated above should be repeated if the operational pressure changes.

    (a)

    (b)

    (c)

    Figure 3 Variation of thermophysical properties for SC water at: (a) 23.5MPa; (b) 25 MPa; (c) 28 MPa

    2.3 Governing equations and turbulence models

    In the present CFD analysis, the finite volume approach was used to discretize and numerically solve

    the Reynolds Averaged Navier-Stokes (RANS) and turbulence model equations simultaneously with

    the continuity and temperature form of the energy balance equations in STAR. The turbulence model

    equations are presented and discussed below regarding their impact on prediction of HTD. However,for the RANS, continuity and energy equations, the reader is referred to the STAR user manual

    (CD-Adapco, 2013) since these equations are standard and their use does not depend on the choice of

    models by a CFD user.

    Although a general consensus on the choice of turbulence models for SC flows is currently notavailable, previous studies (Palko and Anglart, 2008; Zhu, 2010; Liu et al., 2013; Schulenberg and

    Visser, 2013) reported that amongst the existing turbulence models, the SST (shear stress transport)

    model can qualitatively predict HTD. Hence, Menters SST k-model with low Reynolds number

  • 8/10/2019 heat transfer in nuclear reactor

    6/18

    6

    modification as implemented in STAR along with lowy+

    approach for the near wall treatment was used

    in the present study to better capture the occurrence of HTD in the proposed fuel bundle.

    Menter (1994) formulated a hybrid two equation model, the SST k- model, which was designed to

    yield the best behaviour of the k- model by Wilcox (for Wilcoxs model formulations refer

    Wilcox, 2010) and the standard k- model. This model utilizes the original k-model in the near wall

    region (sub and log layer) and gradually switches to the standard k-model in the fully turbulent region

    far away from the wall. This switch was facilitated through a blending function. The transportequations for the SST k-turbulence model are:

    kx

    u

    x

    u

    x

    u

    x

    k

    xku

    xt

    k

    i

    j

    j

    i

    j

    it

    j

    tk

    j

    j

    j

    *

    (1)

    )(

    21

    )(

    2

    )()(

    112

    IV

    jj

    III

    II

    i

    j

    j

    i

    j

    it

    I

    j

    t

    j

    j

    j

    x

    w

    x

    kF

    x

    u

    x

    u

    x

    u

    kxxu

    xt

    (2)

    The first three terms (I to III) on the right hand side of the Eq. 2 represent the conservative diffusion,eddy viscosity production and dissipation of turbulence respectively. The last term (IV) represents the

    additional non conservative cross diffusion term which is not included in Wilcoxs k-model. All the

    coefficients in the model were calculated using a blending function as in Eq. 3. The blending function

    was set to a value of zero close to walls (leading to a standard equation) and a value of one away

    from the walls (which corresponds to a standard equation).

    2111 1 FF (3)

    where, 1 and 2 are the two sets of values for each variable listed inTable 2. The coefficient 1k hasbeen modified by STAR to 0.85 (CD-Adapco, 2013) from 0.5 as originally proposed by Menter. The

    blending function was computed as:

    411 argtanhF (4)

    with:

    )(

    2

    2

    )(

    2

    )(

    1

    4;

    500;

    09.0maxminarg

    III

    k

    III

    yDC

    k

    yy

    k

    (5)

    and, to prevent the build-up of turbulence in the stagnation regions, a turbulence production limiter was

    introduced in the SST model as:

    202 10;

    12max

    jj

    kxx

    kDC

    (6)

    In Eq. 5., the first term represents the turbulent length scale divided by the shortest distance to the nextsurface, the second term becomes important in the viscous sub layer and ensures that F1does not tend

  • 8/10/2019 heat transfer in nuclear reactor

    7/18

    7

    to zero in that region while the last one prevents a possible freestream-dependent solution. The

    turbulent viscosity was computed as:

    SSFkt

    3

    6.0,

    31.0,max

    1min

    2

    *

    (7)

    and S is the modulus of mean strain rate tensor. The function F2was evaluated as:

    222 argtanhF (8)

    22

    500;

    09.0

    2maxarg

    yy

    k

    (9)

    Table 2 Coefficients present in the SST k-turbulence model

    Coefficients

    for Set 1Set 1

    1 Coefficients

    for Set 2Set 2

    2

    1 0.075 2 0.0828

    *

    1 0.09 *

    2 0.09

    1k 0.85 2k 1

    1 0.5 2 0.856

    *

    1 1 *

    2 1

    1 0.41 2 0.41

    1

    *

    1

    2

    1*

    1

    1

    k 2

    *

    2

    2

    2*

    2

    2

    k

    Note that in the SST model by Menter (1994),*

    1 = *

    2 = * ;

    *

    1 = *

    2 = * and 1 = 2 =

    It should be noted that, the wall functions used in the current investigation neglected any

    thermophysical property changes for SC water within the viscous sub-layer. Also, to better capture thenear wall effects, low Reynolds damping functions were considered with the SST model.

    As described earlier, in the current investigation, the low-Reynolds number modification (low-Re

    approach) implemented in STAR was used to capture the occurrence of HTD. The term low

    Reynolds number refers to the turbulent Reynolds number which is related to turbulence kineticenergy and dissipation as:

    t

    t

    k~Re

    2

    (10)

    which is usually low near the wall (Menter, 2009) and should not be confused with the channel

    Reynolds number. In the low-Re model, some of the coefficients previously used ( *

    1 , 1 ,*

    ) for the

    SSTk- model were replaced as:

    4

    4

    *

    2

    *

    Re,1ReRe1

    ReRe154

    t

    t

    low (11)

  • 8/10/2019 heat transfer in nuclear reactor

    8/18

    8

    *

    01Re,1

    1

    ReRe1

    ReRe

    t

    tlow (12)

    kt

    ktlow

    ReRe1

    ReRe3*Re,1

    (13)

    where: 0 = 9

    1

    and =125

    9.

    The three coefficients,Re =8; Re =2.95; kRe =6; in Eq.s 11, 12 and 13 control the rate at which the

    closure coefficients approach their fully-turbulent values.

    2.4 Solution approach and models used

    In all the simulations, the entrance and exit of the flow channel were modeled with uniform velocity

    inlet and pressure outlet boundary conditions. The fuel elements, centre and pressure tubes were set as

    solid walls with no-slip conditions. Symmetry boundary conditions were applied along the sides

    (Figure 2). A uniform heat flux boundary condition was imposed at the surface of the inner- and outer-ring elements. The effect of conjugate heat transfer was not included in the current investigation.

    Following the recommendations made by Kao et al., 2010 and Liu et al., 2013, the gravity (buoyancy

    force) was included in all the simulations to capture a sharp increase in the sheath temperature whenHTD occurs under test conditions considered in the current investigation. However, in order to

    confirm the recommendations made by previous investigators (Kao et al., 2010; Liu et al., 2013), a

    sensitivity analysis to buoyancy was also performed (refer to section 3.2).

    In the current study, the equations were solved using a steady-state segregated solver with Rhie and

    Chow type pressure velocity coupling (CD-Adapco, 2013) combined with a SIMPLEtype algorithm.

    The segregated solver was chosen over the coupled solver because it uses less memory, and improves

    solution convergence for some cases. Considering the recommendation made in ASME CFD

    numerical accuracy guidelines (ASME, 2009), all the equations were solved using second order

    differential schemes. A convergence strategy was adopted for some cases where numerical simulationfailed to converge. First, a fully converged solution was obtained by solving only the flow equations

    with URF set to a value of zero. Using the converged flow solution, the energy equation was enabledand the solution was iterated till convergence was achieved again. The URF values for flow, pressure

    and energy were set to 0.7, 0.3 and 0.99 respectively. However, for some of the run conditions,

    especially at a lower pressure (23.5MPa) and when the inlet temperature is close to the pseudo-criticalpoint, it was found necessary to further lower the flow URF to 0.2 to ensure stability and avoiddivergence of the numerical solution.

    Convergence was monitored for each run and the solution was iterated till the scaled residuals dropped

    at least by three orders of magnitude and fluctuated in a steady manner. In addition, fluid temperatures

    were monitored in the cross-plane sections near the outlet (0.38 m) and in the middle (0.2 m) portion of

    the bundle geometries. The solution was judged to be converged once the residuals had dropped bythree orders of magnitude and steady-state fluid temperatures at the monitored sections were achieved.

    A wide range of test conditions as listed inTable 3 were simulated to test the possibility of occurrence

    of HTD for the current fuel bundle design. An overall of 53 test cases were simulated based on thecurrent test matrix, of which selected results are presented in this paper. In general HTD is more likely

    to occur at lower mass fluxes and prior to pseudo-critical temperature (bulk) (Cheng and Schulenberg,

    2001). Hence, the current paper focuses on the results obtained for low-mass-flow test cases in greaterdetail.

  • 8/10/2019 heat transfer in nuclear reactor

    9/18

    9

    Table 3 Test matrix used for the current investigation

    Operational pressure 23.5, 25, 28 MPa

    Mass flux 0.438, 1.095 Mg/m -s

    Inlet fluid temperature 300-400C

    Heat flux 0.2-1.2 MW/m

    Power 1.97-11.87 MW

    3. Sensitivity Analysis

    3.1 Computational mesh and wall y+treatment

    The grid sensitivity and effect of wally+is investigated as follows. Of the three system (or operational)

    pressures considered for testing (Table 3), the mesh sensitivity was performed at 23.5 MPa to check theaccuracy of the computational mesh size used in the current simulations. As seen inFigure 3,the

    thermophysical properties (especially the specific heat capacity) vary with temperature more

    significantly at lower pressures than at higher pressures. This necessitates the need to examine thesuitability of the mesh size and determine if it can be used for capturing the occurrence of HTD. In

    principle, similar exercise should be performed at 25 and 28 MPa. However, at 25 and 28 MPa, the

    current study utilizes, presumably conservatively, a computational mesh derived from the meshsensitivity analysis of lower pressure condition (23.5 MPa). For the grid-sensitivity study, the mesh

    count was increased by changing the base size of the meshes. The mesh refinement ratio between

    consecutive meshes was maintained at 1.3 as per ASME V&V 20 guidelines (ASME, 2009). Asdiscussed in section 2.1, the computational model used at least ten boundary layers on the elements

    with the first node point set to a corresponding wally+value of ~1.0. The simulations were performed

    using the low-Re SST k-turbulence model, described in section 2.3, to capture the occurrence of

    HTD. The finest mesh used in the current analysis was designated as mesh#1 followed by coarser

    meshes as shown inTable 4. All the three meshes presented a well behaved convergence. As seen inFigure 4,the sheath temperature did not significantlychange for the three meshes. Hence, the

    mesh #3 with 3.5 million cells is used as the base case for the current simulations. In this work, the

    variation of sheath temperature was plotted along the fuel channel at a radial location where themaximum sheath temperature was observed. The radial location for the maximum sheath temperature

    changes case-by-case.

    Table 4 Mesh characteristics and refinement ratios for bare bundle

    Mesh i Total mesh

    count (millions)

    Average

    base meshsize (mm)

    Mesh

    refinementratio, r

    1 11.3 0.22 -

    2 6.0 0.29 1.3

    3 3.5 0.38 1.3

  • 8/10/2019 heat transfer in nuclear reactor

    10/18

    10

    Figure 4Mesh sensitivity analysis for the current investigation at 23.5 MPa

    (G= 0.438Mg/m2-s, Q=3.95MW, q= 0.405 MW/m

    2, Tin=370C)

    As discussed in section 2.1, a computational mesh with y+30) and low y

    + (y

    +

  • 8/10/2019 heat transfer in nuclear reactor

    11/18

    11

    cannot be used directly. Hence, wall temperature was considered to be a key variable to judge the

    occurrence of HTD in the current analysis.

    For the lower operating pressure (23.5 MPa) and lower mass flux, as seen inFigure 6a, buoyancy playsa significant role in the predicted sheath-temperature distribution, resulting in a sharp rise in

    temperature in the upstream section and a steady decrease in the downstream section. In comparison,

    when the buoyancy effect was not accounted for, the sheath temperature increased gradually along the

    length, indicating HTD in the upstream section and HTE in the downstream section due to buoyancy.For a given operating pressure, the temperature difference between cases with and without buoyancy is

    larger for the case of lower mass flux than for the case of higher mass flux. The results clearly indicatethat buoyancy is an important contributor to the occurrence of HTD, especially at lower pressures and

    lower mass fluxes. Similar findings have been reported for flows in smooth pipes and annular channels

    by previous investigators (Palko and Anglart, 2008; Liu et al., 2013; Kao et al., 2013).

    For 25 MPa and 28 MPa at low mass flux condition, the sharp increase in the predicted sheath

    temperatures were found to be lower compared to that of 23.5MPa. In addition, a gradual increase intemperature was observed across the length of the rod bundle (Figure 6a throughFigure 6c) without

    inclusion of the buoyancy force. For all the test cases in Figure 6 and Figure 7, the sharp spike in the

    predicted temperature close to the entrance region of ~3 cm can be regarded as a result of theassumption of uniform flow at the entrance without including a non-heated part that would develop the

    turbulent flow upstream of the entrance. Therefore, the spike (Figures 6b, 6c) is judged to have no

    practical importance to this study. On the other hand, considering buoyancy force for 25 MPa yielded

    a taper off of the temperature increase (at ~0.2m), compared to the case without the buoyancy.Similarly, for 28 MPa, a flatter sheath temperature was predicted with buoyancy than without it.

    For high mass flux conditions at 23.5 MPa, a small step increase in temperature was predicted at

    ~10cm (Figure 7a) for which no explanation is currently available. For all three pressures, no HTD

    was predicted (seeFigure 7bandFigure 7c). However, as a result of neglecting buoyancy, thepredicted sheath temperatures were higher in the downstream half of the test section compared to the

    simulations that included the buoyancy. Based on the six conditions included in Figures 6 and 7, it can

    be inferred that buoyancy cannot be ignored, especially for the lowest pressure combined with thelower mass flux (Figure 6a).

    (a) (b) (c)

    Figure 6 Effect of buoyancy at low mass flux test conditions for the three operating pressures(G= 0.438Mg/m

    2-s, Q=3.95MW, q= 0.405 MW/m

    2)

  • 8/10/2019 heat transfer in nuclear reactor

    12/18

    12

    (a) (b) (c)

    Figure 7 Effect of buoyancy at high mass flux test conditions for the three operating pressures

    (G=1.095 Mg/m2s, Q=9.89MW, q= 1.012 MW/m

    2)

    4. CFD predictions and discussion

    For the test matrix shown inTable 3,simulation results were obtained using mesh#3 (3.5 million cells)

    with wally+

  • 8/10/2019 heat transfer in nuclear reactor

    13/18

    13

    temperatures (Zhang et al., 2012). Also, the conjugate heat transfer (CHT) that considers conduction of

    heat through the fuel rod cladding was not included in the computational model. The incorporation ofthe wall conduction model is expected to result in reduced wall temperatures without impacting the

    occurrence of heat transfer deterioration thereby leaving the qualitative trends unaffected. In an

    attempt to address the limitations of the existing low-Returbulence models in overepredicting the walltemperature (Rosa et al, 2011), Amoako et al., 2013 used high wally

    +wall treatment instead of

    detailed wall treatments along with the low-Returbulence models. However, based on the sensitivityof wall treatment approach discussed earlier in section 3.1, low wally+treatment was used in this study

    to qualitatively determine if HTD could exist for the flow conditions listed in Table3.

    (a) (b) (c)

    Figure 8 Effect of inlet temperature on axial sheath temperature variation (G= 0.438Mg/m2-s,

    Q=3.95MW, q=0.405 MW/m2: (a) 23.5 MPa; (b) 25 MPa; (c) 28 MPa)

    4.2 Effect of operational pressure and mass flux on heat transfer

    The variations with bulk fluid temperature of the maximum sheath temperature and corresponding heat

    transfer coefficients for the low (0.438Mg/m2-s) and high mass fluxes (1.095 Mg/m

    2-s) at the three

    operational pressures are presented inFigure 9 andFigure 10 respectively. In general, deviations from

    the normal heat transfer have been found to occur when the sheath temperature is greater than thepseudo-critical temperature and the bulk temperature is less than the pseudo-critical temperature(Tsh > Tpc>Tb) (Cheng and Schulenberg, 2001).

    At 23.5MPa and at bulk temperature near pseudo-critical temperature the simulations predicted HTD

    for low mass flux (Figure 9a) and HTE for higher mass flux (Figure 10a). For this study, occurrence of

    HTD at 23.5 MPa for low mass flux is in line with the findings by previous investigators (Cheng andSchulenberg, 2001; Gang et al., 2011).

    However, it was observed that the simulations behaved differently at 25 and 28 MPa when compared to

    that of 23.5 MPa. This behaviour was also observed by Kim and co-workers (Kim et al., 2007) thereby

    suggesting the occurrence of HTD and/or HTE depend also on the operational pressure. For the low

    mass flux at 25 and 28MPa, the wall temperature exhibited a dip when approaching the pseudo-criticaltemperature region resulting in an increase in the heat transfer coefficient as predicted inFigure 9b and

    Figure 9c. This change in trend is attributed to the HTE for the simulated test case. In addition, for the

    pressure of 28 MPa at the lower mass flux, HTD was also observed at lower bulk temperatures. On theother hand, for higher mass flux conditions at 25 and 28 MPa, normal heat transfer behaviour was

    observed (Figure 10b andFigure 10c).

  • 8/10/2019 heat transfer in nuclear reactor

    14/18

    14

    (a) (b) (c)

    Figure 9 Variation of maximum sheath temperature and heat transfer coefficient with the bulktemperature (G= 0.438Mg/m

    2-s, Q=3.95MW, q= 0.405 MW/m

    2)

    (a) (b) (c)

    Figure 10 Variation of maximum sheath temperature and heat transfer coefficient with the bulk

    temperature (G= 1.095 Mg/m2s, Q=9.89MW, q= 1.012 MW/m

    2)

    5.

    Limitations and ChallengesThe first step on the application of CFD approach to identify the conditions favorable for theoccurrence of HTD in the newly developed fuel bundle design by AECL is presented in this work. The

    simulations performed were for a short domain length compared to the full length of the fuel element

    (~6m) primarily to reduce the computational run time and fit within the computational resources

    available for this project. The results obtained from this study present trends which provideinformation on the possibility of occurrence of HTD at the test conditions considered.

    One of the uncertainties associated with this work is the use of a single turbulence model for all the

    inlet temperature conditions (i.e. inlet temperatures close to and away from pseudo-critical conditions).

    The CFD predictions, especially of the sheath temperature, may change significantly with the use of adifferent turbulence model and wall treatment approach. However, given the current limited

    understanding of the application of the turbulence models for SC bundle flows, it was decided to use a

    turbulence model that was generally agreed upon by previous investigators. It should be noted that the

    sheath temperature variations predicted in this analysis are only qualitative in nature and may varysignificantly with turbulence models as discussed above. However, the trends obtained in this analysis

    are indicative of the expected heat transfer behaviour for the test conditions considered in the current

    investigation.

  • 8/10/2019 heat transfer in nuclear reactor

    15/18

    15

    6. Summary

    The possibility of occurrence of heat transfer deterioration in the proposed SCWR fuel bundles was

    numerically examined by solving conservation laws of mass, momentum, and energy with the low-Re

    SST k-turbulence model and wally+

  • 8/10/2019 heat transfer in nuclear reactor

    16/18

    16

    THINS Thermalhydraulics of Innovative Nuclear Systems

    URF Under Relaxation Factor

    General symbols

    CDk Production limiter in Eq. 7cp Specific heat (kj/kg-C)

    F1, F2 blending function in Menters SST k-turbulence model (-) in Eq. 3 and 8G Mass flux (Mg/m2-s)

    k Thermal conductivity (W/m-C)k Turbulence kinetic energy (m

    -2s

    -2)

    P Operating pressure (MPa)

    Q Power (MW)q Heat flux (MW/m

    2)

    Re Reynolds Number (-)

    t Time (s)

    T Temperature (C)u velocity field (m/s)

    y Distance to the nearest wall in Eq. 5

    y+ non-dimensional distance from wall (-)

    Greek letters

    Constant in the blending function of the SST model

    Constants 1 and 2 in the blending function of the SST model

    Constant in the SST k-model

    * Constant in the SST k-model

    Constant in the SST k-model

    Dynamic viscosity (Pa-s)

    Kinematic viscosity (m2/s)

    Density (kg/m3) Schmidt number (-)

    Specific dissipation rate (s-1

    )

    Subscripts/ Over- bars

    b Bulk

    in inleti, j, k velocity components

    k Turbulence kinetic energy

    pc Pseudo-critical

    sh Sheath

    t Turbulent Specific dissipation rate

    - over-bar used for average

    8. References

    Ambrosini, W., 2009. Continuing assessment of system and CFD codes for heat transfer and stability

    in supercritical fluids. In Proceedings of 4th

    ISSCWR, Heidelberg, Germany, March 8-11.

  • 8/10/2019 heat transfer in nuclear reactor

    17/18

    17

    Angelucci, M., Ambrosini, W., Forgione, N., 2013. Numerical estimation of wall friction ratio near the

    pseudo-critical point with CFD models. Nuclear Engineering and Design, 264, 71-79.Amoako, E.A., Ambrosini, W., 2013. Developing a CFD methodology for the analysis of flow stability

    in heated channels with fluids at supercritical pressures. Annals of Nuclear Energy, 54, 251-262.

    ASME V&V 20 Committee, 2009. Standard for verification and validation in computational fluid

    dynamics and heat transfer, ASME, New York, USA.

    CASL, last accessed on September 2013. Website:http://www.casl.gov/index.shtmlCD-Adapco, 2013. User Guide, STAR-CCM+ Version 7.04., New York, USA.

    Cheng, X., Schulenburg, T., 2001. Heat transfer at supercritical pressure: literature review and

    application to a HPLWR. Wissenschaftliche Berichte FZKA 6609, Forschungszentrum, Karlsruhe,May 2001.

    De Rosa, M., Guetta, G., Ambrosini, W., Forgione, N., He, S., Jackson, J.D., 2011. Lessons learned

    from the application of CFD models in the prediction of heat transfer to fluids at supercritical pressure.

    In Proceedings of 5th

    ISSCWR, Vancouver, Canada, March 13-16.

    Elligot, D.M, McCreery, G.E., Schultz, R.R., Lee, J., Hejzlar, P., Stahle, P., Saha, P., 2006.Investigation of fundamental thermal-hydraulic phenomena in advanced gas-cooled reactors, Report

    No: INL/EXT-06-11801, Idaho National Laboratory, Idaho, USA.Gang, W., Bi, Q., Yang, Z., Wang, H., Zhu, X., Hao, H., Leung L.K.H., 2011. Experimentalinvestigation of heat transfer for supercritical pressure water flowing in vertical annular channels.

    Nuclear Engineering and Design, 241, 4045-4054.

    He, S., Kim, W.S., Bae, J.H., 2008. Assessment of performance of turbulence models in predicting

    supercritical pressure heat transfer in a vertical tube. International Journal of Heat and Mass Transfer,51, 46594675.

    Jackson, J.D., Anderson, M., Cheng, X., Zahlan, H., Bae, Y.Y., Yamada, K., 2012. Progress on the

    development of new correlations under the framework of the IAEA Coordinated Research Programme

    on heat transfer in SCWRs. In Proceedings of ICAPP 2012, Chicago, USA, June 24-28.

    Kao, M.T., Lee, M., Ferng, Y.M., Chieng, C.C., 2010. Heat transfer deterioration in a supercriticalwater channel. Nuclear Engineering and Design, 240, 3321-3328.

    Kim,H.Y.,Kim,H.,Song,J.H.,Cho,B.H., Bae,Y.Y., 2007. Heat transfer test in a vertical tube using

    CO2at supercritical pressures. Journal of Nuclear Science and Technology, 44, 285293.

    Lemmon, E.W., McLinden, M.O., Friend, D.G., NIST Chemistry WebBook, 2013. Thermophysicalproperties of fluid systems, retrieved September (2013). In: Linstrom P.J. and Mallard W.G. (Eds.),

    National Institute of Standards and Technology, Gaithensburg, Maryland, USA.

    Liu, L., Xiao, Z., Yan, X., Zeng, X., Huang, Y., 2013. Heat transfer deterioration to supercritical waterin circular tube and annular channel. Nuclear Engineering and Design, 255, 97-104.

    Menter, F.R., 1994. Two-equation eddy-viscosity turbulence modeling for engineering applications.

    AIAA Journal, 32, 1598-1605.

    Menter, F.R., 1999. Review of the shear-stress transport turbulence model experience from an

    industrial perspective. International Journal of Computational Fluid Dynamics, 23, 305-316.

    Palko, D., Anglart, H., 2008. Theoretical and numerical study of heat transfer deterioration in high

    performance light water reactor. Science and Technology of Nuclear Installations. 1-5.

    Pioro, I.L., Duffey, R.B., 2007. Heat transfer and hydraulic resistance at supercritical pressures in

    power engineering applications. ASME Press, New York.

    http://www.casl.gov/index.shtmlhttp://www.tandfonline.com/action/doSearch?action=runSearch&type=advanced&searchType=journal&result=true&prevSearch=%2Bauthorsfield%3A(KIM%2C+H+Y)http://www.tandfonline.com/action/doSearch?action=runSearch&type=advanced&searchType=journal&result=true&prevSearch=%2Bauthorsfield%3A(KIM%2C+H)http://www.tandfonline.com/action/doSearch?action=runSearch&type=advanced&searchType=journal&result=true&prevSearch=%2Bauthorsfield%3A(SONG%2C+J+H)http://www.tandfonline.com/action/doSearch?action=runSearch&type=advanced&searchType=journal&result=true&prevSearch=%2Bauthorsfield%3A(CHO%2C+B+H)http://www.tandfonline.com/action/doSearch?action=runSearch&type=advanced&searchType=journal&result=true&prevSearch=%2Bauthorsfield%3A(BAE%2C+Y+Y)http://www.tandfonline.com/action/doSearch?action=runSearch&type=advanced&searchType=journal&result=true&prevSearch=%2Bauthorsfield%3A(BAE%2C+Y+Y)http://www.tandfonline.com/action/doSearch?action=runSearch&type=advanced&searchType=journal&result=true&prevSearch=%2Bauthorsfield%3A(CHO%2C+B+H)http://www.tandfonline.com/action/doSearch?action=runSearch&type=advanced&searchType=journal&result=true&prevSearch=%2Bauthorsfield%3A(SONG%2C+J+H)http://www.tandfonline.com/action/doSearch?action=runSearch&type=advanced&searchType=journal&result=true&prevSearch=%2Bauthorsfield%3A(KIM%2C+H)http://www.tandfonline.com/action/doSearch?action=runSearch&type=advanced&searchType=journal&result=true&prevSearch=%2Bauthorsfield%3A(KIM%2C+H+Y)http://www.casl.gov/index.shtml
  • 8/10/2019 heat transfer in nuclear reactor

    18/18

    18

    Rao, Y.F., Cheng, Z., Waddington, G.M., Nava-Dominguez, A., 2013. ASSERT-PV 3.2: improved

    subchannel thermalhydraulics code for CANDU bundles. In Proceedings of 21stICONE, Chengdu,

    China, July 29-August 2.

    Saha, P., Aksan, N., Andersen, J., Yan, J., Simoneau, J.P., Leung, L., Bertrand, F., Aoto, K.,

    Kamide, H., 2013. Issues and future direction of thermalhydraulics research and development in

    nuclear power reactors, Nuclear Engineering and Design, 264, 3-23.

    Schulenberg, T, Visser, D.C., 2013. Thermal-hydraulics and safety concepts of supercritical watercooled reactors. Nuclear Engineering and Design, 264, 231-237.

    THINS, last accessed on September 2013. Website:http://www.ifrt.kit.edu/thins/

    Wilcox, D.C., 2010. Turbulence Modelling for CFD, DCW Industries, La Caada, California, USA.

    Zhu, Y., 2010. Numerical investigation of flow and heat transfer within the core cooling channel of a

    super critical water reactor. Ph.D. Thesis, University of Stuttgart, Stuttgart, Germany.

    Zhang, G., Zhang, H., Gu, H., Yang, Y., Cheng X., 2012. Experimental and numerical investigation of

    turbulent convective heat transfer deterioration of supercritical water in vertical tube. Nuclear

    Engineering and Design, 248, 226-237.

    http://www.ifrt.kit.edu/thins/http://www.ifrt.kit.edu/thins/